Sie sind auf Seite 1von 12

Construction and Building Materials 32 (2012) 110121

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Review

Mechanical anchorage of FRP tendons A literature review


Jacob W. Schmidt a,c,, Anders Bennitz b, Bjrn Tljsten b, Per Goltermann a, Henning Pedersen c
a

Technical University of Denmark, Division of Structural Engineering, DTU-Building 116, Brovej, DK-2800 Kgs. Lyngby, Denmark
Lule University of Technology, Division of Structural Engineering, SE-971 87 Lule, Sweden
c
COWI A/S, Department 1704, Operation Management & Systems, Parallelvej 2, DK-2800 Kgs. Lyngby, Denmark
b

a r t i c l e

i n f o

Article history:
Available online 2 February 2012
Keywords:
Prestressing
Posttensioning
FRP
Anchorage
Wedge
FEM analysis
Laboratory tests

a b s t r a c t
High tensile strength, good resistance to degradation and creep, low weight and, to some extent, the ability to change the modulus of elasticity are some of the advantages of using prestressed, unidirectional
FRP (Fibre Reinforced Polymer) tendon systems. Bonded and non-bonded versions of these systems have
been investigated over the last three decades with results showing that prestressing systems can be very
efcient when the FRP properties are properly exploited. However, there are often concerns as to how to
exploit those properties to the full and how to achieve reliable anchorage with such systems. This is especially important in external post-tensioned tendon systems, where the anchorage points are exposed to
the full load throughout the life span of the structure. Consequently, there are large requirements related
to the long-term capacity and fatigue resistance of such systems. Several anchorage systems for use with
Aramid, Glass and Carbon FRP tendons have been proposed over the last two decades. Each system is usually tailored to a particular type of tendon. This paper presents a brief overview of bonded anchorage
applications while the primary literature review discusses three methods of mechanical anchorage:
spike, wedge and clamping. Some proposals for future research are suggested. In general, the systems
investigated showed inconsistent results with a small difference between achieving either a successful
or an unsuccessful anchorage. These inconsistencies seem to be due to the brittleness of the tendons,
low strength perpendicular to the bre direction and insufcient stress transfer in the anchorage/tendon
interface. As a result, anchorage failure modes tend to be excessive principal stresses, local crushing and
interfacial slippage (abrasive wear), all of which are difcult to predict.
2012 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.
4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Tendon properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
AFRP systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
GFRP systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
CFRP systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Anchorage and tendon properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.1.
Bonded anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.2.
Mechanical anchorages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.3.
Spike anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.4.
Wedge anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.5.
Clamping anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Slip at the anchorage/tendon interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Evolution of analytical and finite element (FE) models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Finite element (FE) models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author.
E-mail addresses: jws@byg.dtu.dk (J.W. Schmidt), anders.bennitz@ltu.se
(A. Bennitz), bjorn.taljsten@ltu.se (B. Tljsten), pg@byg.dtu.dk (P. Goltermann),
hpe@cowi.dk (H. Pedersen).
0950-0618/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.conbuildmat.2011.11.049

111
111
111
112
112
112
113
114
114
114
116
116
117
118

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

111

Nomenclature
D1
dB
F1
F
FTW
FWB
fu
h
P
Pea
Pfa
RTW
RWB
t

5.

length of region 1 and turning point to region 2


inner diameter of barrel
maximum load of region 1
applied tensile force
resulting force vector from the tensioned tendon
resulting force vector between barrel and wedge
ultimate tensile capacity
height
prestressing or clamping force
load carried by the epoxy
load carried by the bres
friction force between tendon and wedge
friction force between wedge and barrel
thickness

tB
lB
d
m
mea

thickness of the barrel


barrel length
deformation
effective coefcient of friction
coefcient of friction for epoxy in contact with aluminium

mfa

coefcient of friction for bre in contact with aluminium

mT
mWB

coefcient of friction
coefcient of friction between wedge and barrel
stress
circumferential stress
internal pressure

r
rT
rWB

Discussion and future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


5.1.
Reliability and durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.
Material properties and geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
In recent decades, there have been several proposals for the design of external post-tensioned systems for FRPs (Fibre Reinforced
Polymers) plates [110], and NSMR (Near Surface Mounted Reinforcement) bars [11]. These methods all have their origins in
bonded non-prestressed FRP systems where, often, a concrete fracture manifests as intermediate crack debonding (IC debonding) or
end peeling as a result of the FRPs high strength. Prestressing utilises the strength of the FRP material to a higher degree. A failure of
the structure is then more likely to be due to a compressive failure
in the concrete or a tensile failure in the tendon. Another type of
failure that may also occur is slippage in the anchorage area. Proper
anchorage seems to be the decisive factor in these systems, in
ensuring reliable force transfer and interaction with the rest of
the structure. To the authors knowledge, no FRP system has yet
been developed that is economically and practically competitive
compared with existing steel post-tensioning systems. Steel prestressing systems are well documented and tested, providing safe
anchorages that can carry required capacities and be sufciently
reliable. FRP materials may be produced with a range of different
strength and stiffness properties which means it has potential
use in a variety of applications. FRP materials share identical orthotropic properties, linear elasticity and brittleness, unlike steel,
which is isotropic and has plasticity. The literature describes sev-

Fig. 1. Stress/strain curve showing FRP and steel varieties [12,13].

119
119
119
120
120

eral types of tendon and related anchorages that have been developed as alternatives to conventional steel systems. This literature
review describes the most commonly used types of FRP tendon
and charts the advances in the anchorage of FRP tendons over
the years. In addition, this review explains some of the research
that has been carried out on such systems.
2. Tendon properties
In composite prestressing systems, there are three materials that
are mainly used to build tendons: AFRP (Aramid FRP), GFRP (Glass
FRP) and CFRP (Carbon FRP). As shown in Fig. 1, the tensile properties of FRP materials can vary signicantly depending on the bre
material, bre fraction and resin type. As a consequence, the
modulus of elasticity within each material group can be varied,
something that is not possible with steel. However, the yield capability of steel would be advantageous, since it would provide ductility in the structure at the ultimate limit state.
2.1. AFRP systems
One of the rst AFRP tendons manufactured was the Paral
rope, which was developed in the 1960s to moor navigation platforms in the North Atlantic [14]. This type of tendon does not contain any matrix and consequently cannot be bonded to the
structure. Burgoyne conducted extensive research on Paral rope
with the core yarn made of Aramid (Kevlar 49) [15,16]. There are
three versions of Paral ropes, each with a different kind of core:
Type A, Type F and Type G; the latter is normally used for structural applications due to its high modulus of elasticity. The research showed that the tendon itself had excellent fatigue
performance, but was fragile when sheave bending fatigue tests
were carried out, as the area being bent fractured due to stressing
of the outer bres. Research into the long-term behaviour of AFRP
is still ongoing. Long-term stress rupture is one of the main reasons
why engineers are reluctant to adopt AFRP tendons for prestressing
and stay cable purposes. Long-term accelerated testing using both
stepped isothermal and stepped isostress methods is therefore

