Sei sulla pagina 1di 10

Composite Structures 132 (2015) 923932

Contents lists available at ScienceDirect

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

Shear behaviour of masonry walls strengthened by external bonded FRP


and TRC
Thi-Loan Bui a, A. Si Larbi b,, N. Reboul c, E. Ferrier c
a

Institute of Construction Engineering, University of Transport and Communications, 3 Cau Giay, Lang Thuong, Dong Da, Ha Noi, Viet Nam
Universit de Lyon, Ecole Nationale dIngnieurs de Saint-Etienne (ENISE), Laboratoire de Tribologie et de Dynamique des Systmes (LTDS), UMR 5513, 58 rue Jean Parot, 42023
Saint-Etienne Cedex 2, France
c
LGCIE, Universit Claude Bernard LYON 1, 82 bd Neils Bohr, 69622 Villeurbanne, France
b

a r t i c l e

i n f o

Article history:
Available online 2 July 2015
Keywords:
Masonry wall
Strengthening
FRP
TRC
Shear behaviour

a b s t r a c t
This experimental study focuses on the behaviour of hollow concrete brick masonry walls, especially
walls reinforced with composite materials under in-plane loading conditions. This work is a step towards
dening reliable seismic strengthening solutions. Indeed, in France, more stringent seismic design
requirements for building structures have been considered with the replacement of old design codes.
Thus, an experimental program has been performed at the laboratory scale. Six walls have been submitted for shearcompression tests ve walls are reinforced by (1) bre-reinforced polymer (FRP) strips
using E-glass and carbon fabrics and/or (2) a textile-reinforced concrete (TRC), and the last wall acts as a
reference. It is noted that the composite strips are mechanically anchored into the foundations of the
walls to improve their efciency. All of the walls share the same boundary and compressive loading
conditions, which are representative of a seismic solicitation. Nevertheless, masonry wall performances
and anchor efciency are only evaluated under monotonic lateral loadings. A comparative study on global
behaviour and on local mechanisms is performed and, in particular, highlights that the mechanical
anchor systems play an important role in improving the behaviour of reinforced walls (by FRP and
TRC) and that the solutions for strengthening by TRC permit the upgrade of the walls ductility with a
lower strength compared with the solutions with FRP.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Masonry has a long history as a building technique. Even if reinforced concrete and steel prevail in the modern structures,
masonry units are also used. In France, a signicant part of buildings is erected with hollow concrete blocks. However, a relatively
important manufacturing tolerance and a design with large holes
give these blocks and even more to hollow concrete block structures a complex behaviour. Therefore, it is obvious that we
should pay attention to these structures in a seismic context, particularly when a seismic hazard assessment has been revised, leading to a tightening of the safety rules in France.
Indeed, past earthquakes have revealed that unreinforced
masonry structures can suffer extensive damage. Their vulnerability often lays in the weakness of mortar joints in tension and shear,
which are adversely and highly subjected to shear stresses during
earthquakes [1,2].
Corresponding author. Tel.: +33 4 77 43 75 38; fax: +33 4 78 43 33 83.
E-mail address: amir.si-larbi@enise.fr (A. Si Larbi).
http://dx.doi.org/10.1016/j.compstruct.2015.06.057
0263-8223/ 2015 Elsevier Ltd. All rights reserved.

In brief, due to seismic actions, walls in a building can be


subjected to shear forces both in the in-plane and out-of-plane
directions. The in-plane structural walls (i.e., shear walls, subjected
to lateral load along their longitudinal axis) are the primary force
resisting elements [3]. Out-of-plane walls (i.e., exural walls,
subjected to lateral load transverse to their longitudinal axis) are
in turn excited and if they are not resistant enough, their collapse
may disrupt the stability of the building and can result in a major
loss of life and property. Although these out-of-plane failures
should not be overlooked, practitioners (in a broad sense, including
the scientic community) tend to make the in-plane seismic
response of shear walls their rst priority; they indeed appear as
key vertical components to bear seismic loading.
Solutions for repairing or strengthening masonry structures are
many and are varied. Nevertheless, externally bonded brereinforced polymer (FRP) composites are often preferentially chosen by prime contractors [4], mostly because of their lightweight
and their ease of use. However, the reinforcing efciency of FRP
is rarely fully valued when they are only externally bonded to
structural elements. FRP mechanical properties are limited because

