Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

ARTICLE IN PRESS

Composite Structures xxx (2005) xxx–xxx


www.elsevier.com/locate/compstruct

Hazard mitigation and strengthening of unreinforced masonry


walls using composites
a,*
W.W. El-Dakhakhni , A.A. Hamid a, Z.H.R. Hakam b, M. Elgaaly c

a
Department of Civil Engineering, McMaster University Centre for Effective Design of Structures, Hamilton, ON, Canada L8S 4L7
b
Bechtel Power Corporation, Frederick, MD 21703, USA
c
Civil, Architectural and Environmental Engineering Department, Drexel University, Philadelphia, PA 19104, USA

Abstract

An experimental investigation was conducted to study the behavior of unreinforced masonry (URM) walls retrofitted with com-
posite laminates. The first testing phase included testing 24 URM assemblages under different stress conditions present in masonry
walls. Tests included prisms loaded in compression normal and parallel to bed joints, diagonal tension specimens, and specimens
loaded under joint shear. In the second testing phase, five masonry-infilled steel frames were tested with and without retrofit.
The composite laminates increased the stiffness and strength and enhanced the post-peak behavior by stabilizing the masonry walls
and preventing their out-of-plane spalling. Tests reported in this paper demonstrate the efficiency of composite laminates in improv-
ing the deformation capacity of URM, containing the hazardous URM damage, preventing catastrophic failure and maintaining the
wall integrity even after significant structural damage.
 2005 Elsevier Ltd. All rights reserved.

Keywords: Composite masonry; Concrete masonry; Fiber reinforced plastics; Retrofitting; Seismic hazard; Seismic loads; Steel frames

1. Introduction safety is the shaking damage and the collapse of build-


ings and other structures that have been inadequately
Earthquakes have long been recognized as one of the designed or poorly constructed. Major earthquakes
most damaging natural hazards, along with hurricanes, can severely disrupt regional or national economic activ-
tornadoes, floods and fire. No other force in nature ity by damaging social lifelines such as roads, railways,
has the potential to wreak so much havoc in such a short water, power and communications infrastructures and
time. Earthquakes typically strike without warning and, office and residential buildings.
after only a few seconds, leave casualties and damage in A common type of construction in urban centers for
their wake. Although earthquakes cannot be prevented, low-rise and mid-rise buildings is unreinforced masonry
the current state-of-the-art in science and engineering (URM) walls filling the space bounded by the structural
provides new tools that can be used to reduce their dam- framing members. Although considered non-structural
aging effects. Through prudent action, the loss of life, elements, yet under seismic excitation, infill walls tend
serious injury, and property damage as well as social to interact with the surrounding frame and may result
and economic disruptions resulting from earthquakes in different undesirable failure modes both to the frame
can be reduced. The principal threat to human life and and to the infill wall [1]. In general, URM infill walls
have demonstrated poor performance record even in
*
Corresponding author. Tel.: +1 905 525 9140x26109; fax: +1 905
moderate earthquakes. Their behavior is usually brittle
529 9688. with little or no ductility and they, typically, suffer
E-mail address: eldak@mcmaster.ca (W.W. El-Dakhakhni). various types of damage ranging from invisible cracking

0263-8223/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compstruct.2005.02.017
ARTICLE IN PRESS

2 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

Nomenclature

DH horizontal extension g vertical gage length


DV vertical shortening h diagonal tension specimen height
c shear strain n percentage of solid in the masonry unit
AFRP cross-sectional area of the FRP per unit P applied load on the diagonal tension speci-
length of wall mens
An net area of the diagonal tension specimen Ss shear stress based on the net area
As cross-sectional area of the steel reinforcement t diagonal tension specimen thickness
per unit length of wall w diagonal tension specimen width
EFRP modulus of elasticity of the FRP
Es modulus of elasticity of steel

to crushing and, eventually, disintegration and total col- investigate the effects of FRP laminates on altering the
lapse. This behavior constitutes a major source of haz- failure modes and strength and deformation characteris-
ard during seismic events and creates a major seismic tics of different assemblages. Another objective is to
performance problem facing designers today. demonstrate the potential of the FRP on enhancing
Seismic upgrading by adding new structural frames the shear and compressive strength of URM infill walls
or shear walls, have been proven to be impractical, they and preventing brittle collapse by means of stabilizing
have been either too costly or restricted in use to certain the face shell even after excessive damage. This would
types of structures. Other strengthening methods such as also maintain the wallÕs structural integrity and would
grout injection, insertion of reinforcing steel, prestress- reduce the possibility of URM walls collapsing and
ing, jacketing and different surface treatments were sum- spalling, which, in itself, is a major source of hazard dur-
marized elsewhere [2] and specified by the Federal ing earthquakes, even if the whole structure remained
Emergency Management Agency documents [3,4]. Each safe and functioning.
of these methods adds considerable mass and stiffness
leading to higher seismic loads. They also involve the
use of skilled labor and disrupt the normal function of 2. Behavior of infill masonry walls
the building. The use of fiber reinforced polymer
(FRP) laminates for retrofitting and strengthening is a Masonry infill walls in frame structures have been
valid alternative because of their small thickness, high long known to affect the strength and stiffness of the in-
strength-to-weight ratio, high stiffness, and relative ease filled-frame structures. In seismic areas, ignoring the
of application. frame–wall interaction is not always on the safe side,
A strong earthquake introduces severe in-plane and since, under lateral loads, the infill walls dramatically in-
out-of-plane forces to masonry walls which may lead crease the stiffness by acting as a diagonal strut as seen
to catastrophic collapse as seen in Fig. 1 during the in Fig. 2a, thus resulting in possible change in the seis-
1999 Turkey earthquake. However, the majority of mic demand due to significant reduction in the natural
work conducted to date [5–11] has been concentrating period of the composite structural system [1]. Also, the
on the out-of-plane behavior of URM walls strength- composite action of the frame–wall system changes the
ened with externally applied FRPs. Infill panels (or large magnitude and the distribution of straining actions in
portions of wall) may fall out of the surrounding frame the frame members, i.e. critical sections in the infilled-
due to inadequate out-of-plane restraint at the frame–in- frame differ from those in the bare frame, which may
fill interface, or due to out-of-plane flexural or shear fail- lead to unconservative or poorly detailed designs. More-
ure of the infill panel. In undamaged infills, these failures over, these designs may be uneconomical since an
may result from out-of-plane inertial forces, especially important source of structural strength, which is partic-
for infills at higher story levels and with large slender- ularly beneficial in regions of low and, sometimes, mod-
ness ratios. However, it is more likely for out-of-plane erate seismic demand, is wasted. However, URM infill
failure to occur after the masonry units become dis- walls exhibit poor seismic performance under moderate
lodged due to damage from in-plane loading [4]. and high seismic demand. This behavior is due to a ra-
The work presented herein investigates the effects of pid degradation of stiffness, strength and low energy dis-
applying FRP laminates on the in-plane behavior of sipation capacity, resulting from the brittle and sudden
URM assemblages subjected to different stress condi- damage of the URM infill walls.
tions present in masonry infill walls (Fig. 2c). One of The rationale behind neglecting infill walls in the
the objectives of the present experimental study is to design process is partly attributed to incomplete
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 3

Fig. 1. Failure of masonry walls during Turkey earthquake, 1999: (a) out-of-plane failure, (b) in-plane failure and (c) combined in- and out-of-plane
failures.

