Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

The Journal of Adhesion

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

Effect of reinforcements at different scales on


mechanical properties of epoxy adhesives and
adhesive joints: a review

Ali Nemati Giv, Majid R. Ayatollahi, S. Hengameh Ghaffari & Lucas F.M. da
Silva

To cite this article: Ali Nemati Giv, Majid R. Ayatollahi, S. Hengameh Ghaffari & Lucas
F.M. da Silva (2018): Effect of reinforcements at different scales on mechanical properties
of epoxy adhesives and adhesive joints: a review, The Journal of Adhesion, DOI:
10.1080/00218464.2018.1452736

To link to this article: https://doi.org/10.1080/00218464.2018.1452736

Published online: 17 Aug 2018.

Submit your article to this journal

View Crossmark data

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


http://www.tandfonline.com/action/journalInformation?journalCode=gadh20
THE JOURNAL OF ADHESION
https://doi.org/10.1080/00218464.2018.1452736

Effect of reinforcements at different scales on mechanical


properties of epoxy adhesives and adhesive joints: a
review
Ali Nemati Giva, Majid R. Ayatollahia, S. Hengameh Ghaffarib, and Lucas F.M. da Silvac
a
Fatigue and Fracture Lab., Centre of Excellence in Experimental Solid Mechanic and Dynamics, School
of Mechanical Engineering, Iran University of Science and Technology, Tehran, Iran; bSchool of
Mechanical Engineering, Iran University of Science and Technology, Tehran, Iran; cDepartment of
Mechanical Engineering, Faculty of Engineering, University of Porto, Porto, Portugal

ABSTRACT ARTICLE HISTORY


Improvement of the mechanical properties of adhesives and Received 17 December 2017
adhesive joints has been a subject of great interest in recent Accepted 12 March 2018
years. Up to now, several methods have been presented such KEYWORDS
as modifying substrate shapes, adding microparticles (MPs) Adhesive and adhesive
and nanoparticles (NPs), and embedding micro and macrofi- joints; mechanical
bers in the adhesive layer. This review aims to investigate how properties; polymeric
these reinforcements of different scales in the adhesive layer materials; reinforcements
influence the mechanical properties of adhesive joints and
adhesives. Characteristics and applications of reinforcements
are introduced in the first part. In the second part, the effects
of several parameters commonly investigated by researchers
on the strength, stiffness and fracture toughness improvement
of polymeric materials are reviewed for reinforcements of dif-
ferent scales. Finally, damage mechanisms involved in increas-
ing or decreasing the mechanical properties are reviewed and
discussed.

Introduction
Adhesive bonding is one of the material joining methods in which an adhesive
layer is located between substrates. The main role of the adhesive layer in an
adhesive joint is transferring load from one substrate to another one. Adhesively
bonded joints (ABJs) provide much superiority over conventional mechanical
fasteners including design flexibility, ability to join dissimilar materials without
structural changes of the joined materials, electrical and thermal insulation,
inherent sealing characteristics, vibration and sound dampening.[1-3] Moreover,
the advantages of adhesively bonded joints are not limited to the cases mentioned
above. In mechanical properties improvement of ABJs, they are able to uniformly
distribute the internal stresses over the entire bonded region.[4] Fatigue resistance
and high strength to weight ratio can easily be achieved compared to other
conventional joints[5].

CONTACT Majid R. Ayatollahi m.ayat@iust.ac.ir


Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/gadh.
© 2018 Taylor & Francis
2 A. NEMATI GIV ET AL.

Due to these advantages, this type of joint is present in various areas


especially high technology industries. High performance requirements and
the necessity of reducing weight in most engineering applications cause
considerable efforts to be performed for improving the properties especially
the mechanical properties of adhesive joints. There are several methods for
improving the strength of this type of joint which are described below.
The surface roughness of the substrate directly affects the chemical bonds
between the substrate and the adhesive layer at the interface. By increasing the
roughness, the effective contact area between the adhesive and the substrate
increases. Therefore, the number of chemical bonds will be greater and con-
sequently the adhesive joint will become stronger and more durable.[6,7]
One of the main factors which have a significant influence on the strength of
adhesive joints is the adhesive fillet which is commonly formed during the curing
process and assembly of adhesive joints.[8] Tsai and Morton[2] studied the influ-
ence of adhesive fillets on the adhesive peel and shear stress distributions. They
showed that by incorporating an adhesive fillet at the ends of the overlap joint, the
maximum adhesive peel and shear stress values are reduced. Similar researches on
the effect of adhesive fillets on improving adhesive stress distributions and
strength of adhesive joints were carried out by other researchers.[9–11]
The thickness of the adhesive layer notably influences the mechanical
properties of adhesive joints. Gleich et al[12] and da Silva et al[13] observed
that the adhesive interface stresses increase by increasing the thickness of the
adhesive layer and consequently a reduction of adhesive joint strength is
expected. Also, a too thin adhesive layer decreases the adhesive joint strength
because the adhesive layer cannot cover the surface ripples of the substrates.
Therefore, an optimum value is expected for the adhesive thickness.[14]
Another effective method for promoting the strength of adhesive joints is the
incorporation of metallic components into the adhesive fillet.[8] You et al[8]
observed that the shear strength of single lap joints increased by inserting
metallic components into the adhesive fillet due to the reduction of stress
concentration at the ends of the overlap joint.
The modulus of the adhesive layer has an important effect on the strength
of adhesive joints. In an ideal adhesive joint, the flexibility and strength of the
adhesive should vary along the adhesive bondline length. High shear strains
are located at the ends of the overlap joint, therefore, a flexible and ductile
adhesive should be used at this zone while a stiffer and stronger adhesive
could be located in the middle part of the joint.[15–17]
The combination of adhesive with conventional joints leads hybrid joint
systems to perform better under different loading conditions.[18] Sadowski et al[19]
used rivets as reinforcements in adhesive joints in their investigation. They found
that utilizing rivets in the adhesive joint increased the energy absorption of the
hybrid joint (i.e. adhesive and rivet joint) by about 35% compared to the adhesive
joint.
THE JOURNAL OF ADHESION 3

A summary of useful techniques for increasing the strength and perfor-


mance of adhesively bonded joints is given above. These techniques chiefly
depend on changing the geometry of adhesively bonded joint or substrates.
However, these changes are not applicable to all adhesives and may not be
practical to implement. Some adhesives have a weak crosslink structure or
others are inherently brittle with a poor resistance to crack initiation and
propagation.[20,21] In order to eliminate these adhesive drawbacks and
enhance the mechanical properties of adhesively bonded joints, adding rein-
forcements of different scales into the adhesive layer may be simpler and
cheaper. A review concerning these reinforcements is presented in this paper.

Reinforcements at different scales


A common method for improving the aforementioned shortcomings of
adhesives is adding reinforcements to the adhesive layer which results in a
composite material. A composite material is a combination of two or more
materials which have better properties than its individual components.[22]
The reinforcing phase is a fiber or a particle that enhances mechanical,
electrical and thermal properties of the matrix. Reinforcements are usually
harder, stronger, and stiffer than the matrix[23] and according to their sizes
they can be divided into three subdivisions:

(a) Nano scale reinforcements


(b) Micro scale reinforcements
(c) Macro scale reinforcements

Nano scale reinforcements


Nano materials have at least one dimension smaller than 100 nanometers.[24] In
other words, their specific surface area should be more than 60 m2 =cm3 .[25] The
mean state of nano fillers is between bulk and molecular materials.[26] Nano-
sized materials have useful properties such as: large surface area, high surface
energy and reduced number of structural defects.[27] Furthermore, they have
impressive differences in their electronic, kinetic, magnetic and optical proper-
ties compared to their bulk counterparts due to their high specific surface area/
volume ratio.[26,28] Unlike bulk materials, NPs are size dependent. The high ratio
of the atoms on the surface of the particles to their total atoms causes a size-
dependency behavior of the NPs.[28] Due to the unique features of the nano
materials, many scientists made an effort to obtain improved adhesive properties
by incorporating different types of nano fillers. Researchers observed that the
addition of nano particles improved the mechanical and thermal behaviors of
adhesives and polymeric materials.[29–46] Additionally, they are useful in the
4 A. NEMATI GIV ET AL.

manufacture of packaging, cosmetic, biomedical and electronic materials.[26]


There are various types of NPs including nano carbon, nano clays, metal oxides
and nanosilica.[47] As mentioned before, these nano materials can differently
affect the adhesive properties based on their types, therefore, it is important to
understand the properties of each NP. In the following subsections, a brief
description of each group of nanoparticles currently used in adhesives and
adhesive joints are presented.

Carbon based nano materials


There are extensive investigations on carbon nano materials because of their
excellent features such as having thermal and electrical conductivity, high
mechanical strength and optical properties.[48,49] Carbon nano-fillers (CNFs) are
classified into three subdivisions based on their dimensions which include zero
dimensional or spherical particles (e.g. nano diamonds), one dimensional (e.g.
carbon nanotubes) and two dimensional fillers (e.g. graphene nanoplates).[28]
Carbon nanotubes (CNT) are one dimensional fillers that are formed by rolling
up a graphene sheet into a cylinder. The diameter of CNTs is extremely small, in
the order of a few nanometers whereas their length can be less than 100 nm to
several centimeters.[50] On the basis of the number of tubes in the CNTs, they can
be classified into three categories namely single-walled CNTs, double-walled
CNTs and multi-walled CNTs (see. Figure 1). CNTs have extremely large surface
areas, high aspect ratio and great mechanical strength and stiffness.[51] In addition,
they possess good electrical and thermal properties. Due to these unique features,
they are attractive selections for fillers in different polymers and ceramics.[52]
Graphene is a single-atom-thick sheet of sp2-hybridized carbon atoms[53] which
is an example of two dimensional carbon nano filler that exhibits unique proper-
ties, such as large specific surface area ð2620 m2 =g Þ, high mechanical properties
(Young’s modulus of 1 TPa and ultimate strength of 130 GPa) and high electrical
conductivity (above 300 Wmk−1).[51] Furthermore, graphene is structurally
robust and highly flexible, so it is appropriate for engineering thin and flexible
materials.[48] Nano diamond (ND) particles are zero dimensional CNFs.[28] Their

(a) (b) (c)


Figure 1. Schematic of different types of CNTs a) SWCNTs b) DWCNTs c) MWCNTs[54].
THE JOURNAL OF ADHESION 5

size is less than 10 nm[48] and they have excellent mechanical and optical proper-
ties, high surface areas and tunable surface structures. Notwithstanding the afore-
mentioned advantages, there are two main shortcomings for CNFs; the first one is
the inconsistency in their quality and the second one is their high cost.[28]

