Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Building Engineering 29 (2020) 101107

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: http://www.elsevier.com/locate/jobe

Experimental study on axial compressive behavior of rubberized


interlocking masonry walls
Amin Al-Fakih a, **, M. M. A Wahab a, Bashar S. Mohammed a, *, M.S. Liew a,
Noor Amila Wan Abdullah Zawawi a, Sholihin As’ad b
a
Department of Civil and Environmental Engineering, Universiti Teknologi PETRONAS, 32610, Bandar Seri Iskandar, Perak, Malaysia
b
Faculty of Engineering, Department of Civil Engineering, Universitas Negeri Sebelas Maret (UNS) Surakarta, Jalan Ir. Sutami 36A, Kentingan, Surakarta, 57126,
Indonesia

A R T I C L E I N F O A B S T R A C T

Keywords: The behavior of rubberized interlocking masonry walls is substantially influenced by the properties of the waste
Crumb rubber materials used, which makes them behave differently from conventional interlocking masonry systems. Ten each
Interlocking brick of hollow and grouted walls were constructed using rubberized interlocking bricks and then tested under
Compressive loading
compressive loading. Rubberized interlocking bricks are made using 10% crumb rubber and 56% fly ash as a
Walls
partial replacement for fine aggregates and cement by volume, respectively. The structural behavior, including
Failure mode
Response surface methodology (RSM) strength, load-deformation performance, stress-strain relations, and failure mechanisms of the tested load-
bearing walls have been investigated. The results reveal the ability of rubberized interlocking masonry walls
to withstand axial compressive of 3.87 MPa for hollow and 5.75 MPa for grouted specimens, which is approx­
imately 15–20% lower than in conventional interlocking masonry walls. Web splitting, vertical cracking and face
spalling were common failure modes for hollow and grouted rubberized interlocking masonry walls. In contrast
to conventional interlocking walls under compressive loading, rubberized interlocking walls show increased
ductility and undergo measurable post-failure loads with significant displacement due to the presence of crumb
rubber, which permits a large expansion of microcracks inside the specimens after failure. Rubberized inter­
locking walls also tend to have large initial deflections as the bricks settle and the gaps caused by dry joints close.

1. Introduction masonry interlocking brick walls (3.56 MPa). This reduction (53%) may
contribute to the effects of the slenderness ratio (<20). In addition, the
Masonry wall, constructed without mortar, is commonly known as a study reported that a combination of shear and compression cracks with
mortarless or interlocking masonry wall system. Recently, the inter­ bed and head joint failure was found to be a failure mechanism. Fundi
locking masonry system has been well known in the construction in­ et al. [1] investigated the mechanical performance of interlocking
dustry as either load or non-load bearing due to its advantages in laterite soil block walls and found that the ultimate compressive strength
increasing filed productivity and building efficiency with potentially less was reduced compared to individual blocks. They reported that the walls
skilled labor and therefore lower costs [1–6]. Many researchers have had a higher vertical deflection in the initial loading before they could
investigated the quality of mortarless masonry subject to compressive take up the compressive load due to the lack of a mortar layer as the
loading in the last few decades [1,2,7–11]. Ahmad et al. [8] tested the interlocking gaps between the blocks had closed. Failure modes of the
compressive strength of the masonry wall made of mortar-free concrete walls have been observed in the form of diagonal cracking and block
interlocking bricks. Studies have shown that the inherent tension of a spalling.
mortar-free wall allows it to be used for residential buildings. Ahmad The mortarless interlocking masonry walls displayed a significant
et al. [9] studied the structural behavior of the masonry walls of soil deformation with ductile behavior compared to the brittle behavior of
cement. The study concluded that the compressive strength of inter­ the mortared masonry framework, and the horizontal displacement was
locking soil bricks was reduced from single unit bricks (7.5 MPa) to virtually negligible, which indicates that no buckling was found in the

* Corresponding author.
** Corresponding author.
E-mail addresses: amin.ali_g03663@utp.edu.my (A. Al-Fakih), bashar.mohammed@utp.edu.my (B.S. Mohammed).

