Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

Author’s Accepted Manuscript

A review on application of structural adhesives in


concrete and steel–concrete composite and factors
influencing the performance of composite
connections

Pankaj Kumar, Amar Patnaik, Sandeep Chaudhary


www.elsevier.com/locate/ijadhadh

PII: S0143-7496(17)30059-3
DOI: http://dx.doi.org/10.1016/j.ijadhadh.2017.03.009
Reference: JAAD1989
To appear in: International Journal of Adhesion and Adhesives
Received date: 18 June 2016
Accepted date: 6 March 2017
Cite this article as: Pankaj Kumar, Amar Patnaik and Sandeep Chaudhary, A
review on application of structural adhesives in concrete and steel–concrete
composite and factors influencing the performance of composite connections,
International Journal of Adhesion and Adhesives,
http://dx.doi.org/10.1016/j.ijadhadh.2017.03.009
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
A review on application of structural adhesives in concrete and steel–concrete composite and

factors influencing the performance of composite connections

Pankaj Kumara, Amar Patnaikb, Sandeep Chaudharya*

a
Department of Civil Engineering, MNIT Jaipur, Jaipur-302017, India
b
Department of Mechanical Engineering, MNIT Jaipur, Jaipur-302017, India

*
Corresponding Author: Dr. Sandeep Chaudhary, Tel.: +91-94144-75375. E-mail:
sandeep.nitjaipur@gmail.com

Abstract

This paper chronicles the use of structural adhesives in civil engineering construction

since its inception. The usage of structural adhesives as effective and popular strengthening

agents has been discussed. The application of structural adhesives as connecting agents,

especially in cases of steel–concrete composites has also been discussed in detail. Various

factors influence the bond strength of interfaces, such as the physical, mechanical and

chemical properties of structural adhesives and adherends, the shape of adherends, water

immersion, adhesive layer thickness, bonded area geometry, relative humidity and

temperature of the environment during curing and service life, the amount and type of fillers

and surface finishing of adherends.

Keywords: structural adhesive; steel; concrete; composite; shear strength.

1
1. Introduction

1.1 Structural adhesives

Structural adhesives contribute substantially to the integrity of the component or the

product being prepared. Structural adhesives mainly find their use in the construction

industry for repairing, strengthening and reinforcing existing structures, connecting two

similar or dissimilar materials and resisting mechanical and environmental loads. Earlier,

these structural adhesives were used to fill the gaps between precast members. Homo-

polymer Polyvinyl Acetate (PVAC) is among the first constructional adhesives used as filler

in the gaps between precast members. However, owing to its poor resistance against

environmental action, its use has been limited to internal applications [1]. The chemical

composition of structural adhesives usually consists of polymeric chains of epoxy,

polyurethane or acrylic based groups.

Various enhancements in existing structural adhesives and the development of new

adhesives have increased their applicability in various areas such as retrofitting of existing

structures and structural connection. In comparison to the conventional mechanical

connectors, structural adhesives offer increased homogeneity in stress distribution, along with

a reduced formwork, accelerated construction speed, quality assurance and improved fatigue

life of members. Structural adhesives can also be used to join members having thin cross-

sectional elements. They act as binding materials and provide better resistance to corrosion

and water percolation. Owing to these advantages, adhesives are used for bonding similar as

well as dissimilar parts in the aerospace, automotive, marine and construction industries [2,

3]. A systematic research on man-made structural adhesives for construction applications was

started in the 1950s [4, 5].

With time, any structure deteriorates, irrespective of the construction material used.

2
This deterioration usually leads to reduced stiffness and strength of members and structure. In

the case of cement concrete (CC) and reinforced cement concrete (RCC) structures, cracking,

spalling and collapse of large concrete masses are common reasons for the deterioration of

concrete in any structure/member. The selection of an appropriate repairing/retrofitting

method for deteriorated structures is imperative [6], and the use of structural adhesives is

among the most common and efficient techniques to prevent structural deterioration [7-9].

Steel–concrete composite constructions have numerous advantages over non-

composite constructions such as a higher strength to weight ratio, more flexural strength and

stiffness, speedier and more flexible construction, ease in retrofitting and repair, higher

durability and better aesthetics [10-14]. Conventionally, mechanical connectors are used in

steel–concrete composite constructions, but they cause stress concentration and have poor

fatigue life [11, 12]. Another shortcoming of such connections is their inability to provide a

high degree of interaction. Also, the higher density of mechanical connectors may result in

improper placement of concrete. Fig. 1(a) shows a schematic view of a mechanical stud

connector in a steel–concrete composite member.

To overcome the shortcomings of mechanical connectors, the efficacy of structural

adhesives has been thoroughly researched. The new age design requirements with light

weight and a multi-material approach can be effectively met with structural adhesives [15].

Fig. 1(b) shows a schematic view of an adhesive bonded steel–concrete composite member.

Concrete slab
Mechanical connector
Structural adhesive
Steel section

Fig. 1(a). Mechanical stud connected composite member; (b) adhesive bonded composite
member

3
1.2 Scope of current paper

This paper recapitulates the research on structural adhesives in two broad categories, which

are:

(a) an overview of the use of structural adhesives in concrete and steel–concrete composite

structures.

(b) a study of factors affecting the strength of adhesive bonded connections.

A comprehensive discussion on the state of the art is presented in the following sections.

2. Structural adhesive in concrete and steel–concrete composite construction

2.1 Repair and strengthening of existing structures

The repair of CC or RCC members and their joints is commonly done using low

viscosity epoxy resin because adhesives penetrate well, fill the cracks efficiently, and are

capable of restoring the lateral load capacity [4, 14-18].

Thanoon et al. [19] noted that the load-carrying capacity of RCC slab repaired using

structural adhesive was found to increase considerably, with the deflection and stiffness

approximately similar to those of the control slab. The deflection and stiffness of the precast

RCC slab repaired using epoxy were almost the same as those of the control slab. Fig. 2

shows a schematic view of a repaired precast RCC slab, which has two cracks. The stiffening

of hardened concrete may arrest the cracking in structures and may also facilitate the

correction of errors in design and defects in construction. The adhesive bonded structures are

capable of resisting more loads and exhibit slightly increased stiffness [20].

The repaired flat-slabs and columns may sometimes fail below the load-carrying

capacity of the original members, and may also exhibit a reduction in stiffness. However, the

lateral drift capacity of repaired members is nearly equal to that of the original members [17].

Another parameter that affects the strength of a repaired member is the pressure with which

4
epoxy is injected into the member; the member strength is directly proportional to the

injection pressure of the epoxy [21].

Repaired portion Precast concrete slab


(a) Top view

(b) Side view

Fig. 2. Crack repaired using low viscosity adhesive

Systematic studies on the strengthening of RCC structures using adhesives were

started during the 1970s. Various researchers, including MacDonald and Calder [20], Wake

[22], Van Gemert and Maesschalck [23], Tilly [24] and Mays and Vardy [25], carried out

studies on strengthening of precast RCC members. The strengthening of beam elements in

tension, compression and shear zones has recently become a common practice [26-28]. The

primary modes of failure of such strengthened beams are plate end debonding, critical

diagonal cracks and intermediate cracks [29]. However, on beams bonded using polymer, a

premature failure of connections in the tension zone has been observed [28]. Similar studies

also show that externally bonded steel plates at the bottom and side/web faces of precast RCC

members act as additional reinforcement for the member. Fig. 3 shows a schematic view of a

precast RCC beam member strengthened with adhesive bonded external steel plates. Fig. 3(a)

5
depicts the cross-sectional and side views of a shear strengthened precast RCC beam

(structural adhesive is used to connect the steel plates and a precast RCC beam member). Fig.

3 (b) shows a flexurally strengthened precast RCC member, wherein a steel plate is provided

at the bottom face of the precast RCC member, along the length of the beam.

RCC beam Reinforcement bar


Structural adhesive RCC beam
Steel plate Steel plate

Cross-sectional view Side view

(a). Cross-sectional and side view of shear strengthened RCC member using steel plates

Reinforcement bar
RCC beam
Structural adhesive layer
Steel plate
(b). Side view of flexural strengthened RCC member

Fig. 3. RCC beam member strengthened with adhesively bonded external steel plates

6
The effect of strengthening on the flexural strength, shear strength and stiffness of a

member has been thoroughly investigated by various researchers [20, 26, 28, 30-34]. The

strengthened members exhibit a higher enhancement (as high as three times) in stiffness,

compared to the original member. However, this enhanced stiffness is elastic, and the failure

of a member strengthened using adhesive is brittle in nature [10, 16, 21, 35]. Also, it has been

reported that the flexural strength of members strengthened using epoxy resin is higher than

that for the control specimen [10, 19].

The effect of strengthening on precast beams has been investigated by Ali et al. [28].

It has been observed that the shear strength at the compression face is notably higher than that

at the tension face, while the strengthening of the compression face is independent of the

thickness of the steel plate. It has also been observed that, due to the enhanced flexibility of

the strengthened beam, the shear strength of the retrofitted beam varies inversely with the

modulus of elasticity of the strengthening material [27]. Adhikary et al. [30] conducted an

experimental study on precast RCC beams, strengthened using bonded steel plate and found

that the shear strength of web strengthened with thicker plates is higher than that of

unstrengthened (control) web.

Adhikary and Mutsuyoshi [32] observed that the flexural strengthening with thick

plate increases the probability of debonding failure at the beam ends. The use of deeper steel

plates instead of thicker plates has been recommended in order to facilitate the maximum

shear contribution for identical cross-sections [36].

Another experimental study conducted using externally bonded steel plates on precast

RCC beams validates the efficiency of such configurations in achieving a high degree of

crack control. Crushing failure happens in such beams [30]. Along with an increase in the

thickness and width of the web strengthening plates, the ultimate shear strength of the beams

increases. The effectiveness of web strengthening, in terms of the increased shear capacity of

7
rectangular and T cross-sectioned RCC beams, is validated through the close compliance of

experimental results with BS 8110 [33, 34, 37].

The flexural strengthening of precast RCC beams has also been thoroughly

investigated. Changes in the failure modes of such beams, with the thickness of the

strengthening plates, the position of plate curtailment and support conditions, have been

reported by MacDonald and Calder [20]. They recommend the use of bond plates, having a

minimum width to thickness ratio of sixty, to achieve maximum flexural capacity of a

member. The durability of epoxy adhesives through a comparative experimental study has

been reported by Ekenel and Myers [35]. Two sets of similar specimens showing cracks were

strengthened using epoxy; one set was exposed to environmental conditions, while the other

was monitored in the controlled atmosphere of a laboratory. The results of the study

suggested that the environmental exposure led to about a 15% reduction in strength of the

retrofitted specimens.

