Sie sind auf Seite 1von 10

Composite Structures 94 (2012) 26672676

Contents lists available at SciVerse ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Temporary bond strength of partly cured epoxy adhesive for anchoring prestressed
CFRP strips on concrete
Julien Michels a,, Christoph Czaderski a, Raafat El-Hacha b, Rolf Brnnimann c, Masoud Motavalli a,d
a

Structural Engineering Research Laboratory, Swiss Federal Laboratories for Materials Science and Technology (Empa), Dbendorf, Switzerland
Department of Civil Engineering, University of Calgary, Canada
c
Electronics and Metrology Laboratory, Swiss Federal Laboratories for Materials Science and Technology (Empa), Dbendorf, Switzerland
d
Faculty of Engineering, University of Tehran, Iran
b

a r t i c l e

i n f o

Article history:
Available online 9 April 2012
Keywords:
Carbon Fiber Reinforcement Polymer
Concrete strengthening
Prestressing
Gradient anchorage method
Accelerated epoxy curing
Temperature

a b s t r a c t
Externally bonded Carbon Fiber Reinforced Polymer (CFRP) strips have been used for strengthening reinforced concrete structures. This paper presents an experimental study on the debonding of externally
bonded CFRP strips anchored to a concrete substrate by a commercial epoxy adhesive. The study represents the basis for the characterization of an innovative gradient method, giving the possibility to anchor
prestressed CFRP strips to concrete without the use any mechanical anchorage systems such as plates and
bolts. Bond between the two components is achieved by an epoxy adhesive able to carry loads after an
accelerated curing process under elevated temperatures. The effect of heating conguration/duration,
strip thickness and bond length on the temporary bond resistance have been investigated using prestressed and non-prestressed CFRP-strips. Besides the optimization of the heating elements necessary
for the curing process, curing parameters for an optimal temporary bond strength could be determined.
Twenty-ve minutes of heating and curing at 90 C was found to be an optimum heating conguration,
resulting in better short-term mechanical performances than after conventional curing at room temperature for several days. The main reason is a temporarily softer adhesive which allows the use of the full
bond length by reducing shear force peaks.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
Externally bonded (EB) Carbon Fiber Reinforced Polymer (CFRP)
strips have been used extensively as an additional reinforcement for
strengthening reinforced concrete structures. Prestressing EB CFRP
strips generally increases the cracking load, delays yielding and
implicates a better exploitation of the mechanical performances,
such as high tensile strength above 2000 MPa. The prestressed CFRP
strips can be used more efciently and contribute to the load
carrying capacity under both service and ultimate conditions [1].
Combined with the delayed debonding failure [2] compared to
non-prestressed strips, the ultimate load capacity is generally
higher, too.
The following research is part of the development of a novel
anchorage system for prestressed CFRP strips. In contrast to the different developed mechanical anchorage systems described in literature [35], the idea of anchoring prestressed CFRP only with an
epoxy adhesive was proposed by Professor Urs Meier at the Swiss
Federal Laboratories for Materials Science and Technology (Empa)
(refer to [610]). In order to avoid a premature debonding of the
Corresponding author. Tel.: +41 58 765 4339.
E-mail address: julien.michels@empa.ch (J. Michels).
0263-8223/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compstruct.2012.03.037

strip from the concrete substrate, the gradient anchorage method


applies a repetitive local curing of the epoxy adhesive accelerated
by heat followed by a partial prestress release. The locally heated
epoxy is subsequently able to carry a portion of the prestress load.
By repeating the curing-releasing steps, the initial prestress force
can be gradually decreased to zero at the CFRP strip ends, thus
ensuring a proper force transfer from the concrete to the strip.
The aim of the experimental investigation was to determine the
temporary bond/anchorage resistance of CFRPepoxy-concrete
interfaces established by heating and curing under elevated temperatures for a limited time span (up to maximum 1 h) and compare to
results from reference specimens cured at room temperature for
several days. The short-term anchorage resistance was investigated
shortly after the accelerated curing process, whereas the reference
specimens, which did not undergo any external heating, were tested
after 23 days of curing at room temperature.
The mechanical bond properties of CFRP strips connected to concrete through an epoxy cured at room temperature was analyzed
among others by Mazzotti et al. [11], Lu et al. [12], Pellegrino et al.
[13], De Lorenzis et al. [14], Leung et al. [15] and Toutanji et al.
[16]. Temporary stiffness and strength of accelerated epoxy curing
has so far not found any large interest in structural strengthening.
Preliminary axial tensile and pull-off tests and the related analysis