Fig. 2. Behavior of masonry infill walls.

knowledge of the behavior of quasi-brittle materials applicable to the case of solid infill walls (i.e. with no
such as URM and to the lack of conclusive experimental openings) failing in corner-crushing mode [1].
and analytical results to substantiate a reliable design
procedure for this type of structures. On the other hand,
and because of the large number of interacting parame- 3. Experimental program
ters, if the infill wall is to be considered in the analysis
and design stages, a modeling problem arises because The experimental program consisted of two phases.
of the many possible failure modes (Fig. 2b) that need In Phase I, four different types of assemblages (Fig.
to be evaluated with a high degree of uncertainty. This 2c) were tested under different types of loading condi-
is why it is not surprising that no consensus has emerged tions representing critical regions in an infill masonry
leading to a unified approach for the design of infilled- wall.
frame systems, despite more than five decades of
research. It is, however, generally accepted that, under • Axial compression: in-plane, concentric, compressive
lateral loads, the infill wall acts as a diagonal strut loads were applied at 90 (normal to the bed joints)
connecting the two loaded corners. However, this is only and 0 (parallel to the bed joints).
ARTICLE IN PRESS

4 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

Table 1
FRP composite and dry fibers properties
Composite laminate properties Dry fibers properties
Ultimate tensile strength in primary fibers direction (MPa) 309.0 Tensile strength (GPa) 3.24
Elongation at break (%) 1.6 Tensile modulus (GPa) 72.4
Tensile modulus (GPa) 19.3 Ultimate elongation (%) 4.5
Ultimate tensile strength 90 to primary fibers direction (MPa) 309.0 Density (g/cm3) 2.55
Laminate thickness (mm) 0.25 Weight (g/m2) 295.0

• Diagonal tension: this is a standard testing procedure where AFRP is the cross-sectional area of the FRP lam-
used to evaluate the diagonal tensile (or shear) inate per unit length of wall, Es is YoungÕs modulus of
strength of URM and creates a state of stress similar steel, EFRP is the modulus of elasticity of the FRP lam-
to that occurring in infill walls. inate, and As is the cross-sectional area of the steel rein-
• Joint shear: this enabled evaluating the strengthening forcement per unit length of the wall. The numerical
effect of the FRP laminates against the traditionally procedure required to select the FRP laminate is given
weak and brittle horizontal shear slip failure mode. elsewhere [14].

In Phase II, five single-story/single-bay, one-third


3.2. Material properties
scale, moment-resisting, structural steel frames infilled
with unretrofitted and retrofitted hollow block ma-
The one-third-scale true-model blocks [15] used in
sonry walls were tested under displacement controlled
this investigation were replicas of the standard, full-scale
diagonal loading to evaluate the behavior of the
150 mm wide hollow concrete masonry units [16]. The
composite frame–wall system. Two series were
average net-area-based compressive strength of the
considered:
blocks was 27.87 MPa. The masonry assemblages were
constructed using scaled down mortar joints with a
• Weak frames series: including three identical steel
nominal thickness of 3.2 mm. To simulate actual con-
frames tested as a bare (i.e. no infill) frame and two
struction practice, the mortar mix was designed as Type
frames were infilled with URM, one of which was ret-
S mortar [17] and all mortar joints were tooled to a con-
rofitted with FRP.
cave profile. The selected FRP was a bi-directional 0/
• Strong frames series: including two identical steel
90 Glass-FRP with 0.295 kg/m2 of E-glass fibers. The
frames tested as both were infilled with URM and
properties of the GFRP composites, given in Table 1,
one of which was retrofitted with FRP.
were determined according to ASTM D-3039 specifica-
tion [18] and were supplied by the manufacturer. How-
ever, an average strength of 260 MPa (84% of the
3.1. FRP selection
specified strength in the GFRP data sheet) with 8.0%
C.O.V. was determined by testing five specimens accord-
In order to select an FRP laminate for the URM
ing to the ASTM D-3039 [18]. The steel used for the steel
assemblages, an equivalent-stiffness-based approach
frame sections was of A 36 grade (yield strength
[12] was employed. The laminate required was equated
243 MPa) for the Weak, W-Series, frames, and A572-
to the minimum steel reinforcement ratio of 0.2% (based
50 (yield strength 379 MPa) for the Strong, S-Series,
on the gross cross-sectional area of the wall) according
of frames.
to the requirement of the Masonry Standards Joint
Committee [13]. This minimum steel ratio is required
in high seismic zones to be distributed between the ver-
tical and horizontal directions in masonry walls. In 4. Phase I testing: assemblages
other words, the required thickness of FRP laminate
was determined based on the premise that, the stiffness 4.1. Test setup and instrumentation
of the FRP laminate must be at least equal to or greater
than the axial stiffness of the reinforcement in the walls. The test specimens were chosen to represent typical
The required thickness of the FRP laminate was there- loading cases of masonry infill walls as shown in Fig.
fore calculated by direct scaling of the reinforcement 2c. A total of 24 1/3-scale specimens were constructed
area by the ratio of the elastic moduli of the steel and in the laboratory and tested to failure under displace-
FRP material as follows, ment controlled loading. The overall displacement was
measured using Linear Variable Differential Transduc-
Es ers (LVDTs) connected to a PC data acquisition system,
AFRP ¼ As ð1Þ
EFRP which also recorded the applied load on the specimens.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 5

Table 2
Phase I test results
Specimen Results
Test Strength
number
Individual Average C.O.V.
(MPa) (MPa) (%)
90U 1 8.45 7.55 18.5
2 8.25
3 5.94
90R 1 13.10
2 11.75 12.17 6.6
3 11.67
00U 1 5.56
2 8.12 7.22 19.9
3 7.97
Fig. 3. Assemblages test setup and LVDTs configurations: (a) joint 00R 1 11.49
shear and (b) diagonal tension. 2 11.27 11.76 5.6
3 12.51
Typical LVDTs configurations are shown in Fig. 3 for DTU 1 0.85 0.88 12.5
the diagonal tension and the joint shear specimens. 2 0.79
3 1.00
4.2. Preparation of test specimens DTR 1 3.73 4.01 6.3
2 4.23
3 4.07
The two axial compression assemblages were of sim-
ilar dimensions in order to permit direct comparison of JSU 1 0.97 0.82 18.9
their failure loads. Since it was not feasible to cut the 0 2 0.83
3 0.66
assemblages from a built URM walls, the individual
blocks for each assemblage were initially cut to shape JSR 1 6.28 6.53 6.4
2 7.02
using a diamond saw. The head mortar joint between
3 6.30
the two middle blocks in the joint shear specimens was
Examples: DTU2 is the second, unretrofitted assemblage tested under
left unfilled to allow for the specimen to fail in shear.
diagonal tension; 90R3 is the third, retrofitted assemblage tested under
All specimens were constructed with face shell mortar axial compression normal to the bed joint.
bedding. After air curing for at least 28 days, half of
the constructed specimens were retrofitted using two lay-
ers of FRP laminates, one on each surface of the (JS) assemblages. The third character is assigned one
specimens. of two letters, either ‘‘U’’ or ‘‘R,’’ indicating whether
Before applying the FRP laminate, specimen surfaces the assemblage was Unretrofitted or Retrofitted,
were first cleaned from mortar protrusions and dust respectively.
using a wire brush and air blasting, respectively. The
epoxy mixture was then applied using a paint roller to 5.1. Axial compression
both surfaces of the specimen. The pre-cut fabrics were
then placed on the wet surfaces and more epoxy was ap- 5.1.1. Failure modes
plied to insure complete fabric saturation. The assem- The unretrofitted axial compression assemblages 90U
blages were tested in accordance with the ASTM E (see Fig. 4a) and 00U (see Fig. 4b) assemblages exhibited
477-92 [19] and ASTM E 519-81 [20] specifications. typical compression failure modes characterized by ver-
tical splitting along the webs of the two middle blocks
[21]. The splitting cracks left the two face shells to
5. Phase I test results deform individually, as shown in Fig. 4c, with a high
slenderness ratio. Finally, the specimens totally disinte-
The test results are summarized in Table 2, and dis- grated as a result of the out-of-plane buckling and/or
cussed in the following sections in terms of failure spalling of the face shells or a combination of both.
modes, strengths, and deformation characteristics. Each Noticeably, all failure modes were brittle and the assem-
specimen series was assigned a name according to the blages disintegrated almost immediately after reaching
notation shown with the examples at the bottom of Ta- their respective maximum loads.
ble 2. The first two characters refer to the axial compres- In contrast, all the retrofitted assemblages exhibited
sion (90 or 00), diagonal tension (DT), or joint shear one failure mode initiated by vertical splitting of the
ARTICLE IN PRESS