https://doi.org/10.1016/j.jobe.2019.101107
Received 5 September 2019; Received in revised form 2 December 2019; Accepted 2 December 2019
Available online 5 December 2019
2352-7102/© 2019 Elsevier Ltd. All rights reserved.
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

samples tested [12,13]. Marzahn [14] stated that the overall strength of 3. Experimental work
the mortarless interlocking and mortared masonry wall was almost the
same, but the deformation behavior was quite different. The walls built 3.1. Trial mix design
using mortar tended to have an almost linear relationship between stress
and strain from the beginning of the test to failure. However, the Forty-eight mixing designs with variance in crumb rubber content
stress-strain behavior of the mortar-free interlocking was somewhat and fly ash content, including control mix, have been proposed as shown
non-linear in the first part of the curve, extended to approximately one in Table 1. The ratio of cement to aggregate content was set as 1
third of the failure load resulting from the initial settlement of uneven (cement): 2 (fine aggregate). Six specimens of rubberized interlocking
surfaces and block inaccuracies. The second part of the curve was more bricks were prepared, cast and tested for compressive strength for each
dependent on the deformation of the bricks and was linear. mix design. Response data from the experiments conducted were
As reported by several studies [1,15], the interlocking walls analyzed using Design Expert’s Response Surface Method (RSM) to
appeared to fail in the vertical and shear cracks at the mid-height of the calculate the effects of all model terms. In this case, the results of the
wall and the separation of the network joints. The grouted interlocking forty-eight mix designs, as presented in section 3.1.1, were used in the
masonry wall, however, acts in a different way than the non-grouted analysis.
interlocking walls, where the amount of grout used directly affects the
axial strength and failure modes of the walls [16]. As ultimate loads 3.1.1. Compressive strength test for trial mix
were approached, vertical shell cracking developed. This vertical Six rubberized interlocking bricks for each mix (a total of 288) were
cracking was attributed to the lateral tensile stress that develops in the tested at 28 days in accordance with ASTM C140 [34] following the
block as the internal grout expands under compression [17]. Arcady same procedures as set out in section 3.3 of this study. The compressive
et al. [18] studied the behavior of mortar-free interlocking structures strength was calculated by dividing the final load with the bed face
and concluded that stress concentrations on contact surfaces coincide bearing area as prescribed in ASTM C140 [34]. Fig. 2 shows the results
with gaps between blocks and height variations due to block imperfec­ of the compressive strength of the trial mix, where the compressive
tions in production. These concentrations create tensile stress that can strength decreases as the percentage of the replacement of crumb rubber
damage or even break a block. The load-bearing capacity of the increases in both 50 and 60% of the replacement of fly ash.
mortar-free interlocking structure is therefore lower than that of the At 10 vol% of crumb rubber, the average compressive strength of
mortared structure, unless its material is specially reinforced to increase 50% and 60% of fly ash replacement is 10.54 MPa and 20.33 MPa
tensile strength and durability. respectively (Fig. 2). Consequently, the partial replacement of fly ash
The presence of crumb rubber in masonry brick/block samples leads must be greater than 50% and less than 60% while the replacement of
to a reduction in the compressive strength of the specimens due to the crumb rubber must be 10% by volume in order to achieve the minimum
hydrophobic nature of crumb rubber, which engulfs the air around its compressive strength requirement of the load bearing masonry unit
surface and therefore causes a concentration of pressure leading to (13.7 MPa) in accordance with ASTM C90 [35].
failure at low stress [19,20]. Nevertheless, the introduction of fly ash
into brick manufacturing has shown an improvement in strength due to 3.1.2. Analysis of variance
the slower reaction of fly ash during hydration compared to cement, The results of the variance analysis (ANOVA) for the quadratic
which has resulted in increased chemical reactions [21] and decreased compressive strength response model are summarized in Table 2. The F-
hydration heat [22], thus increasing the strength. value of the compressive strength model indicated that the model was
However, the existing behavior of the interlocking masonry wall significant with only 0.01% possibility. The significance of the model
depends not only on the structure as a whole, but ultimately on the and its terms can be checked using the 95% confidence interval
properties of the constituent materials, the brick units and the bonded (P < 0:055). The model and its terms A and A2 were significant, as their
joints. To this extent, few researchers have developed interlocking ma­ P-value was less than 0.05, while the terms B, AB and B2 were insig­
sonry bricks containing different raw and waste materials [1,9,13, nificant. Their P-values were greater than 0.05. This means that the
23–27]. As a result, the action of rubberized interlocking masonry walls terms B, AB and B2 do not have a significant effect on the compressive
is affected by the properties of the waste materials used and the dry bed strength, therefore the compressive strength has a quadratic relationship
joint, which behave differently from traditional interlocking masonry with the crumb rubber.
walls. The structural behavior, including strength capacity, The final equation in terms of actual factors for the compressive
load-deformation behavior, stress-strain relations and failure mecha­ strength of rubberized interlocking brick (RIB) with all the model terms
nisms, of the rubberized interlocking masonry wall (hollow and grouted) is presented in Eq. (1).
subject to compressive loadings has therefore been investigated to pro­
vide a clear understanding of the rubberized interlocking masonry wall. FC ¼ 22:96 1:086A þ 0:025B 0:001A*B þ 0:014A2 (1)