2668

J. Michels et al. / Composite Structures 94 (2012) 26672676

Nomenclature
bf
ft,u
fc,cube
fct,s
lb
lb,eff
uf, uc

strip width
uniaxial tensile strength of CFRP
concrete compression strength on cube
concrete splitting tensile strength
bond length
effective anchorage length
horizontal displacement in x-direction of the CFRP strip
and the concrete
sf
relative displacement (slip) between CFRP strip and
concrete
sf,min, sf,max minimum and maximum slips in horizontal direction
tf
strip thickness
tcool
cooling time prior to the test start
x, y
directions, axis
thold
overheating time during the curing process at elevated
temperature

about the effect of high temperatures on the curing process of epoxy


adhesive were presented in Czaderski et al. [17]. It was found that for
the used adhesive, the rule of thumb given by Mays and Hutchinson
[18] and stating that the necessary curing duration is halved for
every increase of 10 C in curing temperature, can be applied.
2. Research signicance
In order to dene a correct gradient conguration, it is necessary to determine the inuence of a certain number of parameters
on the bond/anchorage resistance. Beside varying the strip thickness and the bond length, it is necessary to optimize the heating
elements and durations to maximize the bond strength. The presented number of tests with different geometrical characteristics
and heating conditions give precise information on the development of a temporary bond strength which can be exploited in
the production of the gradient anchorage force release.

3. Experimental characteristics
3.1. Materials
All CFRP strips used in the presented investigation had a width bf
of 100 mm, whereas three different strip thicknesses tf of 0.6, 0.9
and 1.2 mm were used. Before each test, elastic moduli Ef of the
used CFRP strips (in the ber direction) were determined by
prestressing the strip to several load stages and measuring the corresponding tensile elongation (to derive the strain by dividing the
elongation difference with the measuring length) by means of a
mechanical dial gauge (Table 1).
Direct tensile tests on coupon specimens according to EN ISO
527-5 [19] with strips of 0.6 and 0.9 mm thickness revealed average
uniaxial tensile strengths ftu of 2236 and 2324 MPa, respectively.
According to the quality analysis certicate of the producer, the
uniaxial tensile strength of 1.2 mm strip thickness is 2975 MPa.
Throughout all experiments the epoxy-based S&P epoxy adhesive
220 was used. According to the technical data sheet of the distributor
[20] it offers a compression strength above 90 MPa and bending tensile strength higher than 30 MPa after curing at room temperature.
Its glass transition temperature is higher than 56 C. Further detailed
information is available on the mentioned data sheet.
Concrete substrate, with a maximal aggregate size of 32 mm,
had an average compressive strength fc,cube of 51.1 MPa (tested

Ef
Fl
F1, F2
DF
Fu
Finit
PO
REL
RT
Ta
Tavg
Tn
Th
Tpeak

elastic modulus in ber direction


pulling force
prestress forces in the strip
prestress force release in the strip
ultimate debonding force
force at initiation of debonding
pull-off
releasing
room temperature
temperature in the adhesive
average temperature over the testing durations
nal temperature in the heating element
temperature in the heating element
maximum temperature in the heating element

according to EN 12390-3 [21]) and an average splitting tensile


strength fct,s of 3.3 MPa after 28 days.
3.2. Test setup
All experimental parameters are summarized in Table 1. The
different tests are listed in chronological order as they were performed in the structural laboratory at Empa. A photo of the test
setup is presented in Fig. 1.
The CFRP strip with a width of 100 mm is bonded to the concrete block (anchored against the reaction oor) over the bond
length lb with the epoxy adhesive. After the epoxy was applied
onto the strip, the latter was placed on the concrete block and
nally connected to the hydraulic jacks with clamps at both ends.
Thereafter, the CFRP strip was tensioned and the applied prestressing force was measured using load cells. The experimental campaign differentiates between two main testing congurations,
schematically presented in Fig. 2. The rst type, a Pull-off test
was performed by applying tensile load (on one side only) on the
initially unstressed CFRP strip until failure occurs. In the second
type (Releasing-test), the prestressing force Fp was applied to the
strip (with one jack at each side) prior to adhesive curing, followed
by a progressive force release of one jack until debonding of the
strip from the concrete occured. Prestressing force on the other
jack was kept constant, however, a certain decrease due to the
system stiffness could be observed (refer to Fig. 4b). The tests
started shortly after stopping the heating procedure (approximately 1015 min) in order to partly cool down the adhesive. In
both cases, force increase or decrease was applied at a rate of
approximately 1 kN/s. The mean temperature of the adhesive over
the test duration, measured by inserting thermocouples underneath the CFRP strip in the adhesive layer equally distributed along
the bond length, is given in Table 1. It is noted that the mentioned
mean temperature is not considered as an inuence parameter in
the research presented in this paper.
3.3. Image Correlation System (ICS)
In most research campaigns about CFRP-concrete debonding
phenomena (e.g. [15,16]), strain measurements are only performed
in the central CFRP strip axis, and no information about the
displacement and strain variation in transverse direction is obtained. The Image Correlation System measurements (ICS) with