6 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

Fig. 4. Failure modes of the unretrofitted axial compression assemblages: (a) Series 90U, (b) Series 00U and (c) web splitting mechanism of face shell
mortar bedded masonry [21].

In general, the retrofitted assemblages did not lose


all their strength nor disintegrated upon reaching the
maximum strength. In fact, in the majority of the
tests, a plateau region was attained during which the
compressive stress almost stabilized and began to
gradually decrease with increased displacement. Such
plateau can be regarded as residual strength after fail-
ure of the assemblages, an absent feature in the case
of URM.

Fig. 5. Failure modes of the retrofitted axial compression assemblages:


(a) Series 90R and (b) Series 00R.

interior webs followed by a gradual increase in the load


Stress (MPa)

up to the peak load. After reaching the peak load, a sud-


den bang was heard as a result of the blocks webs com-
pletely breaking off the face shells. The specimens
continued to carry more load under the displacement
controlled loading with a gradual decrease in capacity.
All the retrofitted assemblages were reduced to two in-
tact face shells with all interior webs damaged (see Strain
Fig. 5). (a)

00R3
5.1.2. Strength characteristics 12.0 00R1
00R2
Table 2 gives the variation of the compressive
10.0
Stress (MPa)

strengths of the unretrofitted versus retrofitted assem-


8.0
blages. To facilitate comparison, the compressive 00U2
strengths of the assemblages were calculated as the max- 6.0 00U3

imum load-carrying capacity divided by the gross 4.0


assemblage area perpendicular to the direction of the 2.0 00U1
applied load (6431.0 mm2). It is clear that the values 0.0
of the coefficients of variation for the retrofitted assem- 0.0000 0.0005 0.0010 0.0015 0.0020 0.0025 0.0030 0.0035
blages are generally lower than those of the unretrofitted Strain
(b)
specimens (see Table 2). This demonstrates the lami-
nateÕs role in reducing the inevitable variations in Fig. 6. Stress–strain relationships for the axial compression assem-
URM construction. blages: (a) Series 90U and 90R, and (b) Series 00U and 00R.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 7

5.1.3. Deformation characteristics disintegrated. In two out of the three assemblages


The measured displacement and applied load were (DTU1 and DTU2) the shear slip occurred along the
used to generate the stress–strain relationships shown middle bed joints (see Fig. 7a), while the third specimen
in Fig. 6. In general, a good agreement can be observed (DTU3) failed along the first bed joint (see Fig. 7b). No
between the initial slopes for the prisms tested with signs of cracking or distress were observed prior to fail-
h = 0 and 90 in both the retrofitted and the unretrofit- ure by shear slip.
ted series. For the retrofitting assemblages, using the FRP lam-
inates effectively prevented any tension or shear failure
5.2. Diagonal tension modes. All three retrofitted assemblages failed by local
crushing of their corners contained within the steel load-
5.2.1. Failure modes ing shoes. Fig. 7c illustrates the local crushing failures.
All the unretrofitted diagonal tension assemblages Due to the compressive stress buildup at the toes, verti-
exhibited shear slip failure along the mortar bed joints. cal cracking through the webs was observed and ex-
This is attributed to the mortar joint weak bond strength tended into the first two courses as shown in Fig. 7d.
compared to the tensile strength of the concrete blocks. Other than the local crushing and the minor delamina-
The observed shear slip failure mode was highly brittle tion at the vicinity of the loading shoes, no other
and, as soon as shear slip along a bed joint was initiated, signs of distress or cracking were observed in the
the assemblages split into two parts and subsequently assemblages.

Fig. 7. Failure modes of the diagonal tension assemblages: (a) specimen DTU2, (b) specimen DTU3, (c) Series DTR corner crushing and (d) Series
DTR web splitting.
ARTICLE IN PRESS

8 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

5.2.2. Strength characteristics DTR2 and DTR3 detached due to local delamination
The maximum stresses sustained by the assemblages and crushing at the vicinity of the bottom loading shoe.
are given in Table 2. In accordance with ASTM E The same occurred in assemblage DTR1 towards the
519-81 [20] standard specification, the diagonal tensile end of the test, yet its shear stress versus strain curve
or shear strength is calculated from, showed a load plateau with a slight increase in load.
0:707P In fact, all the assemblages were able to sustain residual
Ss ¼ ð2Þ loads under increasing displacement (even though the
An
resulting load plateaus could not be plotted for DTR2
where Ss is the shear stress based on the net area, P is the and DTR3).
applied load, and An is net area of the specimen calcu-
lated as follows: 5.3. Joint shear
 
wþh
An ¼ tn ð3Þ 5.3.1. Failure modes
2
The unretrofitted specimen failed in a brittle shear
where w, h and t are the specimen width, height and slip debonding mode at very low load and displacement
thickness respectively, and n is the percentage of solid levels. This is a result of the weak mortar joint bond
in the unit, expressed as a decimal. strength and the absence of friction resistance due to
The average strength of the retrofitted diagonal ten- the lack of compressive stresses normal to the mortar
sion specimens was 4.58 times that of the unretrofitted bed joints. The failure was in the form of complete sep-
ones and, as expected, the relatively low coefficient of aration in the top and/or bottom mortar joints vicinity
variation is indicative of the role of the laminates in (Fig. 9a). This failure mode is highly brittle and oc-
reducing the anisotropy and variability of URM. curred without much time elapse between the crack ini-
tiation at the block–mortar interface and the
5.2.3. Deformation characteristics consequential debonding of the blocks.
In accordance with ASTM E 519-81 [20] specifica- In a manner similar to the retrofitted axial compres-
tion, the shear strain was calculated using the vertical sion assemblages, none of the retrofitted assemblages
shortening along the compression diagonal and the hor- failed by shear slip along the block–mortar interface.
izontal extension along the tension diagonal as follows: At most, minor signs of delamination and stretching
DV þ DH of the laminate were observed along a portion of the
c¼ ð4Þ bed joints. All assemblages ultimately failed after one
g
of the middle blocks cracked through the webs and split
where c is the shear strain, DV is the vertical shortening, open, though remaining attached to the assemblage as
DH is the horizontal extension, and g is vertical gage. shown in Fig. 9b. The splitting of the middle blocks is
Fig. 8 illustrates the shear stress versus shear strain attributed to the induced lateral tensile stresses devel-
relationship for the diagonal tension assemblages. Due oped in the laminates, which resisted the closing of the
to the sudden brittle failure mode, obtaining post-peak head joint gap between the top and bottom blocks.
behavior for the unretrofitted assemblages was not The tensile stresses induced in the middle blockÕs face
feasible. shells eventually resulted in cracking through the webs.
For the retrofitted specimens, the LVDTs installed at This failure might also be attributed to the fact that,
the center of the assemblages at the 102 mm gage length, with the presence of the laminates and their ability to
did not record any appreciable deformations in any of transfer shear stresses to the middle blocks, the blocks
the tests. Furthermore, as the load approached its peak, were subjected to a state of stress similar to that occur-
the long compression LVDT brackets on assemblages ring in the specimens under compression parallel to the
bed joint, and thus the observed failure mode was devel-
oped. Nevertheless, this failure mode implies that shear
4.5 DTR2 failure can be eliminated and the wall strength would be
4.0 DTR3
governed by the compressive strength of the composite
Shear Stress (MPa)

