Vous êtes sur la page 1sur 13

Composites: Part B 43 (2012) 398410

Contents lists available at ScienceDirect

Composites: Part B
journal homepage: www.elsevier.com/locate/compositesb

Effect of curing conditions on strength development in an epoxy resin


for structural strengthening
Christoph Czaderski , Enzo Martinelli 1, Julien Michels, Masoud Motavalli
Empa, Structural Engineering Research Laboratory, berlandstrasse 129, 8600 Dbendorf, Switzerland

a r t i c l e

i n f o

Article history:
Received 24 February 2011
Received in revised form 17 May 2011
Accepted 3 July 2011
Available online 13 July 2011
Keywords:
A. Resins
B. Cure behaviour
B. Stress transfer
D. Mechanical testing
FRP (bre reinforced polymer)

a b s t r a c t
Civil structures such as bridges and buildings can be strengthened with prestressed bre reinforced polymer (FRP) strips to enhance both their stiffness and load-bearing capacity. End anchorage is a crucial
issue for prestressed FRP strips. An innovative anchorage procedure, called the gradient anchorage
method and based on the possible accelerated curing of the epoxy-resin in the end region of the FRP
strip, has recently been conceived with the aim of avoiding more invasive mechanical fastening systems.
An in-depth knowledge of the actual development of the key mechanical properties of resins under different curing conditions (i.e., in terms of curing temperature) is of paramount importance for employing
the above mentioned gradient method in practical applications. This paper presents experimental results
and analytical investigations aimed at developing a better understanding of the strength development of
a commercial adhesive under different curing times and temperatures. Firstly, direct tensile tests on
epoxy specimens were performed at different curing temperatures. It was shown that the necessary curing time to reach the maximum tensile strength can be signicantly reduced from several hours at room
temperature to approximately 30 min at 90 C. Furthermore, higher curing temperatures reduced the
activation time after which strength starts to increase. The experimental observations are shown graphically with both the activation time and reaction duration at different curing temperatures. Secondly,
pull-off bond tests were conducted on 100 mm wide and 1.2 mm thick FRP strips bonded to concrete
using epoxy adhesives cured either at 90 C for different durations or at room temperature. An optical
image correlation system (ICS) allowed the load transfer behaviour of the inhomogeneous cured adhesive
between the FRP strip(s) and concrete to be studied. Finally, using the experimental measurements, the
bond shear stressslip interface relationships for the different test specimens were identied in order to
present the effect of elevated curing temperatures and curing durations.
2011 Elsevier Ltd. All rights reserved.

1. Introduction
The ageing of existing civil structures and the enhancement of
safety standards in structural codes are two important reasons
for the increasing interest of the structural engineering community
in developing cost-effective strengthening techniques. For instance, active reinforcement of existing members through prestressing techniques can signicantly enhance both stiffness and
strength of those members, with a more rational and efcient
exploitation of the mechanical properties of the reinforcing materials. Using bre reinforced polymers (FRPs) for external prestressing of existing reinforced concrete (RC) members is one of the most
innovative active techniques for structural strengthening.

Corresponding author. Tel.: +41 58 765 55 11 11; fax: +41 58 765 44 55.
E-mail address: christoph.czaderski@empa.ch (C. Czaderski).
Permanent address: University of Salerno, Dept. of Civil Engineering, via Ponte
don Melillo, 84084 Fisciano, Italy.
1

1359-8368/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesb.2011.07.006

The main advantages derived by employing FRP in structural


strengthening are their low density, ease in mounting and superior
durability. However, similar to the case of external prestressing
with steel cables, FRP prestressed reinforcement needs a sufciently strong anchorage in order to guarantee a correct force
transfer. Besides the conventional mechanical anchorages, which,
for example, use steel plates doweled against the concrete surface,
a fully adhesive anchorage system may be considered to be a more
durable solution, as no permanent steel plates and dowels are
required.
The so-called gradient anchorage method [13] is a possible
technical solution for avoiding mechanical fasteners and connectors in RC members externally prestressed by FRP strips. This
method reduces the prestressing force at the strip ends to zero over
a certain length while the FRP strips are placed by using a special
stressing and heating device. Currently, such a device consisting
of heating elements is being developed in collaboration with an
industry partner for practical applications. In order to achieve an
efcient application of the gradient anchorage method, detailed

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

399

Nomenclature
F
Fu
f(t)
fa
ta
tr
m
n
t
ttot
tn

maximum tensile force during the axial tensile tests


maximum force at failure or at initiation of debonding
during the pull-off bond test
time dependent axial tensile stress of the axial tensile
tests
axial tensile strength of the axial tensile tests
activation time (time which is necessary to start the
curing reaction, see Fig. 15)
reaction duration (time which is necessary to reach the
end strength, see Fig. 15)
reaction rate (see Fig. 15)
virtual intersection of the ascending branch with the yaxis
curing duration
full curing duration, i.e., activation time plus reaction
duration (ta + tr)
normalized duration for fully curing in reference to
room temperature testing at 22 C

information on how to prepare, place and cure the adhesive is


essential. Although plenty of experimental and theoretical
studies have been carried out on the structural behaviour of
FRP-strengthened RC structures (see, for instance, Meier [4] and
Teng et al. [5]), the effect of the curing conditions on the development of the mechanical properties of epoxy resins generally utilised for structural strengthening is one of the less investigated
issues. In this context, quantifying the effect of curing temperature
and duration on the strength development of epoxy resins could
signicantly improve the preparation and installation procedures
for the gradient anchorage or similar applications.
It must be pointed out that the present investigation studied the
effect of curing temperature and duration on the stiffness and
strength evolution of adhesives. As such, it did not consider the effect of test temperature on the properties of fully cured adhesives,
a subject investigated for example in Klamer et al. [6].
In the rst part of this paper, a State-of-the-art section presents information available in the literature related to the
strength development of epoxy adhesives. Following this, the results of two different types of experimental tests are presented.
On one hand, a series of direct tensile tests were performed on
samples of epoxy resin cured under different conditions with
the objective of describing the evolution of the adhesive strength.
On the other hand, six pull-off bond tests were carried out on FRP
strips glued on concrete blocks and cured in different conditions
for investigating how the interface relationship between shear
stresses and interface slip is inuenced by the curing conditions
in terms of temperature and duration (see also [7]). Full-eld
3D-displacements of the test specimens were measured using
an optical image correlation measurement system. Finally, different analytical models are developed and applied for deriving the
quantitative information related to the development of the relevant mechanical properties of both the adhesive and the FRPto-concrete interface.