2669

J. Michels et al. / Composite Structures 94 (2012) 26672676


Table 1
Experimental characteristics of the pull-off and releasing tests.

a
b

Type

tf (mm)

lb (mm)

Heating system

th (min)

tcool (min)

Tavg (C)

Ef (GPa)

Finit or Fu (kN)

sf,min (mm)

sf,max (mm)

PO
PO
PO
PO
REL
REL
REL
REL
REL
REL
REL
REL
REL
REL
REL
REL

0.6
0.6
0.6
0.6
0.6
0.6
0.6
1.2
1.2
1.2
1.2
1.2
0.9
0.9
0.6
1.2

300
300
300
300
300
300
300
300
200
100
300
200
300
200
300
300

A
RT curing
B
B
C
C
C
C
C
C
C
C
C
C
C
RT curing

15

15
20
30
60
15
25
25
25
25
25
25
25
25

6.6
2 days
6.8
10.7
40.0a
7.7
9.8
15.9
9.2
7.5
18.1
10.4
9.8
10.5
13.9
3 days

39.5
19.2
42.8
40
32.5
53.2
46.2
39.6
45.4
47.9
36.7
42.6
41.8
40.5
37.7
20b

138.1
138.2
131.9
132.9
141.7
133.4
134.6
159.3
157.7
159.2
158.4
158.7
148.2
149.5
143.3
156.5

14.8
35.0
52.2
55.6
44.8
40.3
38.3
58.9
45.0
23.0
67.0
51.2
55.7
35.1
44.6
46.8 (Finit)
57.7 (Fu)

1.14
0.00
0.20
0.10
0.02
0.02
0.18
0.15
0.24
0.17
0.04
0.06
0.36
0.09
0.09
0.00
0.04

1.43
0.20
1.37
1.26
1.12
0.92
1.10
0.72
0.53
0.24
0.57
0.38
1.07
0.34
1.00
0.31
0.57

Longer cooling time due to initial issue with image correlation measurement.
Room temperature not measured.

Fig. 2. Schematic description (side view) of the (a) pull-off and (b) releasing tests.

Fig. 1. Test setup for the pull-off and releasing tests.

Aramis software by Gom [22] allows to measure the 3-D full-eld


displacements at various loading stages (see also [2326]).
3.4. Heating procedures
3.4.1. General comments
The target curing temperature of the epoxy adhesive was chosen
to be approximately 90 C [17]. For the heating system the
gradient region is divided in 100  100 mm2 areas which can be
heated individually. The heating system and the different components are shown in Fig. 3. Heating elements of the type TermofoilTM
(Fa. Minco) are used. During tests in the laboratory, thermocouples
sensors are directly embedded in the epoxy adhesive to measure its
temperature Ta. For each test specimen, one sensor is used per
100  100 mm2 of epoxy surface. However, this is impractical for
regular use on site. Therefore, it was decided to measure also the

temperature in the heating element. The conductor in the heating


element is made of Nickel with a temperature coefcient of the
resistance of about half a percent. The dependence is used to
directly measure the temperature Th in the heating element. The
difference between the two temperatures Ta and Th measured during tests can be used to estimate the required temperature in the
heating element for a target temperature in the adhesive. This
difference changes over time. Whilst the temperature in the heating
element can increase quite fast, the adhesive temperature rises
slower due to the heat conductivity and capacity of the various
materials.
During this study on different structural parameters, the heating elements and the related controlling device were improved,
too. Eventually, three different heating congurations have been
applied during the various tests.
The general heating procedures together with the loading and
releasing steps in time are presented in Fig. 4.
3.4.2. Various heating systems
3.4.2.1. System A. System A is the most basic conguration. The
temperature in the heating elements (Th) is manually controlled
by adjusting the current supply. The disadvantages are the slow
temperature rise and the manual temperature adjustment.