The behaviour of adhesive bonded anchor rods embedded in concrete has been

investigated by numerous researchers [38-44]. The effect of epoxy-coated and uncoated bars,

embedded axially in concrete cylinders and exposed to a marine environment, on the bond

strength, was investigated through pull-out tests by El-Hawary [38]. The results of the study

suggest that there is no effect of the epoxy coat on the behaviour of the bond. Another similar

study on steel rods embedded in low strength concrete was performed by Yilmaz et al. [39],

in which the effects of the concrete strength, embedded length, diameter of the anchor bar

and edge distance were analysed. The authors concluded that, with an increase in diameter of

the anchor bars, the failure mode of connection changes from ductile to brittle. The study also

concluded that both the embedment depth and edge distance should be at least fifteen times

the anchor bar diameter for an efficient and economic design. Similar tests were conducted

by Barnaf et al. [40] to determine the bond strength of chemically bonded anchors in high

8
strength concrete. It was noted that the behaviour and failure of the specimen depends on the

characteristics of the adhesive. The changes in the failure pattern, with respect to the diameter

of the steel rod, in a pull-out test were studied by Wang et al. [41]. The study indicated that

with an increase in the diameter of the anchor bar, the failure mode changes from steel bar

pull-out to mixed cone damage of the concrete with pull-out of the rod. The failure mode of

connection also depends on the surface treatment of the anchor bar. Wang et al. [42]

indicated that the grooving in the anchor bar increases the bonded area and thereby improves

the mechanical interlocking force. Upadhyaya and Kumar [43] proposed an analytical method

to predict the pull-out capacity of adhesive bonded anchors in concrete, on the basis of stress

intensity. The strength of the connections was found to be independent of the boundary

conditions of the embedded anchors.

Epackachi et al. [45] investigated the tensile and shear behaviour of post installed

anchors in high strength concrete, using the pull-out tests. It was observed that for a single

anchor subjected to tensile force, fracture of the steel rod is the primary mode of failure,

while for a group of anchors, the failure mode was dominated by concrete core splitting. The

fracture of steel anchors was observed to be the mode of failure in a specimen subjected to

shear forces. The authors concluded that the spacing of anchors has a significant effect on the

tensile behaviour of the specimen. Mahrenholtz and Eligehausen [46] reported a similar

behaviour of anchors in shear and tensile loading for normal strength concrete. Ashour and

Alqedra [47] and Sakla and Ashour [48] observed that the pull-off strength of cast in concrete

and post installed adhesive bonded anchors, under tension, in concrete is proportional to the

anchor diameter, embedded length, concrete strength and type of resin. A similar study

carried out by Alqedra and Ashour [49] noted that the strength of an adhesive bonded anchor

in shear is not significantly influenced by the anchor diameter, embedment length and

concrete strength.

9
The use of structural adhesives for repair and strengthening of load-bearing joints in

timber structures is a quite common, efficient and cost-effective practice [50-55]. Wheeler

and Hutchinson [50] studied the effect of moisture content on the bond strength of single lap

joints bonded using epoxy resin and polyurethane. It was observed that, for joints bonded

using epoxy resins, the bond strength remains unaffected up to a moisture content of 22%,

while for joints with polyurethane, the bond strength varies significantly with moisture

content higher than 10%. Broughton and Hutchinson [52] conducted pull-off strength tests on

epoxy and polyurethane bonded joints in timber members with varying moisture contents.

The strength of the epoxy bonded connections was observed to be higher than that of the

polyurethane bonded connections, and was minimally affected by the moisture content at the

time of bonding and post curing. A review study presented by Custódio et al. [53] suggests

that a single test method is not sufficient to obtain all the necessary information related to

bonding. It has been suggested that, to investigate the instant bond performance of a

connection, tests with peel and cleavage loading should be performed, while to investigate

the long-term performance/service life of connections, tests with shear loading should be

carried out. The behaviour of epoxy bonded rebars under pull-out tests, with variation in

rebar diameter, the bonded anchor length and the adhesive bond layer thickness in wood

specimens, was studied by Cimadevila et al. [54] and Ling et al. [55] under compressive and

tensile loading respectively. Cimadevila et al. [54] concluded that the experimental values of

bond strength under axial compressive load differ significantly from Eurocode 5, while Ling

et al. [55] reported the same behaviour for axial tensile loading.

2.2 Structural adhesive as connector at steel–concrete composite interface

The adhesive connection at the steel–concrete interface is still an evolving concept for

the structural engineering/construction industry. The connection at the interface of two

members is capable of effectively resisting and transferring the stress/force between the

10
members with adequate rigidity. In structural engineering applications, adhesives primarily

act as shear connections in beams, floor systems, bridge deck slabs and beam–column

connections. The composite behaviour of bonded connections in steel–concrete composite

members has been studied by many researchers [11, 12, 56-62].

Generally, for connection purposes, two types of adhesives are used as connection

mediums, namely, epoxy resin based adhesives and polyurethane based adhesives. The type

of adhesives used for composite connections determines the behaviour of such members. The

epoxy adhesives show brittle behaviour with perfect interaction (full interaction), while

polyurethane adhesives exhibit flexible (ductile) behaviour with partial interaction [11, 57].

The interaction states between the two components of adhesive bonded steel–concrete

composite connections are shown in Fig. 4. The degree of interaction is decided by the

stiffness and strength of the connection; an increase in the connection strength might increase

the stiffness of the connection [63]. The stiffness of bonded connections is reported to be five

to fifteen times higher than that of mechanical connections [57, 64].

Souici et al. [12] compared the behaviour of adhesive bonded and mechanical stud

connected steel–concrete composite beams, and found that the adhesively bonded connection

facilitates a continuous transfer of shear force between the steel and concrete, whereas a

perfect connection is not ensured at the interface by mechanical stud connectors. Another

study by Jurkiewiez et al. [61] suggested that the composite behaviour in both mechanically

connected and adhesive bonded connections is almost the same. It has also been observed

that the primary mode of failure of adhesive bonded composite flexural members is crushing

of the concrete instead of failure of the connections [11].

11
Concrete slab
Structural adhesive
Steel section

Full No Partial
interaction interaction interaction

Fig. 4. Full, partial and no interaction condition at interface of steel–concrete composite


cross-section

The variation in shear strength along the longitudinal span of a steel–concrete

composite beam bonded with adhesive is almost insignificant, with a marginally higher

concentration at discontinuous edges. However, the peeling stress is highly concentrated at

the connection edges [56]. Zhao and Li [58] developed a numerical model for an adhesive

bonded composite beam, tested by Bouazaoui et al. [11]. The composite behaviour of the 3.5

m long beam was observed to be similar for both numerical (FE) simulations and the

experimental study. Finite element simulations and the nonlinear beam model developed by

Zhao and Li [58] corroborated the observed experimental behaviour of the composite beam

[11, 58].

The capacity of steel–concrete composite beams primarily depends upon their shear

bond strength. The push-out test (compressive shear test) is a double lap shear test performed

to measure the bond shear strength of connections [65-67]. Ernst et al. [68] demonstrated that

the behaviour of push-out test specimens exhibits a very similar behaviour to the full-scale

composite beams. The push-out tests are able to predict the performance of shear connections

accurately. The shear strength, effective bond length, failure mode and the force transfer

mechanism of steel–concrete composite beams can be efficiently predicted through push-out

tests. A detailed representative geometry of a test specimen and connection at the steel–

concrete interface as per Eurocode 4 [65] is shown in Fig. 5.

12
The strength of a steel–concrete composite specimen depends significantly on the

bonded area geometry of the connection, whereas the specimen geometry has a negligible

effect on its bond strength [60]. Berthet et al. [60] developed an analytical model for push-out

tests, which was capable of predicting the strain variation at the bonded interface of a

composite specimen. It can be clearly established that the stress variation profile always

depends on the bond length and magnitude of load, in the direction of loading.

The use of structural adhesives as shear connectors in a steel–concrete composite

member interface has been discussed in Table 1. Scientific research on bonded interfacial

connections started in the 2000s. Only a few studies were conducted on structural adhesives

as the shear connection between the steel–concrete composite interfaces. The effects of

different parameters such as the bond layer thickness of the adhesive, concrete strength, the

surface treatment of the adherend, shear stress and strain variation along the longitudinal span

and cross-sections have been studied.

Structural adhesive
Steel section

Concrete slab
Structural adhesive
Steel section

(a) (b)

Fig. 5. Compressive shear steel–concrete composite push-out specimen bonded with


adhesive: (a) plan view; (b) elevation

13
Table 1. Summary of studies carried out on structural adhesives as shear connector at the
steel–concrete composite interface
Reference Structural Filler⃰ Nature of Mechanical Parameters/variables
adhesive studies properties studied
considered
Bouazaoui Epoxy resin, No Three-point Ultimate load Adhesive nature,
et al. [11] polyurethane bending test capacity, irregular thickness in
on composite deflection, longitudinal and
beam strain, slip, transverse direction,
neutral axis cross-section strain
position variation
Larbi et al. Epoxy resin, Silica Experimental Ultimate bond Surface treatment,
[57] polyurethane push-out test, strength, adhesive bond
numerical shear stress, thickness, shear
[FE] and peeling stress, connector spacing
analytical tangent effect
study on stiffness,
composite deflection,
beam slip, stress
Zhao and Li Epoxy resin No Numerical Ultimate load, --
[58] study on longitudinal
composite shear strain
beam and stress
variation
along the
longitudinal
and transverse
axis, neutral
axis, crack
position and
propagation,
principal
stresses,
collapse
position
Bouazaoui Epoxy resin Silica Three-point Ultimate load, Deviation in results of
et al. [56] bending test strain simple beam,
on beam, distribution, improved beam, finite
analytical deflection, element model and
modelling slip variation, experimental results
[beam and neutral axis
slipping position,
model] shear stress,
peeling stress
Bouazaoui Epoxy resin No Pull-out test, Ultimate Diameter of
and Li [69] theoretical capacity, reinforcement bar,
model shear stress, embedded length of
crack bar, area of embedded