112

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

used and has been shown to be able to predict the long-term


changes in the properties of Aramid tendons; these testing methods simulate more than 100 years of use of the tendons [17]. Other
types of AFRP tendons used are Arapree, reported on more extensively by Gerritse et al. [1821], FiBRA [22] and Technora [27,29].
2.2. GFRP systems
Research on GFRP started in the 1970s, when several bridges
were prestressed using GFRP cables in Germany and Austria. This
prestressing was carried out to compensate for the low modulus
of elasticity of the GFRP. One of the rst structures constructed
using GFRP technology was a two-span continuous highway bridge
in Dsseldorf, Germany in 1986, which was post-tensioned using
GFRP cables. Each pultruded GFRP rod used in the bridge had a
diameter of 7.5 mm. In total, 19 of these rods were used in the tendon resulting in a post-tensioning force of 600 kN per tendon. The
system used an epoxy-socket anchor, resulting in the tendons
needing to be produced in pre-determined lengths. However, the
type of anchorage that was used led to a large anchorage size
and there were further concerns about the long-term reliability
[23]. Further, Iyer et al. [24,25] described a one-span demonstration bridge that was prestressed using CFRP, GFRP and steel cables.
The span width was notionally separated into thirds and each third
was prestressed with each type of cable. This demonstrated that
CFRP and GFRP cables worked in this application and that any
short-term changes of their properties were predictable. Moreover,
Neptco Inc. has produced a cable consisting of seven GFRP rods
(one central rod surrounded by six others). The individual rods
are twisted together and held in place by plastic straps or with a
resin binder [26].
2.3. CFRP systems
CFRP has superior creep properties, low relaxation, high
strength and can withstand high jacking stress, but at high cost
[27]. One of the rst products developed was the Leadline CFRP
tendon, which is produced with different surface patterns: smooth,
with surface-indents or ribbed. These rods are usually pultruded
and pitch-based (manufactured from resin precursor bres after
stabilization treatment, carbonisation and a nal heat treatment)
and most commonly anchored using a modied wedge anchorage
system [14]. A similar CFRP is the Technora carbon bre rod, a spi-

rally-wound rod that is pultruded and impregnated with a vinyl


ester resin [27,29]. Carbon stress rods are also made by pultrusion,
then epoxy-impregnated and usually coated with sand to increase
their anchorage when embedded in concrete [26]. CFCC (Carbon Fibre Composite Cable) tendons are another type of tendon and are
in its nal stage produced like conventional steel strands, where
several smaller steel strands form one cable [28]. A CFCC tendon
consists of bundles of cables with multiple pieces of prepreg
(semi-hardened tows/strands with a resin precursor) that are treated with a suitable coating [27].
Grace conducted extensive research on CFRP tendons, performing internal prestressing of a double-tee (DT) bridge system that
used CFRP tendons [30]. The DT was reinforced with GFRP bars
and externally prestressed with CFRP tendons. Two deviation
points and a crossbeam were placed at the midspan, where an
externally post-tensioned and draped CFCC strand was applied.
The magnitude of the draping angle and the deviator diameter
were shown to affect the behaviour of the system [30,31]. Also,
there are reports about larger, real life applications that have used
external CFRP prestressing and poststressing of bridges [32,33]. A
DT Beam was in [32] designed with pretensioned Leadline tendons and post-tensioned CFCC strands to simulate the performance
of DT-beams which had been used in the construction of a 3-span
bridge. This design was rst used in a prestressed and poststressed
CFRP bridge in the United States. It was reported that the ultimate
failure mode was due to separation between the ange and the
beam web which led to crushing of the ange and the rupture of
the original prestressed CFRP tendons.

2.4. Anchorage and tendon properties


In order to exploit the properties of an FRP tendon fully, the
stresses that develop in the interface between the anchorage and
the tendon have to be properly managed. When measuring tendon
behaviour, therefore, the anchorage must be sufciently robust. As
shown in Table 1, anisotropic FRP material, coupled with brittleness and weak transverse properties makes the issue of successful
anchorage a difcult one, with the most common outcome being
premature failure. Over the last few decades, there have been several bonded and mechanical anchorage designs, often tailored to a
particular type of tendon. The following section briey reviews
bonded anchorages as an alternative to mechanical solutions,
which are then described in greater depth. A review of the evolu-

Table 1
Examples of GFRP, AFRP, CFRP properties compared to steel, after [27,34]. Shaded rows show where FRP properties are signicantly weaker than steel prestressing steel.

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

113

Fig. 2. Example of contoured and straight sleeve anchorages.

tion of analytical models and nite element models to describe


mathematically the function of an anchorage is also included.
2.4.1. Bonded anchorage
There are numerous types of bonded anchorage systems. Normally, an outer sleeve surrounds either a cementitious or resin
bonding material to anchor the tendon. The main features of such
systems are shown in Fig. 2. Several researchers have devised solutions to issues relating to bonded anchorage systems and how to
deal with the monitoring of the performance of these anchorages
and their mathematical evaluation. Mitchell et al. [35] worked on
a mathematical study involving axisymmetric nite element analysis, using two types of end connections. These end connections
consisted of a metal casing joined to the FRP tendon through a
layer of potting material. Rostsy and Kepp [36] presented the
results of static tensile loading tests and the inuence of bond
length of a steel-sleeved anchorage on the tendons ultimate
load-bearing capacities. The strains in the sleeve and bond stress
distribution were also described. Rostsy was one of the rst
researchers to propose test methods and acceptance requirements
for post-tensioned FRP tendon systems that involved measurements on the anchorages that were used [37,38]. Budelmann
et al. [39] discussed the fatigue behaviour of bond-anchored unidirectional GFRPs where the E-glass bres were embedded in a
quartz sand/PE-resin mortar and anchored in a cylindrical steel
tube. Patrik et al. [40], EMPA (Eidg. MaterialPrfungs- und VersuchsAnstalt fr Industrie, Bauwesen und Gewerbe) developed a
resin-potted anchorage with a stiffness gradient in its resin that
handled 92% of the tendons tensile strength. The common mode
of failure in this potted anchorage was excessive creep deformation
in the resin [41,42]. Lees et al. [43] discussed the problem of induced stress concentrations in an anchorage of FRP tendons and
how to solve this issue with expansive cement couplers. An experimental study was rst carried out on couplers joining steel reinforcement bars and then extended to include the coupling of FRP
materials onto steel prestressing bars. Nanni et al. [14] investigated
some of the available commercial anchorage systems and studied
their ultimate tensile capacity and tensile capacity during shortterm sustained loading. They used grout to anchor a Technora tendon in a 500 mm long sleeve, while the CFCC tendon was anchored
in a 165 mm long sleeve using an epoxy adhesive. Meier and Farshad [44] investigated the connection of high-performance CFRP
cables to suspension and cable-stayed bridges using gradient
materials (soft zone concept) and presented a new, reliable anchor-

ing system, developed with computer-aided materials design and


produced with advanced gradient bond materials based on ceramics and polymers.
Harada et al. [4547] were among the rst pioneers to use
expansive cementitious materials to ll straight metal sleeve
anchorages. The expansive cement was used to place a lateral pressure on the tendon in the sleeve, and thus prevent slippage of the
tendon. It was shown that an internal pressure of between 25 and
40 MPa generated by the expansive cementitious material in the
sleeve was enough to grip FRP tendons with different surface treatments. Benmokrane et al. [48,49] investigated ground anchorages
and described the available AFRP and CFRP tendons, along with
their properties and constituent materials (Arapree, Technora, FiBRA, CFCC and Leadline). They discovered that the grout pullout
capacity was mainly affected by the properties of the cement grout,
surface deformation of the rod, bond length and the modulus of
elasticity of the anchorage tube.
Zhang et al. [50] analysed the mechanism of how bonded
anchorages work with FRP tendons and presented a conceptual
model to calculate the bond stress at the tendongrout interface
and the tensile capacity of the anchorages. This study was carried
out to investigate why the bond distribution of bonded anchorages
is non-uniform along the bonded length and why the point of peak
bond stress shifts from the loaded end of the anchorage to a point
inside the anchorage as the applied load increases. Zhang et al. [51]
focused their investigation of tensile behaviour onto FRP ground
anchorages. Sixteen monorod and four multi-rod grouted AFRP

Fig. 3. Example of a spike anchorage system.