2670

J. Michels et al. / Composite Structures 94 (2012) 26672676

Fig. 3. (a) Heating procedure for System C, (b) bottom view of the heating elements and (c) components necessary for the adhesive heating.

Fig. 4. Heating and loading procedure in time for (a) pull-off and (b) releasing tests.

3.4.2.2. System B. It was found that the heating elements from


System A showed a homogeneous heating over the total element
width. This implicated an inhomogeneous adhesive curing due to
the heat losses at the strip edges [17]. However, new heating elements with the aim of a more homogenous curing over the total
surface are introduced. The heat ow in the center of the heating
element is mainly in the direction normal to its surface, whereas
at the border the heat ows also sideward. The effective heat conductivity is therefore higher at the border resulting in considerably
lower temperature compared to the center and inhomogeneous
curing of the adhesive. A special heating element with substantially higher heating power at the border compared to the center
was developed that compensates these effects. Similar to System
A, the user still has to manually operate the heating device.
3.4.2.3. System C. For this system, an automatic temperature controlling device is used to control the heating elements. As the temperature evolution in the adhesive (Ta) is retarded with regard to
the heating elements themselves and in order to obtain a relatively
constant temperature of 90 C in the adhesive (Ta), the heating

elements have to heat with a temperature clearly above 90 C


(Th). A modied system was developed to control the heating
elements. In order to optimize the heating effect, the rise of the
adhesive temperature (Ta) is accelerated by initial overheating. At
the beginning of the curing process the temperature in the heating
element is set to Tpeak at a holding time thold (10 min). Then, an
exponential decay to the temperature Tn follows (for illustration
see Fig. 3a. Various heating tests revealed satisfying results when
Tpeak was set to 160 C and Tn to 120 C.
4. Results and discussion
4.1. Denitions
The results of the experimental investigation on the mentioned
pull-off and releasing tests are presented in this section. The
parameters affecting the bond behavior are evaluated and analyzed
in detail including the type of heating system and its control
(Systems AC) and the heating duration th. Furthermore, the inuence of the bond length lb and the strip thickness tf is documented

J. Michels et al. / Composite Structures 94 (2012) 26672676

2671

at the end of the section. Besides the force measurements, slips sf at


several load stages are obtained by means of digital image correlation. The slip is dened as the relative displacement between the
CFRP strip and the neighboring concrete surface (see (1)).

sf uf  uC

with sf is the slip between CFRP strip and concrete; uf the horizontal
displacement (x-direction) of the CFRP strip in Section 1; and uc is
the average horizontal displacement (x-direction) of the concrete
in Section 0 and 2 (see Figs. 5 and 6).
For the present analysis, the CFRP displacement in horizontal
direction (pulling or releasing direction) was evaluated along the
central axis of the strip (Section 1, see Fig. 5), whereas the concrete
displacements were measured in two sections (Section 0 and 2) at a
distance of 15 mm from the strip edge on each side. Fig. 6 presents
the calculation method to obtain the slip values from the experimental measurements. For the evaluation of the slip sf, concrete
displacement uc was obtained by averaging two linear regression
slopes through the measured values in Section 0 and 2. In Table 1,
minimal and maximal slip values sf,min and sf,max at the last stage
of failure are presented for each experimental test.
It has to be mentioned at this point that the slips for the specimens cured at elevated temperature were evaluated at the last
stage of failure, whereas the initiation of debonding-state as
described in Czaderski et al. [24] was considered for the specimens
cured at room temperature. The differences in behavior at debonding between the two categories will be discussed in the following
sections.
4.2. Optimizing the heating elements
Accelerated curing of the epoxy adhesive induces signicant
changes in the horizontal displacement behavior over the strip
width. It can be seen from Fig. 7a that, using the heating System
A, the edges of the strip exhibit larger displacements than the inner
part of the strip. This is due to a less cured and thus less stiff adhesive in the border regions as a consequence of lower temperature
development at the strip edges. This phenomenon is because the
used heating elements from System A offer a homogeneous heat
power over the complete surface, but simultaneously have higher
heat loss at the strip border as explained above. This results in a
different slip behavior than for room temperature-cured specimens, generally exhibiting stronger displacement values in the
central axis [24]. The heating elements with higher border heat
power (System B) induce a more homogeneous adhesive curing,
resulting in a more homogeneous strip displacement in the pulling
direction as shown in Fig. 7b.