2. State-of-the-art on strength development of epoxy adhesives


In general, adhesives can be classied according to either their
origin, method of bonding, end use or chemical composition. Detailed information about their possible classication can be found
in Mays and Hutchinson [8]. The present work focuses on the

th
tc
T
Tavg

smax
s
sel
smax
su
sT
Ef
tf
bf
L
kel
nm
x
xel

heating duration
cooling duration
curing temperature
mean temperature during pull-off test
maximum shear strength (see Fig. 17)
FRP-to-concrete interface slip
slip at end of elastic branch (see Fig. 17)
slip at debonding (see Fig. 17)
ultimate slip
the total number of specimens cured at temperature T
strip elastic modulus
strip thickness
strip width
bond length
elastic stiffness (see Fig. 17)
number of displacement measurements
abscissa of the coordinate axis system
point on the abscissa at which s(xel) = sel

strength development of epoxy adhesives, which are the most


commonly used for applications in structural strengthening.
2.1. Curing temperature and duration effects on strength development
The current section gives an overview of previous investigations
on the effect of curing temperature and duration on the strength
development of epoxy adhesives as reported in the scientic literature. A summary of signicant contributions is outlined in Table 1.
Kwan et al. [9] investigated several methods for rapid curing of
structural adhesives. They carried out an experimental campaign
on two kinds of structural adhesives (namely, urethane- and
epoxy-based), considering both ultrasonic and electromagnetic
(radio frequency and microwave) heating techniques. The study
was aimed at demonstrating the feasibility of heating procedures
for adhesive curing other than the traditional ones based on thermal heating. Different behaviours were observed for the two types
of adhesives. In particular, all the rapid curing techniques were
successful for the urethane-based adhesive with a fast enhancement of the joint strength observed within a few seconds, when
tested as a single lap shear joint according to ASTM D3163-96
[10]. On the other hand, the epoxy-based resin behaved worse,
resulting in arcing and smoking while heating.
Several researchers have investigated the effect of elevated curing temperatures on the strength development of structural
adhesives.
Lapique and Redford [11] studied the evolution of some relevant mechanical parameters (i.e., viscosity, strength and stiffness)
of a commercial epoxy-adhesive during the curing period. They
investigated the effect of the curing temperature and quantied
the time evolution of both the strength and the elastic modulus
at room temperature. It was demonstrated that when cured at
room temperature, the strength as well as the elastic modulus increased with time. Furthermore, a faster strength evolution was
observed at higher temperatures, as the same mechanical properties obtained at 23 C after 28 days could be attained at 64 C after
only 4 h.
Dodiuk and Kenig [12] observed increasing exural strength
and exural modulus of breglass epoxy composites with higher
curing temperatures. For instance, the strength achieved after 1 h
curing at 60 C was higher than that obtained after 1020 days of
curing at room temperature.

400

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

Table 1
State-of-the-art summary on curing temperature and time effect on epoxy strength evolution and end strength (RT = room temperature).
Author

Adhesives

Experiments

Curing
temperature

Kwan et al. [9]

Urethane- and epoxy-based

Lap shear test

Lapique and Redford


[11]
Dodiuk and Kenig
[12]
Matsui [13]

Araldite

Uniaxial tension and 3-point bending test

Ultrasonic, radio frequency and


microwave heating
2364 C
128 days

High-temperature epoxies

Bending and lap shear tests

RT, 60, 120 C

Epoxy (Araldite)/instant glue


(cyanoacrylate)
Unlled and silica lled epoxies
Four different epoxies

Lap shear test

18150 C

60 min to
20 days
3.350 h

Slant shear test


Torque tests with circular FRP laminates bonded on concrete

1040 C
34 to 49 C

114 days
2472 h

Hemp bre/epoxy composite/neat


epoxy
Glass ber reinforced epoxy

Fiber pull-out with interfacial shear strength measurements and


uniaxial tension
Bending and shear tests

25120 C

0.0813.8 h

RT to 150 C

5 h step curing

Tu and Kruger [14]


Dutta and Mosallam
[16]
Islam et al. [18]
Cao and Cameron
[17]

Similar observations have been made by Matsui [13], who studied the effects of curing conditions on the development of both
shear strength and stiffness for two kinds of adhesive (namely,
epoxypolyamideamine and cyanoacrylate, only the rst one being
used for structural purposes). In general, the author observed that
the minimum curing duration needed for reaching the maximum
shear strain was clearly related to the curing temperature.
Tu and Kruger [14], after testing two different epoxy resins
(with and without silica ller) in pull-out tests according to ASTM
C882-87 [15], observed an increase in bond strength with higher
curing temperature at a constant curing duration (7 days).
A large number of torque tests with a circular FRP composite
laminate bonded to a concrete cylinder were presented by Dutta
and Mosallam [16]. The investigation presented four different
epoxy-based adhesives tested at various curing temperatures ranging from 34 to 49 C with different curing durations. Similar conclusions as those made by the previous researchers regarding
strength evolution were drawn.
Although all experimental observations pointed out the huge
inuence of the curing conditions on the development of the relevant mechanical parameters, no quantitative formulations have
been proposed so far for expressing this relationship between, for
instance, the actual value of the tensile strength f(t) of epoxy resins
at curing time t and the corresponding curing temperature. Only
the following rule of thumb is given by Mays and Hutchinson
[8], who observed that, for curing temperature higher than 5 C,
the curing duration needed for achieving the nal strength of the
resin is halved for every increase of 10 C in temperature.
2.2. Inuence of curing temperature on the end strength of adhesives
In addition to studying the development of the relevant
mechanical properties of adhesives, various researchers have analysed the end strength of resins and its possible relationship with
the curing temperature. For instance, Dutta and Mosallam [16] observed higher end torque strengths for higher curing temperatures.
Cao and Cameron [17] conrmed this observation by investigating an additional effect. Instead of using isothermal curing, consisting of rapidly heating the resin specimen to the desired
temperature, the heating was carried out in various steps, starting
at a low temperature and increasing it in a step-wise manner for
different curing durations. According to the authors observations,
this technique leads to a more uniform curing of the adhesive layer
and, thus, results in a better mechanical performance.
Dodiuk and Kenig [12] observed both higher exural strength
and stiffness of specimens cured at 60 C and 120 C in comparison
to epoxies cured at room temperature.