114

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

(Arapree and Technora) and CFRP (CFCC and Leadline) anchorages


were tested at different load levels using standard methods to
examine tensile strength and sustained load.
Pincheira and Woyak [52] performed tests on epoxy-lled
sleeve anchorages, between 152 mm and 380 mm long, using different ller types and held by hydraulic grips, a standard wedge
assembly or a custom-made wedge assembly. Matta et al. [53]
developed a new type of swaged stainless steel anchorage/coupler.
In this anchorage, the load was transferred due to the friction and
interlocking produced by the swaging process. A thread adapter
was positioned at the end of the swaged part so that a threaded
bar could be attached and could transfer tension from the turnbuckle used for stressing the tendons. No lengths of the anchorages
or specied strengths of the tendons were reported, although during static tensile testing, they all experienced a rupturing of the
rods. As promising as the system might seem, there were several
drawbacks reported relating to the steel parts of the anchorage
i.e. the steel parts were difcult to protect, the tendons had to be
pre-cut, and the anchorage parts had to be attached before mounting, thus involving tight tolerances.
Issues related to bonded anchorage evidently lay in the long
curing time of the bonding material, creep of the cementitious
bond material and anchorage lengths that make the technique
more suitable for ground anchorage than for structural applications, where instant anchorage is required and there is little
mounting space. However, the bonded anchorage protects the
FRP tendon and is considered useful in internally poststressed
applications where the tendon is stressed and grouted. Still, the
control of the stresses in bonded anchorages is difcult and
demanding, when the soft zone concept is used to prevent high
stresses at the loaded end of the anchorage.
2.4.2. Mechanical anchorages
A mechanical anchorage relies mainly on friction in the interface between the tendon and the inner anchorage surface, so a
compressive force perpendicular to the tendon has to be applied.
Compression is normally obtained through a cone-shaped interface
between a barrel and a wedge, or by clamping the tendon. These
techniques work well for traditional prestressing materials, where
the tendon is made of an isotropic and yieldable material such as
steel, but pose great challenges when using FRP tendons; this is because of the anisotropic properties of FRP that makes the material
weak in a direction perpendicular to the bres. Several ideas have
been developed to address the issues related to mechanical
anchorage of FRP tendons.
2.4.3. Spike anchorage
The spike anchorage, shown in Fig. 3, consists of an internal
spike that presses the bres of the tendon/rope against the barrel

Fig. 4. Example of a wedge anchorage system.

wall, thus holding the single bres and ensuring full anchorage
of the tendon/rope. This system has been used previously with
steel strands, but is also considered suitable as an anchorage for
Paral ropes. Burgoyne [15] described the properties of Paral
ropes with Aramid core yarns together with a presentation of
ongoing research on these materials and related project applications. The method of anchoring Paral ropes was explained and
the results of tests carried out at Imperial College of Science and
Technology, London were detailed. The purpose of the research
was to verify long-term properties, thermal properties and fatigue
resistance when Paral ropes were subjected to tensile bending
and sheave bending. It was reported that the system could be successfully used in structural engineering applications. At an ACI
Convention in 1988, Burgoyne also explained the use of external
FRP prestressing and described anchorage techniques (mainly for
Paral ropes), bre relaxation, durability, fatigue, thermal response and bonding [54]. More tests were described in [16], where
Paral ropes with a nominal tensile capacity of 600 kN were
tested. The specimens were tested under tensile bending with a
cyclic lateral load being carried. Five tests were conducted, where
the cable was preloaded with 300 kN in three tests, 200 kN in one
test and 400 kN in one test. The failure mode was mainly at the
cable deector and the measured life cycles ranged between
69,000 cycles and 1000,000 cycles. Sheave bending tests were also
performed; all ropes failed where the rope was alternately straight
or curved during each cycle. This was due to inter-bre fretting,
which was expected to be highest in these areas.
2.4.4. Wedge anchorage
The split wedge anchorage system is shown in Fig. 4. It consists
of a barrel that is contoured inside, wedges with an external angled
surface that matches the inside of the contoured barrel (the angle
can be changed to control the stresses), and a sleeve to distribute
compressive stresses from the wedges to the tendon and thus prevent premature failure. Mechanical friction anchorages use large
compression forces in order to grip the tendon. This introduces
high principal stresses at the loaded end of the anchorage, where
longitudinal tensile forces in the tendon are at their maximum
and transverse compression forces from the wedges may also be
large. To overcome these problems, the compressive stresses
(essential for the correct operation of the wedge) can be transferred to the back of the anchorage, where the tensile stresses
are lower. Currently, there are two concepts that have been developed to achieve this the differential angle design and curved angle design. Sayed-Ahmed and Shrive [34] were amongst the rst to
use the differential angle concept and developed a new wedge
anchorage system that could be used for both bonded and unbonded Leadline CFRP tendons. The anchorage system consisted of a
steel barrel and four wedges, greased between the barrel/wedge
interface. It was constructed with a differential angle of 0.1 between the barrel and wedge. The external wedge, angled at 2.09,
was smooth on the outside and sandblasted on the inner surface.
The sleeve could be made of either aluminium or copper. The thin
inner sleeve distributes the radial stresses from the wedge more
uniformly around the tendon. Early barrel prototypes were made
of mild steel because it was hypothesised that a long gripping
length (length of 127 mm and outer diameter of 100 mm) would
be required to transfer the load and prevent a concentration of high
stress in that area. However, numerical analyses showed that it
was possible to reduce the dimensions, resulting in a new smaller
design, as shown in Fig. 5.
The new anchorage was made of high performance stainless
steel (a 0.2% proof stress of 862 MPa and a tensile strength of
1000 MPa). The edges of the inner surface of the four-piece wedge
were rounded because tightening of the wedges introduced yielding of the soft metal sleeves that deformed into the wedge gaps.

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

115

Fig. 5. Wedge anchorage system, after [34].