Fig. 6. Slip denition between the CFRP strip and the concrete substrate (values at
failure, heating System A, th = 15 min, lb = 300 mm, tf = 0.6 mm).

In addition to the changes in displacement patterns, considerable difference in the debonding force and failure mode can be
observed. Whereas System A with a heating time of 15 min results
in an adhesive failure (Fig. 7a) at a force of only 14.8 kN over a
bond length of 300 mm, the same curing duration with System B
allows the development of a partial concrete failure (see Fig. 7b)
at a force of 52.2 kN, representing a relative enhancement of 253%.
The mentioned difference between heating Systems A and B can
also be explained by analyzing the slip values measured at failure.
As shown in Table 1, it can be concluded that 15 min of curing time
with heating System A exhibits clearly larger slips (sf,min = 1.14 mm,
sf,max = 1.43 mm) than the same testing conguration with heating
System B (sf,min = 0.2, sf,max = 1.35 mm). This is due to a better curing
process and the related transition from a pure adhesive to a mixed
adhesive/concrete failure.
4.3. Optimization of the heating time and comparison with curing
under room temperature
Fig. 8 presents the debonding force versus the heating duration
of three different testing congurations. For temporary curing at
elevated temperature, the debonding resistance increases with
increasing the heating duration with an optimum ranging between
25 and 30 min. For the pull-off test with tf = 0.6 mm, a slight increase of 6.5% in the debonding force is observed when increasing
the heating duration from 15 to 20 min (Fig. 8a). The releasing test
with the same strip type, the debonding force increases by 16.5%
when curing time increases from 15 to 25 min, however almost
no change in the debonding force was observed when the curing

Fig. 5. Section denition for the data analysis.

2672

J. Michels et al. / Composite Structures 94 (2012) 26672676

Fig. 7. Failure surface and ICS results for displacements in pull-off direction at failure for (a) heating Systems A and (b) heating System B, tf = 0.6 mm, lb = 300 mm, th = 15 min.

time increases from 25 to 30 min (Fig. 8b). However, 1 h of high


temperature exposure decreases the debonding force in the releasing test of about 10% compared to the peak force observed after 25
and 30 min. Fig. 8a and c shows that room temperature curing also
lowers the debonding resistance by 63% and 86% when compared
to the values obtained after 20 and 25 min, respectively. For the
releasing test with a strip thickness tf = 1.2 mm and a bond length
lb = 300 mm, it is specied that the rst test (tagged with  in
Fig. 8c) with a maximum debonding force of 58.9 kN exhibited a
considerable prestress force loss of about 5 kN during the temporary adhesive curing due to a pressure loss in the hydraulic jacking
system. Hence, this test was repeated and a clearly higher resistance (67 kN) was obtained.
4.3.1. Slip behavior for pull-off tests
When analyzing the force-slip relations presented in Fig. 9a,
interesting information is obtained. Firstly, for the pull-off test
with a strip thickness of 0.6 mm, a much stiffer behavior for the
room temperature (RT) cured specimen is observed in comparison
to both 15 and 20 min curing at elevated temperatures. Similar to
the experimental results presented in Czaderski et al. [24], initiation of debonding for the RT cured specimen occurs at a slip of about
0.2 mm, afterwards almost no increase in force can be noticed up
to failure. The additional slip values after initiation of debonding
are due to elastic strip deformation. On the contrary, both temporarily cured specimen at elevated temperature show an increasing
force evolution up to nal failure. By analyzing Fig. 10, presenting
the slip distribution over the total bond length at initiation of debonding (RT curing) and failure level (elevated temperature curing),
respectively, it gets obvious that the softer adhesive, only partly