Curing
duration

Moreover, Tu and Kruger [14] obtained higher bond strengths at


20 C than at 10 C for the same curing duration. With further heating up to 40 C, however, no further enhancement in strength was
obtained. Islam et al. [18], by means of tensile tests, obtained optimal values for both tensile strength and elastic modulus at a curing
temperature of 70 C, while both decreased again approaching
120 C.
To summarize, one can say that the adhesive stiffness and
strength development is highly dependent on the curing temperature and curing duration. In general, stiffness and strength development occurs faster with higher curing temperatures.
Furthermore, some indications of higher end strength with increasing curing temperature can be found in the literature. The main
goal of the current research was to investigate and quantify the
inuence of curing temperature and duration on strength development for the particular epoxy adhesive used in this investigation.
For other adhesive types with different chemical compositions, as
seen in the literature, a different behaviour might be observed.
3. Materials
All experimental tests were performed with a commercially
available epoxy based adhesive called S&P resin 220. Its measured
glass transition temperature Tg was 52 C and 58 C at room and
90 C curing temperatures, respectively. According to the declaration of the manufacturer, the tensile bond strength on steel is larger than 14 MPa.
The FRP strip had a thickness of 1.2 mm and a width of 100 mm
with the mechanical properties listed in Table 2. The elastic modulus was determined in a loading test for each strip before the
bond tests, which are also reported in Table 2. The measured modulus of 174 GPa was almost exactly that specied by the
manufacturer.
Aggregate with particle sizes ranging from 0 to 32 mm was employed for the concrete of the blocks used for the pull-off tests, to
which the FRP strips were glued. At 28 days, a cube compressive
strength fc,cube of 51.5 MPa was measured, whereas the splitting
tensile strength fct was found to be 3.25 MPa.
4. Experiments
Two different types of tests were carried out with the aim of
investigating both the evolution and the nal values of the key
mechanical properties of the adhesives for different curing conditions in terms of both temperature and duration. A commercial
epoxy adhesive which can be employed for the gradient

401

C. Czaderski et al. / Composites: Part B 43 (2012) 398410


Table 2
Mechanical properties of the FRP strips comparison between experimental investigation and product certicate.
Test No.

Experimental investigation

According to S&P certicate

Modulus of elasticity Ef
(GPa)

Modulus of elasticity Ef
(GPa)

Tensile strength fu
(MPa)

Elongation at 2000 MPa e2000


(%)

Width bf
(mm)

Thickness tf
(mm)

1
2
3
4
5
6

169
173
179
174
171
177

173.9

2975

1.142

100

1.23

Mean
value

174

anchorage method in RC beams externally prestressed by FRP


strips was used for all tests. The experimental campaign can be
subdivided into the two following groups of tests:
axial tensile tests on adhesive samples, to measure the development of the adhesive axial tensile strength at different curing
temperatures;
pull-off bond tests on FRP-to-concrete joints, aimed at describing the evolution of the interface behaviour under variable curing conditions.
The former tests were carried out for curing temperatures ranging between approximately 10 and 90 C, while the latter ones
were performed at a constant mean curing temperature of approximately 90 C and variable curing durations. It should be noted that
the tests were performed approximately 510 min after stopping
the heat transfer, which means that the temperature of the test
specimens at testing was still signicantly higher than room temperature (Sections 4.1 and 4.2) and therefore could inuence the
test results. However, this corresponds approximately to the real
situation at the time when the prestressing force is released during
application of the gradient method.
4.1. Axial tensile tests
A large number of axial tensile tests were performed on specimens as those represented in Fig. 1. The uncured resin was applied
between an aluminium cylinder with a diameter of 20 mm and a
square aluminium plate with dimensions of 60  60 mm. These

Fig. 1. Specimens for axial tensile tests, on the left side (Nos. 10 and 11) the
aluminium stamps and plates with applied adhesive in between can be seen, on the
right side (No. 12) the stamp including the xture for the curing process is visible.

Fig. 2. Set-up for axial tensile tests.

test specimens were cured at different temperatures in an oven


and tested with the pull-off tester shown in Fig. 2. The tests were
aimed at measuring the tensile strength of the resin depending
on both curing temperature and duration. It is clear that, due to
the short length of the adhesive sample, this test specimen does
not deliver the real tensile strength. Nevertheless, the actual trend
in the evolution of the tensile strength of the resin can be captured
in these tests, whose specimens are particularly convenient and
can be easily and rather uniformly heated inside a simple oven.
The evolution of the temperature of the air in the oven as well as
the temperature inside the resin layer were monitored in some
of the specimens which underwent an accelerated curing process
by heat transfer (Figs. 3 and 4).

Fig. 3. Axial tensile test specimens inside the oven and cables for temperature
measurements.

402

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

120
Room Temperature
Sensors on the adhesive

Air temp. in the oven

Temperature [ oC]

90

60

30

Oven in testing hall (SH)


0
0

100

200

300

Time [mins]
125
Room Temperature
Sensors on the adhesive

Air temp. in the oven

Temperature [C]

100
75
50
25

Oven room (OR)


0
0

50

100

150

Time [mins]
Fig. 4. Typical temperature development during curing of the axial tensile test
specimens, including the temperature decrease due to removal of the test
specimens from the oven.

A complete overview of the tested resin specimens is reported


in Table 3. The number of specimens cured at room temperature
as well as those which underwent an accelerated curing process
at different temperatures is also shown in the table. For the temperature curing, two different ovens were used. In particular, a total of 200 tests were carried out in several sessions with variable
curing conditions: 136 tests were carried out on specimens with
accelerated curing in an oven with temperatures ranging between
45 and 90 C, and 64 tests were carried out on specimens cured at
different room temperatures, with curing durations ranging from a
few hours to 7 days.
Most of the tests were performed on specimens cured at 90 C,
which is considered as the reference temperature for the present
investigation. However, this was only the nominal temperature
of the air inside the oven. The temperature development in the
adhesive resin was generally slower than that in the air and the nal temperature was slightly lower than the nominal one. A typical
temperature development observed is shown in Fig. 4, emphasising the different measurements inside the adhesive and in the oven
air. Additional tests on specimens cured at 120 C were also carried
out, but the results are not reported herein as their failure was always controlled by the loss of bonding between the resin and the
aluminium plate, possibly as a result of the difference in relative
expansion of the two materials at that temperature.
Fig. 4 also shows the cooling curve for test specimens taken out
of the oven. All the axial tensile tests were carried out after a minimum of 5 min of cooling outside the oven at room temperature.
Therefore, from Fig. 4, it can be concluded that the resin of the
tested specimens had a maximum temperature ranging between
60 and 75 C at the time that the tests were actually performed.
4.2. Pull-off bond tests on FRP strips glued on concrete through a layer
of epoxy resin

Table 3
Overview of the experimental programme for the axial tensile tests on S&P resin 220
(SH: oven in testing hall, OR: oven in oven room).
Series

Tests with heating

Tests w/o heating

[#]

n.