Fig. 6. Displacement behaviour of the wedge anchorage components with 50 kN


presetting load [57].

the tendons, corresponding to a load of 99 kN. Modications after


the initial tests resulted in a range of load capacity from 105 kN
to124 kN. In total, 24 tests were performed: 10 using the prototype
anchorage and 14 using the smaller, rened wedge anchorage. Fatigue testing was also carried out: three tests were performed on
the prototype anchorages and two on the rened anchorage. Tests
were performed with a xed number of total cycles, which were
split up into smaller test series involving different stress ranges
and number of cycles (a summation of the numbers of cycles in
these tests gave the total number of cycles). The stress range appeared to have a signicant effect on the CFRP fatigue strength,
where a narrow stress range was thought to result in a load increase. It was claimed that the system worked at the efciency required by the PTI (Phoenix, Arizona, Post-Tensioning Institute)
[55,56] for steel strands and anchorages.
Experimental and numerical evaluations of a stainless steel
wedge anchorage for CFRP tendons were conducted by Al-Mayah
et al. [57], who tested an anchorage consisting of a stainless steel
barrel (inside angle 1.99), four stainless steel wedges (with a differential angle of 0.1) and an aluminium sleeve. An 800 mm long,
8 mm diameter CFRP Leadline tendon was tested with presetting
(seating of the wedge before tensioning of the tendon) levels of
50 kN (representing 48% of the design tensile capacity), 65 kN
(63% of the capacity), 80 kN (77% of the capacity) and 100 kN
(96% of the capacity). The test specimens were loaded both statically and with controlled deformation, and capacities of approx.
100115 kN were measured. The displacement showed three distinct phases: The displacement in relation to the barrel of the rod
and sleeve and the rods displacement in relation to the sleeve,
see Fig. 6. The rst phase starts when the load reaches the slope
point, F1. Only the rod moves under this load, with a load displacement rate given by slope 1. When the load reaches load level F2,

Fig. 7. Geometric conguration of wedge anchorage with longitudinal curved


wedge [59].

Possible sharp corners of the wedges could dig into the soft-metal
sleeve, creating stress concentrations that can cause premature
failure. In the static tests carried out, one of the aims was to obtain
a 95% anchorage performance of the ultimate tensile capacity of

Fig. 8. Example of clamping anchorage.

116

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

the sleeve starts to slip, which results in a decrease in displacement


rate shown by slope 2. The load then increases to the ultimate load
of the tendon. Slope 3 shows the load/deformation rate of the
sleeve, which after slip is initiated, is almost solely responsible
for further slip in the entire system. It was also shown that the load
levels reached before the slippage of the rod (F1) and the sleeve
(F2) could be sustained with reused anchorages. A similar conclusion was reached when a presetting load of 100 kN was applied,
which is double the load of that used in the tests that produced
the data shown in Fig. 6. Furthermore, Sayed-Ahmed [58] presented research into single and multi-strand steel anchorage systems for CFRP tendons/stays, where the effect on the radial stress
of a differential angle between the anchorage wedge and the CFRP
tendon was investigated. It was proposed that several differential
anchorage wedges in a steel seating plate were used, where nine
wedge systems were placed and jacked. The system was shown
to full the PTI requirements.
Additional research by Al-Mayah et al. [59] showed the development of a new CFRP rod anchorage system using the curved angle concept. The anchorage components were the same as in
former experiments, that is, an outer cylinder (barrel), four wedges,
and a soft metal sleeve. However, the contacting surfaces of the
wedges and barrel had a circular prole along the length of the
anchorage, as shown in Fig. 7. Tensile testing using different presetting loads, geometric congurations, and rod sizes was carried
out. Both aluminium and copper sleeves were used on the
6.4 mm and 9.4 mm diameter rods, with a length of 1000 mm.
The curved wedges were 80 mm long and the barrel 70 mm long.
Tests with and without presetting were performed. The tensile
load was applied at a rate of 0.500.65 mm/min. A number of specimens were also tested at a load rate of 5.0 mm/min. The loading
rate had no effect on anchorage performance. The tangential angle
to the curved interface between the wedges and the barrel was
small so as to reduce the principal stresses at the loaded end. It
was important for the angle to increase smoothly along the length
of the anchorage so that the highest wedge compression value at
the unloaded end of the rod could be handled. It was reported that,
in general, the anchorage performed satisfactorily with presetting
(80 kN) as well as without presetting, which however caused the
largest slippage. Further, the different FRP rod sizes did not affect
performance.
2.4.5. Clamping anchorage
Another type of mechanical anchorage frequently used by
researchers is the clamping anchorage. This is used where size,
ease of use and aesthetic aspects are of less importance. These
anchorages generally consist of two rectangular steel plates, a
sleeve (most commonly made of aluminium or copper) and clamping bolts, as shown in Fig. 8. Each steel plate is manufactured with

Fig. 9. Typical slipping of a CFRP rod, after [63].

Fig. 10. Schematic view of the plates used for the friction experiments:
(a) aluminium parts, and (b) composite part (made of HTA-6376 with a layup of
[45/0/90]3 s), after [65].

a circular longitudinal notch on one side and holes for the bolts.
The aluminium sleeve is thin, with slits along the part clamped
in the anchorage. A short section of the sleeve is left without slits
to hold the parts together; this part is called the sleeve head and
is positioned on the outside of the plates when they are clamped.
It is vital that the right torque is applied to the bolts to obtain
the proper amount of friction. If the torque is too small, the rod
could slip, but too much torque could result in premature failure
of the rod due to stress concentrations. These stress concentrations
are normally caused by the high principal stresses at the loaded
end of the anchorage but can be controlled by applying torque
on the bolts in steps. Thus, less torque is applied to the bolt pair
at the loaded end of the anchorage and then linearly increased until maximum torque is applied at the unloaded end of the
anchorage.

3. Slip at the anchorage/tendon interface


The surface composition of the FRP tendon is critical when
using a mechanical anchorage. When using such an anchorage,
the composition of the matrix surface layer and the smoothness
of the surface are two critical parameters [49,6062]. Al-Mayah
et al. [63,64] addressed this issue and explained the effect of

Fig. 11. Friction force as a function of grip displacement for a friction specimen
after 25 cycles and 5300 cycles, after [65].

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

surface texture, in this case created by sandblasting, between the


contact interfaces in the CFRP couplers. This study examined the
sliding behaviour of a CFRP rod in contact with copper and aluminium sleeves, simulating the components of a wedge anchorage system. The CFRP rod that was tested was a single spiral indented
CFRP rod with a nominal diameter of 9.4 mm, which was then anchored with a clamping anchorage system. A general load displacement relationship was observed across three regions of the graph
and there was a load uctuation amplitude with a height (h), as
shown in Fig. 9. In region 1 (slope 1), from zero to a certain maximum load (F1), the load can be explained as the shear stress, where
the maximum force is divided by the nominal surface area of the
rod. In region 2 (slope 2), a gradually decreasing load over the
remainder of the sliding distance follows the maximum peak load.
Region 3 (slope 3) is reached in some cases, where complete
sticking occurs and the load increases sharply. In conclusion, the
interfacial shear stress capacity between a metal sleeve and a CFRP
rod increased with increased contact pressure, and the local surface area was increased when the surface was sandblasted, again
resulting in an increase in shear stress capacity.
Schn [65,66] investigated the wear of CFRP and the coefcient
of friction for aluminium in contact with the CFRP, conducting
experiments on a double lap joint with an overlapping length of
40 mm, as shown in Fig. 10. The HTA-6376 carbon bre epoxy matrix, with nominal ply thicknesses of 0.13 mm, had a quasi-isotropic stacking sequence corresponding to [45/0/90], meaning a
symmetrical layout where the outer friction surface has a bre
direction of 45. Testing was carried out using displacement-controlled machinery with the application of a 1.5 mm displacement
to the specimen for 15 s. The loading was stopped for 15 s to allow
the contact surfaces to cool and then reversed. The friction coefcient was calculated using:

lT

2P

In the tests, the normal force P was set to 5 kN, which corresponds to a realistic load on bolted joints. Typically, the friction
coefcient initially increases rapidly to a high value and then decreases slightly to a constant level.
Friction force and displacement were measured during the sliding sections of the displacement cycle. A high friction force resulted
in stickslip behaviour, which was also seen after the coefcient of
friction had increased, as shown in Fig. 11. The friction force was
seen to increase from a negative value as a result of the previous
loading in the reverse direction, when the direction of the grip displacement was reversed and almost linear until the sliding restarted. In general, specimens with composite/composite contact
and aluminium/composite contact behaved in a similar way.
The coefcient of friction initially increased and then decreased
slowly. The initial coefcient of friction measured for the composite/aluminium specimens was lower than that measured for the
composite/composite specimens. The effective coefcient of friction was predicted according to the following equation:

Fig. 12. Static model for preliminary anchorage design [70].