3.5
3.0 prisms.
2.5
2.0 DTR1
1.5 5.3.2. Strength characteristics
DTU3
1.0 The strengths of the three unretrofitted joint shear
0.5 DTU2
0.0
DTU1 assemblages are presented in Table 2. In order to deter-
0.0000 0.0005 0.0010 0.0015 0.0020 0.0025 0.0030 0.0035 mine the joint shear strengths shown in the table, the
Shear Strain
failure load for each assemblage was divided by the
Fig. 8. The shear stress versus shear strain relationship for the respective net-mortared area. Since face shell mortar
diagonal tension assemblages. bedding was employed, the actual lengths of the mor-
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 9

Fig. 9. Failure modes of the joint shear assemblages: (a) Series JSU and (b) Series JSR.

tared bed joints less the head joint gap were measured 9.0
using a caliper and multiplied by the average minimum 8.0
bottom face shell thickness of 8.7 mm to determine the
Shear Stress (MPa)

7.0

net joint shear area. 6.0

The FRP laminates were cut precisely and adhered to 5.0 1.0 JSU1
Shear Stress (MPa)

4.0 0.8
cover only the lengths of the mortared bed joints ensur- 0.6
3.0
ing that the head joint gap between the middle blocks 0.4
JSU2
JSU3

2.0
was not obstructed. Therefore, similar to the unretrofit- 1.0
0.2
Average Slip (mm)
0.0
ted specimens, the shear area of the retrofitted assem- 0.0
0.000 0.005 0.010 0.015 0.020

blages is the net-mortared area of the bed joints 0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70

(excluding the head joint gap area). The increase in the Average Slip (mm)
joint shear strength for the retrofitted assemblages was Fig. 10. The shear stress versus slip relationship for the joint shear
8.2 times that of the unretrofitted ones. assemblages.

5.3.3. Deformation characteristics


Fig. 10 illustrates the shear stress versus shear slip ted assemblages exhibited dramatic load drop after
behavior of the joint shear assemblages. The unretrofit- reaching their respective maximum stress at an average
ARTICLE IN PRESS

10 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

slip of 0.011 mm, with a coefficient of variation of 6. Phase II testing: infilled frames
29.0%.
In general, as soon as the retrofitted assemblages This phase focused on testing of one-third scale, mo-
reached their respective maximum loads, a sudden load ment-resisting, structural steel frames infilled with unret-
drop occurred which signified the web splitting of one of rofitted and retrofitted hollow block masonry walls.
the middle blocks as discussed above. This was followed Single-story/single-bay infilled frame subjected to diago-
by a plateau with a gradually descending strength; the nal-compressive loading (as shown in Fig. 11a) were
assemblage at this stage was still able to carry more load used to evaluate the behavior of the composite frame–
than the maximum load reached by the unretrofitted wall system.
specimens. Examination of the failed assemblages re- The interaction between the infill wall and the sur-
vealed that the mortar bond between the top/bottom rounding columns and beams result in unequal contact
and middle blocks was damaged and that the laminates lengths along the boundaries of the infill with each of
were entirely transferring the vertical applied load from the weak and strong frame members. This, in turn, re-
the top to the bottom block through the middle blocks. sults in different infill contribution with different frames.
In all the retrofitted assemblages, the laminate was not Table 3 lists the structural properties of the weak and
entirely torn and the assemblage could have resisted fur- strong frame sections [22] and outlines the five frames
ther loads. tested in Phase II.
The average slip at the maximum shear stress was One-third scaling of the typical clear floor height and
0.379 mm within a coefficient of variation of 17.8%. This column span of the prototype structure was used to ob-
is in excess of 34 times that of the unretrofitted joint tain the dimensions of the model infilled frame [15]. A
shear assemblages, thus indicative of the significant clear height between the beams and columns of
deformation capability gained by using the FRP 1100 mm was used. This is equivalent to a full-scale
laminate. dimension of 3.3 m. The masonry infill walls were built

Actuator and Reaction Frame Bracing connected to Top Girder


Reaction Column
30.04"

W14x90

AMSLER
Loading Jack

Applied Load

Adjustable
Stilt
COMPRESSION
ing

DIAGONAL
Brac

Spherically-seated
=
l

Actuator Head
ame

=
th MC
ng
n Fr

le
if ll "
ctio

n 0
=I 0
Rea

Load Cell ht .5 LC
eig 43
Top Loading Shoe
lh
fil
n
um

Be

In
l/2

am
ol

SC
C
Be
15
6x

am

Out-of-plane bracing
DIAGONAL

DIAGONAL
W

TENSION

TENSION
W
n

plates held via C-clamps


6x
um

15

to the support angles


ol
C

MT LT ST ST LT MT
L6x6x 3/8 Angle

0"
00 on
gh

6. cti =
hi

n
s

Be

SC
se

e
S th um
ur

am

= ep ol
co

C
16

D
"
00

KEY
6.

78.49"
1.99 m LC MC = Main Compression LVDT
MT = Main Tension LVDT
LC = Long Compression LVDT
L6x6x 3/8 Angle MC LT = Long Tension LVDT
SC = Short Compression LVDT
8

15
bl

ST = Short Tension LVDT


Be

oc

6x
am

ks

COMPRESSION = Strain gages (adhered to


w
W

n
id

um

the top of the upper and


6x

DIAGONAL
ol
15

lower flanges)
T ST SETUP FOR
CM & CR FRAMES
Reaction “Strong” Floor Drawing Scale 1:20
LOCATIONS OF STRAIN GAGES AND LVDTs
(a) (b)

Fig. 11. Masonry infilled steel frame specimens: (a) test setup, (b) instrumentation.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 11