T (C)

n.

4
12

90
90

1
2
3
4
5
6
7
8
9
10
11
12

19
7

90
90

16
16
24
24
14

65
45
90
65
90

Total

136

12

48

24
4

24
23

12
8
4

36
168
168

64

Oven

tmax (h)

SH
SH

SH
SH

SH
SH
OR
OR
OR

Six pull-off bond tests were carried out with the aim of measuring the strength of FRP-to-concrete joints. In particular, the time
evolution of the mechanical properties of those joints cured at high
temperature was the key issue of this investigation. Fig. 5 shows a
view of the test layout: a 100 mm wide and 1.2 mm thick FRP strip
was glued to a concrete block and pulled off after either curing at
an elevated temperature (approx. 90 C) or at room temperature.
The bond length was 300 mm and the heating was performed
using three heating elements each with a dimension of
100 mm  100 mm. The pull-off bond tests were prepared following the main phases represented in Fig. 6. The tested specimens
were always equipped with sensors for monitoring the temperature evolution within both the adhesive layer and the heating elements. The preparation of the adhesive layer between the FRP strip
and concrete surface was followed by the heating phase (th), carried out by means of heating elements. After stopping the heating
procedure, a cooling time of 510 min was included before starting

Fig. 5. Test setup for pull-off bond tests.

403

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

Table 5
Results of the axial tensile tests on specimens cured at room temperature (n = number
of tests, T = curing temperature, t = curing time, F = maximum tensile force, fa = axial
tensile strength).

Fig. 6. Adhesive temperature (T1, T2, T3) and force development of Test No. 3 (pulloff bond tests), th: heating duration, tc: cooling duration, Tavg: mean temperature
during pull-off test, Fu: ultimate force, L = bond length, T1, T2, T3 = temperatures in
the adhesive at the positions 1, 2, 3 (see schematic drawing).

Table 4
Overview of the experimental programme for the pull-off bond tests.
Test
No.

Mean
distance
concrete
surfaceupper
strip
surface
(mm)

Strip
thickness
(mm)

Mean
adhesive
thickness
(mm)

Heating
duration
th (min)

Waiting
time to
failure
test tc
(min)

Average test
temperature
Tavg (C)

1
2
3
4
5
6

4.85
5.14
5.00
4.90
4.74
4.93

1.2
1.2
1.2
1.2
1.2
1.2

3.65
3.94
3.80
3.70
3.54
3.73

15
25
20
25
25

5.7
8.6
5.8
7.7
2 days
3 days

45
43
41
42
19
18

Series

T (C)

t (h)

F (kN)

fa (MPa)

1
4
5
7
8a
8b
9

12
24
4
12
4
4
4

2426
22
24
10
10
22
22

4.048.0
4.024.0
7.022.9
6.336.0
48168
48168
48168

0.086.15
0.036.18
1.496.13
0.035.06
4.404.95
5.766.44
5.846.41

0.2519.6
0.119.7
4.719.5
0.116.1
14.015.8
18.320.5
18.620.4

image correlation measurement system (ICS) [19]. A description


of the measurement system used can be found in Czaderski et al.
[20], Czaderski and Rabinovitch [21]. With the ICS, the shape of
the surface was also measured. Fig. 7 shows the isolines of the surface of the strip and concrete before starting the test. It can be seen
that the surface of the strip was not planar, but was curved with a
maximum vertical value at the surface centre. This was the result
of the application process. Hence, a larger adhesive quantity can
be observed in the strip centre. Over the whole bonding area, an
average value can be obtained. The latter can subsequently be
transformed into the mean adhesive thickness by subtracting the
strip thickness. A summary of the measured adhesive thicknesses
is given in Table 4.
The heating duration range of 1525 min was chosen based on
the results observed in the previous campaign, while the waiting
time for cooling was chosen so that the adhesive temperature
was lower than approximately 50 (Table 4). Comparing Figs. 4
and 6, a key observation is the temperature development, and consequently, the rate of the heat transfer in the axial tensile tests and
pull-off bond tests. In particular, the temperature developed much
faster in the second series of tests as a result of greater heat transfer by using heating elements in contact with the CFRP strips compared to heating the air in the oven.
5. Results

the loading test in order to cool down the adhesive (tc). Afterwards,
one can observe the force increasing up to the maximum value Fu.
It can be observed that the temperature during the experiment was
not constant, but slowly decreased. The heating and cooling times,
as well as the average test temperature (Tavg), are given in Table 4.
The cooling time was chosen so that the existing temperature in
the adhesive was safely below its glass transition temperature.
The full-eld 3D displacements of the test specimen were measured during the experimental investigations with an optical

The large number of axial tensile tests produced plenty of


experimental results useful for understanding the development
of the axial strength of the epoxy-resin used in this study at different curing temperatures. The strength was simply obtained by
dividing the failure load by the nominal cross-sectional area of
the adhesive layer.
On the other hand, analysing the pull-off bond tests performed
on the FRP-to-concrete adhesive joints was much more demanding
because the properties of the adhesive interface could not be

Fig. 7. Isolines of the surface of the CFRP strip before the start of the pull-off test determined by using the ICS.