Metal based NPs


Metallic NPs have great potential applications in nanotechnology.[28] Metallic
nano sized particles have some physical properties which are different from
both the ion and the bulk material.[55] Most of the metallic elements of the
periodic table are accessible in the form of nano materials such as nano gold,
nano silver and metal oxides[28] (e.g., zinc oxide, titanium dioxide, and iron
oxide). It has been shown that incorporating metallic NPs improves the
mechanical properties, thermal and electrical conductivity of polymers.[56,57]
Some metallic NPs have unique properties and medical applications. For
instance, silver NPs have antibacterial properties versus various kinds of
bacteria and fungi.[58] Furthermore, metallic NPs have a wide range of applica-
tions such as optics, electronics, catalysis and they have also been used as
polymer reinforcements.[28]

Nano silica
Silicon dioxide or (SiO2) is found in crystalline and amorphous forms and
the color of its powder is white.[25] Silica nanoparticles (SNPs) are porous
and have a large surface area that includes a large number of hydroxyl groups
and unsaturated residual bonds. It has been shown that silica nanoparticles
can improve the strength, durability and also the flexibility of polymers.[28]

Nano clays
Nano clays are natural platelet structured nano materials. They have two
main species including montmorillonite and allophane. The structure of
montmorillonite is a crystalline hydrous phyllosilicate (layer silicate). The
basic structural unit of phyllosilicates (clay minerals) contains two layers of
tetrahedral silica and octahedral alumina. Allophane is a non-crystalline
aluminosilicate and is generated by the weathering of volcanic ash. It is
appropriate for enzyme immobilization.[26] Nano clays are inexpensive
materials[51] which can improve fire, liquid infusion and mechanical proper-
ties of polymers.[28] In addition, they can enhance the mechanical and
thermal properties of cement matrices.[59] Organoclays (hybrid) are organi-
cally-modified montmorillonites that have been utilized as rheological modi-
fiers and additives in paints, inks, greases, cosmetic and as carriers and
delivery systems for the controlled release of drugs.[26] Moreover, they have
shown excellent mechanical, thermal and gas-barrier properties.
Furthermore, they have great role in pollution control and water
treatment.[60]
6 A. NEMATI GIV ET AL.

Micro scale reinforcements


The size of micro structured materials is between 1 to1000 μm and the
surface area to volume ratio in this type of reinforcement is smaller than
nano reinforcements. Therefore, different properties of this type of reinforce-
ments are expected. In the following subsections, some types of micro
reinforcements that have been studied by many researchers for improving
the mechanical properties of adhesives and adhesive joints[21,61–66] are intro-
duced briefly.

Thermally expandable particles (TEPs)


Thermally expansive particles or thermo-expandable microspheres are core/
shell particles which have thermoplastic shells filled by liquid hydrocarbon
(see Figure 2).[67] Their diameter is between 5–50 μm.[68] By heating TEPs,
the shell material becomes softer and the liquid inside it starts to gasify.
Then, the shell expands as the gas inside it pushes the softened shell from
inside out causing it to grow in size. After full expansion, the volume of TEPs
increases from 50 to 100 times.[69] Due to the plastic deformation of the shell
material, the volume of the particles remains constant after cooling.
Therefore, the expansion process of TEPs is irreversible.[68] The expansion
temperature of TEPs is between 70°C to 285°C which is dependent on their
content[69] and it must be higher than the Tg of the resin.[70] It is worth
mentioning that larger particles expand better than smaller ones but when
the surface finish of the final product is important (as, e.g., in printing inks,
putties, and artificial leather), it is necessary to use smaller particles.[68] The
main purpose of using TEPs is for facilitating the dismantling of ABJs. At low
contents of TEPs, dismantling does not occur and at high weight fraction the
strength of the adhesive bond reduces. Thus, the appropriate amount for
weight fraction of TEPs is between 5 to 15%.[70] The expansion capability of
TEPs makes them attractive for various applications, such as weight reduc-
tion, altering product properties (e.g. thermal-, sound-, and electrical insula-
tion properties) and material savings. Moreover they are appropriate for

Figure 2. Schematic of thermally expandable particles[64].


THE JOURNAL OF ADHESION 7

coatings to enable 3D patterns on wall papers, or enhancing surface proper-


ties such as matting or anti-slip.[67]

Cork particles
Ever green oak produces cork that flourishes in particular areas of the Western
Mediterranean countries (Portugal, Spain, Southern France, part of Italy and
North Africa). Cork is a biological material with an alveolar structure which is
similar to a honeycomb with prismatic cells without intercellular distance (see
Figure 3). Cork wall cells are formed by five layers including, cellulose, lignin,
suberin, tannins and waxes. The size of cells depends on the season that they
have been produced and its range is between 10–40 μm. Spring cells are bigger
and their walls are thinner than autumn cells.[21] Cork is very light, very
versatile, elastic[71], innocuous, unaffected by microbial activity[72], flexible,
considerably impermeable to liquids and gases and also is a good electric,
thermal, acoustic and vibration insulator.[71] Owing to the unique properties of
cork, it had vast applications, such as in aeronautical and automobile indus-
tries, oil spills and also it can be used as reinforcement for improving fracture
toughness of brittle resins.[72]

Metallic micro materials


The dimension of metal powders is in the range of 5 to 200 micrometers. The
powder production method can affect the shape, size, surface area, apparent

Figure 3. Corks particles a) without surface treatment b) with surface treatment.[73]


8 A. NEMATI GIV ET AL.

density, flow, angle of repose, compressibility, and green strength of the


powders. Most metals are attainable in the form of a powder. [74] Adding
metal powders into polymers,

(a) Improves thermal conductivity and diffusivity


(b) Stiffens the matrix and makes it more rigid
(c) Reduces internal stresses in injection molded parts
(d) Reduces creep
(e) Lowers the cost of the material
(f) Improves the appearance of the final product

As mentioned before, by incorporating metallic powders into resins, the


thermal and electrical conductivity of polymers can be improved. It is clear
that adhesive thermal and electrical conductivities are dependent on size, shape,
volume fraction and type of fillers.[61] Conductive polymeric composites have
applications in electromagnetic interference shields, discharging static electri-
city, heat conduction, corrosion-resistant paints and antifouling compounds.

Rubber powders
Micro-sized rubber powders (see Figure 4) are typically made of vulcanized
elastomeric materials. These elastomeric materials are mostly produced from
waste tires. However, other industrial rubbers such as butyl, natural and nitrile
rubbers can also be useful in this way. The rubber particle size depends on the
manufacturing process and varies typically between 10 to180 μm. These elas-
tomeric reinforcements are widely used in various applications, especially in
brittle adhesives. The inclusion of rubber particles into brittle adhesives such as
epoxy adhesives can significantly increase their toughness with little effects on
their other desirable engineering properties. Such rubber-toughened adhesives
can be applied in situations where high impact energy is required.

Figure 4. Rubber powders.[75]


THE JOURNAL OF ADHESION 9

Glass fibers
Glass fibers are used in two relevant forms including discontinuous fibers
(short fibers with an aspect ratio ranging from 20 to 60) and continuous
fibers (long fibers with aspect ratios between 200 to 500).[76] High-speed
precision winders (Draw Rolls) produce continues filament. Depending on
the speed of the drawing rolls their diameter can vary from 10–40 μm. Unlike
carbon fibers, they are isotropic, thus, their properties do not change with the
direction of the applied load. Glass fibers have relatively low density, low
cost, high strength, high impact-resistant, low thermal conductivity and great
flexibility. Furthermore, they are good electrical insulator even for low
thicknesses.[77–80] Due to their unique properties they are used as reinforce-
ments and have wide applications in several industries such as aerospace,
defense, automotive and marines.

Carbon fibers
Carbon fibers are formed by graphitic and non-crystalline regions and their
diameter can range from 5 to10 μm. They have excellent mechanical, thermal
and electrical properties including, high stiffness, high tensile strength, low
weight, high chemical resistance, electrical and thermal conductivities, tem-
perature tolerance and low thermal expansion. Moreover, their tensile strength
is not affected by moisture and high temperature. They are very popular and
useful in various industries including aerospace, civil engineering and military.
However, among all reinforcing fibers, they offer the highest price.[79,81–83]

Macro scale reinforcements


In the field of adhesives and adhesive joints, macrofibers are usually metallic such
as steel, aluminum and etc. Their diameter ranges from 0.2 to 1.2 mm.[8,84–88]
Embedding metallic macro fibers is an efficient way for improving the mechanical
behavior of adhesive joints.[84–88] Metal macro fibers have some advantages over
other reinforcements:

(a) Better performance in the adhesive curing process


(b) Improved electrical conductivity of the adhesive layer
(c) Uniform adhesive thickness[87]

Metal macro fibers can enhance the mechanical properties of adhesive


joints[85] and in some researches shape memory alloys have been used. Shape
memory alloys with excellent thermo-mechanical behavior are usually a
combination of two or more metals.[89] The first shape memory alloy was
observed in the combining form of Nickel-Titanium (NiTi) alloy in 1962.
They are used in various applications that need very large recoverable
inelastic strains, shape memory effect and pseudo elastic response.[89]
10 A. NEMATI GIV ET AL.

Mechanical behaviour of reinforced adhesives


Any material, depending on its application, is exposed to different loading
conditions being in many cases more loaded in a particular direction.
Therefore, the improvement in the material structure should be performed
against the dominant load. In order to improve the structure of adhesive
materials and adhesive joints, incorporating reinforcements of different
scales in the adhesive layer has been proposed by many researchers.
Reinforcements in adhesives have some main features including reinforce-
ment size, content and interfacial adhesion with matrix which greatly influ-
ence the mechanical properties of the final product. The effects of these
parameters on the fracture toughness, tensile strength and stiffness of poly-
meric materials are reviewed in the following subsections.