14
initiation portion
stress
Aboobucker Epoxy resin No Push-out test Direct shear Adhesive application
et al. [59] bond strength time, demoulding
time, type and weight
of concrete, type of
epoxy, water content,
super plasticizer dose,
surface preparation of
steel, pressure during
curing
Berthet et Epoxy resin No Push-out test, Ultimate load Connection type, test
al. [60] analytical of bond, shear specimen geometry,
study [beam strain and surface preparation
model] stress
variation with
loading
height.
Jurkiewiez Epoxy resin Silica Push-out test, Ultimate bond Concrete strength,
et al. [61] three- point strength, surface preparation,
bending test, stress-slip section geometry
analytical behaviour,
study [beam beam test:
model and FE ultimate load,
model] deflection,
maximum
strain and
strain
variation in
cross-
sections, load-
deflection for
beam
Luo et al. Epoxy resin, No Push-out test, Shear bond Surface preparation,
[62] polyurethane numerical strength, load- adhesive nature,
[FE] analysis slip adhesive layer
of beam behaviour, thickness, concrete
ultimate load strength, bonding
capacity and area, bonding
deflection of strength, elastic
beam modulus
Souici et al. Mechanical No Four-point Failure mode, Degree of interaction,
[12] shear stud, bending test ultimate load nature of connection
epoxy resin capacity,
maximum
deflection,
strain
distribution,
slip, neutral

15
axis position.
Meaud et al. Epoxy resin Silica Push-out test, Shear stress, Specimen size,
[64] numerical peeling stress, connection geometry,
modelling, strain influence of bonded
FE analysis distribution, area, bottom edge or
average base friction effect
ultimate shear
stress.
Jurkiewiez Epoxy resin Silica Three-point Ultimate load, Concrete strength,
et al. [70] bending test, mid-span cross- section profile
multi-layer deflection, and area of concrete
beam theory, shear stress, and steel,
numerical strain reinforcement
[FE] analysis variation
along the
cross-section
⃰fillers are added at the time of adhesive application.

3. Factors affecting the performance of epoxy adhesive bonded connections

The bond strength of an adhesive bonded connection is usually governed by the

connecting interface. A sound interfacial connection between the two composite materials is

an essential link. The adhesion at the interface between two dissimilar materials, and the

cohesion between the adhesive molecules, is the key to achieve adequate bond strength for a

structural adhesive. These governing factors are in turn dependent on several other

parameters. Numerous studies have been conducted on the factors influencing the bond

strength and these still need further exploration.

The bond strength is likely to depend on various factors such as the type and surface

preparation of adherends, the type and composition of adhesive, concrete placing time, bond

line position, aspect ratio of bonded area (bond geometry), thickness of adhesive layer, shape

of bearing end, geometry of adherends, moisture, freeze-thaw, cyclic loading effects,

weathering actions, elevated temperatures, fillers, and the mechanical and chemical

properties of adhesive and adherends.

3.1 Effect of thickness of connection

16
The bond thickness significantly affects the strength and failure mechanism of a

connection. The effect of adhesive thickness is more dominant when the bond length is

relatively shorter, the thickness of the adherend is high and the nature of the adhesive is

brittle [71]. The traction deformation relation does not depend on the thickness of adherends

[72]. There is no unequivocal opinion on the effect of the adhesive layer thickness on the

strength of connections. Some researchers [73, 74] reported that an increment in thickness

causes an enhancement in strength, while some suggested that an increase in thickness

decreases the strength [15, 75-77]. Colak [76] studied the bond strength of steel–concrete

composite connections subjected to pull-out tests and found that the connections’ bond

strength increased up to a bond thickness of 2 mm and then started to decrease beyond it.

DaSilva et al. [74] reported that a thin layer of bonded adhesive fails at relatively

lower load levels, owing to the high concentration of shear stresses. Thus, the ultimate state

of strain at the bond level is attained before stress dispersion can initiate. The shear and

peeling stresses in the bonded adhesive at the edges of the bonded joint are also affected by

the adhesive thickness itself [78]. At the same load level, higher shear stresses were observed

for lower bond thicknesses [73].

The thickness of the adhesive layer also affects the modes of failure [79]. The

adhesive thickness in lower strength concrete does not affect the bond strength significantly,

while, in higher strength concrete, it leads to a drastic change in the failure mode [80]. The

filler particle size also affects the thixotropy of adhesives [25]. An increase in the thickness of

the adhesive layer in a bonded assembly generally decreases the connection efficiency.

Therefore, joints with a less thick adhesive layer show a positive influence on bond strength

[81, 82], along with an increase in rigidity [61]. The compliance (flexibility) also increases

proportionally with increase in the adhesive bond thickness [83].

17
Martiny et al. [84] reported that with an increase in thickness of the adhesive layers

the fracture energy of a bond increases. Cooper et al. [85] reported that the fracture energy of

a connection increases up to a certain value of bond layer thickness and attains a constant

value beyond that. The authors also stated that the maximum adhesive fracture energy is

obtained when the thickness of the bonded layer (ha) and the plastic zone diameter are almost

equal (2ry).

1 E G IC bulk 
h a  2r y  Eq. (1)
  y 
2

where ha is the thickness of the adhesive; ry is the first order plastic zone size under plane

stress conditions; E is the adhesive’s Young’s modulus; GIC(bulk) is the Mode I fracture energy

of the bulk specimen; σy is the uniaxial yield strength of the adhesive.

Carlberger and Stigh [86] stated that the fracture energy in peel (Mode I) increases

with an increase in adhesive thickness up to a certain level beyond which the peel (Mode I)

fracture energy decreases. The same pattern appears in shear (Mode II) fracture energy.

Mode I Mode II

Fig. 6. Representation of Mode I and Mode II subjected specimen

Fig. 6 represents a schematic view of specimens subjected to simplified Mode I and

Mode II loading. Simplified equations to determine the connection fracture energies in tensile

and shear mode have also been proposed. The Mode I and Mode II fracture energies are

4 P  3a 2 1 
G IC     Eq.(2)
EB 2  h3 h 
and

 v2 h a
G IIC  Eq.(3)
2G a

18
where GIC and GIIC are the fracture energy in Mode I and Mode II respectively. P is the

maximum load; B is the width of the adherend; a is the opening distance; h is the height of

the adherends; τv is shear stress; ha is the thickness of the adhesive layer; Ga is the shear

modulus of the adhesive.

Zhao and Zhang [87] reported that the crack development pattern also changes

significantly beyond a certain (1 mm) adhesive layer thickness in metallic connections. The

failure mode changes from cohesive mode to slant mode and the failure pattern varies with

the changing thickness of the connections.

3.2 Effect of bond geometry

The bond geometry of the connection area and the adherend geometry significantly affect the

bond strength of connections. These aspects play a key role in defining the performance of

the connection. Custódio et al. [53] and Kang et al. [88] chronicle the research on joint

selection and its effect on connection performance. The authors discussed the effect of the

joint geometry on the stress intensity at the bond line, the type of stresses induced, the

effectiveness of the connection and the fracture energy of the connections. Kang and Howell

[88] described the various models available to determine the bond length and strength of a

connection. It has also been stated that the bond performance depends on the joint

configuration and bond length of the connection. Single and double lap shear tests are

common, and suffice for assessing the effectiveness of a connection. ASTM D1002-98 [89]

proposes a simplified equation to determine the maximum permissible bond length of

connection for a single shear test:

Ft y t
L Eq.4

and for a double shear test:

19
Ft y t
L Eq.5
2

where F t y is the yield strength of the adherend; t is the thickness of the adherend; τ is the

average shear stress.

Shear stress distribution

Fig. 7. Double shear test assembly with shear stress distribution

If the bond length and bond area of the connection are sufficient to ensure the

effective distribution and transfer of stresses in the adhesive layer, the connection

performance then depends on factors such as the type of adhesive and adherends, surface

preparation and environmental factors. Yao et al. [90] reported that the bond length of a

connection is a function of the corresponding load level. Any increment in bond length

beyond a certain value does not affect the ultimate strength of the connection. Bizindavyi and

Neale [91] observed that, at an initial load level (up to the cracking load), the strain decreases

exponentially along the bond length up to a certain length, and thereafter it follows a linear

path in the direction of load application. Volnyy and Pantelides [92] concluded that, after

crack initiation, the stress region is transferred towards the unloaded end. The distribution of

shear stress depends on the mechanical properties of the adhesive and follows a triangular

path. The shear strength of the adhesive layer also affects the bond length. If the adherend is

stronger than the adhesive, the connection may fail due to shear strain as it reaches its

ultimate value [93].

Branco et al. [94] observed that the strength of bonding is not only a function of the

bonded area, but is also dependent on the width of the bonded area. For the same bonded area

20
the bond strength increases with an increase in bond width. For high strength concrete, the

bond width has an even higher influence, compared to that for normal strength concrete. The

normal stress concentration along the concrete edge line in an adhesive bonded specimen is

very high. At elevated temperatures, the effect of the bond geometry on the shear strength of

the connection is negligible. A bond width to length ratio of 2.4 or 0.4 offers the maximum

shear strength. If specimens were bonded leaving a certain distance from the concrete edge, a

significant reduction in normal stress concentration was observed, with a sharp increase in

peeling stress [95].

Bizindavyi and Neale [91] demonstrated that the three-dimensional mode of failure

would be prevented by rounding off the bearing section of the concrete block and by starting

the bond line away from the bearing section. Coronado and Lopez [96] concluded that prisms

are more sensitive to the bond width to specimen size ratio, compared to cylinders. The width

of the bearing strip affects the tensile strength of the connection after a width (of bearing

strip) to specimen size ratio of 0.1. However, for this study, the width of the specimen was

kept constant in order to ignore the effect of crack patterns. The specimen size to bearing

strip ratio was varied from 4% to 20%, and the formation of primary cracks was observed up

to a limit of 10%, and thereafter both primary and secondary cracks were observed. It was

advised that, if the specimens are subjected to pure tensile stresses, this ratio must remain less

than or equal to 0.1.

3.3 Effect of surface preparation

The purpose of surface preparation is to remove the weak and unclean layer from the

adherend’s surfaces. The surface preparation of the adherend or surface pre-treatment before

the application of adhesive improves the connection/ bond performance. The surface

roughness depends on the strength of the material, the cleaning method of the interfacial

surfaces, the order and magnitude of cleaning and the size of the elements [97]. The

21
roughness of the concrete substrate has a significant influence on the bond strength of the

interface [98]. Several techniques can be used for surface preparations, such as sand blasting,

mechanical grinding, wire brushing, water jetting, jack hammering and chemical cleaning.

The technique of hammering damages the interfacial surface and leads to the development of

micro-cracks. These micro-cracks reduce the interfacial connection bond strength [99, 100].