Table 3 applied using an AMSLER hydraulic jack with a load


Structural properties and test matrix of the steel frames capacity of 490 kN and a maximum stroke of 125 mm.
Structural property Weak (W) Strong (S) A lateral bracing system consisting of four L 6 · 6 · 3/
frame S3 · 5.7 frame W6 · 15 8 angles were used to prevent accidental out-of-plane
Area 1077 mm2 2858 mm2 deformations during the in-plane loading of the frames.
Depth 76 mm 152 mm Fig. 11b illustrates the typical instrumentation in-
Web thickness 4 mm 6 mm
Flange width 59 mm 152 mm
stalled on the infill wall and bounding frame, as well
Flange thickness 7 mm 7 mm as the locations of the critical sections along the steel
frame where strain gages were placed (an infilled
Strong axis (X-axis)
Elastic moment of inertia 1,049,000 mm4 12,112,000 mm4 W6 · 15 steel frame is shown in the figure for illustra-
Elastic section modulus 27,500 mm3 159,300 mm3 tion). All LVDTs, strain gages, and the load cell used
Plastic section modulus 31,950 mm3 176,980 mm3 to measure the applied compressive load were all con-
Weak axis (Y-axis) nected to a PC data acquisition system.
Elastic moment of inertia 189,400 mm4 3,879,300 mm4
Elastic section modulus 6390 mm3 50,960 mm3
Plastic section modulus 10,700 mm3 77,840 mm3 7. Phase II test results
Infill type Specimen
Bare WB – 7.1. WB frame
Unretrofitted WU SU
Retrofitted WR SR As expected, the frame joints underwent severe rota-
Steel frame section: ‘‘W’’ weak frame S3 · 5.7 or ‘‘S’’ strong frame tion and distortion. In addition, both joints along the
W6 · 15. tension diagonal experienced tearing of the webs as
Infill type: ‘‘B’’ bare (no infill), ‘‘U’’ unretrofitted-masonry infill, or
‘‘R’’ retrofitted-masonry infill.
shown in Fig. 12, although no signs of cracking were
noted in the welds between the columns and the beams
eight blocks wide by 16 courses high. To ensure symme- at these moment-resisting connections (thus indicative
try in the construction, mortar was packed along all the of the strength of the weld). It was evident from the dis-
boundaries between the infill and the confining steel torted shape of the frame and the bent columns and
frame with a nominal mortar joint thickness of 3 mm. beams, that the permanent (i.e., plastic) deformations
To identify the different frames tested in Phase II, propagated from the joints inwards.
each specimen was assigned a name according to the The initial stiffness obtained from the load–deflection
notation in the bottom of Table 3. The first character curve shown in Fig. 12 was determined to be 2.2 kN/mm.
is used to identify the steel section of the bounding A linear behavior was observed up to a load of a 25.8 kN
frame whether S3 · 5.7 (Weak frame) or W6 · 15 corresponding to an average top displacement of
(Strong frame). The second character describes the type 11.9 mm. Subsequently, a load plateau of 27.0 kN was
of infill, if any, ‘‘B’’ refers to no wall (Bare frame), ‘‘U’’ attained at a top displacement of 14.0 mm. Attributed
or ‘‘R,’’ indicating whether the wall was Unretrofitted or to the strain hardening effects, a slight increase in load
Retrofitted, respectively. occurred resulting in an ultimate load of 28.3 kN. The
For the stronger S-frames, the clear height between actual load–deflection behavior closely resembles the ex-
the beams and the clear span between the columns were pected behavior in which an initial linear response occurs
similar to those of the W-frames. Moreover, the beam- until initiation of yielding followed by a plateau then a
column connections were also designed as full-mo- slight load increase to attain the ultimate failure load.
ment-resisting and fabricated using complete-joint-pene-
tration groove welds to weld the beam flanges to the 7.2. WU frame
column flanges while 3.0 mm fillet welds were used to
weld the beam webs to the column flanges. However, The maximum load-carrying capacity attained by the
due to the expected high diagonal-compressive loading infilled frame was 104.0 kN. A stepped diagonal crack
force, 10 mm thick stiffener plates were welded using
3.0 mm fillet welds between the column flanges in order
to prevent premature web buckling at the loaded ends.
To maintain symmetry, similar plates were also welded
at the four corners of the S-frames.

6.1. Test setup and instrumentation

The general test setup and loading assembly are


shown in Fig. 11a. The compressive top load was Fig. 12. Load–deflection relationship for frame WB.
ARTICLE IN PRESS

12 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

100
Applied Load blocks and mortar joints. As the infill panel further
80
WU readjusted indicating redistribution of the transferred
Load (kN)

60 load, the load steadily rose to reach the maximum load


40 WB carried by this infilled frame in spite of the occurrence of
20 local cracking due to crushing of the infill in the vicinity
Applied Load
0 of the bottom loading shoe at approximately 7.6 mm. At
0.0 20.0 40.0 60.0 80.0
Deflection (mm)
12.7 mm, crushing and cracking was observed in the in-
fill panel near the top loading shoe. Subsequently, a
Fig. 13. Load–deflection relationship for frames WU and WB. steady decline in the load-carrying capacity of the frame
occurred as more off-diagonal cracks started appearing
was observed at the center of the infill panel along the on both sides of the first toothed crack in addition to
compression diagonal as shown in Fig. 13, was an- spalling of the block face shells near these cracks. At
nounced by an audible bang and occurred at a load of approximately 15.0 mm, cracks along the infill bed
7.8 kN corresponding to 7.5% of the maximum attained joints were observed and propagated at mid-length of
load. The load–deflection relation obtained for frame the frame columns until a significant portion of the infill
WU is shown in Fig. 13. The load versus deflection panel face shell near the mid-length of the left frame col-
curve of the bare frame tested earlier is reproduced on umn separated (approximately at 28.0 mm) as shown in
the same plot for comparison. Fig. 14a. Face shell spalling continued rapidly (at
Initially, the load–deflection curve was characterized 33.0 mm) resulting in severe deterioration and cracking
by a steady rise as shown in Fig. 13. Prior to attaining in the infill panel until, finally (at 56.0 mm), was attained
the first peak at 97.7 kN at a corresponding deflection signifying the beginning of a load plateau. In turn, this
of 1.9 mm and shortly afterwards, cracking noises were marks the point after which the infill panel ceased to
continuously heard although no visible cracks were ob- contribute in resisting the applied load. Extensive local
served, thus indicating possible damage in the interior crushing and cracking at the toe of the infill panel near
webs of the masonry infill panel resulting in a series of the bottom loading shoe is shown in Fig. 14b. The sever-
quick load descents and ascents. Shortly afterwards, a ity of the damage at the center of the masonry infill is
major off-diagonal crack parallel to the initial toothed illustrated in Fig. 14c. From 56.0 mm onwards, the steel
crack at the center of the infill panel was observed. This frame was entirely carrying the applied load and was de-
crack occurred at a trough in the load–deflection curve formed severely as shown in Fig. 14d. Similar to the WB
at approximately 5.1 mm. Unlike the initial toothed frame, the joints of the steel frame underwent severe
crack, this crack propagated vertically through the plastic rotation at the joints along the tension diagonal,

Fig. 14. Damage of frame WU: (a) at the left column, (b) at the wall toe, (c) at the infill center and (d) at the beam-column joint.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 13