924

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

of the debonding of the composite sheets. To address this issue, an


adequate mechanical anchorage system needs to be set up to
enhance the bond (between a masonry structure and its foundation)
performance. The benets of this solution in terms of the FRP
efciency and lateral load resistance of a masonry wall have
now been widely acknowledged in the case of out-of-plane
actions [5].
In addition, in the context of sustainable development and
health and safety conditions for workers, consideration should
be given to an alternative material to FRP, which is often
manufactured with highly toxic epoxy resins. The idea is to
substitute these resins with cementitious materials while preserving or even improving the dissipative capacity of reinforced
structures. From this perspective, textile-reinforced concrete
(TRC) composites, which combine a suitable ne-grained mortar
with the latest generation of textile fabrics, would benet from
promotion.
The efciency of TRC for strengthening masonry structures has
recently been investigated [611]. Compared with FRP, TRC composites show a nonlinear tensile behaviour with multiple matrices
cracking, giving them a greater deformation capacity, a priori more
suitable for seismic reinforcement [8].
Although instructive, these studies lack diversity for studied
materials, reinforcement congurations, applied normal loads
and slenderness ratio of walls. Sometimes, a small amount of information is known regarding damage and failure mechanisms or
regarding the interaction between masonry material and
reinforcements.
On the one hand, this work is aimed at further developing the
existing experimental database, with special emphasis on
identifying the performances of anchorage devices, particularly
in the framework of a comparative study between FRP and
TRC composites. This comparison will cover criteria at the global
scale and, to a lesser extent, at the local scale. On the other
hand, this paper tries to help identify and clarify damage
dissipative mechanisms and their impact on the failure modes
of the masonry walls.
To attain the aforementioned objectives, an experimental campaign has been performed, based on static monotonic shear tests,
which are a simplied way to simulate stress states resulting from
earthquakes. Certainly, inertial effects and the inherent cyclical
nature of seismic actions are not addressed in the present study.
However, this work can be regarded as a rst step towards the
denition of efcient reinforcement solutions. The approach is to
test some strengthening congurations to have relevant and
valuable information and to offer prospects that would be
appropriate to assess with more realistic loadings in terms of
earthquake hazards.

2. Experimental program
2.1. Masonry walls
A series of six walls has been built with the same dimension
given in the Fig. 1. It should be mentioned that all of the specimens
were built by a professional mason and must be considered to be in
compliance with the practices. The hollow concrete block units,
whose dimensions are 500 mm long, 200 mm high and 75 mm
thick, belong to Group 2 according to Eurocode 6, with a strength
class B40 (characteristic compression strength of 4 MPa).
However, these blocks have been halved lengthwise before being
assembled to make walls dimensions compatible with the limited
means of the laboratory in terms of space and actuator capacity
(Block work size at reduced scale: 250  200  75 mm3).
The compressive strength of the individual masonry blocks has
been determined and ranges from 4 to 10 MPa (6.5 MPa on average
with a standard deviation of 2.33). These blocks are assembled
with a mortar composed of Portland cement (CEM I 52.5) and sand
in the proportion 1:3 with a water/cement ratio equal to 0.5.
Mortar test prisms of 40  40  160 mm3 were tested for compressive and exural strengths. At 31 days, these strengths are 48 MPa
and 10 MPa, respectively.
2.2. Reinforcement
2.2.1. Strengthening materials
Two types of composites have been used: the rst composite is
a bre-reinforced polymer (FRP) while the latter composite is a
textile-reinforced cementitious composite (TRC).
2.2.1.1. FRP composite. The bre-reinforced composite materials
consist of a two-component epoxy matrix and bi-directional fabrics made of either carbon (CFRP) or glass (GFRP). Their mechanical
characteristics have been measured on six specimens according to
ISO 527-1. The obtained results are listed in Table 1.
2.2.1.2. TRC composite. Knowledge on TRC composites is notably
less signicant than knowledge relating to FRP. However, it is
Table 1
Mechanical characteristics of composites.
Composite
strengthening
system

Nominal
thickness
(mm)

Young
modulus
(GPa)

Tensile
strength
(Mpa)

Ultimate
strain (lm/
m)

CFRP
GFRP

0.48
1.7

105
7.2

1700
100

16000
13.800

Concrete loading beam

Reinforced concrete foundation

Fig. 1. Description of unreinforced masonry wall (reference).

T.-L. Bui et al. / Composite Structures 132 (2015) 923932


Table 2
Mix design for TRC composite.
Micro-mortar**

Textile reinforcement

*
**

Nature of bres
TEX
Fibre diameter
Number of
lament/yarn
Knitted grid size
Tensile strength (yarn)

Glass-AR
1200*
19 lm*
1600*

Grain size
Silica-fume
Thixotropy
Shrinkage

<2 mm*
Yes*
Yes*
0*

5  5 mm*
1102 MPa*

5 MPa**
40 MPa*

Young modulus

74,000 MPa*

Tensile strength
Compressive
strength
Young modulus

1700 MPa*

Provided properties.
Laboratory characterisation.

currently well established that various levers (roving diameters,


yarn number, impregnation, etc.) can be mobilised to optimise
TRC with occasionally opposite consequences and many of
them have already been identied [11].
Therefore, in the course of works performed by Contamine et al.
[12], it has been possible to select a composite, whose components
are listed in Table 2, that results from a compromise between all of
the aforementioned parameters, including workability and thixotropy (to apply strengthening materials more easily).
The composite material contains the 4.36% AR-glass bre volume fraction, that is, 2.18% in each principal direction. The textile
used for reinforcement is a bidirectional warp-knitted grid fabric.
Direct quasi-static tensile tests, whose protocol has been validated [13], have been performed to mechanically characterise
the TRC reinforcements. Stressstrain curves are given in Fig. 2.
Two different behaviour laws appear in Fig. 2. Indeed, without
the impregnating resin (latex), TRC exhibits a bilinear behaviour
whereas by using latex, the evolution law is nearly linear because
of a more homogeneous yarn contribution.
2.2.2. Anchorages
To reduce or ideally remove the overturning effects due to
lateral loads on walls and to best maximise the potential of each
reinforcement solution by a priori improving its efciency, a
connector, in the form of an anchorage (see Fig. 3), has been introduced between walls (on their lower part, on both sides) and their
foundation footings.