cured after 15 and 20 min, integrates the total bond length into
the anchorage process before failure. The stiffer adhesive after
room temperature curing on the other hand implicates that only
a reduced efcient anchorage length lb,eff of about 125 mm, smaller
than the total bond length, can be involved.
4.3.2. Slip behavior for releasing tests
Fig. 9b shows the force-slip relations for the releasing tests with
a strip thickness of 1.2 mm. Similarly to the presented pull-off tests,
the room temperature cured specimen reveals a clearly stiffer
behavior than its counterpart, temporarily cured for 25 min at high
temperature. However, an important difference between pull-off
and releasing for specimens cured at room temperature can be
pointed out: whilst no further force increase in pull-off test is possible anymore after initiation of debonding, a signicant additional
resistance can be activated after rst crack appearance (see Fig. 9b).
Fig. 11 presents several load stages recorded with the ICS of the
slips sf in the horizontal direction plotted against the total bond
length. The results were recorded with a measurement rate of
1 Hz, which means that there is a time span of 1 s between two consecutive load stages. It can be seen that slip curves for load stages 91
and 92 almost coincide, whereas the transition to load stage 93 is
characterized by a sudden vertical shift in the curves. At this moment, the rst crack appears. The maximum shift in displacement
Ds is pointed out in Figs. 9 and 11. Further anchorage resistance
after rst cracking can be observed. This phenomenon is due to
the particular crack evolution taking place deeper in the concrete
substrate and the resulting crack opening mechanisms. Hence, the
force value of 46.8 kN at loadstage 92 can be increased by10.9 kN
up to the maximum force of 57.7 kN at loadstage 162. Another

J. Michels et al. / Composite Structures 94 (2012) 26672676

2673

Fig. 9. Forceslip relations for the (a) pull-off with tf = 0.6 mm and (b) releasing
tests with tf = 1.2 mm.

Fig. 10. Slips sf at initiation of debonding or failure level for pull-off with
tf = 0.6 mm and lb = 300 mm.
Fig. 8. Inuence of the heating time th on the anchorage resistance and comparison
with the specimen cured at room temperature ( = prestress force loss of about 5 kN
during heating procedure).

interesting observation from Fig. 11 is the constantly growing participation of the bond length to the shear stress transfer with growing releasing force.
For the releasing specimen cured at elevated temperatures, no
sudden shape change in the slips curve was observed, a continuous
transition between the different load stages was noticed (see

Fig. 9b). As an example in Fig. 12, two intermediate load stages


(50 and 100) and the correspondent force levels are presented in
addition to the nal load stage 115 with Fu. In this case, no sudden
slip increase (see also Fig. 9b) at rst crack appearance is noticed,
but rather a continuous force rising up to failure is observed. In
Fig. 13, the stages at rst crack appearance and failure level for
the specimen cured at room temperature are compared with the
slip at ultimate failure for the specimen cured for 25 min at elevated temperature. Even though differences are not as signicant

2674

J. Michels et al. / Composite Structures 94 (2012) 26672676

Fig. 11. Slips sf for releasing test cured at room temperature at different load stages,
tf = 1.2 mm, lb = 300 mm.

Fig. 13. Slips sf at rst crack appearance (RT curing) and failure state (RT curing and
25 min accelerated curing) for releasing tests/tf = 1.2 mm, lb = 300 mm/red curve of
specimen with 25 heating time corresponds to stage 115 in Fig. 12. (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)

much as possible the total necessary application time. The enhanced short-term anchorage resistance for a dened bond length
allows anchoring higher total prestress forces during the application period. It is worth noticing that, as the curing process in the
adhesive continues, stiffness of the epoxy (initially cured at high
temperature!) rises and eventually reaches the same level than
for the corresponding tests at room temperature [17]. To conclude,
the present investigation exhibits the double advantage the gradient method offers: adhesive curing at high temperature considerably shortens the necessary time span for obtaining sufcient
anchorage capacity and simultaneously, due to the temporary lower epoxy stiffness, even enhances the total resistance values.
4.4. Gradient design

Fig. 12. Slips sf for releasing test temporarily cured at elevated temperature,
tf = 1.2 mm, lb = 300 mm, th = 25 min.

as for the earlier presented pull-off tests, temporary curing at high


temperatures also induces larger slips in the releasing direction
compared to room temperature curing.
To summarize, one can conclude that temporary heating of the
epoxy at elevated temperature accelerates its curing and induces a
faster strength development. However, the obtained stiffness after
approximately 25 min of curing at 90 C is still clearly lower than
after 3 days curing at room temperature (Fig. 9). Similar tendencies
with pull-off tests on strips with a thickness of 1.2 mm are presented in Czaderski et al. [17]. The partly cured adhesive has the
advantage of being able to better distribute the bond stresses over
the total bond length and thus provoke a higher anchorage resistance. After 1 h of heating, the stiffness has already passed a certain
threshold value after which the maximum debonding force already
starts decreasing (Fig. 8). After the mentioned 25 min of curing
however, the adhesive has developed a sufciently high shear
strength to initiate an almost complete failure in the concrete
substrate (Fig. 14) and simultaneously offers a stiffness low enough
to maximize the efcient anchorage length and anchorage resistance. This effect is particularly benecial when applying the gradient anchorage method for instance. The goal of this technique
is to maximize anchorage capacity by simultaneously reducing as