2. Materials

Two types of cementitious materials have been used, namely cement


and fly ash. Class F-Fly ash was used as a partial replacement of cement
by volume. The chemical composition, specific gravity and loss of
ignition of OPC and category F-fly ash were found to comply with the
requirements of ASTM C150 [28] and ASTM C618-17a [29] as stated by
Al-Fakih et al. [30,31].
Local washed river sand with a maximum size of 1,2 mm was used as
a natural fine aggregate in the production of rubberized interlocking
bricks. The fine aggregate was partly replaced by 10% crumb rubber
(waste material collected from scrap tires) by size. The grading of
washed river sand and crumb rubber was carried out in accordance with
the fine aggregate procedures laid down in ASTM C136 [32]. The
gradation of the crumb rubber (Fig. 1) was found to be between the
upper and lower limits specified in in ASTM C33 [33]. Fig. 1. Grading curves of fine aggregate.

2
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Table 1
Trial mixture ratio proportion of RIB (FA denotes Fine Aggregate (sand)).
Mix Cementitious Aggregates Aggregates Aggregates Aggregates Aggregates Aggregates
materials

Cement Fly FA Crumb FA Crumb FA Crumb FA Crumb FA Crumb FA Crumb


Ash Rubber (0%) Rubber (10%) Rubber (20%) Rubber (30%) Rubber (40%) Rubber (50%)

Control 1 0 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1


F10 0.9 0.1 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F20 0.8 0.2 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F30 0.7 0.3 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F40 0.6 0.4 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F50 0.5 0.5 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F60 0.4 0.6 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1
F70 0.3 0.7 2 0 1.8 0.2 1.6 0.4 1.4 0.6 1.2 0.8 1 1

Fig. 2. Compressive strength of RIB trial mix (FA: Fly ash and CR: Crumb rubber).

Table 2 Table 3
ANOVA for Response Surface Quadratic model. Compressive strength model validation.
Source D:F Mean square F-value P-value Significant Response R2 Adj. Pred. μ S. COV, Adeq.
Prob > F R2 R2 D % R2

Model 5 614.51 44.79 <0.0001 Yes Compressive 0.84 0.82 0.77 8.99 3.7 41.15 18.4
A - Crumb rubber 1 2464.12 179.62 <0.0001 Yes strength
B - Fly ash 1 2.92 0.21 0.6470 No
Where R2: degree of correlation, Adj. R2: adjusted correlation, Pred. R2: pre­
AB 1 9.42 0.69 0.4120 No
dicted degree of correlation, μ: mean, S.D: standard deviation, COV: coefficient
A2 1 595.69 43.42 <0.0001 Yes
B2 1 0.4 0.029 0.8657 No of variance, and Adeq. R2: adequate precision.

Where A2 and B2: second order effect of crumb rubber and fly ash, respectively.
significant loss in compressive strength. It was observed that in order to
AB: interaction effects.
D.F: degree of freedom, F-values: Fisher statistical test values, P- values: Prob­ acquire compressive strength of 15 MPa, the crumb rubber amount must
ability values. be around 10% and the fly ash content is to be 56%.