Fig. 2. Tensile behaviour of TRC materials during uniaxial tensile tests.

Fig. 3. Description of a MAPEI anchor.

925

Given the existing solutions and their well-established performances (easy application and high strength), only connection
solutions based on carbon bres and an epoxy resin have been considered. The anchorage system (MAPEI) is an anchor made from
monodirectional carbon bres with at least 36 yarns, each including 12,000 bres. The anchorage strength given by the manufacturer is 30 kN at the ultimate limit state.
2.2.3. Strengthening congurations
The denition of strengthening patterns must be part of a
strategy aimed at nding a balance between lateral strength
and energy dissipation capacity [14]. Thus, the objective in reinforcing a structure is to improve its strength capacity, to enhance
its ductility, or both.
According to this strategy, different strengthening congurations with TRC and FRP composites have been proposed. The
reinforcing material is always applied symmetrically on both wall
surfaces.
The experimental program consists of testing six masonry walls
to failure, including an unreinforced wall (as the reference specimen) and ve TRC or FRP-reinforced specimens (see Fig. 4). With
FRP composites, three strengthening patterns have been proposed.
The rst wall, referenced CGRW, has been reinforced with both carbon bre-reinforced polymer (CFRP) and glass bre-reinforced
polymer (GFRP) to signicantly improve strength capacity. Each
side of both faces is reinforced by a continuous sheet, is comprised
of two glass layers over a width of 400 mm along the entire height
of the masonry and is combined with two discontinuous carbon
sheets, which are 1410 mm long and 60 mm wide; the horizontal
distance between these two carbon sheets is 100 mm. The two
remaining walls, for their part, have been reinforced by either
CFRP sheets (CRW wall) or GFRP sheets (GRW wall) to combine
strength capacity and ductility.
Concerning TRC solutions, embedding glass fabrics (rather than
carbon fabrics) in an epoxy matrix tends to produce composites
with better energy dissipation capacities. Thus, it is appropriate
to take advantage of this property in the context of reinforcing
masonry walls. This choice to use only glass bres is more appropriate because it adds value to TRC materials (compared with
carbonepoxy composites), which achieves a smaller ecological
footprint and has fewer problems with the hygiene and safety conditions for workers.
As a consequence, it is clear that without overlooking the
loadbearing capacity, it is desirable to increase the ductility of reinforced walls, which is the main motivation of our choices. In the
rst pattern (TRCW1 wall), only one TRC layer, 1410 mm high
and 200 mm wide, is applied on each side of both faces.
However, in the second selected pattern, the main objective is to
improve the dissipation capacities, so a vertical strip is applied in
the middle of the wall. Each strip is made of three TRC-layers,
which are 1410 mm high and 200 mm wide.
2.2.4. Application of strengthening systems and anchor placement
FRP/TRC reinforcements have been laid up. First of all, wall
surfaces have been cleaned to remove dust and loose materials,
which could disturb the bond between the masonry wall and its
reinforcements. The application of a strengthening system rst
involves covering the wall with a layer of epoxy resin (FRP) or mortar (TRC). Then, the fabric (FRP) or textile (TRC) is placed along the
wall and is pressed against the resin or mortar. A second layer of
resin or mortar is eventually applied to ensure fabric impregnation.
For walls reinforced with several layers, the last two steps are
repeated as necessary.
These reinforcements are anchored to the foundation because of
the CFRP structural connections. An anchorage anchor is composed
of two parts: the anchorage strictly speaking and the whip. The