Optimization of the heating elements and the necessary heating


time show that 25 min of curing duration at 90 C represents an
optimum value in order to maximize the anchorage resistance
when using this specic type of epoxy adhesive. For practical applications of the gradient anchorage method, CFRP strips with 0.9 and
1.2 mm thickness can be considered. With both strip types, the
force gradient at the strip ends will be established with 200 and
300 mm bonded sections, respectively. Fig. 15 presents the temporary anchorage resistances of the releasing tests with a strip thickness tf = 0.9 mm (a) and tf = 1.2 mm (b). It can be concluded from
the graphs in Fig. 15a and b that both increasing strip thickness
and bond length enhance the anchorage resistances of the partly

Fig. 14. Debonding failure in the concrete substrate, tf = 1.2 mm, lb = 300 mm,
releasing test, th = 25 min.

J. Michels et al. / Composite Structures 94 (2012) 26672676

2675

Fig. 15. Inuence of the bond length on the temporary anchorage resistance in releasing for (a) tf = 0.9 mm and (b) tf = 1.2 mm/th = 25 min ( = tests performed after a strong
prestress loss of about 5 kN during the heating process).

cured epoxy after temporary heating. Further experiments on longer bond lengths could be useful in order to investigate possible
asymptotic values for the temporary anchorage resistance (and
its corresponding anchorage length) after which an increase in
the bond length does not inuence the anchorage resistance anymore. Eventually, the experiments for the two different strip thicknesses of 0.9 and 1.2 mm and the respective bond length of 200
and 300 mm show the possible anchorage resistances for the practical application of the gradient method.
5. Conclusions and outlook
Locally accelerated curing of epoxy adhesives is essential for the
proposed gradient method for anchoring prestressed CFRP strips
and eliminating the use and or keeping the mechanical anchorage
systems (plates and bolts) at the ends of the strips. The curing by
sequentially heating the individual sectors must be completed
before ambient temperature curing progresses too far. However,
the temperature is limited to avoid heat induced deterioration of
the involved materials. With the developed heating elements, the
temperature gradients over the strip width could be signicantly
reduced. This allows heating the adhesive layer more uniformly,
reducing the curing time. The developed temperature control system accelerates heating the adhesive, further reducing the curing
cycle time.
In order to dene a reasonable force release for the gradient
anchorage, different pull-off and releasing tests allow to summarize adequate application congurations. With optimized heating
elements, it was shown that 25 min of curing time (at 90 C adhesive temperature) represents an optimum time interval regarding
temporary bond strength. After this heating duration, the adhesive
possesses enough strength to transfer the failure from the adhesive
to the concrete substrate with a considerably higher anchorage
resistance, and simultaneously offers a stiffness low enough to
activate the whole bond length for increasing the anchorage resistance. After a longer curing process, the stiffness of the adhesive
would increase and therefore the effective anchorage length decreases, resulting in a lower anchorage resistance. Eventually, the
temporary anchorage resistances for the used strips with 0.9 and
1.2 mm thickness were determined for two different bond lengths
of 200 and 300 mm and thus the force gradient can be designed.
Attention is drawn to the fact that the presented conclusions
are based on one specic concrete strength type. Lower tensile