Horgnies et al. [98] reported that the mould material used for high strength concrete casting

affects the bond strength, due to a change in the associated surface energy. During

demoulding, the surface of the concrete sticks out from the interface and increases its

porosity. This may result in an increase in the roughness of the connection and also in the

fracture energy of the interface. The failure bond strength is a function of the roughening

material, concrete strength, texture depth, shape factor and the plot ratio of the adherend

[101].

Tensile strength tests on butt jointed steel to steel were performed by Mays and Vardy

[25] for many adhesives. The surface preparation was found to have a significant effect (more

than 150%) on the tensile strength of some adhesives. It was concluded that the strength of an

interfacial connection depends more on the surface energy of the adherend than on the

surface energy of the adhesive. Surface preparation done using grit-blasting with large

diameter aggregates gave rougher surfaces, which in turn affected the bonding, due to the

increased bond area. Fernando et al. [102] could not observe any significant effect of the

chemical composition of the bonded surface on the bond strength. Ekenel and Myers [35]

conducted tests and reported that surface roughness combined with crack injections

significantly increased the flexural capacity of specimens and reduced the crack width

openings, as compared to other CFRP-strengthened specimens.

Julio et al. [100, 103] used five roughness techniques, namely, casting against steel

formwork, wire brushing, partially chipped, partially chipped and pre-wetted, and sand

22
blasting on the concrete interface. The sand blasting technique gave the highest strength in

shear and tension. Benzarti et al. [104] reported that for FRP–concrete bonds, the

sandblasting gives greater bond strength than hand grinding. Santos and Julio [105] used

three surface techniques, namely, casting in steel against steel formwork, wire brushing and

sand blasting, to analyse the effect of surface roughness on bond strength in a specimen

subjected to shear and tension. It was observed that the bond strength is a function of surface

roughness. These aforementioned surface techniques were employed to connect hard to hard

or hard to fresh concrete. It was observed that the bond strength in shear loading, at the

bonding interface of concrete, is significantly affected by substrate moisture and roughness.

The surface roughness changes the failure modes from adhesive to cohesive or mixed.

According to Santos et al. [106], the bond strength of a composite specimen is

influenced by the surface roughness and moisture content of the substrate. In the case of the

specimen consisting of dry hardened substrate and fresh concrete, the bond strength is

significantly influenced by the surface preparation/roughness of the hardened substrate.

However, in the case of a saturated hardened substrate with fresh concrete, the bond strength

decreases with an increase in surface moisture content. They also observed that the specimen

with a saturated substrate attained lower bond strength, compared to the composite specimen

with dry substrate. At the same shear stress level, the higher roughness (peak to valley height)

stores more load [107].

3.4 Effect of aggregate size

Only a few studies are available concerning the effect of aggregate size on adhesion.

Mays and Vardy [25] studied the aggregate size effect and concluded that, when the concrete

slab was cast over fresh-lain epoxy resin, then the coarse aggregates (of larger size and

angular shape) were observed to be puncturing the adhesive layer.

23
3.5 Effect of filler

Fillers are added to adhesives to improve their physical and mechanical properties.

The type and properties of the filler used play a key role in determining the bond strength of

the adhesive bonded connections. The addition of the filler leads to an improved adhesion

between the particle matrix of adhesives and also ensures an effective transfer of stresses

between adhesive and filler. Tüzün and Tunalıoğlu [108] concluded that the strength of the

bond is found to be higher in small diameter filler. Nano silica fillers, when added to epoxy,

lead to a significant increase in the durability of epoxy (unadulterated adhesive). This mixture

also enhances the bond strength of the metallic connection, when subjected to quasi-static and

cyclic loading, by almost 20% [109]. Fillers such as silicon carbide and carbon nanotubes,

when added in a certain proportion, improve the shear bond strength of adhesive layers at

elevated temperatures of up to 700°C. The effect is primarily due to the polymerisation of

silicon carbide at temperatures up to 200°C, which results in the formation of a condensed

matrix that fills the pores on the adhesives surface, leading to an improved whiskers

reinforcing effect [110].

A study conducted by Bowditch [111] explored the effect of filler having high surface

energy. It was noted that when the adhesive was administered with untreated silica filler, the

reduction was significantly less than when the filler material was treated with silane (SiH4).

In a water environment, the failure will occur near the interface due to delaminating from the

resin matrix, which results in the development of the weaker zone. Gao et al. [112] concluded

that a rubber-modified adhesive enhances the structural performance by increasing the

maximum load-carrying capacity and the corresponding ductility. Singla and Chawla [113]

observed that the addition of fly ash to epoxy resin leads to an increment in the compressive

strength. This is because of the hollowness of fly ash particles and a strong adhesion between

the fly ash and resin. However, the enhanced compressive strength was accompanied by a

24
reduction in impact energy. The same percentage of fly ash with a small amount of glass fibre

led to a further indication of a higher increment in compressive as well as impact strength due

to the reinforcing effect of the glass fibre. The experimental investigations conducted by

Colak [76] revealed that when the epoxy adhesive was mixed with up to 46% (by volume) of

quartz sand, the shear strength of the connection increased. However, a further increase in

filler addition shows a decrease in bond strength. According to Wang et al. [110], the addition

of multi-walled carbon nanotubes or short Kevlar fibres improves the bond strength of the

adhesive by enhancing the reinforcing effect. Owing to this reinforcing effect, a delay in

crack initiation in the adhesive layer was observed.

The crack propagation rate for Mode II fracture associated with an end-notched

flexure specimen is very high with a higher percentage of filler [114]. Kong et al. [115]

indicated that adulterated polyblended adhesive showed a better degree of curing in

comparison to the unadulterated adhesive, and has no noticeable effect on the bond strength.

Moussa et al. [116] reported that the filler added (adulterated) adhesive achieved a higher

glass transition temperature at an early stage of curing. Kahraman et al. [75] concluded that

the addition of filler in structural adhesive, for a metallic joint, reduces the strength of the

bonded connection. Frigione et al. [81] and Colak et al. [117] indicated that the addition of

inorganic fillers in structural adhesive increased the bond strength, durability, glass transition,

temperature and bond performance for concrete-to-concrete connections.

3.6 Effect of water immersion

An aggressive environment may severely harm the durability of a bonded connection.

The strength of connections is adversely affected if the specimen is immersed in water up to

the bond level. The durability and strength of a connection immersed in water may depend on

the nature of the adhesive, temperature, amount and size of filler, time and bonded area to

edge line ratio.

25
The bond strength reduction due to water immersion is higher in a filler added

adhesive, because the addition of filler allows water seepage inside the adhesive layer,

thereby leading to the formation of cavities. It causes rapid plasticization of the adhesive. An

increase in temperature during water immersion leads to changes in the failure pattern from

adhesion to cohesion or mixed failure [95, 118]. The adhesive oxide used at the interface

plays a major role in assuring the durability of the connection. The hydration of adhesive

oxides and plasticization of the adhesive leads to bond failure. The durability of water-

immersed adhesive connections also depends on the chemical composition of the adhesive

[25, 111, 119-122].

Mays and Vardy [25] studied the effect of six different adhesives on connection

strength when immersed in water. Among all the adhesives, the aromatic polyamine type

hardener adhesive showed sufficient resistance against dampness. It was also concluded that

the chemical composition of the adhesive affects the durability of the connection. Gasparini

et al. [123] reported that the conventional sealant allows the diffusion of water through the

bond line, and metallizing the bond line area is preferred over the use of conventional filler to

increase the bond resistance. The durability studies conducted on modified acrylic and epoxy

adhesive indicated that modified acrylic adhesives show an insignificant change in strength,

whereas epoxy adhesives were found to be less durable.

The diffusion of water depends on interfacial properties, as well as depending on the

bond area to edge line ratio and diffusion coefficient of the adhesive [119]. The diffusion of

water in the specimen with steel-bonded adhesive is higher than the specimen with bulk

polymer adhesives due to seepage of the aggressive medium near the metal–polymer

interface [124]. The bond strength of a filler added adhesive in a water immersed specimen

depends on the nature and size of the filler. There may be both adverse and favourable effects

on the connection durability depending on the size of the filler particles. In water immersion,

26
treated (coupling agent) fillers have been found to show higher resistance [111]. Separation

from the interface matrix along with pure adhesive failure at the interface has been observed

in the case of filler added adhesive connections, with time. The failure may take place due to

weakening at the interface and the higher surface energy of the filler. A filler modified

adhesive with a thicker adhesive layer shows less degradation in strength [81]. When filler

modified adhesive bonded connections are exposed to moisture, the interfacial fracture

toughness and energy of the bonded connections decrease significantly [125].

Nguyen et al. [126] suggested that, on immersion in sea water, an adhesive bonded

specimen shows a degradation in strength and stiffness. The reduction in strength is of the

order of about 14% at the end of 4 months and a further 3% by 12 months. Meanwhile, the

reduction in stiffness is noted to be 15% in 2 months, followed by another 15% at the end of

4 months and a total of 39% reduction in stiffness was noted at the end of 12 months. The

effect of elevated temperature on a seawater-immersed specimen is a higher rate of

degradation in stiffness and strength.

3.7 Effect of relative humidity

Relative humidity is a vapour pressure and time-dependent phenomenon. The level of

humidity slowly affects the durability of the bonded connection. Datla et al. [127] conducted

a study on a structural adhesive subjected to elevated temperature and relative humidity. They

concluded that at room temperature, the strength of bonded connections is significantly

affected by the relative humidity. However, at elevated temperatures, the crack growth rate is

high, which reduces the effect of relative humidity on the connection. At lower temperatures

and high humidity levels, the moisture diffusion rate at the adhesive interface is high. The

fatigue performance of the bonded connection degrades predominantly owing to excessive

moisture. It has also been reported that 100% relative humidity is more harmful for the

specimen durability than water immersion, even though they are very similar phenomena.

27
However, according to Datla et al. [127] and Mikami et al. [128], the connection strength of

structural adhesive bonded specimens subjected to a combination of higher temperature

(100°C) and lower humidity (0%) degrades severely (the degradation in bond strength is

about 40–50% ), compared to specimens subjected to high temperatures and high humidity.

Nguyen et al. [126] concluded that seawater immersed bonded specimens show more

degradation in strength than a specimen placed in high relative humidity (90%). The

degradation in strength, stiffness and glass transition temperature is significant at high

temperature and humidity levels [129, 130]. Lettieri and Frigione [131] stated that the

physical properties of bulk adhesive such as the glass transition temperature are not affected

by a lower level of humidity (15%). However, the strength of adhesive in a bonded

connection depends on the water immersion time, which in turn is higher in the case of

elevated temperature. Water (moisture) uptake increases at higher humidity levels (above

75%), which leads to plasticization of the adhesive. This plasticization slightly increases the

elastic modulus in the early stage and reduces the ultimate interfacial adhesion [126, 129].