Effects of reinforcements of different scales on adhesive and adhesive joint


strength
In rigid micro and nanoparticle reinforcements, the strength of polymeric
materials is dependent on three main parameters including particle/matrix
interfacial adhesion, particle size and content. The effect of the particle size
on the tensile strength improvement of polymeric materials has been inves-
tigated by many researchers.[90–92] Figure 5 shows the change in the tensile
strength of epoxy/silica particles composites as a function of particle loading
and size with 55wt% and 64wt% particle contents.[92] The tensile strength of
the composite increases by a reduction of particle size. A similar observation
on the tensile strength improvement of epoxy/silica microparticles was
obtained. The enhancement of tensile strength of polymeric materials rein-
forced with micro and nanoparticles have been also corroborated by other
researches.[93–95] Therefore, it can be concluded that the tensile strength of
rigid particulate filled composites increases with decreasing particle size
because smaller particles have higher total surface area promoting a more
effective stress transfer between the polymer and the particle.
Particle/matrix interfacial adhesion is another important parameter which
has a prominent role in the tensile strength improvement of particle-rein-
forced polymers. Spanoudakis and Young[96] observed that the tensile strength
of epoxy/glass beads composites decreases continuously with an increase of
untreated particle content. However, the addition of treated particles caused
the tensile strength of reinforced epoxy to increase (see Figure 6). Similar
results about the effect of particle/matrix interfacial adhesion on the tensile
strength of composites were obtained from other researchers.[97–99] From the
above discussion, it seems that a strong interfacial adhesion enhances the
tensile strength of composites because for well-bonded particles, the applied
stress can be effectively transferred to the particles from the matrix.
THE JOURNAL OF ADHESION 11

Figure 5. Tensile strength of epoxy/spherical silica composites with different particle sizes and
contents.[92]

Figure 6. Tensile strength of epoxy composite filled with modified and unmodified glass beads.[96]

The effect of particle content on the tensile strength of composites is


unclear.[100–104] This is due to the interplay between these three parameters
that cannot always be separated.[90]
Brittle particles can affect the strength of composites in two ways a) by
weakening because they produce stress concentration b) by reinforcing
because they may serve as barriers to crack growth. Both mentioned effects
can be dominant and play a significant role in the strength improvement or
degradation of polymeric materials. Predicting the strength of reinforced
polymers is difficult because it depends on different parameters such as
interface adhesion, stress concentration and defect size/spatial distributions.
However, many phenomenological and semi-empirical models have been
12 A. NEMATI GIV ET AL.

produced to predict strength against any specific parameter. Some of the


valuable studies conducted in this way are presented below.
Neilson[105] proposed an equation (Eq. 1) for strength prediction of
particle-reinforced polymers with poor interfacial adhesion:
 
σc ¼σm 1νp 2=3 Q (1)

According to Eq. 1, the strength of composite materials ðσc Þ is dependent on


the strength of the unmodified polymerðσm Þ, particle volume fraction νp
and a factor Q. The Q factor is related to the discontinuities in the load
transfer and generation of stress concentration at particles-matrix interface.
The maximum value of factor Q is equal to unity when there is no stress
concentration at the particles/matrix interface. As it is evident in Eq. 1, the
strength of polymeric materials reinforced with weak-bonded particles
decreases with an increase of particles volume fraction. For strong particle–
matrix interfacial bonding, Pukanszky et al.[106,107] introduced a semi-empiri-
cal equation (see Eq.2) for predicting the strength of nanocomposites as a
function of nanoparticles volume fraction and interfacial bonding between
the nanoparticles and host polymer:
 
1  νp
σc ¼ σm eβνp (2)
1 þ 2:5νp

According to Eq. 2, β factor is a comparative indication of the level of


bonding between the particle and polymer. σc and σm are the strengths of
modified and unmodified polymer, respectively and νp is the volume fraction
of the nanoparticles. Figure 7 illustrates the normalized tensile strength of
particulate composites versus the volume fraction of particles for various
values of β. As can be seen in Figure 7, the strength of polymeric materials
reinforced with poor-bonded particles ð0 < β < 3Þ diminishes by increasing
the particle content because the particles are not able to tolerate internal
stresses. Therefore, the particles act as voids in the structure of composite.
When the particle-polymer interaction becomes stronger (i.e. β > 3), the
strength of the composite increases by increasing the particle content.
Young and Beaumont[108] also proposed an equation (Eq. 3) for predicting
the strength of composites as a function of average inter-particle dis-
tance ðDs Þ:
σc ¼ σm þ S=Ds (3)

According to Eq. 3, S is a constant parameter which depends on the particle/


matrix adhesion, and Ds can be expressed as follows:

Ds ¼ 2dp 1  νp =3νp (4)
THE JOURNAL OF ADHESION 13

Figure 7. Normalized tensile strength of polymeric composites versus the volume fraction of
particles with different particle-polymer interaction.[106]

where dp is the mean particle size. Eqs. 3 and 4 illustrate that the reduction of
particle size and increase of particle content increase the strength of
composites.
As previously discussed, the strength estimation of reinforced polymeric
materials is not simple due to its dependency on several factors; however, it is
a useful tool for comprehending the effect of each parameter on the strength.
The effect of microfiber reinforcements on the strength of polymeric materials
has been chiefly investigated by two main parameters including the fiber volume
fraction and the fiber/matrix interfacial adhesion. Several studies have been con-
ducted to clarify the relationship between the fiber/matrix interfacial adhesion and
the tensile strength of composites. Zhao et al[109] studied the effect of interfacial
adhesion on the tensile strength of unidirectional epoxy/glass composites. They
utilized five different kinds of surface treatments: (a) unsized and untreated; (b)
γ  glycidoxypropyltrimethoxysilane (γ  GPS); (c) g-methacryloxypropyltri-
methoxysilane (γ  MPS); (d) mixture of g-aminoxypropyltrimethoxysilane
(γ  APS), film former (urethane) and lubricant (paraffin); and (e) urethane-
sized. The results indicate that the ultimate tensile strength of the composite
with the strongest fiber/epoxy interfacial adhesion (γ  GPS) is not the highest
(Figures 8 and 9). This contradiction is attributed to the interfacial failure mode.
The micro damage of γ  MPS and γ  GPS was debonding along the interface
14 A. NEMATI GIV ET AL.

Figure 8. Effect of different surface treatments on the ultimate strength of reinforced epoxies.[109]

and matrix-crack-controlled, respectively. According to the aforesaid failure


modes, the matrix strength of γ  MPS treated composites is higher than the
fiber/matrix interfacial strength, thus the matrix can transfer the load to the fibers.
However, for γ  GPS treated composites, the matrix fails before transferring the
load to the fibers. Therefore, it is believed that micro-failure behavior has a
predominant role to determine the efficacy of fiber/matrix interfacial adhesion
on composites tensile strength.[110,111]
From the above studies, it can be concluded that the stronger fiber/matrix
interfacial adhesion does not necessarily provide higher tensile strength for
reinforced polymers due to the interdependency of the micro-failure beha-
vior and fiber/matrix interfacial adhesion.
There are also several studies concerning the effect of fiber content on the
tensile strength of polymeric materials. Similar to particulate reinforced
composites, an unclear relationship between the fiber content and the tensile
THE JOURNAL OF ADHESION 15

Figure 9. Ultimate tensile strength of reinforced epoxies as a function of glass fiber volume
fraction with different surface treatments.[109]

Figure 10. Effect of glass fiber content on the tensile strength of reinforced epoxies.[112]

strength of reinforced polymers was observed.[109,112] Figure 9 shows the


effect of filler content with water-sized for γ  GPS and γ  MPS-treated
fibers on the tensile strength of epoxy/glass composites. It indicates that
16 A. NEMATI GIV ET AL.

the ultimate tensile strength of γ  MPS-treated composites increases linearly


with increasing fiber volume fraction, while the ultimate tensile strength of
epoxies with water-sized and γ  GPS-treated fibers increases slightly from
54.2 to 65% of fiber volume fraction.[109]
Swapnil et al[112] observed different relationships among the fiber loading
and the tensile strength of reinforced epoxies. As shown in Figure 10, the
longitudinal tensile strength of glass/epoxy composites improves until 50% of
volume fraction and then decreases with increasing the fiber content.
Embedding macrofiber reinforcements into the adhesive layer is another
effective way for the adhesive joint strength improvement. The effect of fiber
diameter, stiffness and horizontal distances (fiber horizontal distance is the
minimum distance between the two fibers next to each other) was investi-
gated by several researchers.[85,87] They observed that increasing the fiber
diameter and stiffness and decreasing the fiber horizontal distances improved
the adhesive stress distribution leading to an increase in the adhesive joint
strength (see Figure 11).

Effect of reinforcements of different scales on adhesive and adhesive joint


stiffness
Stiffness has a significant effect on the performance of adhesively bonded
joints in high technology industries. Numerous studies reported an incre-
ment in the stiffness of polymers by incorporation of particles and fibers to
various polymeric materials.[113,114] The incorporated particles have much
greater stiffness in comparison with their host polymer. They stay among the
polymer chains and reduce the chain flexibility, thus the stiffness of the
polymer increases.[115] The improvement in the elastic stiffness of rigid
micro and nano reinforced polymers such as adhesives depend on several
factors including particle size, particle/matrix interfacial adhesion and parti-
cle loading.[90] The effect of particle size on the stiffness of polymeric
materials has been studied by several researchers. The elastic modulus varia-
tions of epoxy/spherical glass particles are presented as a function of particle
loading and size (4.5–62 μm) in Figure 12.[96] As can be seen from Figure12,
for higher glass filler contents (i.e. 30–46 vol%), the stiffness of the reinforced
polymer decreases slightly with an increase of glass particle size, while at
lower glass filler volume fractions (i.e. 10–18 vol%), the stiffness of the
reinforced polymer is insensitive to the particle size.[96]
Similarly, an independent relationship between the elastic modulus of
epoxy/silica composites and particle size with diameters of 240, 560 nm and
1.56 μm was observed.[116] However, when the size of particles changes from
micro to nanoscale, the elastic modulus of the reinforced polymers behaves
differently from others. Cho et al[117] experimentally investigated the effect of
nano and microparticles size on the Young’s modulus of epoxy/alumina
THE JOURNAL OF ADHESION 17

Figure 11. Effect of fiber horizontal distances, diameter and stiffness on the adhesive stress
distribution.[85]
18 A. NEMATI GIV ET AL.