926

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

anchorage that is rst impregnated with epoxy-resin is inserted


into holes that have been drilled before in the reinforced concrete
footing and is lled with resin. The remaining part of the anchor
(the whip) is deployed similar to a fan on the wall, and resin is
again applied so that the adhesive completely penetrates into the
bres. This splayed part ensures the connection of strengthening
materials with the masonry wall.
2.3. Testing procedure
As mentioned above, monotonic in-plane combined compressionshear tests have been selected.
The vertical axial load (N) is centred and applied by means of a
hydraulic actuator with 200 kN capacity. This actuator is combined
with a force sensor that is located on a stiff steel plate (100 mm
wide  100 mm long  20 mm thick) and is placed on the concrete
loading beam. These devices are held by a steel prole connected
to the strong oor through steel rods, which allows us to control
the applied vertical load (Fig. 5).
The vertical load is should represent the weight of the upper
oors. In literature, this value varies greatly and depends on the
masonry compression strength. Tomazevic [15] has proposed to
adopt a constant value equal to 20% of the compression strength
for conned brick masonry. Papanicolaou et al. [9] have worked
with large ranges from 2.5% to 10% for rectangular walls and from
10% to 25% for slender walls. This value is between 11% and 16% in
F.da Portos study [10] regarding steel-reinforced masonry. For a
given wall geometry, the axial compression load magnitude inuences the overall performances, in particular, the efciency of
strengthening solutions. Papanicolaou et al. [9] have noted that
reinforcement efciency is reduced as vertical compression load
rises. Because this study is motivated by assessing the shear contribution of the reinforcements, a relatively low value of approximately 6% of the masonry compression strength has been
adopted. It corresponds to a mean vertical stress of 0.2 MPa
(15 kN).
In practical terms, the vertical load is very slowly applied up to
the target value equal to 15 kN. It is kept constant during the test
because of a force control of the vertical actuator. At this time, horizontal load is imposed under quasi-static monotonic conditions.
This loading step is performed under displacement control at a rate
of 0.015 mm/s to better capture post-pic behaviour. Lateral loading
is stopped when masonry walls have obviously failed, that is, when
lateral force drops signicantly. According to Tomazevic [15], failure is ascertained when horizontal load falls by 20%. This criterion
will be adopted herein.
In addition to the force sensors that have been presented before
in the description of the test set-up, for the sake of clarity, several
displacement transducers and strain gauges complete the instrumentation. Their locations and purposes will now be discussed.
The displacement transducer C1 (LVDT 100 mm) measures
the lateral displacement at the top of the wall and will enable us
to characterise the overall behaviour of the wall through loaddisplacement curves. The assumption that the foundation footing has
no slips is controlled because of the displacement transducer C3
(LVDT 100 mm). To have information at the local scale and to
assess the contributions of different reinforcements, strain gauges
(120 X), numbered from J1 to J9, are bonded on only one face of
the walls (Fig. 4).

walls do not heave. Concerning measures from LVDT C3, they


emphasise that the concrete footing relative to the rigid oor does
not slide for all six walls. As a consequence, the obtained relationships of lateral loads versus horizontal displacements at the top of
the wall can be considered without caution towards boundary
conditions.
All walls show a nonlinear behaviour after an initial linear
elastic branch. A signicant deviation is experienced in the
length of the linear zone although stiffnesses are close. Until
the end of this rst phase, wall integrity is preserved at the global scale. Next, a nonlinear phase starts, which differs depending
on the nature of the walls and reinforcements. This nonlinear
zone can be related to the damage (in tension or in shear) of
one or more masonry components, the strengthening materials
or even the reinforcement/block (or mortar joint) interface
(interphase).
A dramatic increase in the ultimate strength and in the second
phase stiffness is clearly observed, even if it is conditioned by reinforcing materials and strengthening congurations. Similarly, dissipation capacities increase overall in varying degrees that must
be assessed and quantied. It must be underlined that only the
wall CGRW behaves as brittle (as a result of a sudden and premature failure) (Fig. 6).
To evaluate the performances of these walls in terms of appropriately chosen and unbiased indicators, the experimental
loaddisplacement curves will be idealised according to the trilinear model proposed by Tomazevic [15] (Fig. 7). Given the experimentally observed behaviours, this approach seems to be more
consistent than the idealised bilinear relationship proposed by
Magenes and Calvi [16].
The three phases of the conventional trilinear diagram are
bounded by three characteristic points, which facilitate discussion
and comparison on wall behaviours. Thus, energy dissipation, calculated from the idealised diagrams, will be useful to assess the
suitability of strengthening solutions with dissipation needs in
the context of earthquakes. The rst zone, called elastic, ends at
lateral load Vcr and displacement dcr. They mark the formation of
the rst signicant cracks in the wall, which entail a change in
stiffness [3]. As masonry walls exhibit highly nonlinear behaviours,
a conventional Vcr value is generally adopted. According to
Tomazevics study [15], Vcr is equal to 70% of the maximum resistance Vmax. The second zone extends to the maximum lateral load
Vmax and displacement dVmax. At last, the ultimate zone is characterised by the ultimate load Vdu, corresponding to 80% of the maximum load and the maximum displacement du on the softening
branch.
Kel, lu and Ediss are the initial stiffness (the slope of the rst
phase, considered as elastic), ductility coefcient and dissipation
energy, respectively. This last parameter provides information on
resistance and deformation capability. Indeed, it is determined as
the area below the idealised loaddisplacement diagram and is
given by the following equation:

Ediss

1
dcr :V cr dVmax  dcr V max V cr du  dVmax V max
2
V du 

The ductility coefcient, obtained by dividing the ultimate horizontal displacement (du) by the elastic displacement (dcr), reects the
deformation capacity of shear walls in the post-elastic zone. This
parameter is a decisive criterion for paraseismic construction.