1
P

l  lea Pea lfa Pfa

117

where l is the effective coefcient of friction, P is the total normal


force, lea and lfa are the coefcients of friction of epoxy and bre in
contact with the aluminium, Pea is the load carried by the epoxy and
Pfa is the load carried by the bre. During the reciprocating sliding,
this matrix was worn away and the bres in the 45 ply layer became visible. The initial coefcient was calculated to be approximately 0.23, whereas the peak coefcient after wear was found to
be approximately 0.68. The coefcient of friction was described as
being independent of the normal load.
Al-Mayah et al. [64,67] reported the same phenomenon in pullout experiments, which investigated the interfacial behaviour of
CFRP metal couples under different contact pressures, with different composite surface proles and different ultimate CFRP tensile capacities. The composite rods were, in these experiments, in
contact with either annealed or as-received aluminium or copper
sleeves. The capacity to transfer shear stress increased signicantly
when a smooth machined CFRP rod was used. It was observed that
patches of bres and epoxy were formed when the sliding started
and increased as the sliding continued. The contact shear stress
was seen to increase with increased contact pressure when the aluminium sleeve, particularly when made with annealed aluminium,
was used. In addition, soft annealed sleeves combined with higher
tensile strength CFRP rods resulted in higher contact shear stress,
whereas a lower shear stress was recorded when using the higher
strength rod with the harder, as-received aluminium, sleeve. As a
result of these observations, it is recommended that soft sleeves
should be used in the design of a wedge anchorage, regardless of
the strength or prole of the rod.
4. Evolution of analytical and nite element (FE) models
Early literature on the subject of wedge anchorages for FRP tendons attempted to use basic analytical and FE models to evaluate
their behaviour. Over the years, these models have become more
advanced. The static 2D rigid body analytical models which had
been applied to thin walled cylinder theory are now used with
thick walled cylinder theory. FE models have evolved from being
two dimensional and axisymmetric to being three dimensional.
Static 2D rigid body models are exemplied in the calculations performed within the work of [41,42,6870]. The models are typically
built upon a static equilibrium between the forces, such as in
Fig. 12, where the forces acting on the different parts are illustrated, as described in [70]. The barrel vector, RWB, is often of interest if a minimum barrel thickness is to be calculated and the
resulting equation is described in [3], where lWB is the coefcient
of friction between the wedge and the barrel.

Fig. 13. Contact-pressure distribution on the rod using mathematical and axisymmetric nite element models, after [72].

118

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

RWB

lWB  cosh sinh

Using values of lWB = 0.1 and h = 2.1 results in a value of


RWB = 7P. This value is used by Campbell et al. [69] and Reda Taha
and Shrive [41] together with the thin walled cylinders theory,
equation [4], to calculate the minimum thickness of a barrel made
of UHPC (Ultra High Performance Concrete). rWB is the internal
pressure, rT is the circumferential stress, dB is the inner diameter,
tB is the thickness of the cylinder and lB is the length of the barrel.

rT

rWB  dB
2  tB

) tB

rWB  dB
2  RWB
where rWB
2  rT
p  dB  lB

Using the previous example, with an ultimate tensile capacity of


the concrete equal to 9 MPa, a maximum prestressing force, P, of
104 kN and a barrel length of 180 mm, a barrel thickness of
140 mm would be required.
Shaheen and Shrive [70] compared results from tests where the
barrel failed at a prestressing load of 40 kN. Using a value of 40 kN
in equation [4], the stress in the concrete when it fails should be
8 MPa. When compared to the expected 25 MPa, the thin walled
cylinder theory obviously does not adequately describe the stresses within the barrel. This is a common conclusion in literature
on the subject of shells and cylinders. For example, Roark and
Young [71] stated that the theory is only applicable when the cylinders thickness is less than one-tenth of its radius. Shaheen and
Shrive [70] advanced this theory when they used the thick walled
cylinder theory as well as the thin walled cylinder theory to nd
the required ultimate tensile strength of a UHPC barrel. No comparison with other experiments was given however and so the
model cannot be evaluated. Note that these calculations were
based upon forces calculated using the static 2D rigid body model,
which is a highly simplied representation of the real stresses involved. Al-Mayah et al. [57,72] described a further development
of the application, combining calculations for the assembled rod,
internal sleeve, wedge and barrel, and thus obtained a better
understanding of the radial stress distribution within the anchorage. The results, compared with results from an axisymmetric nite element model, are shown in Fig. 13. The anchorage
modelled was designed using the curved angle concept and so
the highest stresses were reached at the back of the anchorage.
According to Al-Mayah et al., differences between the analytical
and numerical models were due to the orthotropy of the rod,
something that is excluded in the analytical model. However, the
thick walled cylinder theory also has some limitations. As mentioned by Al-Mayah et al., all materials are modelled as being isotropic, and are also assumed to be linearly elastic with the same
mechanical properties under tension as well as under compression.
The space between the wedges is not considered, since the theory
assumes axial symmetry. All longitudinal forces are ignored and
either plane stress or plane strain conditions are required to solve
the system of equations. The fundamentals of the theory can be
found in references such as [73,74].

Fig. 14. FE model used in [72].