where FC is the compressive strength (MPa), A is crumb rubber (%), and 3.2. Production of rubberized interlocking bricks
B is fly ash (%).
The degree of correlation was used to validate the developed model. As a result of the compressive strength of the trial mix and the RSM
As shown in Table 3, the model has significant correlation value (R2 > analysis, 10% by volume of crumb rubber as partial replacement of sand
0:8) of 0.84, therefore only 16% of the experimental data of the and 56% fly ash as partial replacement of cement by volume was chosen
compressive strength cannot be correlated. Adjusted and predicted de­ for the production of rubberized interlocking brick (RIB) with the
gree of correlation were approximately in good agreement as the dif­ quantity shown in Table 4.
ference between them was less than 0.1. Furthermore, the standard The mix ratio of 1:2 (cement: sand) and a water of 10% of total
deviation (SD) of the model showed that the experimental data fitted the weight were followed to obtain the targeted strength. Due to the hy­
developed model. drophobic nature of crumb rubber (expels water and entrap air on its
Fig. 3 shows the contour and the 3-D response surface plots obtained surface), mix exhibited inhomogeneity which resulted in reduction of
from the simulation results. As it can be seen in Fig. 3, an increase of adhesion between crumb rubber and cement paste. Therefore, firstly,
crumb rubber replacement causes a decrease in compressive strength sand and crumb rubber has been mixed for 5 min to obtain a uniform
where 50% crumb rubber replacing fine aggregates shows the most mix. This has been followed by the addition of cement and fly ash into
the mix and ensuing dry mix for 4 min. In the end, the water has been

3
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Fig. 3. Response surface for compressive strength of RIB.

surface were removed and capped with a 5 mm thick layer of


Table 4
SikaGrout-215 to facilitate the test and allow the applied load to be
Mix proportion and quantity of one unit of RIB.
evenly distributed over the loaded areas. Digital compressive test ma­
Mix. Cement Fly Ash Sand Crumb Rubber Water (%) chine with a capacity of 3000 kN as shown in Fig. 5b. The unit was
Mix proportion (%) 44 56 180 20 10 placed at the center of the test machine and the load was applied and
Mix quantity (g) 931 790 3200 142 507 increased steadily from 14.8 kN/s at a rate of 0.5 N/mm2 per second
until failure. The compressive strength (fb ) of the RIB was then calcu­
lated by dividing the final load achieved (Fmax ) over the load bearing
slowly added to the mix and mix thoroughly for about 5 min. Then, the
area of the brick (A). The loaded area of Fig. 5a is the load bearing area
dry mix was placed into the mold of a semi-automatic machine (Fig. 4)
of the rubberized interlacing brick.
to produce two standard stretcher bricks at a 10 s of time, with a pressing
In addition, a total of 6 grout cylinders (300 � 150 mm) were tested
pressure of 17 MPa. Finally, the bricks have been removed from the
at 28-day age using a compression test machine (Fig. 5c) as specified in
molds and left in room temperature and moist curing for 28 days
ASTM C1019 [36]. The ingredients of the grout mix consisted of ordi­
(Fig. 4c). The identical dimensions of the developed RIB are 250 mm
nary Portland cement and washed river sand with a mix ratio of 1:2.5 by
length, 105 mm height, and 125 mm thickness.
weight. The water to cement ratio was 0.8 by weight to meet the flow
cone test (zero slump test) as reported by Satyarnoa et al. [37].
Wire strain gauges; rosette type, were mounted in the middle of the
3.3. Compressive strength test of RIB and concrete grout
face of the stretcher unit and in the center of the grout cylinder to
The test code for rubberized interlocking bricks under compressive is measure the vertical and lateral strain. In accordance with ASTM E111
not yet available. The code of practice for the concrete masonry unit [38], the stress-strain curves of the stretcher unit and the cylinder
ASTM C140 [34] was therefore used as a guideline for the determination grout were plotted and the modulus of elasticity (MOE) was calculated
of the compressive strength of the specimens tested. Ten stretcher from the slope of the straight-line portion of the curves between 10%
specimens were prepared and tested for compressive strength at the age and 33% of the peak compressive stress.
of 28 days. The vertical interlocking keys (protrusions) on the top

Fig. 4. RIB production a) Semi-automatic machine b) Final prototype c) RIB prototype.

4
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Fig. 5. Compressive strength test set-up (unit: mm).

3.4. Construction and testing set-up of hollow and grouted masonry wall ratio of 2 (height): 1 (length), therefore the RIB wall (1260 mm in height
and 650 mm in length) in the current study was followed the same
As stated in BS5628: Part 1:1992 clause A.2.2 [39] and as required by aspect ratio (2:1). Thus, the obtained results will be therefore applicable
ASTM E72 [40], the identical wall panel dimensions range from 1200 to for the full-scale real wall.
1800 mm in length with a minimum cross-section area of 0.125 m2 and Three each of hollow and grouted wall have been constructed using
from 2.4 to 2.7 m in height where, in special cases, it was necessary to rubberized interlocking bricks assembled in stretcher running bond with
test panels having dimensions outside these limits. Considering the dimensions shown in Fig. 6. The test walls have been made of twelve
rubberized interlocking brick as a special case. Moreover, it can be seen courses and each course contains two stretcher bricks (SB) and one half
that the aspect ratio (height to length) of the tested wall should be in a brick (HB) as shown in Fig. 6. The first course was positioned drily in a

Fig. 6. Masonry wall test setup and instrumentations (Unit: mm).