3.8 Glass transition temperature

The strength of adhesives also depends on the surrounding temperature. The type of

adhesive employed determines the effect of variation in temperature on the bond strength.

Usually, the adhesive bond strength decreases with increase of temperature. However, at a

certain elevated temperature called the glass transition temperature, the nature of the adhesive

changes from a glassy state to a viscous/ rubbery state. The glass transition temperature

depends on the amplitude and frequency of loading. The glass transition temperature is a

property of the adhesive material, and generally ranges from 45°C to 120°C for structural

thermosetting adhesives [81, 107, 128, 132-139].

28
Some researchers observed that at temperatures above 65°C, the epoxy resin adhesive

loses its shear bond strength by approximately 90% [140, 141]. Branco et al. [94] observed

that at temperatures above 55°C, the bond strength is only 10% of its strength at 20°C.

Salman [142] reported that the bond strength of epoxy resin reduces considerably, in

comparison to polyester, with an increase in temperature. It has also been suggested that at a

fixed temperature (glass transition temperature), the bond strength of epoxy resin shows a

considerable reduction. Bowditch [111] observed that adhesive molecules are the weakest

link in a connection at elevated temperatures. The degradation of the composite’s behaviour

becomes predominant when the specimens are also exposed to high temperatures for long

durations. Coronado and Lopez [96] found that the temperature increment with time shows a

considerable change in bond strength and also in the fracture energy of the connection.

According to Tadeu and Branco [95], Ahmed and Kodur [135], Palmieri et al. [143]

and Arruda et al. [144], when a connection is subjected to elevated temperatures and the bond

temperature approaches the glass transition temperature, the reduction in mechanical

properties such as strength, deflection, stiffness, strain and stresses is phenomenal. The

mechanical properties of bulk adhesives such as strength and stiffness also depend on the

curing temperature [115, 116, 139].

3.9 Effect of elevated temperature and fire

The effect of elevated temperatures and fire on the concrete–adhesive interface has

been explored by only a few researchers. Adhesive bonded specimens subjected to elevated

temperatures or fire exhibit a steep degradation in bond performance, leading to weakened

bond links, thereby putting the construction at risk [145].

The curing temperature of a specimen affects its strength and stiffness. At low curing

temperatures (about 5°C), initially both the strength and stiffness follow a similar increasing

29
trend; however, later, the gain in strength is slower than that in stiffness. On the other hand, at

higher temperatures, the maximum stiffness gain is slow, compared to the strength, owing to

increased random cross-linking of adhesive molecules. A slight change in temperature does

not degrade the mechanical properties of adhesives significantly [146]. Ferrier et al. [139]

conducted an experiment on concrete–adhesive–FRP bonds and observed that any variation

from room temperature (increase or decrease) leads to a degradation in bond strength.

However, the effect of temperature reduction on bond strength is more pronounced. At low

temperatures, the adhesives become brittle, thereby causing a reduction in the Young’s

modulus of elasticity, leading to a negligible shear strength of connections [147].

Tadeu and Branco [95] conducted experiments to determine the ultimate shear

strength at varying temperatures on three different grades of concrete. At 60°C, a substantial

reduction (about 50%) was observed in bond strength, leading to a cohesion failure in the

adhesive layer. At 120°C, the strength of the bond was much less (10% of the unheated

sample). This has been attributed to the deterioration of the epoxy adhesive and also to the

presence of a free water layer over the concrete. It was also found that high strength concrete

is more sensitive to elevated temperatures than low strength concrete.

Pinoteau et al. [145] carried out pull-out tests (adhesive coated steel rod embedded in

concrete specimens) in fire, and observed that at elevated temperatures, water migrates

toward the inner core of the concrete (towards the steel rod), creating pressure on the internal

core. This pressure leads to an observable layer of water over the adhesive surface, in turn

leading to bond failure at lower strength. The bond strength slip behaviour of epoxy-coated

and uncoated reinforcement bars embedded in concrete under elevated temperatures was

studied by Mohaisen [148]. He concluded that the bond strength of coated reinforced bars

depended on the temperature level. It was also found that at elevated temperatures, for epoxy-

30
coated reinforcement bars, the rate of slip increment was faster, and the residual bond

strength was less, than that for uncoated reinforcement at the same temperature.

Rashid et al. [149] investigated the bond performance of polymer cement mortar laid

over a strengthened reinforced concrete beam at elevated temperatures, under the four-point

bending test. The study was conducted at 20°C, 40°C and 60°C. The authors concluded that

with an increase in temperature, the failure mode changes from flexural-shear to debonding

failure. Plecnik et al. [150] performed the four-point bending test on an epoxy-repaired

concrete beam and observed that the performance of the repaired concrete beam under fire

depends on the type of cracks and also on the extent of repair. When subjected to fire, a beam

having shear cracks, repaired using epoxy resin, sustains damage due to the failure of the

epoxy, whereas, in the case of a beam having flexural cracks (without crushing of concrete in

the compression zone), and repaired using epoxy, the degradation in strength is negligible in

comparison to the degradation in stiffness.

4. Bonding and microstructure

Kumar [151] carried out experimental investigations to understand the physical

bonding of composite interfaces and the feasibility of adhesive-bonded composite

connection. Fig. 8 shows the variation in shear strength with change in the bond layer

thickness. It was observed that the bonding remained inadequate up to a bond layer thickness

of 2 mm, beyond which a perfect bond was obtained. Fig. 9 shows the various stages during

the testing of a steel–concrete composite specimen.

31
12

Shear strength (MPa)


9

0
0 1 2 3 4
Adhesive layer thickness (mm)

Fig. 8. Change in shear strength with bond layer thickness for steel–concrete composite

connection

(a) (b) (c)

32
(d) (e) (f)
Fig. 9. (a) Adhesive mixing process; (b) Laying of adhesive layer over concrete surface; (c)
Bonded steel–concrete interface and bonding mechanism; (d) Compressive shear push-out
test specimen; (e) Interface failure from concrete–epoxy interface (f) Adhesive bonded steel–
concrete composite bridge

The chemical bonding at the concrete, epoxy and composite interface was also

investigated by the authors through Fourier transform infrared spectrometry (FT-IR). The

spectrometry provides an insight on the chemical bonds in the concrete, epoxy and their

interface. The variations in absorbance with theh wave number, obtained by FT-IR, are

shown in Fig. 10. The chemical bonds present in the materials have been distinguished

through intensity peaks present at a certain wave number. Fig. 10 also shows the chemical

bonds present in the epoxy, concrete and epoxy–concrete composite interface with their

respective intensity peaks.

The relative shifts of the intensity peaks provide a valuable understanding of the bonding

behaviour at the concrete–epoxy composite interface. The spectral data, as shown in Fig. 10,

illustrate the presence of –OH bond in the concrete and concrete–epoxy interface samples, at

wave numbers 3602.78 cm-1 and 3484.61cm-1 respectively. A slight decrease in the wave

number of the interface sample can be clearly observed in this figure. This phenomenon can

be interpreted as desolation of –OH bond in the case of the concrete–epoxy interface sample.

A band of –CH2 bond has been observed at wave numbers 2925.38 cm-1 and 2867.14 cm-1 for

the epoxy adhesive and composite interface respectively. The relative shift of −58.24 cm-1

was observed in the case of the concrete–epoxy interface, for the -CH2 band bond. The

shifting of intensity peaks towards lower wave numbers represents an increase in bond

length, which signifies the weaker strength of the bond at the interface.

Wave numbers 1414.50 cm-1, 1430.18 cm-1 and 1435.06 cm-1 represent the peaks for the

concrete (-CO), epoxy and concrete–epoxy composite interface (-CH) respectively. However,

33
the relative shift in the composite interface wave number is not significant. It was concluded

that this particular chemical bond band has a minor influence on the bond strength.

The formation of –SiO and -CH bond bands was observed at wave numbers 990.06 cm-1 and

1081.33 cm-1 for concrete and epoxy respectively. For the composite interface, a bond band

at wave number 1008.95 cm-1 was observed. The chemical bond formed due to the reaction

between concrete and epoxy adhesive is present at wave number 1008.95 cm-1, which leads to

the development of bond strength at the epoxy–concrete composite interface.

In addition to the abovementioned bond bands, the presence of -POC, -COC, and benzene

ring in the epoxy adhesive and –Si-O-Si bond in the concrete element was also observed. The

peaks observed at the composite interface had a similar intensity and wave number to those

observed in the concrete and epoxy adhesive. The comparable characteristics of the peaks

show that the bond bands have some influence on the reaction mechanism, leading to bond

strength at the composite interface.

Interface
0.2
Epoxy
Concrete

-CH
-CO
c-o-c
p-o-c
Absorbance

-CH2 -CHO -CH


0.1

-si-o
-CO
si-o-si
-OH

0.0

4000 3500 3000 2500 2000 1500 1000 500


-1
Wavenumber (cm )

34
Fig. 10. FT-IR spectroscopy of concrete, epoxy and concrete–epoxy interface

Scanning electron microscopy (back-scattered electron technique) was used to further

analyse the concrete–epoxy interface. The interface was analysed at different magnification

levels to efficiently examine the interface bonding. Fig. 11 shows the interface at the

magnification levels of 150×, 1000× and 5000×. The interface layers can be clearly

distinguished as concrete and epoxy at higher magnification levels. The brighter area

observed in the 150× image corresponds to the concrete element, while the dark area

corresponds to the epoxy adhesive layer. Another observation drawn from the image is the

near-perfect bonding (absence of voids and cracks) between the concrete element and the

adhesive layer. At 5000× level, the bonding due to the chemical reaction between the

concrete element and epoxy adhesive layer is clearly visible.

Fig. 11. Back-scattered electron image of concrete–epoxy composite interface

35
5. Conclusion

The published literature shows a broad consensus on structural adhesives. The

exhaustive literature review manifests that the use of structural adhesives in construction has

increased rapidly. The advances in structural adhesives may have promise as replacements

for mechanical connectors in steel–concrete composite construction. The instantaneous and

long-term bond strength of composite connections depends on the physical, mechanical and

chemical properties of the adhesive and adherend, the type and surface preparation of the

adherend, the type and composition of the adhesive, the concrete placing time, bond line

position, aspect ratio of the bonded area, thickness of the adhesive layer, shape of the bearing

end, geometry of the adherend, moisture, freeze-thaw, weathering action, elevated

temperatures and filler. The specific effects of important parameters are given below:

1. Varied observations have been reported throughout the literature about the effect of

the thickness of the adhesive layer on bond strength. The thickness of the adhesive

layer significantly affects the bond strength, efficiency, and failure mode of bonded

composite connections. An increase in bond layer thickness reduces the effectiveness

of connections.