in addition to local web buckling in the columns and load-carrying capacity, some hairline cracks were ob-
beams at the loaded joints. served in the blocks near the vicinity of the top loading
The infilled WU frame attained a maximum load-car- shoe. These cracks were visible underneath the clear
rying capacity of 104.0 kN which represents an increase laminate adhered on the exterior of the masonry infill
of 267.5% compared to the capacity of the bare WB panel. As the load began to decrease, cracking noises
frame. The initial stiffness of the WU frame measured and clicks were heard until suddenly at a load of
as the secant stiffness at 50% of the maximum load-car- 175.1 kN, corresponding to 80.0% of the ultimate load
rying capacity was determined as 55.7 kN/mm which is on the descending branch of the load curve, the interior
25 times that of the bare frame. The significant increase webs near the top portion of the infill panel were dam-
in WU frameÕs load-carrying capacity and initial stiff- aged causing the separated retrofitted face shells near
ness compared to the WB frame was expected, particu- the top loading shoe to snap outwards and moved out-
larly in view of the fact that the infill panel is side the flanges of the frame members (Fig. 16a). Unfor-
relatively stiff compared to the frame. Ultimately, upon tunately, shortly before and after this outward ‘‘burst,’’
attaining a load plateau at a displacement of 57.4 mm the buckled face shell brushed against the main and infill
signifying the end of the infillÕs participation in in-plane compression LVDTs, thus preventing further recording
load resistance, the plastic load-carrying capacity of the of displacement.
WU frame was determined from the test as 28.2 kN. Minor signs of delamination along the second bed
This is comparable to the bare frameÕs ultimate load joint on the backside of the infill panel were evident.
capacity obtained previously from the WB frame test. There were no signs of distress in the steel frame, clearly
indicating that the applied load was primarily endured
7.3. WR frame by the retrofitted infill panel with minimal contribution
from the steel frame.
The load versus deflection curves for the WR frame is A thorough understanding of the behavior and re-
shown in Fig. 15a. The maximum load attained by the sponse of the retrofitted infill panel was further facili-
frame was 218.9 kN which represents increases of 7.7 tated upon its removal from the bounding steel frame
times and 2.2 times the maximum loads attained by thereby enabling a closer inspection. Fig. 16b shows
the WB and WU frames respectively. At an applied load the wall separation in the left side of the frame. The ret-
of 182.4 kN corresponding to 83.3% of the maximum rofit technique using FRP laminates was very successful

Fig. 15. Frame WR: (a) load–deflection relationship, (b) delamination zone and (c) damaged webs region.
ARTICLE IN PRESS

14 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

in preserving the integrity of the highly brittle masonry. tension cracks or shear slip along the bed joints in
The fact that the panel, simulating a story-high wall, was the infill. As shown in the various assemblage tests
removed in one intact piece (in spite of some damage to discussed in Phase I of the experimental program,
the interior webs) is testimony to the beneficial effect of the FRP laminate basically suppressed any tension
retrofit with FRP overlay. At the toes of the infill panel and shear failure modes in the masonry by reinforcing
within the vicinity of the loaded corners of the frame, all the weak mortar joints. Thus, the resulting retrofitted-
interior webs were damaged. As shown in Figs. 15c and masonry infill wall has been transformed into two
16c, the web damage extended inwards towards the cen- very strong face shells connected by masonry webs
ter of the panel to approximately one-quarter of the which are considered as the weak elements in the
diagonal length. Web damage was also evident along assembly.
the perimeter of the panel which was in contact with The secant stiffness at 50% of the maximum load, the
the frame members. The web damage was minimal near initial stiffness of the BR frame is 131.4 kN/mm which
the corners of the tension diagonal. represents increases of 58.7 times and 2.4 times the ini-
However, the separated face shells were held intact tial stiffness values of the WB and WU frames respec-
by the strong laminate and, in general, there was min- tively. The peak load was reached at a compressive
imal (if any) delamination between the overlay and deflection of 5.6 mm. Similar to the WU frame, the re-
the block face shells (except in the location shown corded deflections along the compression diagonal in
in Fig. 15b). No web damage was evident at the cen- the direction of the applied load were greater than those
ter of the infill. Furthermore, unlike in the unretrofit- along the tension diagonal. This is attributed to the local
ted-masonry infill wall of the WU frame, the laminate cracking at the infillÕs loaded toes resulting in a reduced
successfully prevented the occurrence of any diagonal stiffness along the loading direction.

Fig. 16. Damage of frame WR: (a) out-of-plane wall burst, (b) infill wall separation and (c) extent of web splitting.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 15

Ultimately, as the infill panel was no longer in any within the infill panel as it adjusted to bear against the
effective contact with the bounding steel frame, a load deforming shape of the steel frame. Suddenly, face shell
plateau was attained which represents the bare frameÕs spalling occurred in the tension corner regions (to the
plastic load capacity. The load stabilized at a value of left and right of the central crack where extensive off-
28.9 kN which is comparable to the WB frame capacity. diagonal cracks were occurring) as shown in Fig. 18a.
This occurred at an approximate SU deflection of
7.4. SU frame 9.0 mm. In fact, this served as a major indicator of the
shift in load resistance between the infill panel and the
Plot of the applied top load versus deflections for the steel frame. As the frame was still being pushed diago-
SU frame is shown in Fig. 17. The initial secant stiffness nally, due to incompatible deformations between the
of the SU frame was 91.4 kN/mm. The ultimate load- steel frame and the masonry wall, the infill wall was
carrying capacity of the SU frame determined in the sec- quickly losing structural integrity accompanied with ra-
ond test was 284.4 kN. The unretrofitted-masonry infill pid face shell spalling and collapse of massive ‘‘chunks’’
panel remained crack-free up to an applied diagonal of the upper region of the wall (Fig. 18b).
load of 122.4 kN corresponding to 43.0% of the ultimate At the end, only the lower three courses of the ma-
load-carrying capacity of the SU frame. Thereafter, a sonry wall remained standing on the lower beam and
longitudinal crack at the middle of the panel occurred column of the frame as shown in Fig. 18c. As the load
similar to the WU frame. However, unlike the toothed increased, the SU frame was simply behaving as a bare
crack in the WU frame, which propagated along the W6 · 15 moment-resisting frame and yielding com-
head and bed joints around the masonry units, the mid- menced at the beam-column joints.
dle crack in the SU frameÕs panel extended through both The formation of plastic hinges, eventually leading to
the units and the mortar joints. As loading progressed, a plastic collapse mechanism, characterizes the failure
the middle crack extended further in addition to the for- mode of the CM frame in which the masonry infill panel
mation of some off-diagonal hairline cracks and a short did not increase its load-carrying capacity. However,
bed joint crack above the first masonry course. Near the failure of the masonry infill panel is attributed to a com-
peak load, signs of crushing of the boundary mortar bination of toe crushing (characterized by local com-
joint between the steel frame and the infill panel in the pressive crushing of the masonry at the vicinity of the
vicinity of the loaded corners were observed. Moreover, loaded corners) and diagonal-compression failure at
a hairline separation crack between the panel and the the center of the panel (characterized by the formation
frame at the tension corners was observed to extend of extensive diagonal and off-diagonal longitudinal
approximately three courses long. cracks). The frame was unloaded after the full stroke
The cracking pattern of the masonry infill wall resem- of the loading actuator was consumed.
bles that encountered in the WU frame test in which a
central crack is first initiated along the loaded diagonal 7.5. SR frame
of the wall followed by the formation of some off-diag-
onal cracks. In the second test attempt, the existing but Fig. 19a shows the load–deflection relationship of the
closed hairline cracks resulting from the first test wid- SR frame which had a maximum load capacity of
ened as the frame reached a first peak load at 343.0 kN and an initial stiffness of 262.7 kN/mm.
266.9 kN at a corresponding deflection of 6.8 mm. At 0.8 mm, an audible bang was heard although no
Shortly before reaching the peak load, minor cracking crack was detected visually. Characterized by a small
was observed in the infillÕs toe near the bottom loading shift in the load–deflection curve, the bang suggested
shoe. As the frame was pushed further in spite of the de- the occurrence of a crack in the interior webs, possibly
creased load resistance, small off-diagonal cracks began at one of the loaded toes of the infill wall. At 2.3 mm,
forming on the left and right sides of the central crack. the crack location most probably occurred in the top
These cracks assisted in the redistribution of the load toe. At a deflection of 1.98 mm, the load increased fur-
ther until it reached the first peak of 270.1 kN. At this
point, a greater load bang was heard and crushing at
300.0 the top loaded toe of the infill panel was observed. As
250.0 in the retrofitted assemblages, crushing at the top toe
Load (kN)

200.0 of the retrofitted-masonry infill panel involved damage


150.0
of the interior webs leading to the laminated-face shells
100.0
50.0 snapping outwards. The damage extended along the in-
0.0 fill-frame boundary for a length of approximately six
0.00 20.00 40.00 60.00 80.00
courses (three block-lengths) and only one course wide
Deflection (mm)
(one half a block-length) as shown in Fig. 19b. A conse-
Fig. 17. Frame SU load–deflection relationship and diagonal cracking. quential loss in load capacity occurred but gradually in-
ARTICLE IN PRESS

16 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

Fig. 18. Damage of frame SU: (a) face shell spalling in the diagonally cracked region, (b) collapse of the upper infill region and (c) infill wall
remnants.