5
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

straight line on steel I-beam without any mortar. Once the base is researchers [43–47]. Second, a linear behavior for a small strain is
properly laid, the bricks were stacked dry and locked into place. The two observed and, finally, a non-linear behavior is initiated up to the final
protruding parts on the top shell-face surface positioned closely into load leading to the failure of the brick as shown in Fig. 7. The non-linear
cavity part on the bottom shell-face surface of the brick. These pro­ relationship between stress and strain has revealed a peak strain of
trusions and cavities allow the bricks to interlock with other bricks 0.0039, which corresponds to the ultimate stress. This finding is
placed above and below. For grouted walls, the vertical holes were filled consistent with the peak strain of the rubberized hollow brick reported
using grout and left undisturbed to cure in the laboratory for 28 days. by Silva et al. [48] and Sadek and El-Attar [45]. The elastic modulus of
Prior to the test, for the top course of the walls, the protrusions and the stretcher brick (Eb ) is found to be 7416 MPa, which is 403 times
grouts were grinded and eliminated from the top shell-face of the bricks greater than the compressive strength of stretcher brick (f ’b ) as presented
to smooth the surface. As shown in Fig. 6, hollow and grouted wall in Eq. (2).
specimens were capped with 5 mm thick of SikaGrout-215 and then a 25
mm spreader steel plate was placed on the top of the wall to ensure Eb ¼ 403 f ’b (2)
uniform vertical load distribution. The bottom support of the walls was The concrete grout has an adequate compressive strength of 21.1
restrained to avoid any movement during the testing. In addition, the MPa with a corresponding average strain of 0.0022 mm/mm. The
top end of the wall was set to be pinned support, while the bottom end modulus of elasticity (Eg ) of grout calculated by measuring the slope of
was fixed support to represent the wall assemblage in the actual single- stress-strain curve (Fig. 8) between 10% and 33% of the ultimate
storey building. The wall panels were loaded vertically using a 500 kN strength of the specimens as per ASTM C469 [49] and found to be 15827
manual hydraulic jack with a loading rate of 0.5 kN per second. MPa. This value is 750 times greater than the compressive strength of
In order to measure the applied compression load, the load cell was concrete grout (fg ) which can be express in Eq. (3).
attached to the hydraulic jack. Vertical displacement was measured
using six Linear Variable Differential Transformers (LVDT1–LVDT6) on Eg ¼ 750 fg (3)
the left, middle and right sides of the wall. The strain was calculated by
dividing the change in the height of the wall, measured by LVDTs, by the 4.2. Compressive strength and deformation behavior of wall
original height, and then the stress-strength curve was plotted. The
stiffness of the wall was obtained from the slope of the load-deformation In the present study, the load transmitted from the single-and dou­
curve in the region between 0.05 and 0.4 of the maximum stress. Lateral ble-story building to the load bearing wall system was primarily calcu­
deformation was measured by one Linear Variable Differential Trans­ lated. The design of the imposed and dead load specifications and the
formers (LVDT 7) placed in the middle of the compression face of the concrete density were chosen on the basis of BS 6399: Part 1: 1996 [50]
walls tested. During the test, the force values, where cracks appeared, and BS8110-1- 1992 [51] to meet at least the minimum structural design
were observed and recorded. The test was carried out in accordance with requirements for the concrete system components. The calculated loads
the procedures laid down in ASTM E7 [40]. on the wall, at the ultimate limit state (ULS), was found to be 20.8 kN/m
for single storey and 58.0 kN/m for double storey which are transmitted
4. Results and discussion from the floor, roof and self-weighting elements of the storey. The
comparison between these calculated load values and the experimental
4.1. Characteristic of rubberized interlocking brick and grout load results (Table 7, column 4) was therefore clearly demonstrated by
the rationale for the actual behavior of the rubberized interlocking
4.1.1. Compressive strength masonry wall prototype tested under axial compressive loads and the
Table 5 shows the results of the RIB compressive strength test (based actual case of the house wall assembly.
on the load bearing area) and the grout cylinder test. The results showed
that the average compressive strength of the RIBs was 18.40 MPa, which 4.2.1. Hollow walls
is 25% higher than the minimum compressive strength required for the Wall mean compressive strength (f ’mw ), wall efficiency (load bearing
load-bearing masonry unit as stated in ASTM C90 [35]. Therefore,
efficiency, f ’mw /f ’b ), where f ’b is the mean compressive strength of brick,
rubberized interlocking bricks are categorized as load-bearing bricks.
wall splitting stress and wall stiffness are calculated and tabulated in
The compressive strength of the grout mix was found to be 21.1 MPa,
Table 6. Hollow walls reached their compressive capacity at an average
which is somewhat close to the compressive strength of the rubberized
load of 181 kN and showed an average ultimate compressive strength
interlocking brick units in order to avoid a sharp decline in the final load
capacity of the specimen as recommended by Ko €ksal et al. [41] and (f ’mw ) of 3.87 MPa. This ultimate compressive strength provides only
Martins et al. [42]. Due to its lack of necessity, the weight of the grout 21.43% of the load-bearing efficiency. This can be attributed to the ef­
cylinder was not recorded. fects of the slenderness of the wall (ratio of height to thickness (h=t ¼