2.4. Experimental results


2.5. Strength capacity
2.4.1. Global behaviour
The number of steel rods, their high axial rigidity and their suitable tightening, without overlooking the high exural rigidity of
transverse steel girders, are many arguments for assuming that

From these results, it appears that reinforced walls achieve substantially higher ultimate loads than the reference unreinforced
wall, regardless of reinforcement type. Shear strength increases

927

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

URW

TRCRW2

TRCRW1

(One layer without latex)

(3 layers with latex)

CRW

GRW

CGRW

Fig. 4. Unreinforced wall and FRP/TRC-reinforced walls.

Spreader beam
Load cell
Hydraulic jack

Load cell

Vertical load actuator

Post-tension rod
Wall
Beams

Reaction wall

Threaded rod

Fig. 5. Test set-up.

from 110% (TRCW1) to 450% (CGRW). TRC reinforcements are


slightly less efcient (175% on average) than FRP reinforcements
in terms of strength capacity. A more detailed analysis allows a
quantitative comparison of the different congurations with TRC
and FRP materials and highlights that strength capacity gains grow
proportionally to reinforcement axial stiffness (qv :ER ) (The case of
the CGRW wall has been eluded because it collapsed when the
anchorages prematurely failed) Fig. 9. This relationship is consistent with Mahmood et al. [17].

It must be noted that the TRCW1 wall appears as an exception


because its reinforcement has a slight effect on the ultimate load. It
can be explained by the excessively low stiffness of this reinforcement, only comprising one TRC layer.
2.6. Initial stiffness
Initial stiffness is marginally affected by reinforcements. This
can stem from the low thickness of strengthening systems, in

928

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

2.8. Dissipation energy


In addition to the ductility factor, the dissipation energy (Ediss)
will help us position, at least in order of magnitude, the potentials
of FRP and TRC as masonry reinforcements. It is clear that for the
adopted congurations, both with TRC (with the exception of the
TRCRW1 wall with a very low reinforcement ratio) and FRP, results
are conclusive because increases in the range of 200500% can be
expected. To better understand, it is important to correlate the
above performance indicators with observed failure modes, damage mechanisms and kinematics and with data at the local scale.
This is the purpose of the rest of this paper.

Fig. 6. Curves of load versus horizontal displacement at the top of the wall.

particular with FRP materials and to a lesser extent with TRC materials. By contrast, the size of the linear zone is strongly subordinated to the adopted reinforcements nature and conguration.
Indeed, according to experimental results (Vcr indicator seems
irrelevant or at least unsuitable for results obtained in the present
case and presented as loaddisplacement curves), overall, wall
macroscopic damage is delayed, except for the TRCW1 wall whose
reinforcement ratio is low.
In the rst stage, the discernible increase in initial stiffness may
be attributed to the ability of the reinforcement (mainly when its
axial stiffness is sufcient) to bridge emerging microcracks and
cracks because of a suitable load transfer.
2.7. Ultimate displacement and ductility
Furthermore, it is important to note that the ultimate displacement is a minimum (except for the CGRW wall) maintained when
compared with the reference wall. This emphasises the interest of
chosen reinforcement solutions and a priori attests to the usefulness of anchorages between footings and walls.
It is also worth noting that two reinforced walls (TRCW2 and
GRW) see a tangible increase, of approximately 36% in their ultimate displacements. Walls are likely subject to exural mechanisms, which induce normal tensile stresses in reinforcements.
Thus, the ultimate displacement capacity of strengthened masonry
walls is, among other factors, certainly conditioned by the axial
stiffness of reinforcements, which can explain at least partially
the increase in stiffness for these reinforced walls.
The decrease in ultimate displacement is peculiar to the CGRW
wall because anchorages have certainly been involved in the wall
failure (see below). As mentioned previously, the ductility factor
(l = du/dcr) plays a crucial role in assessing the seismic behaviour
of masonry walls in particular because it reduces the number
of elastic seismic design actions. Table 3 highlights that
TRC-reinforced walls exhibit nearly the same ductility factor as
an unreinforced wall. In contrast, for the FRP-reinforced walls case,
if lateral strength clearly increases, the ultimate ductility coefcient is low compared with an unreinforced wall (Fig. 8d and
Fig. 10). This may be related to the capacity that TRC composites
have, unlike FRP composites, to crack, to follow masonry wall
displacements and potentially to absorb wall damage, which
thus remains controlled.
Needless to say, the above assumptions require a more elaborate experimental campaign to be validated, but some of the proposed explanations will also be reviewed regarding available
data at the local scale.