10 vol%

Elastic modulus (GPa)


18 vol%

30 vol%

40 vol%

46 vol%

Particle diameter ( )

Figure 12. Predicted elastic modulus of epoxy/spherical glass particles composite as a function
of particle size with different volume fractions.[90,96]

Figure 13. Young’s modulus of epoxy/alumina composites as a function of particle size and
volume fraction.[117]

composites. They observed that the elastic modulus of the reinforced epoxy is
not significantly affected by microparticles size. Moreover, a considerable
increase in Young’s modulus occurred with decreasing the particle size from
50 to 15 nm (see Figure 13). Similar results about the effect of micro and
nanoparticles reinforcements on the elastic modulus of composites were
obtained from other researchers.[100,118–122]
The efficacy of particle/matrix interfacial adhesion on the elastic modulus
of polymeric materials has been also studied by researchers. The role of
strong and poor interfacial adhesion between glass beads and epoxy matrix
on the tensile modulus of reinforced epoxy is shown in Figure 14. As is
evident from Figure 14, the strong or poor interfacial adhesion between the
nanoparticles and the polymer matrix has no influence on improving the
tensile modulus of nanocomposites.[123,124] Similar results were obtained
THE JOURNAL OF ADHESION 19

Figure 14. Elastic modulus of epoxy–Glass bead composites with strong and poor interfacial
adhesion.

from other studies.[125,126] Therefore, it can be concluded that the interfacial


adhesion might be an ineffective factor in improving the stiffness of rein-
forced polymeric materials because the elastic modulus of materials is
obtained at the initial part of the load-displacement curve with small amount
of deformation.[90]
The particle loading factor compared to other factors mentioned above might be
more influential in enhancing the modulus of reinforced polymeric materials.
Adachi and coworkers [116] studied the effect of particle size and volume fraction
on the elastic modulus of epoxy composites filled with spherical silica particles. The
size of particles varies from 240 nm to 1.56 μm and the volume fraction ranges
from 0 to 0.35. Figure 15 shows the trend of the elastic modulus as the filler loading

Figure 15. Young modulus of epoxy/silica composites with different particle volume fractions.[116]
20 A. NEMATI GIV ET AL.

increases. It can be concluded that for particles having a greater modulus than the
neat epoxy, the increase of particle content leads to a continuous enhancement in
the modulus of the reinforced epoxy until a critical filler content. For higher
volume fractions, this trend reverses due to agglomeration of nanoparticles.[127]
This trend of improvement in the elastic modulus of polymeric materials rein-
forced with rigid micro and nanoparticles was confirmed by many researches
mentioned in the above explanations.[96,119,122,124–126,128–130]
The tensile elastic modulus of microfiber-reinforced polymers was mostly
investigated by two parameters including fiber volume fraction and fiber/
matrix interfacial adhesion. Researchers introduced the “rule of mixture”
(ROM), an empirical method permitting to estimate the properties of com-
posites as a function of reinforcement and matrix volume fractions
(Eq.5).[131]
EIc ¼ Ef Vf þ Em Vm (5)

EIC is the longitudinal modulus of the composite,Vf is the volume fraction of


the fiber,Ef is the stiffness of the fiber and Vm ; Em are the corresponding
quantities for the matrix. According to Eq. 5, it can be concluded that the
longitudinal modulus of reinforced polymers increases constantly with
increasing the fiber volume fraction. Gershon and Marom[132] experimentally
studied the effect of glass fiber volume fraction on the longitudinal modulus
of a reinforced epoxy and the results were in accordance with Eq. 5. It is
evident that the values predicted by ROM are always higher than the experi-
mental results due to ignoring the variations resulting from imperfect fiber
alignment, non-uniform fiber distribution, local non-homogeneities and void
content (see Figure 16).

70 Ef

60

50

40

30

20

10
Em
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fiber volume fraction (Vf)
Figure 16. Elastic modulus of glass fiber-epoxy composites with various volume fractions.[132]
THE JOURNAL OF ADHESION 21

Madhukar et al[111] conducted an experimental investigation to determine


the effect of fiber/matrix adhesion on the tensile modulus of graphite/epoxy
composites. The result is similar to that mentioned above for rigid particle-
reinforced polymers. In other words, the tensile modulus of graphite/epoxy
composites is approximately insensitive to the fiber/matrix adhesion.
The effect of metal macrofiber reinforcements on the elastic modulus of
polymeric materials has not yet been investigated. However, the initial stiffness
of metal macrofiber-reinforced adhesive joints obtained from the load-displace-
ment curve of joints was first reported by Khoramishad and Razavi.[85] They
showed that the reduction of horizontal distances between the metal fibers in the
adhesive layer caused the initial stiffness of reinforced adhesive joints to improve
while this improvement was not observed in the elastic modulus of adhesive
joints studied by other researchers.[87]

Effect of reinforcements on adhesive and adhesive joint fracture toughness


As previously discussed, the main drawback of epoxy polymers is their notice-
able brittleness, low fracture resistance and impact strength due to their highly
crosslinked structure.[133] Fracture toughness of an adhesive layer has a signifi-
cant influence on the performance and durability of adhesively bonded joints in
critical structures such as aerospace, automotive, marine and defense.[28,134,135]
Therefore, toughening polymers is a vital and challenging issue that was studied
by many researchers. The addition of micro and nanoparticles is a well-estab-
lished modification method which improves the mechanical properties and
toughness of thermoplastic polymers.[136,137] Toughened adhesives by incorpor-
ating reinforcement materials have two main advantages: 1) reduced cost by
incorporating a high percentage of a low-cost material 2) enhanced composite
with improved properties.[138] Enhancement in fracture toughness of polymers
depends on various parameters including particle size, particle-matrix interfacial
adhesion, matrix ductility and particle loading (content). The significant effect of
particle size on the fracture toughness of polymeric materials has been reported
by several researchers.[90,121,139,140] The fracture toughness of a cured epoxy resin
filled with 55 and 64 wt% angular-shaped silica particles is plotted in Figure 17 as
a function of particles size (2-47 μm). The toughness increases by increasing the
particle size and loading due to the crack deflection around large particles.[140]
Similarly, the fracture toughness of alumina trihydrate filled epoxies improves
by increasing the particle size from 2 to12 μm as displayed in Figure 18.[121] The
fracture toughness of particulate composites typically increases with an increase
of the particles size. Nevertheless, this statement is not very pervasive especially
for nano-sized particles. In this case, the crack deflection mechanism has a
negligible effect because the micron-dimensioned crack tip is not deflected by
too small nano-sized particles. Therefore, the effect of other toughening
mechanisms should be considered such as plastic deformation and plastic void
22 A. NEMATI GIV ET AL.

Figure 17. Effect of particle size on epoxy filled silica composite fracture toughness with two
different particle loadings. Broken line indicates the fracture toughness of unfilled cured epoxy
resin.[90,140]

Figure 18. Effects of particle size and volume fraction on the fracture toughness of alumina
trihydrate powder filled epoxy composites at room temperature.[90,121]

growth mechanisms (these mentioned mechanisms are reviewed in Section. 5).


Huang and Kinloch[141] obtained an equation (Eq. 6) which models the mechan-
isms of plastic deformation and plastic void growth and it can be used for
nanosilica modified epoxy:

GC ¼ GCU þ φ (6)

GC is the fracture energy of the reinforced epoxy polymer, GCU is the fracture
energy of the neat epoxy polymer and φ is the improvement in the fracture
THE JOURNAL OF ADHESION 23

toughness of the reinforced epoxy due to the presence of particulate phase. It


is evident that φ contains the effect of the two aforesaid toughening
mechanisms:
φ ¼ CðΔGS þ ΔGV Þ (7)
C is a constant coefficient which is used to make good agreement between the
equation and experimental data, ΔGS and ΔGV represent the contributions to the
total increase in the fracture energy from the plastic deformation and plastic void
growth mechanisms, respectively. ΔGS is independent of particle size while ΔGV is
sensitive to particle size.[142–144] Zamanian et al[145] found that coefficients C and
ΔGV increase by decreasing the silica nanoparticle size (see Figure 19). Therefore,
according to Eqs. 6 and 7, the fracture toughness of reinforced epoxies improves
with a reduction in particle size. Some researchers believe that nano-sized reinfor-
cements have a better effect on the fracture toughness of composites in comparison
to microscale reinforcements. The fracture toughness of 20 nm SiO2 particulate
composites increased by about 79% at 5 vol% over the neat epoxy, while the
enhancement in fracture toughness of 10 μm SiO2 particles was only about
23.5% at 10 vol% (see Figure 20).[139]Particle loading is an important parameter
that should be considered carefully. The effect of particle loading on the fracture
toughness of composites is directly related to the nature of the particles. Generally,
for rigid particles (for a given particle size), the fracture toughness increases
monotonically with particles loading, while for ductile particles, the fracture
toughness increases slightly and then reaches a plateau.[138] According to most
observations, it can be concluded that the fracture toughness of particle-reinforced
polymers improves monotonically by increasing the particle content, but for
higher volume fractions the trend reverses due to the agglomeration and settlement
of the particles during curing of composites which leads to the reduction of
material performance by the creation of stress concentration sites (see
Figure 21).[116,120,138,146] Particle-matrix interfacial adhesion as a third factor
influences the fracture toughness of polymers, especially thermoplastic materials
because well-bonded particles act as crack stoppers not as defects in the matrix and
promote load transfer between particles and matrix (see Figure 22). In order to
enhance the interfacial adhesion between reinforcements and a polymer, particle
surface modification is performed by different useful techniques such as coupling
agents, alkyl succinic anhydride, stearic acid and so forth.[125,130,147] This surface
modification, especially in the case coupling agents, causes thermosetting materials
to behave ineffectively because crack growth is dominated by matrix failure and
particle breakage.[90]
Fracture toughness of microfiber-reinforced polymers is mainly influenced
by two main parameters including fiber volume fraction and fiber/matrix
interfacial adhesion. Some researchers believe that the fracture toughness of
reinforced composites increases with an improvement of the fiber/matrix
interfacial adhesion.[148] For the fiber volume fraction parameter, researchers
24 A. NEMATI GIV ET AL.

Figure 19. Effect of nanoparticle size and content on the fracture toughness of epoxy/silica
composites comparing experimental data (points) and the calculated model (line) (a) 12, (b) 20,
and (c) 40 nanosilica modified epoxies.[145]

presented contradictory results. Forte[149] observed that the fracture tough-


ness of double cantilever beam (DCB) adhesive joints remarkably decreased
by inclusion of glass fibers. However, this reduction becomes less by increas-
ing the fiber volume fraction. Yoon et al[150] studied the effect of glass fiber
volume fraction on the fracture toughness of DCB adhesive joints at cryo-
genic temperature. They found that the fracture toughness of DCB adhesive
joints increased up to certain level and then fell down.Considerably higher
THE JOURNAL OF ADHESION 25

Figure 20. Fracture toughness of epoxy filled 20 nm and 10 μm SiO2 particles at various volume
fractions.[139]

Figure 21. General trend of toughness properties as a function of particle volume fraction.[138]

fracture energy improvements could be obtained by embedding metallic


macrofibers into the adhesive layer. Razavi et al[84] incorporated metallic
fibers with different horizontal distances and determined the mixed mode
fracture behavior of a metallic fiber-reinforced epoxy adhesive. They con-
sidered the distance between the reinforcing fibers as the influencing para-
meter. The results revealed that the fracture energies of the reinforced
adhesive increased by decreasing the fiber distances. The fracture energies
of reinforced adhesives under mode I, mode II and mixed mode loading were
26 A. NEMATI GIV ET AL.