404

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

Table 6
Results of the axial tensile tests on specimens cured in an oven (n = number of tests,
T = curing temperature, t = curing time, F = maximum tensile force, fa = axial tensile
strength).
Series

T (C)

t (h)

F (kN)

fa (MPa)

2
3
5
6
8
9
10
11
12

4
12
19
7
16
16
24
24
14

90
8890
8889
90
65
45
90
65
90

1.01.3
0.84.0
0.73.7
0.40.6
0.72.7
0.74.0
0.252.5
0.522.5
0.30.7

2.873.88
1.365.66
1.287.14
0.232.03
0.056.41
0.06.10
0.07.2
0.315.94
03.4

9.112.35
4.318.0
4.122.7
0.76.5
0.1620.4
0.019.4
0.022.9
1.018.9
010.8

directly measured by those tests, but had to be derived by means of


an indirect identication procedure [20,22].
This section presents the raw results obtained by the experimental tests; only a few phenomenological observations are reported. An analytical model for deriving the key parameters of
the corresponding interface bond law is discussed in Section 6.
5.1. Axial tensile tests
The results of the axial tensile tests carried out on the adhesive
specimens cured at room temperature are reported in Table 5. The
table reports both the minimum and maximum nominal curing
temperatures and the duration (in hours) of each series. Furthermore, the minimum and maximum axial forces at failure F and
the corresponding mean axial strengths fa are given for each series.
Here, axial strengths were evaluated under the assumption of a
uniform distribution of axial stresses on the circular transverse
section of the adhesive layer. As discussed before, this mean axial
strength is not the correct tensile strength because of the short
length of the test specimen. However, it allows the strength development at different curing temperatures to be studied. The results
for the specimens cured at higher temperatures are reported in
Table 6 following the same procedures explained above.
Several observations may be made on the basis of these experimental results. Firstly, the maximum values of the axial strength fa
obtained from the specimens cured either at room temperature or
within the oven are quite close to one another, being in both cases
about 1820 MPa. However, a closer comparison between the different curing conditions reveals slightly lower end strength
reached at 10 C. This temperature resulted in an end strength of
16 MPa, whereas higher temperatures led to higher resistances of
about 1822 MPa. This is in accordance with the ndings of the
researchers mentioned above, especially Tu and Kruger [14], who
found a very similar bond strength enhancement from about 15

to 18.5 MPa in the same curing temperature range of 1020 C.


However, in order to verify this nding, further detailed experimental investigation is needed. In general, the presented strength
range corresponds to the declaration of the distributor of the material, as previously described in Section 3.
On the other hand, the key difference between the various curing temperatures lies in the time span necessary for reaching the
end strength. For the case of specimens cured at room temperature
it is in the order of 12 h to a few days, whereas it drops to a few
hours in the case of specimens cured in the oven at higher temperatures (4590), see Fig. 12. Consequently, the effect of heat transfer on the curing process of resins is quantitatively relevant and the
relationships between the curing temperature and the development of strength in the resin should be approximated with exponential or linear functions, as reported in detail in Section 6.
In all the tests, failure developed throughout the adhesive layer
with no signicant loss of adhesion observed at the aluminium-toresin interface. Consequently, the strength values strictly refer to
the adhesive and are not signicantly affected by the testing system. Fig. 8 illustrates two relevant failure modes of the resin layer,
the rst observed either at low curing temperature or short curing
duration (Fig. 8a), and the second observed after the activation of
the chemical reactions resulted in the hardening of resin (Fig. 8b).
5.2. Pull-off bond tests
The results of the pull-off bond tests carried out on the six FRPto-concrete specimens are given in Table 7 as a function of the
heating duration and waiting time before the test started. Fig. 9
shows the maximum measured force or the force at initiation of
debonding as a function of the curing duration. The characteristic
initiation of debonding stage was dened as the decisive stage
to determine the bond shear stressslip relationship, although
the force can slightly increase afterwards. A discussion related to
that topic can be found in Czaderski et al. [20]. The result of Test
No. 6, carried out on a specimen cured at room temperature for
3 days, is considered as a reference. As indicated in Fig. 9 and Table
7, the absolute strength and displacement values of Test No. 2 have
to be excluded from a detailed analysis, as problems occurred with
the test setup (complete concrete block displacement occurred instead of a relative movement between the concrete and FRP strip).
It can be observed that the specimens cured at an elevated temperature for 15 or 20 min exhibit signicantly lower strengths. Moreover, all the specimens cured for 25 min (tested either a few
minutes after the heating procedure or after 2 days) failed at forces
higher than that observed for the reference specimen cured at
room temperature.
The failure modes observed in the various tests conrm the
above quantitative observations. Fig. 10 shows the interfaces of

Fig. 8. Axial tensile tests: typical failure modes.

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

405

Table 7
Results of the pull-off bond tests (Fu = maximum force at failure or at initiation of
debonding, smax = maximum bond strength, sel = slip at maximum bond strength,
smax = maximum slip at full debonding, wt: waiting time to failure test 2 or 3 days).

Test
No.

Heating
duration (min)

Fu
(kN)

smax
(MPa)

sel
(mm)

smax
(mm)

Failure
mode

1
2a
3

15
25
20

18.4
60.4
34.4

0.8
5.1
1.4

0.602
0.205
0.803

0.652
0.408
0.803

4
5
6

25
25 + wt
wt

75.7
62.0
57.6

4.4
7.7
7.9

0.426
0.020
0.031

0.712
0.248
0.218

Adhesive
Concrete
Mainly
adhesive
Concrete
Concrete
Concrete

Fu is not the force at failure, due to problems with the test set-up.

60

15

25

25*

Test No.6

20

Test No.5

Test No.2*

Test No.4

20

Test No.3

40

Test No.1

Maximum force Fu [kN]

80

25 + 2 days no heating +
waiting time
3 days
waiting time

Curing duration [min]


Fig. 9. Maximum force at failure or at initiation of debonding according to the
applied curing duration and waiting time before testing (Fu of Test No. 2 is not the
force at failure, due to problems with the test set-up).

the test specimens after failure. For test specimen No. 1, the resin
was not completely cured after only 15 min, which resulted in a
failure process in the adhesive layer without any cracks developing
in the concrete. A substantially similar behaviour was observed for
the specimen cured for 20 min (Test No. 3), although a small concrete failure area is visible. On the contrary, the other specimens
failed with a fracture developing throughout the concrete substrate (Test Nos. 2, 4 and 5) as was the case for the reference test
(No. 6).
Another interesting observation is presented in Fig. 11, where
the isolines of the displacements in the direction of the applied
pulling force (x-direction) determined with the ICS for specimens
Nos. 4 and 6 are shown. One can observe a signicant difference
between the behaviours in terms of displacements measured.
Whereas the displacement throughout the central axis is signicantly lower than the one on the outer borders for Test No. 4, the
opposite occurs for the reference Test No. 6. A reason for this discrepancy lies in the inhomogeneous heating process over the strip
width. During curing with the heating elements, a lower temperature develops at the strip edges leading to a lower stiffness and
strength compared to the central part of the strips. Therefore, higher displacements can develop at the strip edges. The reference Test
No. 6 shows the typical behaviour given in the literature, e.g., [23],
whereby the centre part of the strip exhibits higher displacements
than the edge parts of the strip.
In addition to the direct measurement of the ultimate load leading to the failure of the FRP-to-concrete joint, the pull-off bond
tests are analysed in the next section with the aim of deriving
the bond shear stressslip interface law of the FRP-to-concrete
interface and their evolution depending on the heating duration.