2.8.1. Failure modes


This section is devoted to providing information regarding failure modes based on reinforcement types and patterns, so that the
efciency of strengthening materials in the global lateral behaviour
and their impact on damage mechanisms can be precise.
2.8.1.1. Unreinforced wall. An unreinforced wall exhibits a exural
failure mode (Fig. 11) characterised by horizontal cracks on the left
part of the wall (on the side where lateral load is applied) due to
tensile stresses in bed mortar joints and by toe crushing (on the
right part) at the compressed corner.
To better understand the failure process, the order in which
damage has been visually detected is as follows. Horizontal cracks
start to open on the left side probably because the tensile bond
strength is exceeded and gradually develops as the crack opening
grows. These exural cracks involve the rst and second rows of
mortar bed joints. It is not excluded that high shear stresses at
the left bottom corner have contributed to the initial failure of
the rst line of bed joints rst. The increasing lateral displacements
therefore induce the spreading of cracks through the second row of
bed mortar joints. At this point, the wall tends to rotate about the
right bottom corner, thereby inducing the horizontal spread of
cracks (secondarily, vertical joints between the rst and second
rows are damaged). As the resistant cross section is reduced, the
right bottom corner is compressed. It eventually fails under a complex stress state as evidenced by concentrated crushings together
with cracks on an approximately 45-degree angle.
It is worth noting that observation of this failure mode has
given valuable information for designing strengthening congurations in the present study.
More specically, the objective is to counteract or at least defer
damage of the horizontal joints that are subjected to predominantly tensile stresses because of vertical strips at the ends of walls

Fig. 7. Idealised tri-linear diagram.

929

T.-L. Bui et al. / Composite Structures 132 (2015) 923932


Table 3
Summary of experimental results.
Walls

URW
CGRW
CRW
GRW
TRCRW1
TRCRW2

CFRP
GFRP
CFRP
GFRP
TRC
TRC

t R  bR
(mm2)

AR
qv lt

qv :ER
(MPa)

Vcr = (70%Vmax)
(kN)

dcr
(mm)

Vmax
(kN)

dVmax
(mm)

Vu = (80%Vmax)
(kN)

du
(mm)

0.48  60
1.7  400
0.48  60
1.7  400
3  200
9  200

3.82

82,020

7.88
35.79

0.88
3.66

11.25
51.13

4.59
8.39

9.00
40.90

12.98
8.70

0.3
3.52
3.31
13.98

31,316
50,702
4854
22,135

22.40
35.53
8.61
18.92

2.50
3.06
0.75
1.07

32.00
50.75
12.30
27.03

9.94
8.89
5.85
7.32

25.60
40.60
9.84
21.62

13.20
17.70
12.80
17.60

l = du/
dcr

Ediss
(kN mm)

Failure modes

8.97
9.78

14.78
2.38

123.90
285.30

Flexural
Flexural + shear

8.96
11.61
11.48
17.68

5.28
5.78
17.07
16.45

324.26
708.24
133.49
403.80

Flexural + shear
Flexural + shear
Flexural + shear
Flexural + shear

Kel
(kN/
mm)

P
tR - thickness of composite band; bR - width of composite band; AR - total cross section area of strengthening (= bR :tR ); l and t- width and thickness of masonry wall; qv vertical reinforcement ratio.

Fig. 8. Comparative diagrams of the different indicators (a-strength capacity, b-stiffness, c-ultimate displacement, d-ductility coefcient, e-dissipation energy).

unreinforced zones are maintained as concentrated damage zones


in which energy dissipation can advantageously take place.
Of course, wall behaviour in terms of lateral strength or
energy dissipation capacity highly depends on the strength of
anchorages.

Fig. 9. Relationship between strength capacity and reinforcement axial stiffness


(qv :ER ).

(areas where tensile stresses mainly develop during an earthquake


loading cycle).
Moreover, the use of wide strips (or even central strips) aims at
covering a large surface to bridge cracks, while at the same time,

2.8.1.2. FRP-reinforced walls. Failure modes of FRP-reinforced walls


are presented in Fig. 12. First, it must be noted that they are different and depend, among other things, on adopted reinforcement
congurations. Walls predominantly exhibit shear failures with
different crack patterns that also correlate with the damage nature
either brittle failure or progressive degradation which has
major implications on the dissipation capabilities of walls.
The sudden failure of the CGRW wall is singular and results
from the premature failure of anchorages, which are subject to tensile forces in the left part (side where lateral load is applied). This
wall has been reinforced over a large surface area on which both
glass and carbon textiles have been laid up, resulting in a signicant increase in the wall stiffness (compared with an unreinforced
wall but also to other strengthened specimens). Under loading, the
wall begins a rigid body motion that induces high shear stresses at

930

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

for the CRW wall (which is a typical shear failure pattern), even
along the edge of the unreinforced zone (in particular for GRW
wall), thus reecting that reinforcements can bridge cracks and
also inuence their propagation.
Nevertheless, the collapse of these walls (or very marked deterioration) occurs by the crushing of the lower right corner (highly
subject to compression) and ends and by the splitting of the
extreme unit block. With the failure of this block, the wall tends
to turn about the right toe, as found for an unreinforced wall
(Fig. 12(b) and (c)).
Fig. 10. Reassessment of ultimate ductility factors by using the experimental elastic
limit rather than the conventional one.