Figure 22. SEM micrographs of fracture surfaces of PC/glass bead composites (a) excellent
interfacial adhesion; (b) poor interfacial adhesion.[138]

respectively 12.46, 2.54 and 3.75 times higher than the non-reinforced adhe-
sive for a fiber distance of d=t ¼ 1(d/t parameter is the ratio of fiber diameter
to thickness of adhesive layer).

Effects of reinforcements on mechanical properties of adhesive and


adhesive joints exposed to water and moisture
Moisture absorption has unfavorable effects that cause the performance of
polymeric materials to be greatly affected. Long exposure time against moist-
ure results in irreversible damage of polymeric materials owing to hydrolysis
reactions, degradation of molecular structure and formation of microcavities
into the polymers.[151,152] For the case of adhesives and adhesive joints, these
irreversible damages result in loss of cohesive and bond strength of adhesives
and adhesive joints.[153,154] Hence, attempts to find a way for reducing the
effects of moisture absorption on the degradation of mechanical properties of
adhesives and adhesive joints are important. A method for improving the
mechanical properties of adhesives and adhesive joints exposed to moisture
is incorporating reinforcements. The effect of each reinforcement scale on
improving the mentioned properties are separately discussed in following
subsections and at the end, a summary of the most notable results studied by
researchers in this area are presented in Table 1 to 2.

Nanoscale reinforcements
In general, nano sized particles can play a significant role as moisture barriers
in reducing the moisture permeability of polymers.[155,156] Several parameters
such as the morphology and hydrophobicity[157,159] of nanoparticles decrease
THE JOURNAL OF ADHESION 27

Table 1. Strength improvement of nanoreinforced adhesives and adhesive joints under different
hygrothermal conditions.
Adherends NPs Percentage increase of
/type of Hygrothermal (wt strength at saturated
Adhesive adhesive joint condition Nanoparticles %) point Ref
Epoxy -/ Bulk RH = 100% SiC 1 16 [157]
specimen
Temperature: 60 C For 3 −2
time period of 360
hour
5 −7.4
Epoxy -/ Bulk RH = 100% MWCNTs 0.1 31.8 [157]
specimen
Temperature: 60 C For 0.3 29.9
time period of 360
hour
0.5 22.7
Epoxy Aluminum/ Lap According to standard CNH 0.5 133.3 [158]
shear joint EN ISO 9142
For time period of 12 CNTs 0.5 150
weeks
GNPs 0.5 100
Epoxy Stainless steel/ RH = 100% SNPs 10 20 [166]
Lap shear joint
Temperature: 60 C For 20 20
time period of 3
weeks

Table 2. Strength improvement of microreinforced adhesive and adhesive joint under different
hygrothermal conditions.
Adherends MPs Percentage increase of
/type of Hygrothermal MPs (wt% (wt strength at maximum time
Adhesive adhesive joint condition or vol %) %) of immersion Ref
Epoxy Aluminum/ Lap RH = 100% Aluminum 10 −17 [165]
shear joint powders
Immersed in 25 ~0
distilled water for 6
month
50 −11.7
Epoxy Aluminum/ Lap RH = 100% Aluminum 10 ~5 [165]
shear joint powders
Immersed in sea 25 ~0
water for 6 month
50 ~0
Epoxy -/ Bulk RH = 100% Cork 1 0 [163]
specimen particles
Temperature: 60 C
Immersed in
deionized
Epoxy -/ Bulk RH = 100% Cork 1 ~0 [163]
specimen particles
Temperature: 60 C
Immersed in
deionized
28 A. NEMATI GIV ET AL.

moisture diffusion through polymers. Plate-like nanoparticles[158] such as


graphene nanoplates and nano clays are the most efficient of all nano
morphologies for reducing moisture penetration acting as obstacles in the
path of moisture diffusion. Bharadwaj[160] presented a simple mathematic
model which predicts the moisture permeability through polymers reinforced
by nanofillers (see Eq. 8). According to Eq. 8, the relative moisture perme-
ability (the ratio of the moisture
 permeability of polymer nano-composites
ðPc Þto the pure polymer Pp ) is dependent on the volume fraction of the
filler ðΦf Þ (calculated from weight fractions) and a tortuosity factor ðτÞ. The
tortuosity factor (see Eq. 9) was defined as the ratio of the real distance ðd0 Þ
that moisture must travel to the shortest distance in the presence of nano-
fillers it would have travelled in the absence of nano fillers. The tortuosity
factor was also expressed in term of length (L), width (W), volume fraction of
the nanofiller and order parameter (S). The number in range of  12 to 1 was
dedicated for this parameter in which S =  12 , S = 1 and S = 0 indicate the
vertical, horizontal and random orientation of nanofillers (see Figure 23),
respectively.
Pc 1  Φf
¼ (8)
Pp τ
0
  
d L 2 1
τ¼ ¼1þ Φf Sþ (9)
d 2W 3 2
By replacing Eq. 9 in Eq. 8, Eq. 8 was rewritten as in Eq. 10:
Pc 1  Φf
¼   (10)
Pp 1 þ 2W Φf 23 S þ 12
L

According to Eq. 10, it can be concluded that without considering other


factors such as hydrophobicity, polymers reinforced with spherical nanopar-
ticles tend to absorb more moisture and are less efficient than other types of
nanofillers due to having equal length to width ratio.
The hydrophobic nature of reinforcements is another main factor that
can effectively prevent moisture penetration into reinforced polymers. The
family of carbon materials is inherently hydrophobic and due to this useful

Figure 23. Order parameter values in three different orientations (vertical, random and horizon-
tal orientations) of nanofillers.[60]
THE JOURNAL OF ADHESION 29

feature they are much regarded in the ship building industry.[161] For
instance, Bal and Saha[162] studied the effect of moisture absorption on
the mechanical properties of epoxy polymers reinforced with CNTs. They
found that the hydrophobic nature of CNTs caused the moisture penetra-
tion through the epoxy polymers to decrease. Therefore, the degradation
in mechanical properties of CNT-reinforced polymers was less than
unreinforced polymers.

Microscale reinforcements
Micro reinforcements including fibers and particles unlike nano reinforcement
generally have no tendency to decrease the moisture penetration and conse-
quently improve the mechanical properties of reinforced polymers.[163–165] The
ineffective behavior of micro reinforcements on improving these properties
when they are exposed to moisture and water can be the result of two main
factors including loss of strength of the micro reinforcement and loss of adhe-
sion and bond strength of the interfaces between the micro reinforcements and
the polymer matrix. Strength degradation of micro reinforcements exposed to
moisture and water chiefly occurs in microfiber reinforcements, especially the
natural ones. This type of micro reinforcement probably swells or dissolves with
water and moisture and consequently loses its functionality. Moreover, micro
reinforcements have large bond interfaces with the polymer matrix which leads
moisture and water to penetrate quickly and easily into the interfacial bond. This
provides conditions for moisture and water to bound with polar groups of the
polymer matrix, attach onto the hydrophilic groups of the fibers establishing
intermolecular hydrogen bonding with the fibers and consequently reducing the
interfacial adhesion.

Damage mechanisms in reinforced adhesives and adhesive joint


The addition of reinforcements into a polymer matrix as a second phase
can help the matrix behave differently against micro cracks and cracks
which are probably created by imperfections and voids. This change in
fracture surface allows some mechanical properties depending on type,
shape, size of reinforcements to improve significantly. In this section,
several possible mechanisms based on three scale size of reinforcements
(macro, micro and nano reinforcements) are introduced and discussed
briefly.

Crack pining and bowing mechanism


Researchers by experimental investigation found that cracks in brittle
materials undergo local velocity and shape changes in the surrounding
area of second-phase particles. In such cases, when the propagating
30 A. NEMATI GIV ET AL.

crack interacts with second-phase particles, particles can act as obstacles


to crack propagation causing the primary crack front to bow out and
enabling energy to be dissipated through the matrix (see Figure 24). This
mechanism has a significant effect on improving the fracture energy.
The effect of this mechanism on the fracture toughness is considerable
when micron-sized particles are used. This mechanism depends on two
parameters of the micron-sized particles: the particle size and the con-
tent. For a low volume fraction of particles, decreasing the particle size
and increasing the particle content increase the fracture toughness.[167]
For a high volume fraction of micron-sized particles, this trend will
increase non-linearly which implies there may be another mechanism
that competes with crack pinning. The effect of crack pinning on
improving the fracture toughness of ductile materials was found to be
less important.[168]

Bridging mechanism
The bridging mechanism is one of the important mechanisms that contribute
to enhancing the mechanical properties (especially the fracture toughness) of
polymers reinforced by nano, micro and macro reinforcements (see
Figure 25). The effect of this mechanism on improving the fracture tough-
ness can be dominant depending on the geometry and size of the reinforce-
ments. For instance, the use of larger particles and reinforcements with
cylindrical shapes cause the improvement in fracture toughness and energy
dissipating mechanism to be greater. Moreover, the interfacial bond strength
between the reinforcements and the polymer and the angle that the

(b)
Bowed section
Particle reinforcement Secondary crack front

(a)

Direction of crack propagation


Primary crack front

Figure 24. Schematic of crack-bowing mechanism (top view), a) crack approaches second-phase
particles (primary crack front), b) the crack bows at the second-phase particles (secondary crack
front).
THE JOURNAL OF ADHESION 31

Figure 25. Participation of nano reinforcements in bridging mechanism in adhesive, a) bridging


fibers b) pulled out fibers.[169]

reinforcements make with the fracture surface lead the reinforcements to


break or pull out from one surface to another.

Microcracking mechanism
This mechanism by formation of multiple cracks in front of the main crack tip
(see Figure 26) causes the fracture toughness of reinforced polymers, especially
polymers reinforced with microparticles, to improve. The microcracking
mechanism increases strain energy absorption by the creation of new surfaces
and particles debonding. Particle debonding and new cracks formation around
the main crack tip decrease the modulus and the stress intensity factor near the
crack tip resulting in an increase of fracture toughness of reinforced polymers.
In this mechanism, the particle size effect on the fracture toughness is contra-
dictory to that of bridging and pining mechanisms such that by decreasing the
particle size the fracture toughness will increase.[170]

Crack Micro cracks Particle reinforcement


Figure 26. Schematic of microcracking mechanism.
32 A. NEMATI GIV ET AL.

Crack deviation mechanism


The effect of crack deviation mechanism on improving the toughness of poly-
mers reinforced with particles was first investigated by Faber and Evants.[171]
They found that some features including particle morphology, particle aspect
ratio, volume fraction of particles and the distribution of particles into the
matrix had a significant role in increasing the fracture toughness. In the crack
deviation mechanism, the crack collides with the particles and then deviates
from its original path (see Figure 27). Therefore, this change in crack direction
causes an increase of the fracture roughness and consequently postpones the
final failure and increases the fracture toughness of reinforced polymers.