2. The bond geometry of the connected interface and the overall connection type have a

major influence on bond strength. The intensity of stress can be reduced by careful

selection of the position and width of the bond line at the connected interface. A bond

width to length ratio of around to 2.4 or 0.4 leads to a high shear strength. The effect

of bond geometry when the specimens are subjected to high temperature is reported

not to have any effect on the shear strength of the connection.

3. Surface preparation can be done through several techniques like chemical cleaning,

wire brushing, mechanical grinding, water jetting, sand blasting and jack hammering.

Cleaning with water jetting and jack hammering induces microcracks on the adherend

36
surface, which degrade the bond strength of the connections. Sand blasting and

mechanical grinding improve the efficiency of connections to some extent.

4. The performance of bonded connections can be improved through the addition of

fillers. The ratio of filler size to bond layer thickness is a fundamental aspect that

decides the connection durability. The filler addition may exhibit adverse or

favourable effects on the long-term performance of connections.

5. Water immersion and relative humidity affect the durability of connections. The

diffusion of water depends on the bond area to edge line ratio. The interfacial

chemical composition also influences the durability of water-immersed bonded

connections. The addition of fillers may improve the performance of connections

under water and in the presence of relative humidity.

6. The curing temperature of the adhesive significantly affects the connection

performance at room as well as elevated temperatures.

7. Most of the structural adhesives used in the construction industry, cured at room

temperature, cannot withstand temperatures above 150°C.

8. Environmental factors like moisture, water immersion, weathering action and

temperature affect the durability of the connection predominantly, and not the

instantaneous behaviour.

6. Future scope

The durability of composite interfacial connections is still a new area for the structural

engineering community. The instantaneous and long-term aspects of composite connection

are being studied, but a clear statement could not be drawn from the existing studies for

aspects such as bond thickness, bonded area geometry and aggregate size studies. Further and

detailed studies examining the aforementioned characteristics are required for an enhanced

clarity and in-depth understanding.

37
Acknowledgement

The authors express sincere gratitude to The Institution of Engineers (India)

[RD2015006; April, 2015] for having funded and supported this work.

References

[1] BS 5270 Bonding agents for use with gypsum plasters and cement. Specification for
polyvinyl acetate (PVAC) emulsion bonding agents for indoor use with gypsum
building plasters., BS, British Standards Institution London, 1989.
[2] Mays GC. Hutchinson A.R., Adhesives in civil engineering, Cambridge University Press,
New York, 2005.
[3] Täljsten B. The importance of bonding—A historic overview and future possibilities,
Advances in Structural Engineering, 2006; 9(6): 721-736.
[4] Tremper B. Repair of damaged concrete with epoxy resins, Journal Proceedings, 1960, 57
(8): 173-182.
[5] Shaw JDN. Adhesives in the construction industry: Materials and case histories,
Construction and Building Materials, 1990; 4(2): 92-97.
[6] Kalyanasundaram P., Rajeev S., Udayakumar H., REPCON: expert system for building
repairs, Journal of computing in civil engineering, 1990; 4(2): 84-101.
[7] Johnson S.M., Deterioration, maintenance, and repair of structures, McGraw Hill, New
York, 1965.
[8] Chand S. Cracks in building and their remedial measures., Indian Concrete Journal 1979;
268-272.
[9] Shash AA. Repair of concrete beams–a case study, Construction and Building Materials,
2005; 19(1): 75-9.
[10] Ekenel M., Rizzo A., Myers JJ., Nanni A. Flexural fatigue behavior of reinforced
concrete beams strengthened with FRP fabric and precured laminate systems, Journal
of Composites for Construction, ASCE 2006; 10(5): 433-442.
[11] Bouazaoui L., Perrenot G., Delmas Y., Li A. Experimental study of bonded steel
concrete composite structures, Journal of Constructional Steel Research, 2007; 63(9):
1268-1278.
[12] Souici A., Berthet J.F., Li A., Rahal N., Behaviour of both mechanically connected and
bonded steel-concrete composite beams, Engineering Structures, 2013; 49; 11-23.
[13] Wang YH., Nie JG., Li JJ. Study on fatigue property of steel–concrete composite beams
and studs, Journal of Constructional Steel Research, 2014; 94 :1-10.
[14] Ellobody E., Finite element analysis and design of steel and steel–concrete composite
bridges, Butterworth-Heinemann, MA, USA, 2014.
[15] Mette C., Stammen E., Dilger K. Challenges in joining conductive adhesives in
structural application–Effects of tolerances and temperature, International Journal of
Adhesion and Adhesives, 2016; 67: 49-53.
[16] Chung H., Epoxy repair of bond in reinforced concrete members, Journal Proceedings,
1981; 78(1): 79-82.
[17] Pan A.D., Moehle J.P., An experimental study of slab-column connections, Structural
Journal, 1992; 89(6): 626-638.
[18] Robertson I.N., Johnson G., Repair of slab-column connections using epoxy and carbon
fiber reinforced polymer, Journal of Composites for Construction, ASCE 2004; 8(5):
376-383.

38
[19] Thanoon W.A., Jaafar M.S., Kadir M.R.A., Noorzaei J., Repair and structural
performance of initially cracked reinforced concrete slabs, Construction and Building
Materials, 2005; 19(8): 595-603.
[20] MacDonald M., Calder A., Bonded steel plating for strengthening concrete structures,
International Journal of Adhesion and Adhesives, 2 (1982) 119-27.
[21] Farhey DN., Adin MA., Yankelevsky DZ. Repaired RC flat-slab-column
subassemblages under lateral loading, Journal of Structural Engineering, 1995; 121(11):
1710-20.
[22] Wake WC. Theories of adhesion and uses of adhesives: a review, Polymer, 1978, 19(3):
291-308.
[23] Van Gemert D., Maesschalck R. Structural repair of a reinforced concrete plate by epoxy
bonded external reinforcement, International Journal of Cement Composites and
Lightweight Concrete, 1983; 5: 247-55.
[24] Tilly GP. Fatigue of land-based structures, International Journal of Fatigue, 1985; 7(2):
67-78.
[25] Mays GC., Vardy AE. Adhesive-bonded steel/concrete composite construction,
International Journal of Adhesion and Adhesives, 1982; 2(2): 103-7.
[26] Oehlers DJ., Park SM., Ali MM. A structural engineering approach to adhesive bonding
longitudinal plates to RC beams and slabs, Composites Part A: Applied Science and
Manufacturing, 2003; 34(9): 887-97.
[27] Ali MM., Oehlers DJ., Bradford MA. Shear peeling of steel plates adhesively bonded to
the sides of reinforced concrete beams, Proceedings of the Institution of Civil
Engineers-Structures and Buildings, 2000; 140(3): 249-59.
[28] Ali MM., Oehlers DJ., Bradford MA., Debonding of steel plates adhesively bonded to
the compression faces of RC beams, Construction and building materials, 2005; 19(6):
413-422.
[29] Chen JF., Teng JG. Shear capacity of FRP-strengthened RC beams: FRP debonding,
Construction and Building Materials, 2003; 17(1): 27-41.
[30] Adhikary BB., Mutsuyoshi H., Sano M., Shear strengthening of reinforced concrete
beams using steel plates bonded on beam web: experiments and analysis, Construction
and Building materials, 2000; 14(5): 237-244.
[31] Oehlers DJ. Development of design rules for retrofitting by adhesive bonding or bolting
either FRP or steel plates to RC beams or slabs in bridges and buildings, Composites
Part A: applied science and manufacturing, 2001; 32(9): 1345-1355.
[32] Adhikary BB., Mutsuyoshi H. Numerical simulation of steel-plate strengthened concrete
beam by a non-linear finite element method model, Construction and Building
Materials, 2002; 16(5): 291-301.
[33] Barnes RA., Mays GC. Strengthening of reinforced concrete beams in shear by the use
of externally bonded steel plates: Part 1–Experimental programme, Construction and
Building Materials, 2006; 20 (6): 396-402.
[34] Barnes RA., Mays GC. Strengthening of reinforced concrete beams in shear by the use
of externally bonded steel plates: Part 2–Design guidelines, Construction and Building
Materials, 2006; 20(6): 403-411.
[35] Ekenel M., Myers JJ. Durability performance of RC beams strengthened with epoxy
injection and CFRP fabrics, Construction and Building Materials, 2007; 21(6): 1182-
1190.
[36] Adhikary BB., Mutsuyoshi H. Shear strengthening of RC beams with web-bonded
continuous steel plates, Construction and Building Materials, 2006; 21(6): 296-307.
[37] BS 8110. Structural use of concrete, Part 1, Code of practice for design and construction,
British Standards Institution, London, (1997).