4.1. Finite element (FE) models


Early attempts to model the behaviour of an anchorage using FE
calculations are described in [34]. This model and the models following in [57,69,75] are axi-symmetric. The axisymmetric representation of the wedge anchorage is not perfect because even
though the axisymmetric FE model has the ability to include orthotropic material properties and longitudinal stresses, it cannot take
account of the effect of the spaces between the wedges. This was
demonstrated by Mitchell et al. [35], where the variation of radial
stresses throughout the thickness of the wedge was shown after a
complete axisymmetric analysis. The analysis showed high shear
stresses at the loaded end of the anchorage during seating and
again when the anchorage was fully loaded. A uniform compression from the wedges was also shown, whereas the axial stress
in the tendon linearly decreased from the full load at the loaded
end of the anchorage to zero load at the unloaded end of the
anchorage. Axisymmetric models have given researchers a sound
understanding of how the longitudinal distribution of normal
stresses on the rod vary with differences in the angle between
the wedge and the barrel, with curved interfaces and with different
presetting distances. Campbell et al. [69] showed that there could
be a signicant redistribution of the radial stresses on the rod by
simply changing the difference in the angle between the wedge
and the barrel from 0 to 0.2, with the latter value resulting in a
favourable change of peak pressure towards the unloaded end of
the anchorage.
In Al-Mayah et al. [76,72] an improvement is made to the theory
with the inclusion of 3D modelling. The two papers describing 3D
models focus on two attempts to improve the design of the wedge
anchorage. Al-Mayah et al. [76] altered the outer diameter of the
barrel such that the thick end was made thinner. Again, this was
to transfer a larger portion of the radial stresses towards the unloaded end of the anchorage. Loaddisplacement curves from the
model showed good agreement with experimental data. A more detailed model was presented in [72]. This model aimed to reproduce
the behaviour of the anchorage with the longitudinally curved inner
surfaces as shown in Fig. 13. The meshed model is shown in Fig. 14.
The spaces between the wedges were assumed to be symmetrical to
minimise computational costs; this should not affect the accuracy of
the model. The model described in [72] used the material properties
given in Table 2, and all parts were modelled as being isotropic except for the rod, which was modelled with an orthotropic composition. The coefcient of friction was 0.02 for the barrelwedge
interface, 0.24 for the sleeverod interface and 0.4 for the wedge
sleeve interface. The resulting longitudinal distribution of radial
stresses onto the rod for an outer longitudinal curvature of the
wedge with radius 1900 mm and different presetting distances is
shown in Fig. 15a. By achieving this distribution, stress concentrations at the loaded end of the anchorage can be avoided and the tensile force in the rod is smoothly transferred into the anchorage at a
successively increasing rate. An impressive agreement between the
model and experimental results was also demonstrated, as shown in
Fig. 15b. In all of Al-Mayahs models, the experimental and FE analysis results show good agreement with regard to loaddisplacement
behaviour. However, it should be noted that no plastic behaviour of
the metals was included in the models and the coefcients of friction were all given as constants. In reality, friction is a function of
several variables, such as normal force and sliding distance and,
with extreme stresses acting on the components; it is obvious that
yielding must occur. The model is consequently not a perfect representation of the anchorage, but a representation that can produce
the same loaddisplacement curve as derived from experiments.
For even better calibration, other measurements should also be
compared to the results from the FE calculations. These other measurements might be the circumferential, longitudinal stresses on

119

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121


Table 2
Material properties used in the 3D FE model presented in [72].
Property

Rod

Sleeve

Wedge

Barrel

Material ()
E-modulus, E1 (GPa) (Longitudinal direction)
E-modulus, E2 (GPa) (Transverse direction)
Shear modulus, G12, G13 (Transverse direction)
Major Poissons ratio n12, n13, n23
Minor Poissons ratio n21, n31

CFRP
124
7.4
7.0
0.26
0.02

Copper
117
117.0
44.7
0.31
0.31

Steel
200
200.0
77.0
0.30
0.30

Steel
200
200.0
77.0
0.30
0.30

Fig. 15. (a) Radial stresses onto the rod for different presetting distances; and (b)
comparison of FE and experimental results, after [72].

the barrel, the relative motion between the components within the
anchorage and shear stress distribution in the interfaces between
adherent and adhesives.

(i) premature failure at the loaded end of the anchorage due to high
principal stresses, (ii) local crushing and (iii) sliding of the contact
interface between the FRP rod and sleeve where the tendon resin
layer has worn off. The literature studied suggests that anchorages
using soft aluminium are benecial in transferring shear stresses to
the wedges without causing any premature failure. Controlling,
understanding and thus reducing the number of possible failure
modes are some of the main areas to study and solving these problems might solve the FRP anchorage issue. Analytical models are
limited, ranging from static, 2D rigid body models using both the
thin walled cylinder theory and the thick walled cylinder theory,
to FE models, which have evolved from 2D evaluations and axisymmetric models to 3D models. However, there is huge complexity in modelling all the anchorage parts with their properties,
shapes, tendon forces, sliding distances, extreme stresses and
yielding effects and, as these are specic for each type of anchorage, the task is demanding. It is evident that the development of
a reliable, easy-to-produce and mount anchorage still remains to
be achieved, though several anchorage designs have been proposed. For FRP prestressing applications, such anchorages have to
be developed to meet the same requirements as those for steel prestressing systems and thus compete at the same level. Critical issues relating to FRP anchorages and the use of proper test
methods to verify their performance have to be addressed more
consistently in the future. Furthermore, the testing of FRP tendons
has shown that, often, there is brittle behaviour that causes a sudden failure at high levels of force. It can therefore be difcult to distinguish between a tendon and an anchorage failure. This means
that a fracture which seems to have been caused by a tendon failure could easily have been caused by the anchorage failing instead.
The difculty is whether to focus on the level of stress on the tendon or on the whole system (anchorage + tendon).
The literature study has revealed some areas of FRP anchorage
that could be focussed on in the future:
5.1. Reliability and durability

5. Discussion and future research


Currently, there are no efcient and competitive prestressing
systems for FRP tendons despite research over the last two decades. A great challenge lies within the anchorage of the FRP tendon
itself and there have been several attempts to develop FRP
anchorage systems. Different types of anchorages grip in different
ways. A mechanical anchorage is preferred because it is easy to
mount and control the stress through it, whereas a bonded anchorage requires the instant curing of bonding agents and needs longer
anchorage lengths. Several mechanical anchorages have been
developed to anchor FRP tendons such as wedge anchorages or
clamping anchorages. However, the stability of the anchorages is
unreliable, and experimental results suggest that only small differences in the mounting procedure greatly affect the performance of
the anchorage, that is, the difference between a successful anchorage and an unsuccessful anchorage is small. The way in which the
closed mechanical anchorage system works seems to be very complex to model and verify through tests. The main failure modes are

 Anchorage requirements to provide for a safe failure, which


are comparable with existing requirements for steel
systems.
 Denition of stable and reliable friction surface between the
FRP tendon and the anchorage.
 Monitoring of stresses in the anchorage in tests and eld
applications.
 More testing of the life span of FRP anchorages as, so far,
there have been only limited investigations into the life
span (short and long-term effects).
 Development of working deviators and anchorageconcrete
connections with respect to the maximum curvature and
transverse capacities of the tendons.
 Standards and guidelines relevant to FRP anchorage.
5.2. Testing
 Addressing failure modes and their acceptability.
 Long-term resistance and fatigue resistance.

120

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121

 Consideration of whether the stress level in the tendon or in


the anchorage should be taken into account in the design.
 Fire resistance.
 Thermal resistance.
5.3. Material properties and geometry
 What effect does the material and geometry of the wedge
and barrel have on the performance of the anchorage?
 How are the anchorage and the system as a whole best
modelled and veried?
 Could materials other than metals be used as anchorage
parts (cement or bre based)?
 Can the ease of mounting match that of conventional steel
prestressing?