Fig. 19. Frame SR: (a) load–deflection relationship, (b) damage progress and (c) final damaged zones.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 17

creased as the frame was further loaded. Similar crush- that, other than minor delamination at the infill-frame
ing also occurred at the bottom loaded toe of the infill boundaries and toe crushing and in spite of the separa-
panel (Fig. 19b). tion between the infill and the frame at the tension diag-
Separation between the infill panel and the steel onal corners, the central region of the wall was intact
frame at the tension diagonal corners occurred as load- without any cracking or damage. A schematic diagram
ing progressed. At approximately 22.0 mm, the separa- illustrating the state of the retrofitted infill wall at the
tion gap was clearly visible as shown in Fig. 20a. As end of the test is shown in Fig. 19c.
the frame was pushed further beyond 22.0 mm, the sep- Even though the load–deflection curves indicated that
aration between the frame and the infill increased. The a load plateau was reached, it was decided to further
extent of toe crushing, which is defined as splitting of load the SR frame. This decision was triggered by the
the face shell and at times accompanied by minor delam- fact that the plateau occurred at load of 339.0 kN which
ination between the overlay and the face shell itself, also is 21.0% greater than the expected plastic load-carrying
increased (Fig. 20b and c). The steel frame was consid- capacity of the bare W6 · 15 frame of 280.4 kN as deter-
erably deformed with significant plastic rotation at the mined experimentally from the prior SU frame test, thus
tension joints as shown in Fig. 20d. Examination of suggesting that the retrofitted infill wall still contributed
the infill panel at the end of the test (Fig. 20e) indicated to load resistance.

Fig. 20. Frame SR damage: (a) at left side, (b) at top loaded corner, (c) at bottom loaded corner, (d) at beam-column joint and (e) final damaged
configuration.
ARTICLE IN PRESS

18 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

8. Summary of Phase II test results point at 50% of the ultimate load using the applied diag-
onal load versus in-line compressive displacement
Beside local toe crushing, secondary signs of distress curves. Although a bare W6 · 15 steel frame was not
resulting from severe face shell splitting such as tearing tested as a separate frame, its stiffness which is shown
of the laminate or minor delamination between the in Table 4 was experimentally determined from the sec-
block face shells and the laminates were the only ob- ond test of the SR frame whose infill panel was sepa-
served damages as the FRP retrofitted infilled-frames rated from the frame along the majority of its
were pushed to severely deformed configuration. The perimeter and sustained local damage at its loaded toes
frames with the retrofitted infill walls depicted similar whereas the steel frame did not experience any distress in
behavior throughout the entire loading history. In both the prior test. For each frame within the two main steel
the WR and SR tests, as soon as local crushing occurred frame types, the increases in stiffness compared to the
at the wallÕs corners, clearly visible and wide separation bare frame and the first test of the unretrofitted-masonry
gaps between the panel and the frame constantly in- infilled frame are computed and presented in the table.
creased unlike in the WU and SU frames. Unlike the
unretrofitted-masonry infill walls which disintegrated
soon after the infilled-frame system reached its peak 9. Conclusions
load, the retrofitted infill walls remained supported with-
in the bounding steel frame until the end of the loading This paper presents an experimental investigation on
and even after attainment of load plateau which signaled the retrofitting of concrete masonry infill walls using
that the frames reached the plastic load-carrying capac- FRP laminates, which provides a strengthening alterna-
ity of the bare steel frame. This behavior demonstrates tive for URM infill walls. The relative ease with which
the superior contribution of FRP laminates in altering FRP laminates can be installed on the walls makes this
the brittle hazardous behavior of URM infill walls to form of strengthening attractive to the owner, consider-
a ductile and damage-tolerant wall with apparent post- ing both reduced installation cost and down time of the
peak capacity and energy dissipation capabilities. occupied structure. Another reason is to comply with
Table 4 summarizes the maximum diagonal-compres- new seismic codes requirements without the need to
sive load sustained by the five tested frames. In addition, demolish the whole wall and rebuild it. The following
the plastic load capacity of the bare W6 · 15 which was conclusions resulted from Phase I of the investigation:
experimentally determined through testing the SU frame
after the remains of the infill panel were removed (the 1. The laminates significantly increased the load-carry-
third test of the SU frame), is also included in the table. ing capacity of the masonry assemblages exhibiting
The percentage and the corresponding multiple in- shear failures along the mortar joints (joint shear
creases in the load-carrying capacity compared to that and diagonal tension). The average joint shear
of the bare frame and the unretrofitted-masonry infilled strength of the retrofitted specimens was equal to
frame for each of the two steel frame types are also cal- eight times that of their unretrofitted counterparts.
culated and presented in Table 4. 2. The unretrofitted axial compression assemblages
Table 4 also compares the initial secant stiffness of the failed suddenly and disintegrated totally upon reach-
five tested frames in this study. The stiffness values were ing peak stress. However, the FRP supplied the ten-
calculated as the slope of line joining the origin and the sile strength required to stabilize the out-of-plane

Table 4
Phase II test results
Frame Maximum % Increase compared to XÕs increase compared to Initial secant XÕs Increase Compared to
load (kN) Bare frame Unretrofitted Bare frame Unretrofitted stiffnessa (kN/mm) Bare frame Unretrofitted-
infilled frame infilled frame masonry
infilled frame
W-Frames WB 28.3 2.2
(S3 · 5.7) WU 104.0 267.5% 3.67 55.7 24.92
WR 219.0 673.9% 110.6% 7.74 2.11 131.4 58.74 2.36
S-Frames SB 280.4b 20.9c
(W6 · 15) SU 284.4 1.4% 1.01 91.4 4.38
SR 343.1 22.4% 20.6% 1.22 1.21 175.1 8.40 1.92
a
Initial secant stiffness determined from the load–deflection curve of the tested frames as the slope of the line joining the origin and the point at
50% of the maximum load-capacity.
b
Plastic load-carrying capacity of the bare W6 · 15 frame was experimentally determined from the third test of the SU frame after the remains of
the infill wall were removed entirely.
c
Initial secant stiffness of the bare W6 · 15 frame was experimentally determined from retesting the SR frame.
ARTICLE IN PRESS

W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx 19

buckling of the individual face shells, thus preventing maintain the wallÕs structural integrity and would
brittle failure after webs splitting and allowing the reduce the possibility of URM walls collapsing and
specimen to carry more loads. This prevented cata- spalling, which, in itself, is a major source of hazard
strophic failure of the masonry–FRP composite during earthquakes, even if the whole structure
assemblages compared to their URM counterparts. remained safe and functioning.
3. The FRP laminates resulted in a gradual prolonged 7. The masonry–FRP composite walls do not fail cata-
failure, and a stable wall with noticeable structurally strophically as their URM counterparts. The FRP
integrity and residual strength even after failure. laminates resulted in a gradual prolonged failure
Thus, the long known hazard problem associated and a stronger wall with more energy dissipation
with URM can be eliminated using the proposed ret- and apparent post-peak strength. This should result
rofit technique. In seismic zones, the prevention of in a higher response modification factor than that
the brittle failure mode is highly desirable since it pro- typically selected for the analysis of URM structures.
vides a means for energy dissipation and conse- 8. By supplying the shear strength at the mortar joints,
quently reduces the seismic forces on the frame FRP laminates can serve as external reinforcement
structure. for unreinforced or under-reinforced masonry walls,
thus providing a quick and cost-effective solution to
The following conclusions resulted from Phase II of conform to the more restrict emerging seismic codes
the investigation: requirements.