Crushing of the
compressive brick

Cracks of mortar
joints

Fig. 11. Flexural failure for unreinforced wall (URW).

the bottom of the wall which could explain, even partially, the
observed failure mode.
As a consequence, the use of anchorage devices certainly
improves the lateral strength of reinforced walls. However, to
avoid sudden ultimate failures of walls and to ensure sufcient
energy dissipation capabilities, an anchorage design cannot be
decoupled from the stiffness or reinforced walls (taking into
account both reinforced surface area and strengthening materials).
The other two FRP-reinforced walls are not affected by anchor
failure and show similar damage mechanisms.
These walls exhibit coupled shearexure failure modes. In
both specimens, shear cracks initiate in the middle of the wall
(unreinforced zone) and propagate along the compressed diagonal

2.8.1.3. TRC-reinforced walls. As indicated above for FRP-reinforced


walls, the addition of TRC reinforcements changes the failure
mode. TRC-reinforced walls have failed by combined shearexure
(Fig. 13). Both reinforcement patterns lead to cracks at the left bottom part (side where load is applied) and in strengthening strips
(unlike FRP) over horizontal joints.
In the case of the TRCRW1 wall (low reinforcement ratio and
low reinforced surface area), cracks develop along a horizontal
joint and go through a block unit in the centre with an inclined
crack that suggests a reaction to a shear solicitation.
At the ultimate limit state, a macro crack has been observed
that corresponds to the failure of mortar used in TRC without
any textile degradation. In scientic literature, this failure mode
is commonly referred to as the peeling-off failure [18]
(Fig. 13a).
In the case of the TRCRW2 wall (reinforced by three TRC layers
on a larger area), damage and failure mechanisms are different.
Cracks grow horizontally and spread over the height of the
strengthening strips. Finally, multi-crack initiation contributes,
beyond what is observed on the wall, to the global damage of
the specimen. It must be noted that until failure, which ultimately
occurs by crushing the right bottom corner, crack opening displacement (without having been measured) remains limited
(Fig. 13b).
In light of the obtained results, it is important to emphasise
that:
- Anchors can be considered useful and efcient systems, but
their use must be correlated at this time with the global stiffness of the wall (reinforcements nature and surface area) to
avoid their sudden and premature failure.
- Dissipation mechanisms are improved and concentrated for all
reinforced walls, regardless of the adopted reinforcements type
and pattern. However, the dissipation process is controlled
because it occurs on the reinforcement itself for TRC materials,
whereas it is uncontrolled with FRP-debonding.

Crushing of
the
compressive
brick

Crack of FRP
Local debonding
s
propa
gation
Crack Crushing of
the
s
propa compressive
brick
gation

Failure of anchors

(a)

(b)

(c)

Fig. 12. Failure modes and damage mechanisms for FRP-reinforced walls: (a)- sudden failure of CGRW wall; (b) and (c)- coupled shearexure failure of CRW and GRW walls.

931

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

Macro-crack of TRC

Multi-cracks of TRC

(a)

(b)

Fig. 13. Failure modes and damage mechanisms for TRC-reinforced walls (a)- TRCRW1 wall and (b)- TRCRW2 wall.

Fig. 14. Evolutions of strains in FRP reinforcements along wall length for CGRW wall (a): GRW wall (b) and CRW wall (c).

2.8.2. Local behaviour


Only strains measured in FRP-reinforced walls are displayed
because gauges applied on TRC have failed early, without giving
valuable information. Moreover, because of complex local effects
(in zones where anchorage cords and FRP composites are held
together), only strain gauges (3; 6; 9) that are located above
anchorage cords are considered. The evolution of strains in vertical
FRP reinforcements along the wall length is given in Fig. 14 for different loading levels ranging from 0 to Vmax, the maximum lateral
strength. In general, maximum tensile strains are measured with
gauges located at the tensile side of the wall, and strains gradually
decrease, quasi linearly, to become negative at the compressed
side. These evolutions give credibility to the assumption that plane
sections remain plane, despite wall heterogeneity and until lateral
loading values are close to failure load.
Furthermore, at maximum load, maximum strains in FRP reinforcements reach values of 4000 lm/m in CFRP strips and
9000 lm/m in GFRP strips, respectively, for CRW and GRW walls.
These values are very close to the ultimate strain values for glass
bre-reinforced composites and account for only 50% of the ultimate strain value for carbon-based reinforcements. Thus, the usefulness of glass bres is evident. In contrast, in the case of the
hybrid solution with both glass and carbon bres, strains remain
small (2500 lm/m CGRW wall) because of anchor failure. This
limited strain highlights that without efcient anchorage, reinforcements cannot signicantly contribute to load transfer. This
is the reason why the global performance of this wall is limited.