References
[1] Stcklin I, Meier U. Strengthening of concrete structures with pre-stressed
gradually anchored CFRP strips. In: Proc., ber-reinforced polymer
reinforcement for concrete structures, FRPRCS-5, Cambridge, UK; 1618 July,
2001. p. 2916.
[2] Czaderski C, Motavalli M. 40-Year-old full-scale concrete bridge girder
strengthened with prestressed CFRP plates anchored using gradient method.
Composites Part B 2007;38(78):87886.
[3] Sika. Prestressing systems for structural strengthening with Sika CarboDur
CFRP Plates. Technical report, Sika Services AG, Zrich, Switzerland; 2004. 7 p.
[4] VSL and Sika. Strengthening with post-tensioned CFRP. Technical report, Sika
Services AG, Zrich, Switzerland; 2007. 4 p.
[5] Wu Z. Integrated strengthening of structures with bonded prestressed FRP
reinforcement. In: Proc., rst Asia-Pacic conference on FRP in structures,
APFIS, Hong Kong, China; 1214 December, 2007. p. 4352.
[6] Monti G, Liotta MA. Pretensioning of externally bonded FRP fabrics for exural
reinforcement of RC beams: method and design equations. In: Proc., rst AsiaPacic conference on FRP in structures, APFIS, Hong Kong, China; 1214
December, 2007. p. 37992.
[7] El-Hacha R, Wight RG, Green MF. Anchors for prestressing FRP sheets. In: Proc.,
CSCE annual conference, Regina, Saskatchewan, Canada; 25 June, 1999. p.
23544.
[8] El-Hacha R, Green MF, Wight RG. Strengthening of concrete members with
prestressed FRP sheets: behaviour at room and low temperatures. ACI Spec
Publ 2000;188(64):73749.
[9] El-Hacha R, Wight RG, Green MF. Innovative system for prestressing berreinforced polymer sheets. ACI Struct J 2003;100(3):30513.
[10] El-Hacha R, Wight RG, Green MF. Prestressed carbon ber reinforced polymer
sheets for strengthening concrete beams at room and low temperatures. J
Compos Constr 2004;8(1):313.
[11] Nordin H, Tljsten B. Concrete beams strengthened with prestressed surface
mounted CFRP. J Compos Constr 2006;10(1):608.
[12] Tljsten B. FRP strengthening of existing concrete structures. Design guideline,
Lule University of Technology, Division of structural engineering, Lule,
Sweden; 2006. 228 p.
[13] Collins MP, Mitchell D. Prestressed concrete structures. Englewood Cliffs, New
Jersey, USA: Prentice Hall Inc.; 1991. 766 p.
[14] Nanni A, Bakis CE, ONeil EF, Dixon TO. Performance of FRP tendon-anchor
systems for prestressed concrete structures. PCI J 1996;41(1):3444.
[15] Burgoyne CJ. Structural use of paral ropes. Constr Build Mater
1987;1(1):313.
[16] Burgoyne CJ, Hobbs RE, Strzemiecki J. Tension bending and shave bending
fatigue of parallel lay aramid ropes. In: Proc., eighth int. conf. on offshore
mechanics and arctic engineering, The Hague, The Netherlands, 1923 March,
1989. p. 6918.
[17] Giannopoulos IP, Burgoyne CJ. Design criteria for aramid bres. In: Proc., int.
conf. on advanced composites in construction, ACIC, Edinburgh, Scotland, UK;
2009. p. 156167.
[18] Gerritse A, Schrhoff H. Prestressing with aramid tendons, Technical
contribution. FIP 10th Congress, New Delhi; 1986. 7 p.
[19] Gerritse A, Werner J. ARAPREE: the prestressing element composed of resin
bonded twaron bres. Rijswijk, The Netherlands: Holland Beton Group; 1988.
8 p.
[20] Gerritse A, Werner J. First application of Arapre. In bre reinforced cement and
concretes: recent developments. London, UK: Elsevier Applied Sciences; 1989.
p. 509.
[21] Gerritse A, Werner J. ARAPREE a non-metallic tendon: performance and design
requirements. In: Proc., spec. conf. on advanced composite materials in civil
engineering structures, Las Vegas, USA, 31 January1 February, 1991. p. 14354.
[22] Tsuji Y, Kanda M, Tamura T. Applications of FRP materials to prestressed
concrete bridges and other structures in Japan. PCI J 1993;38(4):508.

[23] Dolan CW. Developments in non-metallic prestressing tendons. PCI J


1990;35(5):808.
[24] Iyer SL, Kumarswamy C. Performance evaluation of glass ber composite cable
for pre-stressing concrete units. In: Proc., 33rd int. SAMPE symposium,
Anaheim, California, USA; 710 March, 1988. 12 p.
[25] Iyer SL, Sripathy G. First post tensioned deck bridge with composite cables. In:
Proc., fourth materials engineering conference, Washington, DC, USA; 1014
November, 1996. p. 37585.
[26] Dolan CW, Hamilton HR. Design recommendations for concrete structures
prestressed with tendons. University of Wyoming, FHWA, CONTRACT,
DTFH61-96-C-00019; 2001.
[27] ACI. Prestressing concrete structures with FRP tendons. ACI 440.4R-04,
Farmington Hills, MI; 2004. 35 p [ISBN 0-87031-166-2].
[28] Santoh N. CFCCs (carbon ber composite cables). In bre-reinforced-plastic
reinforcement for concrete structures: properties and applications. New
York: Elsevier Science Publishers; 1993. p. 22348.
[29] Noritake K, Kakihara R, Kumagai S, Mizutani J. Technora, an Aramid FRP Rod. In
bre-reinforced-plastic reinforcement for concrete structures: properties and
applications. New York: Elsevier Science Publishers; 1993. p. 26790.
[30] Grace NF, Abdel-Sayed G. Double tee and CFRP/GFRP bridge systems. Concr Int
1996;18(2):3944.
[31] Grace NF, Abdel-Sayed G. Behaviour of externally draped CFRP tendons in
prestressed concrete bridges. PCI J 1998;43(5):88101.
[32] Grace NF, Abdel-Sayed G, Tang BM. Construction, evaluation and ultimate load
test of CFRP/CFCC prestressed concrete DT-beam. SAMPE J 2002;38(5):2833.
[33] Grace NF. Bridge street bridge the rst CFRP prestressed concrete bridge in
the United States. Concr Eng Int 2003;7(2):337.
[34] Sayed-Ahmed EY, Shrive NG. New steel anchorage system for post-tensioning
applications using carbon bre reinforced plastic tendons. Can J Civ Eng
1998;25(1):11327.
[35] Mitchell RA, Woolley RM, Halsey N. High-strength end ttings for FRP rod and
rope. J Eng Mech Div 1974;100(4):687703.
[36] Rostsy FS, Kepp B. Zum verhalten dynamisch beanspruchter GFKspanngleider. Research report, Institut fr Baustoffe, Massivbau und
Brandschutz, TU Braunschwieg, Germany; 1986.
[37] Rostsy FS. New approach in the already published recommendations for
anchorage assembly. In: Proc., 9th congress of federation internationale de la
precontrainte, Stockholm, Sweden; 1982. p. 958.
[38] Rostasy FS. Draft guidelines for the acceptance testing of FRP post-tensioning
tendons. J Compos Constr 1998;2(1):26.
[39] Budelmann H, Kepp B, Rostsy FS. Fatigue behaviour of bond-anchored
unidirectional glass-FRPs. In: Proc., rst materials engineering congress,
Denver, Colorado, USA; 1315 August, 1990. p. 114251.
[40] Patrick K, Meier U. CFRP cables for large structures. In: Proc., of spec. conf. on
advanced composite materials in civil engineering structure, Las Vegas, USA;
1991. p. 23344.
[41] Reda Taha MM, Shrive NG. New concrete anchors for carbon ber-reinforced
polymer post-tensioning tendons Part 1: State-of-the-art review/design. ACI
Struct J 2003;100(1):8695.
[42] Reda Taha MM, Shrive NG. New concrete anchors for carbon ber-reinforced
polymer post-tensioning tendons Part 2: Development/experimental
investigation. ACI Struct J 2003;100(1):96104.
[43] Lees JM, Gruffydd-Jones B, Burgoyne CJ. Expansive cement couplers a means
of pre-tensioning bre-reinforced plastic tendons. Constr Build Mater
1995;9(6):41323.
[44] Meier U, Farshad M. Connecting high-performance carbonber-reinforced
polymer cables of suspension and cable-stayed bridges through the use of
gradient materials. J Comp-Aid Mater Des 1996;3(13):37984.
[45] Harade T, Idemitsu T, Watanabe A, Khin M, Soeda K. New FRP tendon anchorage
system using highly expansive material for anchoring. In: Modern prestressing
techniques and their applications, proceedings, vol 2, FIP93 symposium, Japan
Prestressed Concrete Engineering Association; 1993. p. 7118.
[46] Harada T, Matsuda H, Khin M, Tokumitsu S, Enomoto T, Idemitsu T. Development
of non-metallic anchoring devices for FRP tendons. In: Proc., second int. RILEM
symposium, FRPRCS-2, Ghent, Belgium, 2325 August, 1995. p. 418.
[47] Harada T, Soeda M, Enomoto T, Tokumitsu S, Khin M, Idemitsu T. Behaviour of
anchorage for FRP tendons using highly expansive material under cyclic
loading. In: Proc., third int. symposium on non-metallic (FRP) reinforcement
for concrete structures, FRPRCS-3, Sapporo, Japan, 1416 October, 1997. p.
71926.
[48] Benmokrane B, Xu H, Nishizaki I. Aramid and carbon bre-reinforced plastic
prestressed ground anchors and their eld applications. Can J Civ Eng
1997;24(6):96885.
[49] Benmokrane B, Zhang B, Chennouf A, Masmoudi R. Evaluation of aramid and
carbon bre reinforced polymer composite tendons for prestressed ground
anchors. Can J Civ Eng 2000;27(5):103145.
[50] Zhang B, Benmokrane B, Chennouf A. Prediction of tensile capacity of bond
anchorages for FRP tendons. J Compos Constr 2000;4(2):3947.
[51] Zhang B, Benmokrane B, Chennouf A, Mukhopadhyaya P, El-Safty A. Tensile
behavior of FRP tendons for prestressed ground anchors. J Compos Constr
2001;5(2):8593.
[52] Pincheira JA, Woyak JP. Anchorage of carbon ber reinforced polymer (CFRP)
tendons using cold-swaged sleeves. PCI J 2001;46(6):10011.
[53] Matta F, Nanni A, Abdelrazaq A, Gremel D, Koch R. Externally post-tensioned
carbon FRP bar system for deection control. Constr Build Mater
2007;23(4):162839.