4. Retrofitting the infill panel with externally, epoxy-


bonded FRP laminates resulted in an increase in
load-carrying capacity of 2.1 and 1.2 times that of
the corresponding unretrofitted-masonry infilled Acknowledgments
frames for the W-frames and the S-frames
respectively. The work presented herein was supported under
5. Even in the S-frames whose load capacity was not sig- Grant No. CMS-9730646 from the National Science
nificantly increased due to retrofit of the infill panel, Foundation (NSF). The results, opinions, and conclu-
the laminates were able to completely alter the defor- sions expressed in this paper are solely those of the
mation characteristics and behavior of the wall itself. authors and do not necessarily reflect those of the
In the unretrofitted-masonry infilled frames, the walls NSF. The authors would like to gratefully acknowledge
were completely destroyed and the blocks fell out-of- assistance of Edward Fyfe, Peter Milligan and Sarah
plane which in real life poses a hazard to buildingsÕ Cruickshank, Fyfe Co. LLC, California, for providing
occupants. The failure mode of the two unretrofitted the FRP, and John Sabia, D.M. Sabia Co., Pennsylva-
frames was characterized as corner-crushing and nia for providing the mason. The authors would also
diagonal-compression respectively. In the retrofit- like to thank Mr. Omar El-Dakhakhni for his
ted-masonry infilled frames, no signs of diagonal assistance.
cracking were observed and both frame types failed
due to local crushing at the loaded corners. Examina-
tion of the retrofitted panel indicated that the central References
region remained intact and that the majority of the
damage occurred at the outermost perimeter and at [1] El-Dakhakhni WW, Elgaaly M, Hamid AA. Three-Strut model
the loaded corners where the inner webs of the blocks for concrete masonry-infilled steel frames. ASCE J Struct Eng
cracked resulting in the formation of separate lami- 2003;129(2):177–85.
[2] Hamid AA, Mahmoud ADS, El Magd SA. Strengthening and
nated-face shells. repair of unreinforced masonry structures: State-of-the-art. Proc
6. The retrofitting technique maintained the walls struc- 10th Int Brick and Block Masonry Conf, vol. 2. London: Else-
tural integrity and prevented collapse and debris fall- vier Applied Science; 1994. p. 485–97.
out. The FRP laminates contained and localized the [3] FEMA-273. NEHRP guidelines for the seismic rehabilitation of
damage of the URM walls even after ultimate failure buildings, Federal Emergency Management Agency, Washington,
DC, 1997.
and no signs of distress were evident throughout the [4] FEMA-274. NEHRP commentary on the guidelines for the
wall except at the vicinity of the corners and around seismic rehabilitation of buildings, Federal Emergency Manage-
the openings. In contrast to the URM walls, the ment Agency, Washington, DC, 1997.
strengthened walls were stable after failure. In a real [5] Triantafillou TC. Strengthening of masonry structures using
epoxy-bonded FRP laminates. J Compos Construct, ASCE
building, this can reduce the seismic hazard associ-
1998;2(2):96–104.
ated with the wall tipping off or falling out of the [6] Velazquez-Dimas JI, Ehsani MR. Modeling out-of-plane behavior
frame, and eliminate injuries or loss of lives and of URM walls retrofitted with fiber composites. J Compos
properties due to the wall collapse. This would also Construct, ASCE 2000;4(4):172–81.
ARTICLE IN PRESS

20 W.W. El-Dakhakhni et al. / Composite Structures xxx (2005) xxx–xxx

[7] Albert ML, Elwi AE, Cheng JJR. Strengthening of unreinforced [15] Harris HG, Sabnis A, Gajanan M. Structural modeling
masonry walls using FRPs. J Compos Construct, ASCE and experimental techniques. 2nd ed.. New York: CRC Press;
2001;5(2):76–84. 1999.
[8] Hamoush SA, McGinley MW, Mlakar P, Scott D, Murray K. [16] American Society for Testing and Materials, ASTM C 90-92b
Out-of-plane strengthening of masonry walls with reinforced Standard specification for load-bearing concrete masonry units.
composites. J Compos Construct, ASCE 2001;5(3):139–45. Annual Book of ASTM Standards, vol 04.05, West Conshohoc-
[9] Hamilton III HR, Dolan CW. Flexural capacity of glass FRP ken, Pennsylvania.
strengthened concrete masonry walls. J Compos Construct, ASCE [17] American Society for Testing and Materials, ASTM C 270-92a.
2001;5(3):170–8. Standard specification for mortar for unit masonry. Annual Book
[10] Kuzik MD, Elwi AE, Cheng JJR. Cyclic flexure tests of masonry of ASTM Standards, vol 04.05, West Conshohocken,
walls reinforced with glass fiber reinforced polymer sheets. J Pennsylvania.
Compos Construct, ASCE 2003;7(1):20–30. [18] American Society for Testing and Materials, ASTM D 3039/D
[11] Tan KH, Patoary MKH. Strengthening of masonry walls against 3039M-93. Standard test method for tensile properties of polymer
out-of-plane loads using fiber–reinforced polymer reinforcement. matrix composite materials. Annual Book of ASTM Standards,
J Compos Construct, ASCE 2004;8(1):79–87. vol 15.03, West Conshohocken, Pennsylvania.
[12] Seible F. Repair and seismic retest of a full-scale reinforced [19] American Society for Testing and Materials, ASTM E 447-92b.
masonry building. In: Proceedings of the Sixth International Standard test methods for compressive strength of masonry
Conference on Structural Faults and Repair, vol 3, 1995. p. 229– prisms. Annual Book of ASTM Standards, vol 04.05, West
36. Conshohocken, Pennsylvania.
[13] MSJC. Masonry Standards Joint Committee ‘‘Building Code [20] American Society for Testing and Materials, ASTM E 519-81.
Requirements for Masonry Structures (ACI 530-04/ASCE 5-04/ Standard test method for diagonal tension (shear) in masonry
TMS 402-04)’’, reported by the Masonry Standards Joint Com- assemblages. Annual Book of ASTM Standards, vol 04.05, West
mittee, 2005. Conshohocken, Pennsylvania.
[14] Hamid AA, El-Dakhakhni WW, Hakam ZR, Elgaaly M. Behav- [21] Drysdale RG, Hamid A, Baker LR. Masonry structures: behavior
ior of composite unreinforced masonry—fiber–reinforced polymer and design. 2nd ed. Boulder, CO: The Masonry Society; 1999.
wall assemblages under in-plane loading. J Compos Construct, [22] AISC. LRFD manual of steel construction. 3rd ed. Chicago,
ASCE 2004;9(1):73–83. IL: American Institute of Steel Construction; 2003.

You might also like