39
[38] El-Hawary MM., Evaluation of bond strength of epoxy-coated bars in concrete exposed
to marine environment, Construction and Building Materials,1999; 13(7): 357-62.
[39] Yilmaz S., Özen MA., Yardim Y. Tensile behavior of post-installed chemical anchors
embedded to low strength concrete, Construction and Building Materials, 2013; 47:
861-6.
[40] Barnaf J., Bajer M., Vyhnankova M. Bond strength of chemical anchor in high-strength
concrete, Procedia Engineering, 2012; 40: 38-43.
[41] Wang D., Wu D., He S., Zhou J., Ouyang C. Behavior of post-installed large-diameter
anchors in concrete foundations, Construction and Building Materials, 2015; 95: 124-
32.
[42] Wang D., Wu D., Ouyang C., Zhai M. Performance and design of post-installed large
diameter anchors in concrete, Construction and Building Materials, 2016; 95: 142-50.
[43] Upadhyaya P., Kumar S. Pull-out capacity of adhesive anchors: an analytical solution,
International Journal of Adhesion and Adhesives, 2015; 60: 54-62.
[44] Omar HA., Yusoff NIM., Ceylan H., Sajuri Z., Jakarni FM., Ismail A. Investigation of
the relationship between fluidity and adhesion strength of unmodified and modified
bitumens using the pull-off test method, Construction and Building Materials, 2016;
122: 140-148.
[45] Epackachi S., Esmaili O., Mirghaderi SR., Behbahani AAT. Behavior of adhesive
bonded anchors under tension and shear loads, Journal of Constructional Steel
Research, 2015; 114: 269-280.
[46] Mahrenholtz P., Eligehausen R. Post-installed concrete anchors in nuclear power plants:
Performance and qualification, Nuclear Engineering and Design, 2015; 287: 48-56.
[47] Ashour AF., Alqedra MA. Concrete breakout strength of single anchors in tension using
neural networks, Advances in Engineering Software, 2005; 36: 87-97.
[48] Sakla SS., Ashour AF. Prediction of tensile capacity of single adhesive anchors using
neural networks, Computers & structures, 2005; 83(21): 1792-1803.
[49] Alqedra MA., Ashour AF. Prediction of shear capacity of single anchors located near a
concrete edge using neural networks, Computers & Structures, 2005; 83(28): 2495-
2502.
[50] Wheeler AS., Hutchinson AR. Resin repairs to timber structures, International journal of
adhesion and adhesives, 1998 18(1): 1-13.
[51] Davis G. The performance of adhesive systems for structural timbers, International
Journal of Adhesion and Adhesives, 1997; 17(3): 247-255.
[52] Broughton JG., Hutchinson AR. Adhesive systems for structural connections in timber,
International journal of adhesion and adhesives, 2001; 21(3): 177-186.
[53] Custódio J., Broughton J., Cruz H. A review of factors influencing the durability of
structural bonded timber joints, International journal of adhesion and adhesives, 2009;
29(2): 173-185.
[54] Cimadevila JE., Rodríguez JV., Chans MO. Experimental behaviour of threaded steel
rods glued into high-density hardwood, International journal of adhesion and adhesives,
2007; 27(2): 136-44.
[55] Ling Z., Yang H., Liu W., Lu W., Zhou D., Wang L. Pull-out strength and bond
behaviour of axially loaded rebar glued-in glulam, Construction and Building
Materials, 2014; 65 440-9.
[56] Bouazaoui L., Jurkiewiez B., Delmas Y., Li A. Static behaviour of a full-scale steel–
concrete beam with epoxy-bonding connection, Engineering Structures, 2008; 30(7):
1981-1990.
[57] Larbi AS., Ferrier E., Jurkiewiez B., Hamelin P. Static behaviour of steel concrete beam
connected by bonding, Engineering structures, 2007; 29(6): 1034-1042.

40
[58] Zhao G., Li A., Numerical study of a bonded steel and concrete composite beam,
Computers & Structures, 2008; 86(19): 1830-8.
[59] Aboobucker MAM., Wang TY., Liew JR. An experimental investigation on shear bond
strength between steel and fresh cast concrete using epoxy, The IES Journal Part A:
Civil & Structural Engineering, 2009; 2(2): 107-115.
[60] Berthet JF., Yurtdas I., Delmas Y., Li A. Evaluation of the adhesion resistance between
steel and concrete by push out test, International Journal of Adhesion and Adhesives,
2011; 31(2): 75-83.
[61] Jurkiewiez B., Meaud C., Michel L. Non linear behaviour of steel–concrete epoxy
bonded composite beams, Journal of Constructional Steel Research, 2011; 67 (3): 389-
397.
[62] Luo Y., Li A., Kang Z. Parametric study of bonded steel–concrete composite beams by
using finite element analysis, Engineering Structures, 2012; 34: 40-51.
[63] Oehlers DJ., Bradford MA., Elementary behaviour of composite steel and concrete
structural members, Elsevier, MA, 1999.
[64] Meaud C., Jurkiewiez B., Ferrier E., Steel–concrete bonding connection: An
experimental study and non-linear finite element analysis, International Journal of
Adhesion and Adhesives, 2014; 54: 131-142.
[65] EC4, Eurocode 4: Design of composite steel and concrete structures. Part 1-1: General
rules and rules for buildings., EN 1994-1-1, Brussels, Belgium., 2004.
[66] BS 5950, Code of practice for design of simple and continuous composite beams. Part 3:
Section 3.1 Structural use of steelwork in building. London., 1990.
[67] IS 11384, Code of practice for composite construction in structural steel and concrete,
Bureau of Indian Standards, New Delhi, India., 1985.
[68] Ernst S., Bridge RQ., Wheeler A. Correlation of beam tests with pushout tests in steel-
concrete composite beams, Journal of structural engineering, 2010; 136(2): 183-192.
[69] Bouazaoui L., Li A. Analysis of steel/concrete interfacial shear stress by means of pull
out test, International Journal of Adhesion and Adhesives, 2008; 28 (3): 101-108.
[70] Jurkiewiez B., Meaud C., Ferrier E., Non-linear models for steel–concrete epoxy-bonded
beams, Journal of Constructional Steel Research, 2014; 100: 108-121.
[71] Da Silva LFM., das Neves PJ., Adams R., Spelt J. Analytical models of adhesively
bonded joints—Part I: Literature survey, International Journal of Adhesion and
Adhesives, 2009; 29: 319-330.
[72] Leffler K., Alfredsson K., Stigh U. Shear behaviour of adhesive layers, International
Journal of Solids and Structures, 2007; 44(2): 530-545.
[73] Derewonko A., Godzimirski J., Kosiuczenko K., Niezgoda T., Kiczko A. Strength
assessment of adhesive-bonded joints, Computational Materials Science, 2008; 43(1):
157-164.
[74] Da Silva LFM., das Neves PC., Adams RD., Wang A., Spelt JK. Analytical models of
adhesively bonded joints-Part II: Comparative study, International Journal of Adhesion
and Adhesives, 2009b; 29(3): 331-341.
[75] Kahraman R., Sunar M., Yilbas B., Influence of adhesive thickness and filler content on
the mechanical performance of aluminum single-lap joints bonded with aluminum
powder filled epoxy adhesive, Journal of Material Processing Techechnology, 2008;
205(1): 183-189.
[76] Colak A. Parametric study of factors affecting the pull-out strength of steel rods bonded
into precast concrete panels, International Journal of Adhesion and Adhesives, 2001;
21(6): 487-493.

41
[77] Çolak A., Çoşgun T., Bakırcı AE. Effects of environmental factors on the adhesion and
durability characteristics of epoxy-bonded concrete prisms, Construction and Building
materials, 2009; 23(2): 758-767.
[78] Schnerch D., Dawood M., Rizkalla S., Sumner E. Proposed design guidelines for
strengthening of steel bridges with FRP materials, Construction and building materials,
2007; 21(5):1001-1010.
[79] Buyukozturk O., Hearing B. Failure behavior of precracked concrete beams retrofitted
with FRP, Journal of composites for construction, 1998; 2(3): 138-144.
[80] López-González JC., Fernández-Gómez J., González-Valle E. Effect of adhesive
thickness and concrete strength on FRP-concrete bonds, Journal of Composites for
Construction, 2012; 16(6): 705-711.
[81] Frigione M., Aiello MA., Naddeo C. Water effects on the bond strength of
concrete/concrete adhesive joints, Construction and building materials, 2006; 20(10):
957-970.
[82] Cognard JY., Davies P., Gineste B., Sohier L. Development of an improved adhesive
test method for composite assembly design, Composites Science and Technology,
2005; 65(3): 359-368.
[83] Stigh U., Biel A., Walander T. Shear strength of adhesive layers–models and
experiments, Eng Fract Mech, 2014; 129: 67-76.
[84] Martiny P., Lani F., Kinloch A., Pardoen T. A multiscale parametric study of mode I
fracture in metal-to-metal low-toughness adhesive joints, Int J Fracture, 2012; 173(2):
105-133.
[85] Cooper V., Ivankovic A., Karac A., McAuliffe D., Murphy N. Effects of bond gap
thickness on the fracture of nano-toughened epoxy adhesive joints, Polymer, 2012;
53(24) 5540-5553.
[86] Carlberger T., Stigh U. Influence of layer thickness on cohesive properties of an epoxy-
based adhesive—an experimental study, The Journal of adhesion, 2010; 86(8): 816-
835.
[87] Zhao XL., Zhang L. State-of-the-art review on FRP strengthened steel structures,
Engineering Structures, 2007; 29(8): 1808-1823.
[88] Kang THK., Howell J., Kim S., Lee DJ. A state-of-the-art review on debonding failures
of FRP laminates externally adhered to concrete, International Journal of Concrete
Structures and Materials, 2012; 6(2): 123-134.
[89] D1002, Standard Test Method for Apparent Shear Strength of Single-Lap-Joint
Adhesively Bonded Metal Specimens by Tension Loading (Metal-to-Metal), American
Society for Testing Materials, West Conshohocken, PA,US., 2010.
[90] Yao J., Teng J., Chen JF. Experimental study on FRP-to-concrete bonded joints,
Composites Part B: Engineering, 2005; 36(2): 99-113.
[91] Bizindavyi L., Neale KW. Transfer lengths and bond strengths for composites bonded to
concrete, Journal of composites for construction, 1999; 3(4): 153-160.
[92] Volnyy VA., Pantelides CP. Bond length of CFRP composites attached to precast
concrete walls, Journal of Composites for Construction, ASCE, 1999; 3(4): 168-176.
[93] Nozaka K., Shield CK., Hajjar JF. Effective bond length of carbon-fiber-reinforced
polymer strips bonded to fatigued steel bridge I-girders, Journal of Bridge Engineering,
ASCE 2005; 10(2): 195-205.
[94] Branco FJ., Tadeu AJ., Nogueira JA. Bond geometry and shear strength of steel plates
bonded to concrete on heating, Journal of materials in civil engineering, ASCE 2003;
15(6): 586-593.
[95] Tadeu AJ., Branco FJ. Shear tests of steel plates epoxy-bonded to concrete under
temperature, Journal of Materials in civil Engineering, ASCE 2000; 12(1): 74-80.