Particle debonding and subsequent plastic void growth mechanism


Debonding of particles followed by plastic void growth is an effective
mechanism for improving the mechanical properties especially the toughness
and of polymers reinforced with nano-sized particles (see Figure 28). The
process consists in two parts beginning with particle debonding due to local
stress concentration at particle-matrix interfaces. After debonding of the
particles from the matrix, sufficient conditions for local plastic void growth
through the matrix are provided. This can also be dependent on the degree of
interfacial adhesion that is provided from the particle-matrix interfaces. The

Crack Crack path deflection Particle reinforcement

b
Microcrack
deviation

Figure 27. Micro crack growth deviation a) schematic b) for an adhesive reinforced with
GNPs.[172]
THE JOURNAL OF ADHESION 33

Figure 28. Particle debonding with subsequent plastic void growth a) for aluminum oxide-
reinforced epoxy polymer b) for nano silica-reinforced epoxy polymer.[173]

formation of large amounts of plastic void through the matrix causes more
dissipated energy and finally increases the fracture resistance.

Shear banding mechanism


Shear banding which is originated from local stress concentrations due to the
presence of reinforcements into the matrix (see Figure 29) can play a
significant role on the mechanical properties improvement especially the
toughness of reinforced polymers. For the case of polymers reinforced by
nanoparticles, the formation of shear bands can be facilitated by particle
debonding with subsequent void nucleation (see Figure 30) such that
debonding at the particle–matrix interface reduces the degree of triaxial
stresses acting on the polymer matrix next to the particles. This reduction
of stress enables the shear bands to grow further. Shear yielding initiates
from the point of the maximum stress concentration. As the applied load is

a b

Figure 29. Shear banding mechanism a) for an adhesive reinforced with rubber particles (micro
scale) b) for an adhesive reinforced with metallic fibers (macro scale).[84,174]
34 A. NEMATI GIV ET AL.

Crack Debonding

Particle Shear banding induced by


reinforcemet particle debonding with
subsequent void growth

Figure 30. Schematic of shear banding mechanism.

further increased, formation of shear bands occurs with an approximate


angle of 45 degree to the applied stress.[174]

Conclusion
Adhesively bonded joints like other conventional joints have some disadvan-
tages that should be improved to expand their use in high technology
industries. In order to improve the mechanical properties of adhesive joints,
many techniques such as changes in the geometry of adhesive joints and
substrates have been studied by researchers and reported in the introduction
section of this paper. The adhesive is the most important part of an adhesive
joint system which role is to transfer the load from one substrate to another.
A weak performance of the adhesive layer causes premature joint failure.
Therefore, the adhesive needs to be reinforced like composite materials or
refined by rearranging the polymer chains. The addition of reinforcements at
different scales into the adhesive layer was considered in this study. Before
adding reinforcements into polymeric materials like adhesives, attention
should be paid to whether the selected reinforcement is suitable for the
desired application or not. In order to achieve this goal, the characteristics
of reinforcements at different scales and their applications recently used in
the field of adhesives and adhesive joints were presented in Section 2.
Mechanical properties improvement in any material should be performed
purposefully. Sometimes without major changes in the material structure this
improvement can be achieved. Researchers have studied parameters such as
reinforcement size, content and interface reinforcement/matrix for improving
the mechanical properties of composite materials. The effects of these para-
meters on strength, elastic modulus and fracture toughness were discussed in
Section 3. Finally, several damage mechanisms which are commonly observed in
reinforced adhesives and adhesives joints were introduced briefly. As discussed
THE JOURNAL OF ADHESION 35

in section 3, depending on the type of mechanical properties, some parameters


can be more influential than others. Therefore, effective parameters of reinfor-
cements should be well recognized for each scale. Micro and macro fiber
reinforcements compared to other reinforcements have shown better strength
and stiffness improvements in reinforced polymers. However, this improvement
is only relevant to specific fiber orientations and consequently this type of
reinforcements might not be applicable because sometimes loads applied to
structures are complex. The effect of rigid nano and microparticles on mechan-
ical properties of reinforced polymeric materials was also reviewed. The results
showed that nanoparticles with high specific surface area to volume ratio have
better performance for improving strength, stiffness and fracture toughness
properties. However, it is difficult to obtain a unique influence from nanopar-
ticles on improving mechanical properties of reinforced polymers because this
influence depends considerably on the manufacturing procedure. The effects of
reinforcements of different scales on water barrier properties and subsequently
the strength reduction of reinforced polymeric materials was also studied in this
review. Nano reinforcements unlike micro reinforcements act as obstacles in the
path of the moisture and cause the moisture not to penetrate easily into poly-
mers. Researchers believe that the ineffective behavior of micro reinforcements
against the moisture penetration is primarily related to large interfacial bond
which results in reducing interfacial adhesion.

References
[1] Kožuh, Z., Kralj, S., Cvirin, Ž. PROMET-Traffic&Transportation. 9(1–2), 33–40 (1997).
[2] Tsai, M. Y., Morton, J. Compos. Struct. 32(1–4), 123–131 (1995).
[3] Da Silva, L. F. M., Campilho, R. D. S. G. Advances in Numerical Modelling of Adhesive
Joints. In Advances in Numerical Modeling of Adhesive Joints, Springer Science &
Business Media: Verlag Berlin Heidelberg: (2012), pp. 1–93.
[4] Ghoddous, B. Lat. Am. J. Solids Struct. 14(2), 256–276 (2017).
[5] Budhe, S., Banea, M. D., De Barros, S., Da Silva, L. F. M. Int. J. Adhes. Adhes. 72,
30–42 (2017).
[6] Sancaktar, E., Gomatam, R. J. Adhes. Sci. Technol. 15(1), 97–117 (2001).
[7] Teixeira, F. G., Da Silva, L. F. M. J. Adhes. 87(7–8), 671–687 (2011).
[8] You, M., Zheng, Y., Zheng, X.-L., Liu, W.-J. Int. J. Adhes. Adhes. 23(5), 365–369 (2003).
[9] Deng, J., Lee, M. M. K. Compos. Part B Eng. 39(4), 731–739 (2008).
[10] Lang, T. P., Mallick, P. K. Int. J. Adhes. Adhes. 18(3), 167–177 (1998).
[11] Adams, R. D., Comyn, J., Wake, W. C. Structural Adhesive Joints in Engineering,
Springer Science & Business Media, Netherlands, (1997).
[12] Gleich, D. M., Van Tooren, M. J. L., Beukers, A. J. Adhes. Sci. Technol. 15(9),
1091–1101 (2001).
[13] Da Silva, L. F. M., Rodrigues, T., Figueiredo, M. A. V., De Moura, M., Chousal, J. A. G.
J. Adhes. 82(11), 1091–1115 (2006).
[14] Arenas, J. M., Narbon, J. J., Alía, C. Int. J. Adhes. Adhes. 30(3), 160–165 (2010).
[15] Da Silva, L. F. M., Adams, R. D. Int. J. Adhes. Adhes. 27(3), 227–235 (2007).
[16] Fitton, M. D., Broughton, J. G. Int. J. Adhes. Adhes. 25(4), 329–336 (2005).
36 A. NEMATI GIV ET AL.

[17] Da Silva, L. F. M., Lopes, M. J. C. Q. Int. J. Adhes. Adhes. 29(5), 509–514 (2009).
[18] Da Silva, L. F. M., Pirondi, A., Ochsner, A. Hybrid Adhesive Joints, 6, (Springer Science
& Business Media, Springer-Verlag Berlin Heidelberg, 2011).
[19] Sadowski, T., Kneć, M., Golewski, P. Int. J. Adhes. Adhes. 30(5), 338–346 (2010).
[20] Barbosa, A. Q., Da Silva, L. F. M., Abenojar, J., Figueiredo, M. Appl. Adhes. Sci. 5(1), 9
(2017).
[21] Barbosa, A. Q., Da Silva, L. F. M., Öchsner, A., Abenojar, J., Del Real, J. C. J. Adhes. 88
(4–6), 452–470 (2012).
[22] Khayal, O. M. E. S. J. Microsc. Ultrastruct. 5, 119–122, (2017).
[23] Roylance, D. Introduction to Composite Materials, (Dep. Mater. Sci. Eng. Massachusetts
Inst. Tech., Cambridge, (2000).
[24] Sekunowo, O. I., Durowaye, S. I., Lawal, G. I. Int. J. Mech. Aerospace, Ind. Mechatron.
Manuf. Eng. 9(1), 1–7 (2015).
[25] Napierska, D., Thomassen, L. C. J., Lison, D., Martens, J. A., Hoet, P. H. Part. Fibre
Toxicol. 7(1), 39 (2010).
[26] Calabi Floody, M., Theng, B. K. G., Reyes, P., Mora, M. L. Clay Miner. 44(2), 161–176
(2009).
[27] Razavi, S. M. J., Ayatollahi, M. R., Giv, A. N., Khoramishad, H. Int. J. Adhes. Adhes. 80,
October 2017, 76–86 (2018).
[28] Shadlou, S., Ahmadi-Moghadam, B., Taheri, F. Rev. Adhes. Adhes. 2(3), 371–412
(2014).
[29] Zheng, Y., Zheng, Y., Ning, R. Mater. Lett. 57(19), 2940–2944 (2003).
[30] Tutunchi, A., Kamali, R., Kianvash, A. J. Adhes. 91(9), 663–676 (2015).
[31] Hsiao, K.-T., Alms, J., Advani, S. G. Nanotechnology. 14(7), 791 (2003).
[32] Park, S. W., Lee, D. G. J. Adhes. Sci. Technol. 23(4), 619–638 (2009).
[33] Park, S. W., Kim, B. C., Lee, D. G. J. Adhes. Sci. Technol. 23(1), 95–113 (2009).
[34] Li, J., Lumpp, J. K. Electrical and Mechanical Characterization of Carbon Nanotube
Filled Conductive Adhesive. Aerospace Conference, 2006 IEEE, (2006), p. 6.
[35] Khalili, S. M. R., Tavakolian, M., Sarabi, A. J. Adhes. Sci. Technol. 24(11–12), 1917–
1928 (2010).
[36] Gorbatkina, Y. A., Ivanova-Mumzhieva, V. G., Ul’yanova, T. M. Polym. Sci. Ser. C. 49
(2), 131–134 (2007).
[37] Ayatollahi, M. R., Nemati Giv, A., Razavi, S. M. J., Khoramishad, H. J. Adhes. 93, 896–
913 (2016).
[38] Kaboorani, A., Riedl, B. Compos. Part A Appl. Sci. Manuf. 42(8), 1031–1039 (2011).
[39] Zhou, Y. X., Wu, P. X., Cheng, Z. Y., Ingram, J., Jeelani, S. Express Polym. Lett. 2(1),
40–48 (2008).
[40] Tutunchi, A., Kamali, R., Kianvash, A. J. Adhes. Sci. Technol. 29(3), 195–206 (2015).
[41] Saha, M. C., Kabir, M. E., Jeelani, S. Mater. Sci. Eng. A. 479(1), 213–222 (2008).
[42] Prolongo, S. G., Gude, M. R., Sanchez, J., Urena, A. J. Adhes. 85(4–5), 180–199 (2009).
[43] Saeed, M. B., Zhan, M.-S. Int. J. Adhes. Adhes. 27(4), 306–318 (2007).
[44] Sprenger, S., Eger, C., Kinloch, A. J., Lee, J. H., Taylor, A. C., Egan, D. Spec. Issue
Adhaes. Kleb. Dicht. 48(3), 17–21 (2004).
[45] Buchman, A., Dodiuk-Kenig, H., Dotan, A., Tenne, R., Kenig, S. J. Adhes. Sci. Technol.
23(5), 753–768 (2009).
[46] Kaboorani, A., Riedl, B. J. Ind. Eng. Chem. 18(3), 1076–1081 (2012).
[47] Uddin, F. Metall. Mater. Trans. A. 39(12), 2804–2814 (2008).
[48] Cha, C., Shin, S. R., Annabi, N., Dokmeci, M. R., Khademhosseini, A. ACS Nano. 7(4),
2891–2897 (2013).
THE JOURNAL OF ADHESION 37