Fig. 10. Pull-off bond tests at different curing temperatures: observed failure
modes at the concrete-FRP interface.

6. Analytical calculations
6.1. Axial tensile tests
The experimental observations derived by the axial tensile tests
show that the strength fa is negligible in the rst stages of the curing phase. The chemical reactions resulting in the hardening of the
resin and the development of its mechanical properties actually
begin after a certain time span, referred to here as the activation
time ta. After this time, the mechanical properties (i.e., the axial
strength f(t)) develop quickly, with rates depending on the curing
temperature, up to a maximum asymptotic value (the end
strength) which is almost independent of the curing conditions.
The following trilinear relationship was considered, as it clearly
represents the key features of the observed behaviour in terms of
activation time (ta = n/m), reaction rate (m) and asymptotic
strength value fa (end strength):

f t m  t n with 0 6 f t 6 fa :

406

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

Fig. 11. ICS image of displacement in x-direction (pulling direction) of the FRP strip and the concrete surface.

A least-squares calibration of Eq. (1) was performed for every set of


sT experimental results reported in Tables 5 and 6 and derived at the
various temperatures T reported in Table 8:
Table 8
Results of the calibration of the tri-linear model for the different temperatures
(T = curing temperature, m = reaction rate, n = virtual intersection of the ascending
branch with the y-axis, ta = activation time, tr = reaction duration, ttot = full curing
duration, tn = normalized duration for full curing relative to room temperature testing
at 22 C, fa = axial tensile strength).
T (C)

m
(MPa/h)

n
(MPa)

ta = n/m
(h)

tr
(h)

ttot
(h)

tn
()

fa
(MPa)

10
22
25
45
65
90

1.12
2
2.32
9.37
13.06
18.99

12.72
11
10.57
14.13
8.03
5.36

11.37
5.49
4.56
1.51
0.61
0.28

13.53
9.77
8.11
2.00
1.34
0.88

24.90
15.26
12.67
3.51
1.95
1.16

1.63
1.00
0.83
0.23
0.13
0.08

15.15
19.57
18.82
18.8
17.68
16.79

"
 n
 T arg min
m;
m;n

#
sT
X
f ti ; m; n  fa;i 2 ;

i1

where ti is the duration of curing and fa,i the observed axial strength
of the ith specimen cured at temperature T (out of the total sT cured
at the same temperature).
The results of the calibrations in terms of m, n, and other related
parameters, are reported in Fig. 12 for the six relevant temperatures T considered in the curing process of the test specimens.
The diagrams are sorted in Fig. 12 from lower to higher curing temperatures and the effect of the heating process on the time evolution of the axial strength fa is clearly visible. For instance, the
activation time, which is about 6 h in the case of room temperature
(T = 22 C), is sharply reduced to less than half an hour in the case
of the higher oven temperature (T = 90 C). Fig. 13 shows the exponential reduction of the activation time ta as a function of the curing temperature T; the corresponding exponential interpolation is

407

25

25

20

20

Strength [MPa]

Strength [MPa]

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

15
10

Nominal Temperature
~10 oC

Test results
Tri-Linear Model

15
Room Temperature
22 oC

10
5

Test results
Tri-Linear Model

0
0

10

20

30

40

50

10

25

25

20

20

15
Room Temperature
25 oC

10
5

Test results
Tri-Linear Model

30

40

50

15
10

Nominal Oven Temp.


T=45C

Test results
Tri-Linear Model

10

20

30

40

50

Curing Duration [h]

Curing duration [h]

25

25

20

20

Strength [MPa]

Strength [MPa]

20

Curing Duration [h]

Strength [MPa]

Strength [MPa]

Curing Duration [h]

15
10

Nominal Oven Temp.


T=65C

Test results

15
10

Nominal Oven Temp.


T=90C

Test results
Tri-Linear Model

Tri-Linear Model

0
0

Curing duration [h]

Curing duration [h]

Fig. 12. Axial tensile strength evolution and calibration of the tri-linear law for the specimens cured in an oven at different temperatures.

100.0

100.0

tstart = 20.828e-0.0574 T

m = 0.0332 T1.4143
R2 = 0.9545

R2 = 0.9843

ta [h]

m [MPa/h]

10.0

1.0

10

20

30

40

50

60

70

80

90

10.0

1.0
1

100

10

100

0.1

0.1

T [oC]
Fig. 13. Axial tensile tests: calibrated relationship between curing temperature and
activation time.

T [oC]
Fig. 14. Axial tensile tests: calibrated relationship between curing temperature and
reaction rate.

408

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

Fig. 15. Calibrated relationship for strength development depending on heating duration at different curing temperatures (the activation time and reaction duration as
indicated in the graph are only valid for T = 60 C).

drawn as well as its analytical expression and the corresponding


coefcient of determination R2 with respect to the experimental
data. In addition, Fig. 14 displays the relationship between the
reaction rate m and the curing temperature T, suggesting the exponential interpolation shown along with the corresponding value of
the coefcient R2. In both cases, the proposed interpolations are in
very good agreement with the observed data and the corresponding coefcients of determination are rather close to unity.
Finally, the relationship (1) between the actual strength f(t)
developed in the resin at time t and the curing temperature can
be utilised for generating a complete chart representing the evolution of the ratio f(t)/fa as a function of time t at different curing
temperatures T (see Fig. 15).
The duration necessary for developing the nal asymptotic
strength can be obtained by adding the activation time to the reaction duration. Fig. 16 presents the normalised duration for complete curing (room temperature of 22 C is taken as the reference
value) compared to the earlier mentioned rule of thumb by Mays
and Hutchinson [8]. A good agreement between the experimental
results and the suggested rule of thumb found in the literature
can be observed.