strength would induce lower debonding force and thus also


shorten the necessary curing duration. Further investigation would
be necessary in order to determine the respective optima in curing
time at high temperatures dependent on the substrate quality.
However, the authors recommend not applying the presented
anchorage method on a low strength concrete.
Acknowledgments
The authors would like to express their gratitude to the industrial partner S&P Clever Reinforcement from Switzerland, to the
Swiss innovation promotion agency (CTI Project No. 10493.2
PFIW-IW) and to the staff of the Structural Engineering Testing
Laboratory at Empa.
References
[1] El-Hacha R, Wight RG, Green MF. Prestressed bre-reinforced polymer
laminates for strengthening structures. Prog Struct Eng Mater J 2001;3:
111121.
[2] Michels J, Czaderski C, Motavalli M. Prestressed CFRP for structural
strengthening. In: SMAR rst middle east conference on smart monitoring,
assessment and rehabilitation of civil structures, Dubai, UAE; 2011.
[3] El-Hacha R, Wight RG, Green MF. Innovative system for prestressing berreinforced polymer sheets. ACI Struct J 2003;100(3):30513.
[4] Pellegrino C, Modena C. Flexural strengthening of real-scale RC and PRC beams
with end-anchored pretensioned FRP laminates. ACI Struct J 2009;106(3):
319328.
[5] Kim YJ, Wight RG, Green MF. Flexural strengthening of RC beams with
prestressed CFRP sheets: development of nonmetallic anchor systems. J
Compos Construct 2008;12(1):3543.
[6] Meier U, Stcklin Y. A Novel carbon ber reinforced polymer (CFRP) system for
post-strengthening. In: International conference on concrete repair,
rehabilitation and retrotting (ICCRRR), Cape Town, South Africa; 2005.
[7] Czaderski C, Motavalli M. 40-Year-old full-scale concrete bridge girder
strengthened with prestressed CFRP plates anchored using gradient method.
Compos B: Eng 2007;38(78):87886.
[8] Aram MR, Czaderski C, Motavalli M. Effects of gradually anchored prestressed
CFRP strips bonded on prestressed concrete beams. J Compos Construct
2008;12(1):2534.
[9] Motavalli M, Czaderski C, Pfyl-Lang K. Prestressed CFRP for strengthening of
reinforced concrete structures: recent developments at Empa, Switzerland. J
Compos Construct 2011;15(2):194205.
[10] Kotynia R, Walendziak R, Stoecklin I, Meier U. RC slabs strengthened with
prestressed and gradually CFRP strips under monotonic and cyclic loading. J
Compos Construct 2011;15(2):16880.
[11] Mazzotti C, Savoia M, Ferracuti B. A new single-shear set-up for stable
debonding of FRP-concrete joints. Construct Build Mater 2009;23(4):1529
1537.
[12] Lu XZ, Teng JG, Ye LP, Jiang JJ. Bond-slip models for FRP sheets/plates bonded to
concrete. Eng Struct 2005;27(6):92037.

2676

J. Michels et al. / Composite Structures 94 (2012) 26672676

[13] Pellegrino C, Tinazzi D, Modena C. Experimental study on bond behavior


between concrete and FRP reinforcement. J Compos Construct 2008;12(2):
180189.
[14] De Lorenzis L, Miller B, Nanni A. Bond of ber-reinforced polymer laminates to
concrete. ACI Mater J 2001;98(3):25664.
[15] Leung CKY, Klenke M, Tung WK, Luk HCY. Determination of nonlinear
softening behavior at FRP composite/concrete interface. J Eng Mech
2006;132(5):498508.
[16] Toutanji H, Saxena P, Zhao L, Ooi T. Prediction of interfacial bond failure of FRPconcrete surface. J Compos Construct 2007;11(4):42736.
[17] Czaderski C, Martinelli E, Michels J, Motavalli M. Effect of curing conditions on
strength development in an epoxy resin for structural strengthening. Compos
B: Eng 2011;43(2):398410.
[18] Mays GC, Hutchinson AR. Adhesives in civil engineering. Cambridge University
Press; 1992.
[19] EN ISO 527-5/Kunststoffe Bestimmung der Zugeinschaften Teil 5:
Prfbedingungen fr unidirektional faserverstrkte Kunststoffverbundwerkstoffe, Brussels; 1997.

[20] S&P Clever Reinforcement. <www.sp-reinforcement.ch>.


[21] EN 12390-3 DE. Testing hardened concrete Part 3: compressive strength of
test specimens. German version; 2011.
[22] Gom. Software ARAMIS, version v. 6.2, Braunschweig, Germany: Gom GmbH;
2009.
[23] Czaderski C, Rabinovitch O. Structural behavior and inter-layer displacements
in CFRP plated steel beams optical measurements, analysis, and comparative
verication. Compos B: Eng 2010;41(4):27686.
[24] Czaderski C, Soudki K, Motavalli M. Front and side view image correlation
measurements on FRP to concrete pull-off bond tests. J Compos Construct
2010;14(4):45163.
[25] Subramaniam KV, Carloni C, Nobile L. Width effect in the interface fracture
during shear debonding of FRP sheets from concrete. Eng Fract Mech
2007;74(4):57894.
[26] Carloni C, Subramaniam KV. Direct determination of cohesive stress transfer
during debonding of FRP from concrete. Compos Struct 2010;93(1):18492.

Das könnte Ihnen auch gefallen