J.W. Schmidt et al. / Construction and Building Materials 32 (2012) 110121


[54] Burgoyne CJ. External prestressing with polyaramid ropes. In: Int. symposium
on external prestressing in bridges ACI convention, Houston, USA; November,
1988. 19 p.
[55] PTI. Acceptance standards for post-tensioning systems. 1st ed. Post-Tensioning
Institute, Phoenix, USA; 1998. 53 p.
[56] PTI. Post-tensioning manual. 6th ed. Post-Tensioning Institute, Phoenix, USA;
2006. 354 p.
[57] Al-Mayah A, Soudki K, Plumtree A. Experimental and analytical investigation
of a stainless steel anchorage for CFRP prestressing tendons. PCI J
2001;46(2):88100.
[58] Sayed-Ahmed EY. Single- and multi-strand steel anchorage systems for CFRP
tendons/stays. In: Proc., 4th structural speciality conference of the canadian
society for civil engineering, Montral, Qubec, Canada; 58 June, 2002. p.
138190.
[59] Al-Mayah A, Soudki K, Plumtree A. Development and assessment of a new
CFRP rod-anchor system for prestressed concrete. Appl Compos Mater
2006;13(5):32134.
[60] Tepfers R. Cracking of concrete cover along anchored deformed reinforcing
bars. Mag Concr Res 1979;31(106):312.
[61] Mochida S, Tanaka T, Yagi K. The development and application of a ground
anchor using new materials. In: Proc., 1st Int. conf. on advanced composite
materials in bridges and structures, Sherbrooke, Qubec, Canada; 79 October,
1992. p. 393402.
[62] Nanni A, Al-Zaharani MM, Al-Dulaijan SU, Bakis CE, Boothby TE. Bond of FRP
reinforcement to concrete experimental results. In: Proc., second int. RILEM
symposium, FRPRCS-2, Ghent, Belgium, 2325 August, 1995. p. 13545.
[63] Al-Mayah A, Soudki K, Plumtree A. Effect of sandblasting on interfacial contact
behavior of carbonber-reinforced polymermetal couples. Comp Constr
2005;9:28995.
[64] Al-Mayah A, Soudki K, Plumtree A. Effect of sandblasting on interfacial contact
behavior of carbonber-reinforced polymermetal couples. J Compos Constr
2005;9(4):28995.

121

[65] Schn J. Coefcient of friction for aluminum in contact with a carbon ber
epoxy composite. Tribol Int 2004;37(5):395404.
[66] Schn J. Coefcient of friction and wear of a carbon ber epoxy matrix
composite. Wear 2004;257(34):395407.
[67] Al-Mayah A, Soudki K, Plumtree A. Effect of sleeve material on interfacial
contact behavior of CFRP-metal couples. J Mater Civ Eng 2006;18(6):82530.
[68] Campbell TI, Keatley JP, Barnes KM. Analysis of an anchorage for CFRP
prestressing tendons. In: Proc., annual conf. canadian society for civil
engineering, Halifax, Canada; 1998. p. 55160.
[69] Campbell TI, Shrive NG, Soudki KA, Al-Mayah A, Keatley JP, Reda MM. Design
and evaluation of a wedge-type anchor for bre reinforced polymer tendons.
Can J Civ Eng 2000;27(5):98592.
[70] Shaheen E, Shrive NG. Reactive powder concrete anchorage for post-tensioning
with
carbon
ber-reinforced
polymer
tendons.
ACI
Mater
J
2006;103(6):43643.
[71] Roark RJ, Young WC. Formulas for stress and strain. New York, USA: McGrawHill Book Company; 1975.
[72] Al-Mayah A, Soudki K, Plumtree A. Novel anchor system for CFRP rod: niteelement and mathematical models. J Compos Constr 2007;11(5):46976.
[73] Wang CT. Applied elasticity. New York, USA: McGraw-Hill Book Company;
1953.
[74] Burr AH, Cheatham JB. Mechanical analysis and design. Englewood Cliffs, New
Jersey, USA: Prentice Hall Inc.; 1995.
[75] Al-Mayah A, Soudki K, Plumtree A. Mechanical behavior of CFRP rod anchors
under tensile loading. J Compos Constr 2001;5(2):12635.
[76] Al-Mayah A, Soudki K, Plumtree A. Variable thickness barrel anchor for CFRP
prestressing rods. ACI Spec Publ 2005;231:23752.

Das könnte Ihnen auch gefallen