42
[96] Coronado CA., Lopez MM. Experimental characterization of concrete-epoxy interfaces,
Journal of Materials in Civil Engineering, ASCE 2008; 20(4): 303-312.
[97] Siewczyńska M. Method for determining the parameters of surface roughness by usage
of a 3D scanner, archives of civil and mechanical engineering, 2012; 12(1): 83-89.
[98] Horgnies M., Willieme P., Gabet O. Influence of the surface properties of concrete on
the adhesion of coating: characterization of the interface by peel test and FT-IR
spectroscopy, Progress in Organic Coatings, 2011; 72(3): 360-379.
[99] Hindo KR. In-place bond testing and surface preparation of concrete, Concrete
International, 1990; 12(4): 46-48.
[100] Julio EN., Branco FA., Silva VD. Concrete-to-concrete bond strength. Influence of the
roughness of the substrate surface, Construction and Building Materials, 2004; 18(9):
675-681.
[101] Wang F., Li M., Hu S. Bond behavior of roughing FRP sheet bonded to concrete
substrate, Construction and Building Materials, 2014; 73: 145-152.
[102] Fernando D., Yu T., Teng J.G. Behavior of CFRP laminates bonded to a steel substrate
using a ductile adhesive, Journal of Composites for Construction, ASCE 2013; 18(2):
04013040-1-10.
[103] Julio EN., Branco FA., Silva VD., Lourenco JF. Influence of added concrete
compressive strength on adhesion to an existing concrete substrate, Building and
environment, 2006; 41(12): 1934-1939.
[104] Benzarti K., Chataigner S., Quiertant M., Marty C. Aubagnac C. Accelerated ageing
behaviour of the adhesive bond between concrete specimens and CFRP overlays,
Construction and building materials, 2011; 25(2) 523-538.
[105] Santos PM., Julio EN., Correlation between concrete-to-concrete bond strength and the
roughness of the substrate surface, Construction and Building Materials, 21 (2007)
1688-95.
[106] Santos DS., Santos PM., Dias-da-Costa D. Effect of surface preparation and bonding
agent on the concrete-to-concrete interface strength, Construction and Building
Materials, 2012; 37: 102-110.
[107] Papastergiou D., Lebet JP. Experimental investigation and modelling of the structural
behaviour of confined grouted interfaces for a new steel–concrete connection,
Engineering Structures, 2014; 74: 180-192.
[108] Tüzün FN., Tunalıoğlu MŞ. The effect of finely-divided fillers on the adhesion
strengths of epoxy-based adhesives, Composite Structures, 2015; 121: 296-303
[109] Zhou H., Liu HY., Zhou H., Zhang Y., Gao X., Mai YW. On adhesive properties of
nano-silica/epoxy bonded single-lap joints, Materials & Design, 2016; 95: 212-218.
[110] Wang B., Bai Y., Hu X., Lu P. Enhanced epoxy adhesion between steel plates by
surface treatment and CNT/short-fibre reinforcement, Compos Sci Technol, 2016; 127:
149-157.
[111] Bowditch MR. The durability of adhesive joints in the presence of water, International
Journal of Adhesion and Adhesives, 1996; 16(2) 73-79.
[112] Gao B., Kim JK., Leung CK., Experimental study on RC beams with FRP strips
bonded with rubber modified resins, Compos Sci Technol, 2004; 64(16): 2557-2564.
[113] Singla M., Chawla V. Mechanical properties of epoxy resin–fly ash composite, Journal
of minerals and Materials characterization and engineering, 2010; 9(3): 199-210.
[114] Dessureautt M., Spelt J. Observations of fatigue crack initiation and propagation in an
epoxy adhesive, International journal of adhesion and adhesives, 1997; 17(3): 183-195.
[115] Kong X., Xu Z., Guan L., Di M. Study on polyblending epoxy resin adhesive with
lignin I-curing temperature, International Journal of Adhesion and Adhesives, 2014; 48:
75-79.

43
[116] Moussa O., Vassilopoulos AP., de Castro J., Keller T. Early-age tensile properties of
structural epoxy adhesives subjected to low-temperature curing, International Journal of
Adhesion and Adhesives, 2012; 35: 9-16.
[117] Colak A., Cosgun T., Bakirci AE. Effects of environmental factors on the adhesion and
durability characteristics of epoxy-bonded concrete prisms, Construction and Building
Materials, 2009; 23(2) 758-767.
[118] Da Silva MA., Biscaia H. Degradation of bond between FRP and RC beams,
Composite structures, 2008; 85(2): 164-174.
[119] Brewer D., Gasparini D., Andreani J. Diffusion of water in steel-to-steel bonds, Journal
of Structural Engineering, ASCE 1990; 116(5): 1180-1198.
[120] Armstrong KB. Long-term durability in water of aluminium alloy adhesive joints
bonded with epoxy adhesives, International Journal of Adhesion and Adhesives, 1997;
17(2): 89-105.
[121] Karbhari VM., Navada R. Investigation of durability and surface preparation associated
defect criticality of composites bonded to concrete, Composites Part A: Applied
Science and Manufacturing, 2008; 39(6): 997-1006.
[122] Djouani F., Connan C., Delamar M., Chehimi MM., Benzarti K. Cement paste–epoxy
adhesive interactions, Construction and Building Materials, 2011; 25(2): 411-423.
[123] Gasparini D., Nara H., Andreani J., Boggs C., Brewer D., Etitum P. Steel-to-steel
connections with adhesives, Journal of Structural Engineering, ASCE 1990; 116(5):
1165-1179.
[124] Zanni-Deffarges MP., Shanahan MER., Diffusion of water into an epoxy adhesive:
comparison between bulk behaviour and adhesive joints, International Journal of
Adhesion and Adhesives, 1995; 15(3): 137-142.
[125] Subramaniam KV., Ali-Ahmad M., Ghosn M. Freeze–thaw degradation of FRP–
concrete interface: impact on cohesive fracture response, Engineering Fracture
Mechanics, 2008; 75(13): 3924-3940.
[126] Nguyen TC., Bai Y., Zhao XL., Al-Mahaidi R. Durability of steel/CFRP double strap
joints exposed to sea water, cyclic temperature and humidity, Composite Structures,
2012; 94(5): 1834-1845.
[127] Datla NV., Papini M., Ulicny J., Carlson B., Spelt JK. The effects of test temperature
and humidity on the mixed-mode fatigue behavior of a toughened adhesive aluminum
joint, Engineering Fracture Mechanics, 2011; 78(6): 1125-1139.
[128] Mikami C., Wu HC., Elarbi A. Effect of hot temperature on pull-off strength of FRP
bonded concrete, Construction and Building Materials, 2015; 91: 180-186.
[129] De Neve B., Shanahan MER. Effects of humidity on an epoxy adhesive, International
Journal of Adhesion and Adhesives, 1992; 12(3): 191-196.
[130] Park CE., Han B., Bair H. Humidity effects on adhesion strength between solder ball
and epoxy underfills, Polymer, 1997; 38(15): 3811-3818.
[131] Lettieri M., Frigione M. Effects of humid environment on thermal and mechanical
properties of a cold-curing structural epoxy adhesive, Construction and Building
Materials, 2012; 30: 753-760.
[132] Mouritz AP., Gibson AG. Fire properties of polymer composite materials, Springer
Science & Business Media, Netherland, 2007.
[133] Chowdhury EU., Bisby LA., Green MF., Kodur VK. Investigation of insulated FRP-
wrapped reinforced concrete columns in fire, Fire Safety Journal, 2007; 42(6): 452-460.
[134] Williams B., Kodur V. Green MF., Bisby L. Fire endurance of fiber-reinforced polymer
strengthened concrete T-beams, ACI structural Journal, 2008; 105(1): 60-67.
[135] Ahmed A., Kodur V., The experimental behavior of FRP-strengthened RC beams
subjected to design fire exposure, Engineering Structures, 2011; 33(7): 2201-2211.

44
[136] Baran E., Topkaya C. An experimental study on channel type shear connectors, Journal
of Constructional Steel Research, 2012; 74: 108-117.
[137] Moussa O., Vassilopoulos AP., Keller T. Experimental DSC-based method to
determine glass transition temperature during curing of structural adhesives,
Construction and Building Materials, 2012; 28(1): 263-268.
[138] Andrew JJ., Arumugam V., Santulli C. Effect of post-cure temperature and different
reinforcements in adhesive bonded repair for damaged glass/epoxy composites under
multiple quasi-static indentation loading, Composite Structures, 2016; 143: 63-74.
[139] Ferrier E., Rabinovitch O., Michel L. Mechanical behavior of concrete–resin/adhesive–
FRP structural assemblies under low and high temperatures, Construction and Building
Materials, 2015, http://dx.doi.org/10.1016/j.conbuildmat.2015.12.127.
[140] Hollaway LC., Leeming MB. Strengthening of reinforced concrete structures,
Woodhead Publishing Limited, Cambridge, England, 2001.
[141] Sasse HR. Adhesion between polymer and concrete, Springer Science Business Media.
BV, 1986.
[142] Salman H. Temperature effects on the bond of resin anchored reinforcement, MSc
Dissertation, Department of Civil and Structural Engineering, University of Sheffield,
1992.
[143] Palmieri A., Matthys S., Taerwe L. Experimental investigation on fire endurance of
insulated concrete beams strengthened with near surface mounted FRP bar
reinforcement, Composites Part B: Engineering, 2012; 43(3): 885-895.
[144] Arruda MRT., Firmo JP., Correia JR., Tiago C. Numerical modelling of the bond
between concrete and CFRP laminates at elevated temperatures, Engineering
Structures, 2016; 110: 233-243.
[145] Pinoteau N., Pimienta P., Guillet T., Rivillon P., Rémond S. Effect of heating rate on
bond failure of rebars into concrete using polymer adhesives to simulate exposure to
fire, International Journal of Adhesion and Adhesives, 2011; 31(8): 851-861.
[146] Moussa O., Vassilopoulos AP., de Castro J., Keller T. Time–temperature dependence
of thermomechanical recovery of cold-curing structural adhesives, International Journal
of Adhesion and Adhesives, 2012; 35: 94-101.
[147] Adams RD., Coppendale J., Mallick V., Al-Hamdan H. The effect of temperature on
the strength of adhesive joints, International Journal of Adhesion and Adhesives, 1992;
12(3): 185-190.
[148] Mohaisen SK. Experimental bond force-slip relationship for epoxy-coated reinforcing
vbars under elevated temperature conditions, Diyala Journal of Engineering Sciences,
05 2012; 5: 160-171.
[149] Rashid K., Zhang D., Ueda T., Jin W. Investigation on concrete-PCM interface under
elevated temperature: At material level and member level, Construction and Building
Materials, 2016; 125: 465-478.
[150] Plecnik J.M., Plecnik J.M., Fogarty J.H., Kurfees J.R. Behavior of epoxy repaired
beams under fire, Journal of Structural Engineering, 1986; 112(4): 906-922.
[151] Kumar P. Experimental investigations for shear bond strength of steel and concrete
bonded by epoxy, M. Tech. Dissertation, Department of Civil Engineering, Malaviya
National Institute of Technology Jaipur, India, 2013.

45

You might also like