3. Conclusions
The present experimental study has focused on masonry walls
reinforced by TRC and FRP composites that are subject to monotonic in-plane combined shearcompression tests. The main ndings are as follows:
- Reinforcements, regardless of their nature and the adopted layout diagram (provided that reinforcement ratio is sufcient),
enable us to extend the structural integrity eld of masonry
walls.
- TRC reinforcements lead to lower performance levels than FRP
reinforcements in terms of lateral strength capacity, but they
signicantly increase their ductility capacity.
- TRC and GFRP seem to be more appropriate than CFRP in terms
of ultimate displacement capacity.
- A low TRC reinforcement ratio only marginally modies global
masonry wall performances.
- Dissipation mechanisms, which differ between FRP and
TRC-reinforced walls, have been claried; in particular, they
diffuse dissipation processes (micro-cracks) in TRC
composites.
- Anchorage systems are appropriate (and technologically possible) to improve the in-plane performances of reinforced
masonry walls.
- Reinforcement design is limited by the compressive strength of
concrete hollow blocks.

932

T.-L. Bui et al. / Composite Structures 132 (2015) 923932

- The rigid body motion of a strengthened wall depends on the


reinforcement pattern (reinforcement ratio, axial rigidity and
reinforced surface area) and is likely to substantially limit the
contributions of anchorage systems.

References
[1] Fallahi A, Alaghebadian R, Miyajima M. Microtremor measurements and
building damage during the Changureh-Avaj, Iran earthquake of June 2002. J
Nat Disaster Sci 2003;25(3):3746.
[2] Klingner R. Behavior of masonry in the Northridge (US) and Tecoman-Colima
(Mexico) earthquakes: lessons learned, and changes in design provisions.
6_Congresso Nacional de Sismologia e Engenharia Sismica, Guimares,
Portugal; 2004.
[3] Capozucca R. Experimental analysis of historic masonry walls reinforced by
CFRP under in-plane cyclic loading. Compos Struct 2011;94:27789.
[4] Lanivschi CE. State of the art for strengthening masonry with Fibre Reinforced
Polymers. Buletinul Institutului Politehnic Din IASI Publicat de Universitatea
Tehnica, Gheorghe Asachi din Iasi Tomul LVIII (LXII), Fasc. 2, 2012 Sectia
ConstrucTII. Arhitechtura.
[5] Lunn Dillon S, Rizkalla SH, Maeda S, Ueda T. FRP anchorage systems for inll
masonry structures; APFIS 2012 Hokkaido University Japan, 24 february
2012.
[6] Bernat E, Gil L, Roca P, Escrig C. Experimental and analytical study of TRM
strengthened brickwork walls under eccentric compressive loading. Constr
Build Mater 2013;44:3547.
[7] Carozzi FG, Milani G, Poggi C. Mechanical properties and numerical modeling
of fabric reinforced cementitious matrix (FRCM) systems for strengthening of
masonry structures. Compos Struct 2014;107:71125.

[8] Papanicolaou CG, Triantallou TC, Karlos K, Papathanasiou M. Textilereinforced mortar (TRM) versus FRP as strengthening material of URM walls:
in-plane cyclic loading. Mater Struct 2007;40:108197.
[9] Papanicolaou CG, Triantallou T, Lekka M. Externally bonded grids as
strengthening and seismic retrotting materials of masonry panels. Constr
Build Mater 2011;25:50414.
[10] da Porto Franca, Mosele Flavio. Claudio Modena in-plane cyclic behaviour of a
new reinforced masonry system: experimental results. Eng Struct
2011;33:258496.
[11] Prota A, Marcari G, Fabbrocino G, Manfredi G, Aldea C. Experimental in-plane
behavior of tuff masonry strengthened with cementitious matrix-grid
composites. J Compos Constr ASCE 2006;10(3):22333.
[12] Contamine R, Si Larbi A, Hamelin P. Contribution to direct tensile testing of
textile reinforced concrete (TRC) composites. Mater Sci Eng: A
2011;528:858998.
[13] Contamine R, Si Larbi A, Hamelin P. Identifying the contributing mechanisms
of textile reinforced concrete (TRC) in the case of shear repairing damaged and
reinforced concrete beams. Eng Struct 2013;46:44758.
[14] Davidovici: Risque sismique: renforcer des btiments existants; ENIT
October 2008.
[15] Tomazevic M, Lutman M, Petkovic L. Seismic behavior of masonry walls:
experimental simulation. J Struct Eng ASCE 1996;122(9):10407.
[16] Magenes G, Calvi GM. In-plane seismic response of brick masonry walls.
Earthquake Eng Struct Dyn 1997;26:1091112.
[17] Mahmood H, Ingham JM. Diagonal compression testing of FRP-retrotted
unreinforced clay brick masonry wallettes. J Compos Constr ASCE
2011;15(5):81020.
[18] Banholzer B. Bond behaviour of a multi-lament yarn embedded in a
cementious matrix [Philosophy doctoral thesis]. Schriftenreihe Aachener
Beitrge zur Bauforschung. Institut fr Bauforschung der RWTH Aachen
2004; 12.

Potrebbero piacerti anche