The maximum observed displacement at the last stage before


failure or initiation of debonding was taken as smax. In particular,
three load stages were considered in the present study at three different load levels F(i) (i = 1, 2, 3) (see Fig. 18).

6.2. Pull-off bond tests


The key objective of this part of the investigation was to identify
the general bilinear bond shear stressslip relationship represented in Fig. 17 [24] . Three mechanical parameters completely
dene the relationship between the shear stresses sf and the interface slip s:

Fig. 16. Normalized duration for complete curing (activation time plus reaction
duration) dependent on the curing temperature: experimental measurements in
comparison with the rule rule of thumb according to Mays and Hutchinson [8].

the maximum shear strength smax;


the interface slip sel at the end of the elastic branch of the interface relationship;
the interface slip smax at debonding.
Displacements were monitored during the tests by means of an
advanced optical measurement system [19]. The slip was
calculated from the difference between the displacements of the
CFRP strip (section 0 in Fig. 11) and the concrete surface (Sections
5 and 6 in Fig. 11) beside the strip.

Fig. 17. General bi-linear relationship for the FRP-to-concrete interface.

409

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

0.3

Interface slip - s [mm]

Experimental
results

0.2

Calibrated
model

62.0 kN

62.0 kN

40.5 kN

40.5 kN

20.1 kN

20.1 kN

Czaderski et al. [20] or Faella et al. [22]; it can be written symbolically as follows:

Pulling direction

Strip

Test no.5

s sx; smax ; su ; xel ;

Concrete

0.1

load increase

0.0
0

50

100

150

200

250

300

x [mm]

Interface slip - s [mm]

indicating that the value of the interface slip s at a point on the abscissa x throughout the bonded length L depends on both the postelastic branch of the bi-linear law represented in Fig. 17 (namely, on
the maximum shear strength smax and the ultimate slip su) and the
point on the abscissa xel throughout the adhesive interface at which
s(xel) = sel.
The resulting external force F can be evaluated through another
relationship whose expression is reported below as a function of
the same variables:

F Fsmax ; su ; xel :

0.3
Test no.6
Experimental Calibrated
model
results

0.2

57.6 kN

57.6 kN

40.1 kN

40.1 kN

20.1 kN

20.1 kN

the interface slips evaluated using Eq. (6) result in values as


close as possible to the observed values si,exp;
the external force F evaluated using Eq. (7) is as close as possible
to that applied during the experimental test.

load increase

0.0
0

50

100

150

200

250

300

x [mm]
Fig. 18. Measured relative displacements for different loading steps in x-direction
(longitudinal to the strip direction) between the strip and the concrete surface
referred to as interface slip and simulated slips after the calibration procedure.

The above conditions can be mathematically stated through the


following least-squares optimization problem with reference to
the third load stage F(3) (namely, resulting in specimen failure)
[22]:

(
nm
X
sxi ; smax ; su ; xel  si;exp 2
s2u
smax ;xel
ii
9
"
#2
Fsmax ; su ; xel  F 3 =

:
;
F 3

max ; xel arg min


s

The following equations and descriptions explain the modelling


approach for both the elastic and the post-elastic stages.

F
cosh xx

;
xEf bf tf sinh xL

where Ef is the Youngs modulus of the composite strip, bf and tf are


the width and thickness of its transverse section and the parameter
x is dened as follows:

s
kel
x
:
Ef t f

Considering the rst load level F(1) and assuming that the behav (and the
iour under this load remains elastic, the optimal value of x
corresponding value of the elastic stiffness kel) can be derived by
solving the following least-squares minimization problem:

(
)
nm
X
 arg min
x
sxi ; x; F i  si;exp 2 ;

ii

where nm is the number of displacement measurements si,exp available at the same load stage.
6.2.2. Post-elastic stage
The post-elastic response of an adhesive joint can be modelled
by a more complicated analytical relationship, where the dependent variables are smax and smax = su. The explicit analytical expression of that relationship is omitted here, but can be found in

The results of the optimization procedure are reported in Table


7 in terms of the resulting values of smax, sel, and smax = su.
The bilinear interface relationships corresponding to the above
calibrations are presented in Fig. 19, which demonstrates the effect
of accelerated curing on the mechanical properties of the FRP-toconcrete interface. The relationships derived from Test Nos. 1
(15 min heating duration) and 3 (20 min) are characterized by a
very low stiffness and no decreasing branch, as the resin is not fully
cured. On the other hand, a signicantly higher value of the
stiffness is observed in Test No. 4 (25 min heating duration), but
the stiffness is still clearly lower than that of the fully cured pulloff tests. Also, the value of the strength smax in Test No. 4 is lower
than that for Test Nos. 5 and 6. Finally, the interface relationships
derived for Test Nos. 5 and 6 are rather close to one another, conrming the fact that the accelerated curing procedure (carried out
8

Bond shear stress - [MPa]

6.2.1. Elastic stage


The elastic stiffness kel = smax/sel (Fig. 17) can be identied by
means of a numerical calibration of the following relationship, representing the interface displacement eld due to an axial force F
applied at the end of a bonded joint behaving elastically throughout its length L [22]:

sx; x; F

Consequently, the optimal values of the parameters smax and xel


can be derived imposing the following two conditions:

0.1

fully cured

Test. no.1
Test. no.3

Test. no.4

Test. no.5

25 mins

Test. no.6

3
2
20 mins

1
15 mins

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Interface slip - s [mm]


Fig. 19. Pull-off bond tests: resulting interface relationships between shear stresses
and slips.