[49] Jariwala, D., Sangwan, V. K., Lauhon, L. J., Marks, T. J., Hersam, M. C. Chem. Soc. Rev.
42(7), 2824–2860 (2013).
[50] De Volder, M. F. L., Tawfick, S. H., Baughman, R. H., Hart, A. J. Science (80-.). 339
(6119), 535–539 (2013).
[51] Sadigh, M. A. S., Marami, G. Mater. Des. 92, 36–43 (2016).
[52] Ibrahim, K. S. Carbon Lett. 14(3), 131–144 (2013).
[53] Rafiee, M. A., Rafiee, J., Wang, Z., Song, H., Yu, -Z.-Z., Koratkar, N. ACS Nano. 3(12),
3884–3890 (2009).
[54] Labulo, A. H., Martincigh, B. S., Omondi, B., Nyamori, V. O. Advances in carbon
nanotubes as efficacious supports for palladium-catalysed carbon–carbon cross-cou-
pling reactions, J. Mater. Sci. 52(16), 9225–9248 (2017).
[55] Morones, J. R., Elechiguerra, J. L., Camacho, A., Holt, K., Kouri, J. B., Ramírez, J. T.,
Yacaman, M. J. Nanotechnology. 16(10), 2346 (2005).
[56] Gilbert, E. N., Hayes, B. S., Seferis, J. C. Polym. Eng. Sci. 43(5), 1096–1104 (2003).
[57] Ekrem, M., Ataberk, N., Avcı, A., Akdemir, A. J. Adhes. Sci. Technol. 31(7), 699–712
(2017).
[58] Sweet, M. J., Chessher, A., Singleton, I. Metal-Based Nanoparticles; Size, Function, and
Areas for Advancement in Applied Microbiology. In Advances in Applied Microbiology,
Gadd, G. M., and Sariaslani, S., Eds. (Elsevier, Amsterdam, Netherlands. 2012), Vol. 80,
pp 113–142.
[59] Assaedi, H., Shaikh, F. U. A., Low, I. M. J. Asian Ceram. Soc. 4(1), 19–28 (2016).
[60] Nazir, M. S., Kassim, M. H. M., Mohapatra, L., Gilani, M. A., Raza, M. R., Majeed,
K. Characteristic Properties of Nanoclays and Characterization of Nanoparticulates
and Nanocomposites. In Nanoclay Reinforced Polymer Composites, Jawaid, M.,
Qaiss, A. K., and Bouhfid, R., Eds. (Springer, Singapore, 2016), pp 35–55.
[61] Kahraman, R., Sunar, M., Yilbas, B. J. Mater. Process. Technol. 205(1), 183–189 (2008).
[62] Khalili, S. M. R., Shokuhfar, A., Hoseini, S. D., Bidkhori, M., Khalili, S., Mittal, R. K.
Int. J. Adhes. Adhes. 28(8), 436–444 (2008).
[63] Kilik, R., Davies, R. Int. J. Adhes. Adhes. 9(4), 224–228 (1989).
[64] Banea, M. D., Da Silva, L. F. M., Carbas, R. J. C., Campilho, R. Int. J. Adhes. Adhes. 54,
191–199 (2014).
[65] Hunter, R., Möller, J., Vizán, A., Pérez, J., Molina, J., Leyrer, J. J. Adhes. 93(11), 879–
895 (2017).
[66] Banea, M. D., Da Silva, L. F. M., Carbas, R. J. C. Int. J. Adhes. Adhes. 59, 14–20 (2015).
[67] Jonsson, M. Thermally Expandable Microspheres Prepared via Suspension
Polymerization-Synthesis, Characterization, and Application, (KTH, Stockholm,
Sweden, 2010).
[68] Jonsson, M., Nordin, O., Kron, A. L., Malmström, E. J. Appl. Polym. Sci. 117(1), 384–
392 (2010).
[69] Banea, M. D., Da Silva, L. F. M., Carbas, R. J. C., Campilho, R. D. S. G. J. Adhes. 91(10–11),
823–840 (2015).
[70] Pesquet, G., Da Silva, L. F. M., Sato, C. Frat. Ed Integrità Strutt. 16, 18 (2011).
[71] Gil, L. Materials (Basel). 2(3), 776–789 (2009).
[72] Barbosa, A. Q., Da Silva, L. F. M., Öchsner, A., Marques, E. A. S., Abenojar, J. Micro
Cork Particles as Adhesive Reinforcement Material for Brittle Resins. In Materials
Design and Applications, da Silva, L. F. M., Ed., (Springer, Cham, Switzerland, 2017),
pp 399–418.
[73] Barbosa, A. Q.; da Silva, L. F. M.; Abenojar, J.; Figueiredo, M.; Öchsner, A. Toughness
of a brittle epoxy resin reinforced with micro cork particles: Effect of size, amount and
surface treatment, Compos. Part B Eng., 2017), 114, 299–310 (2017).
38 A. NEMATI GIV ET AL.

[74] James, W. B., Corporation, H. Powder Metallurgy Methods and Applications, ASM
International: Novelty, Ohio, USA, 7.
[75] Crumb Rubber Powders. [Online]. Available: https://www.indiamart.com.
[76] Serope Kalpakjian, S. S. Manufacturing Engineering and Technology, 7th ed.; Pearson
Publications: Singapore. 2013.
[77] Bagherpour, S. Fibre Reinforced Polyester Composites, (Intech, Rijeka, Croatia 2012).
[78] Wallenberger, F. T., Watson, J. C., Li, H. Glass Fibers. PPG Industries; Inc., ASM Int.:
Ohio, USA, 2000.
[79] Prashanth, S., Km, S., Nithin, K., Sachhidananda, S. Fiber Reinforced Composites - A
Review. J. Mat. Sci. & Eng. 6(3), 341 (2017). doi:10.4172/2169-0022.1000341
[80] Comparison of Carbon Fiber, Kevlar® (Aramid) and E Glass used in Composites for
Boatbuilding. [online]. (2015) https://www.christinedemerchant.com/carbon-kevlar-
glass-comparison.html.
[81] MInus, M., Kumar, S. Jom. 57(2), 52–58 (2005).
[82] Solomon, B., George, D., Shunmugesh, K., Akhil, K. T. Polym. Polym. Compos. 25(3),
237 (2017).
[83] Dorey, G. J. Phys. D. Appl. Phys. 20(3), 245 (1987).
[84] Razavi, S. M. J., Ayatollahi, M. R., Esmaeili, E., Da Silva, L. F. M. Eur. J. Mech. (2017),
65, 349–359.
[85] Khoramishad, H., Razavi, S. M. J. Int. J. Adhes. Adhes. 55, 114–122 (2014).
[86] Kaji, M., Farahani, M., Ansari, M. J. Adhes. Sci. Technol. 31(17), 1963–1975 (2017).
[87] Nemati Giv, A., Ayatollahi, M. R., Razavi, S. M. J., Khoramishad, H. J. Adhes. 1–21
(2017). DOI:org/10.1080/00218464.2017.1305270
[88] Esmaeili, E., Razavi, S. M. J., Bayat, M., Berto, F. J. Adhes. 1–20 (2017).
[89] Khalili, S. M. R., Fathollahi, M. R. Int. J. Adhes. Adhes. 54, 21–29 (2014).
[90] Fua, S.-Y., Fengb, X.-Q., Laukec, B., Maid, Y.-W. Compos. Part B. 39, 933–961 (2008).
[91] Pukanszky, B., Vörös, G. Compos. Interfaces. 1(5), 411–427 (1993).
[92] Nakamura, Y., Yamaguchi, M., Okubo, M., Matsumoto, T. J. Appl. Polym. Sci. 45(7),
1281–1289 (1992).
[93] Reynaud, E., Jouen, T., Gauthier, C., Vigier, G., Varlet, J. Polymer (Guildf). 42(21),
8759–8768 (2001).
[94] Sumita, M., Shizuma, T., Miyasaka, K., Ishikawa, K. J. Macromol. Sci. Part B Phys. 22
(4), 601–618 (1983).
[95] Buggy, M., Bradley, G., Sullivan, A. Compos. Part A Appl. Sci. Manuf. 36(4), 437–442
(2005).
[96] Road, M. E. Crack propagation in a glass particle-filled epoxy resin Part 2 Effect of
particle-matrix adhesion. J. Mater. Sci., 19(2), 487–496 (1984).
[97] Rong, M. Z., Zhang, M. Q., Pan, S. L., Lehmann, B., Friedrich, K. Polym. Int. 53(2),
176–183 (2004).
[98] Wu, C. L., Zhang, M. Q., Rong, M. Z., Friedrich, K. Compos. Sci. Technol. 65(3), 635–
645 (2005).
[99] Thio, Y. S., Argon, A. S., Cohen, R. E. Polymer (Guildf). 45(10), 3139–3147 (2004).
[100] Zhu, Z., Yang, Y., Yin, J., Qi, Z. J. Appl. Polym. Sci. 73(14), 2977–2984 (1999).
[101] Tjong, S. C., Xu, S. A. J. Appl. Polym. Sci. 81(13), 3231–3237 (2001).
[102] Jiang, L., Zhang, J., Wolcott, M. P. Polymer (Guildf). 48(26), 7632–7644 (2007).
[103] Maazouz, A., Sautereau, H., Gerard, J. F. J. Appl. Polym. Sci. 50(4), 615–626 (1993).
[104] Chen, Y., Zhou, S., Yang, H., Wu, L. J. Appl. Polym. Sci. 95(5), 1032–1039 (2005).
[105] Nielsen, L. E. J. Appl. Polym. Sci. 10(1), 97–103 (1966).
[106] Turcsanyi, B., Pukanszky, B., Tüdõs, F. J. Mater. Sci. Lett. 7(2), 160–162 (1988).
THE JOURNAL OF ADHESION 39