410

C. Czaderski et al. / Composites: Part B 43 (2012) 398410

for Test No. 5) does not affect the nal behaviour of the resin (as
obtained in Test No. 6, the reference test).
7. Conclusions
The presented experimental and analytical results lead to several conclusions. As expected, the axial tensile tests revealed a considerable inuence of the curing temperature on strength evolution
of the epoxy adhesive. The necessary curing duration for completely developing the end strength of the adhesive, composed of
the reaction duration and the activation time, was signicantly reduced at higher curing temperatures. On the basis of the axial tensile tests, a complete chart representing the evolution of the
normalized tensile strength f(t)/fa as a function of time t at different
curing temperatures T was presented (Fig. 15). The tests demonstrated that to develop almost the same end strength, curing durations ranging from approximately 1 day at 10 C to a slightly more
than one hour at 90 C were required. Moreover, the values of curing duration obtained at the various temperatures were found to be
in good agreement with a rule of thumb estimate available in the
scientic literature (see Section 2.1). No signicant variation in end
strength was found at higher curing temperatures. However, further experiments are needed in order to conrm this observation.
The investigations on the FRP-to-concrete interface at 90 C curing temperature showed that 1520 min of curing resulted in an
adhesive failure, which means that the strength of the adhesive
layer had not completely developed for that duration and temperature. On the other hand, the specimen cured for 25 min at 90 C
exhibited the same failure mode observed in the pull-off test on
the reference specimen, which was cured for 3 days at room temperature. The experimental results in terms of the distribution of
interface slip throughout the bond length at different load levels
were used to identify the shear stressinterface slip relationships
for the various specimens cured at 15, 20 and 25 min and for the
reference one cured in the usual way. Besides the above mentioned
adhesive failure at the shortest two curing durations, curing for
25 min induced a lower elastic stiffness, a lower maximum bond
strength and a larger total slip displacement as compared to the
fully cured specimens. The displacement measured over the strip
width by optical image correlation revealed the need for optimising the performance of the heating elements in order to obtain a
more homogeneous heat transfer and temperature within the
adhesive layer.
Finally, the quantitative results obtained in the present study
are of key importance for enhancing the practical implementation
of the gradient anchorage method for prestressed FRP strips in RC
beams (Czaderski and Motavalli [1], Aram et al. [2]). Dening the
optimal anchorage length, the number of steps and the heating
temperature and duration to be used for implementing the above
mentioned procedure are among the next objectives of this
research.
Acknowledgements
The nancial support of the Swiss innovation promotion agency
(CTI) of Switzerland is acknowledged (Project Number KTI Nr.

10493.2 PFIW-IW). Furthermore, the nancial support and delivery


of the test materials of the industrial partner of the project, S&P
Clever Reinforcement from Seewen, Switzerland, is also
appreciated.
References
[1] Czaderski C, Motavalli M. 40 years-old full-scale concrete bridge girder
strengthened with prestressed CFRP plates anchored using gradient method.
Compos Part B: Eng 2007;38(78):87886.
[2] Aram MR, Czaderski C, Motavalli M. Effects of gradually anchored prestressed
CFRP strips bonded on prestressed concrete beams. J Compos Constr ASCE
2008;12(1):2534.
[3] Motavalli M, Czaderski C. FRP composites for retrotting of existing civil
structures in Europe: state-of-the-art review. In: ACMA, 2007. Tampa, USA,
October 1719; 2007.
[4] Meier U. Strengthening of structures using carbon bre/epoxy composites.
Constr Build Mater 1995;9(6):34151.
[5] Teng JG, Chen JF, Smith ST, Lam L. FRP strengthened RC structures. Chichester
(England): John Wiley & Sons, Ltd.; 2002. p. 245.
[6] Klamer EL, Hordijk DA, Hermes MCJ. The inuence of temperature on RC beams
strengthened with externally bonded CFRP reinforcement. Heron
2008;53(3):15785.
[7] Czaderski C, Martinelli E, Michels J, Motavalli M. Effect of partly cured
adhesives on bond-slip-relationship of FRP-concrete interface. In: SMAR, rst
middle east conference on smart monitoring, assessment and rehabilitation of
civil structures, 2011. Dubai, UAE, February 810; 2011. p. 8.
[8] Mays GC, Hutchinson AR. Adhesives in civil engineering. Cambridge University
Press; 1992.
[9] Kwan KM, Cheng K, Benatar A. Feasibility study of raid curing of structural
adhesives. In: Proceedings of the ANTEC 98. Atlanta; 1998. p. 108994.
[10] ASTMD3163-96. Standard test method for determining strength of adhesively
bonded rigid plastic lap-shear joints in shear by tension loading.
[11] Lapique F, Redford K. Curing effects on viscosity and mechanical properties of a
commercial epoxy resin adhesive. Int J Adhes Adhes 2002;22(4):33746.
[12] Dodiuk H, Kenig S. Low temperature curing epoxies for structural repair. Prog
Polym Sci (Oxford) 1994;19(3):43967.
[13] Matsui K. Effects of curing conditions and test temperatures on the strength of
adhesive-bonded joints. Int J Adhes Adhes 1990;10(4):27784.
[14] Tu L, Kruger D. Engineering properties of epoxy resins used as concrete
adhesives. ACI Mater J 1996;93(1):2635.
[15] ASTMC882-87. Standard test method for bond strength of epoxy-resins used
with concrete.
[16] Dutta PK, Mosallam A. A rapid eld test method to evaluate concrete
composite adhesive bonding. Int J Mater Prod Technol 2003;19(12):5367.
[17] Cao Y, Cameron J. The effect of curing conditions on the properties of silica
modied glass ber reinforced epoxy composite. J Reinf Plast Compos
2007;26(1):4150.
[18] Islam MS, Pickering KL, Foreman NJ. Curing kinetics and effects of bre surface
treatment and curing parameters on the interfacial and tensile properties of
hemp/epoxy composites. J Adhes Sci Technol 2009;23(16):2085107.
[19] Gom, Software ARAMIS, version v.6.2. Gom GmbH: Braunschweig, Germany;
2009.
[20] Czaderski C, Soudki K, Motavalli M. Front and side view image correlation
measurements on FRP to concrete pull-off bond tests. J Compos Constr ASCE
2010;14(4):45163.
[21] Czaderski C, Rabinovitch O. Structural behavior and inter-layer displacements
in CFRP plated steel beams optical measurements, analysis, and comparative
verication. Compos Part B: Eng 2010;41(4):27686.
[22] Faella C, Martinelli E, Nigro E. Interface behaviour in FRP plates bonded to
concrete: experimental tests and theoretical analyses. In: Proceedings of the
international conference on advanced materials for construction of bridges
and other structures III. Davos (CH); 2005.
[23] Pellegrino C, Tinazzi D, Modena C. Experimental study on bond behavior
between concrete and FRP reinforcement. J Compos Constr 2008;12(2):1809.
[24] Faella C, Martinelli E, Nigro E. Direct versus indirect method for identifying
FRP-to-concrete
interface
relationships.
J
Compos
Constr
ASCE
2009;13(3):22633.

Vous aimerez peut-être aussi