[107] Pukánszky, B., Turcsanyi, B., Tudos, F. Interfaces Polym. Ceram. Met. Matrix Compos.
Amsterdam Elsevier. 467–477 (1988).
[108] Young, R. J., Beaumont, P. W. R. J Mater Sci. 12(4), 684–692 (1977).
[109] Zhao, F. M., Takeda, N. Compos. Part A Appl. Sci. Manuf. 31(11), 1203–1214 (2000).
[110] Norita, T., Matsui, J., Matsuda, H. S. Compos. Interfaces. 123–132 (1986).
[111] Madhukar, M. S., Drzal, L. T. J. Compos. Mater. 25(8), 958–991 (1991).
[112] Swapnil, A. S., SatheSandip, B., ChaudhariBapu, P., Vishal, S. J. Mater. Today Proc. 4
(9), 9487–9490 (2017).
[113] Ayatollahi, M. R., Shadlou, S., Shokrieh, M. M. Eng. Fract. Mech. 78(14), 2620–2632
(2011).
[114] Shokrieh, M. M., Kefayati, A. R., Chitsazzadeh, M. Mater. Des. 40, 443–452 (2012).
[115] Ayatollahi, M. R., Shokrieh, M. M., Shadlou, S., Kefayati, A. R., Chitsazzadeh, M. Iran
Polym J. 20(10), 835–843 (2011).
[116] Adachi, T., Osaki, M., Araki, W., Kwon, S.-C. Acta Mater. 56(9), 2101–2109 (2008).
[117] Cho, J., Joshi, M. S., Sun, C. T. Compos. Sci. Technol. 66(13), 1941–1952 (2006).
[118] Verbeek, C. J. R. Mater. Lett. 57(13–14), 1919–1924 (2003).
[119] Radford, K. C. J. Mater. Sci. 6(10), 1286–1291 (1971).
[120] Al-Turaif, H. A. Prog. Org. Coatings. 69(3), 241–246 (2010).
[121] Lange, F. F., Radford, K. C. J. Mater. Sci. 6(9), 1197–1203 (1971).
[122] Singh, R. P., Zhang, M., Chan, D. J. Mater. Sci. 37(4), 781–788 (2002).
[123] Amdouni, N., Sautereau, H., Gerard, J. F. J. Appl. Polym. Sci. 46(10), 1723–1735 (1992).
[124] Dekkers, M. E. J., Heikens, D. J. Appl. Polym. Sci. 28(12), 3809–3815 (1983).
[125] Levita, G., Marchetti, A., Lazzeri, A. Polym. Compos. 10(1), 39–43 (1989).
[126] DiBenedetto, A. T., Wambach, A. D. Int. J. Polym. Mater. 1(2), 159–173 (1972).
[127] Yasmin, A., Luo, -J.-J., Daniel, I. M. Compos. Sci. Technol. 66(9), 1182–1189 (2006).
[128] Ji, X. L., Jing, J. K., Jiang, W., Jiang, B. Z. Polym. Eng. Sci. 42(5), 983–993 (2002).
[129] Suprapakorn, N., Dhamrongvaraporn, S., Ishida, H. Polym. Compos. 19(2), 126–132
(1998).
[130] Ou, Y., Yang, F., Yu, Z. J. Polym. Sci. Part B Polym. Phys. 36(5), 789–795 (1998).
[131] High Strength Composites. [Online]. http://www.mech.utah.edu/~rusmeeha/labNotes/
composites.html
[132] Gershon, B., Marom, G. J. Mater. Sci. 10(9), 1549–1556 (1975).
[133] Barbosa, A. Q., Da Silva, L. F. M., Abenojar, J., Del Real, J. C., Paiva, R. M. M.,
Öchsner, A. Thermochim. Acta. 604, 52–60 (2015).
[134] Banea, M. D., Da Silva, L. F. M., Campilho, R. D. S. G., Sato, C. J. Adhes. 90(1), 16–40
(2014).
[135] Machado, J. J. M., Marques, E. A. S., Da Silva, L. F. M. J. Adhes. 1–32 (2017).
[136] Prolongo, S. G., Gude, M. R., Ureña, A. J Mater Sci Eng. 1(2), 22–2169 (2012).
[137] Domun, N., Hadavinia, H., Zhang, T., Sainsbury, T., Liaghat, G. H., Vahid, S.
Nanoscale. 7(23), 10294–10329 (2015).
[138] Barbosa, A. Q., Silva, L. F. M., Banea, M. D., Öchsner, A. Materwiss. Werksttech. 47(4),
307–325 (2016).
[139] Jang, J. Particle Size Effect on Mechanical and Thermal Properties of SiO2 Particulate
Polymer Composites, (University of Delaware, 2012).
[140] Nakamura, Y., Yamaguchi, M., Kitayama, A., Okubo, M., Matsumoto, T. Polymer
(Guildf). 32(12), 2221–2229 (1991).
[141] Huang, Y., Kinloch, A. J. J. Mater. Sci. 27(10), 2763–2769 (1992).
[142] Hsieh, T. H., Kinloch, A. J., Masania, K., Lee, J. S., Taylor, A. C., Sprenger, S. J. Mater.
Sci. 45(5), 1193–1210 (2010).
40 A. NEMATI GIV ET AL.

[143] Hsieh, T. H., Kinloch, A. J., Masania, K., Taylor, A. C., Sprenger, S. Polymer (Guildf).
51(26), 6284–6294 (2010).
[144] Liang, Y. L., Pearson, R. A. Polymer (Guildf). 50(20), 4895–4905 (2009).
[145] Zamanian, M., Mortezaei, M., Salehnia, B., Jam, J. E. Eng. Fract. Mech. 97, 193–206
(2013).
[146] García, M. G., Marchese, J., Ochoa, N. A. J. Appl. Polym. Sci. 118(4), 2417–2424 (2010).
[147] Kinloch, A. J. Structural Adhesives: Developments in Resins and Primers, (Elsevier
Applied Science Publishers, 1986).
[148] Madhukar, M. S., Drzal, L. T. J. Compos. Mater. 26(7), 936–968 (1992).
[149] Forte, M. S., Whitney, J. M., Schoeppner, G. A. Compos. Sci. Technol. 60(12), 2389–
2405 (2000).
[150] Yoon, S. H., Kim, B. C., Lee, K. H., Lee, D. G. J. Adhes. Sci. Technol. 24(2), 429–444
(2010).
[151] Alessi, S., Conduruta, D., Pitarresi, G., Dispenza, C., Spadaro, G. Polym. Degrad.
Stab. 96(4), 642–648 (2011).
[152] Xiao, G. Z., Shanahan, M. E. R. J. Appl. Polym. Sci. 69(2), 363–369 (1998).
[153] Molitor, P., Barron, V., Young, T. Int. J. Adhes. Adhes. 21(2), 129–136 (2001).
[154] Mubashar, A., Ashcroft, I. A., Critchlow, G. W., Crocombe, A. D. Int. J. Adhes. Adhes.
29(8), 751–760 (2009).
[155] Ladhari, A., Ben Daly, H., Belhadjsalah, H., Cole, K. C., Denault, J. Polym. Degrad. Stab.
95(4), 429–439 (2010).
[156] Pramoda, K. P., Liu, T. J. Polym. Sci. Part B Polym. Phys. 42(10), 1823–1830 (2004).
[157] Khoramishad, H., Alizadeh, O. Polym. Compos. (2017). DOI:10.1002/pc.24303.
[158] Jojibabu, P., Ram, G. D. J., Deshpande, A. P., Bakshi, S. R. Polym. Degrad. Stab. 140,
84–94 (2017).
[159] Gude, M. R., Prolongo, S. G., Urena, A. Int. J. Adhes. Adhes. 40, 179–187 (2013).
[160] Bharadwaj, R. K. Macromolecules. 34(26), 9189–9192 (2001).
[161] Kootsookos, A., Mourotz, A. P., St John, N. A. Comparison of the Seawater Durability
of Carbon and Glass Polymer Composites, ID-1200. (Proc. ICCM-13, Beijing, 2001).
[162] Bal, S., Saha, S. Int. J. Damage Mech. 26(5), 758–770 (2017).
[163] Barbosa, A. Q., Da Silva, L. F. M., Öchsner, A. J. Adhes. Sci. Technol. 29(16), 1714–1732
(2015).
[164] Azwa, Z. N., Yousif, B. F., Manalo, A. C., Karunasena, W. Mater. Des. 47, 424–442
(2013).
[165] Al-Harthi, M., Kahraman, R., Yilbas, B., Sunar, M., Aleem, B. J. A. J. Adhes. Sci.
Technol. 18(15–16), 1699–1710 (2004).
[166] Zhou, H., Liu, H.-Y., Zhou, H., Zhang, Y., Gao, X., Mai, Y.-W. Mater. Des. 95, 212–218
(2016).
[167] Lange, F. F. Philos. Mag. 22(179), 983–992 (1970).
[168] Owen, A. B. J. Mater. Sci. 14(10), 2521–2523 (1979).
[169] Gude, M. R., Prolongo, S. G., Gómez-Del Río, T., Urena, A. Int. J. Adhes. Adhes. 31(7),
695–703 (2011).
[170] Dittanet, P. Fracture Behavior of Silica Nanoparticle Filled Epoxy Resin, (Lehigh
University, Bethlehem, Pennsylvania, USA 2011).
[171] Faber, K. T., Evans, A. G. Acta Metall. 31(4), 565–576 (1983).
[172] Khoramishad, H., Ebrahimijamal, M., Fasihi, M. Fatigue Fract. Eng. Mater. Struct.
(2017).
[173] Quaresimin, M., Schulte, K., Zappalorto, M., Chandrasekaran, S. Compos. Sci. Technol.
123, 187–204 (2016).
[174] Kinloch, A. J. MRS Bull. 28(06), 445–448 (2003).

You might also like