4.1.2. Stress-strain behavior


The observed stress-strained behavior of the RIB is characterized by;
first, a significant initial concave section (nonlinear region) due to weak
interfacial ties between cement paste and crumb rubber resulting in
lower compressive strength as reported and confirmed by several

Table 5
Average compression test results of brick and grout specimens.
Specimen Weight Bearing Max. Compressive Elastic
(kg) area load strength, f ’b Modulus, Eb
(mm2) (kN) (MPa) (MPa)

Stretcher 5.39 18750 343.68 18.40 7416


COV (%) 2.39 0.51 5.56 5.72 8.87
Grout NA 7854 165.5 21.1 15827
COV (%) NA 0 9.36 9.36 7.20
Fig. 7. Longitudinal stress-strain curve for rubberized interlocking brick (RIB).

6
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

the grout. In contrast to the case of hollow walls, there is no significant


transfer of load at dry joints at the initial loading stage due to the seating
effect. In addition, the re-distributed axial stress in the brick shells was
unevenly distributed due to contact imperfections in the dry interfaces
that caused premature local failures. In contrast to the hollow walls, the
grouted walls showed higher efficiency.
The average load-displacement curves are plotted in Fig. 11 and the
average axial displacement value shown is 5.48 mm. The figure showed
that grouting allows rubberized interlocking masonry walls to behave in
the same way as drilled conventional walls, particularly in-service
loading phases. However, the existence of dry joints between masonry
bricks will have an impact on post-cracking behavior. As a result, the
occurrence of initial seating deformations is delayed. This behavior is
consistent with the load-vertical displacement reported by Sadek and Al-
Fig. 8. Longitudinal stress-strain curve for concrete grout (CG). Ettar [45] on rubberized masonry walls and with the same pattern of
interlocking bricks studied by Ali et al. [13]. The lateral displacement of
10.08) as well as the geometric imperfections of the dry contact in­ the grouted walls is shown in Fig. 12. It was found to be 0.69 mm on
terfaces that cause the initial deformation of the seating and premature average, which means that there was no buckling on the specimens
failure of the specimens, thus reducing the strength of the walls. How­ tested so that they could be neglected.
ever, this reduction in compressive strength is consistent with what
previous researchers have reported for interlocking masonry walls [1,
15,52].
The HW1-HW3 load-axial deformation curves shown in Fig. 9 reveal
similar behavior with ultimate axial deformation values ranging from
6.13 to 7.73 mm (6.98 mm average). Higher axial deformation (average
1.5 mm) was observed at lower applied loads (average 30 kN) in all
hollow walls compared to axial deformation observed at higher applied
loads (6.98 mm at 181 kN). This can be due to the interlocking mech­
anism and the irregular interface joints that caused the initial joint
closure and higher seating deformation [1] as well as the presence of
crumb rubber because it has great flexibility and ability to stretch and
rotate around its axes, making it able to withstand large deformations
[45,53]. A similar trend in the load-displacement curve was also
observed in masonry walls constructed using interlocking soil-cement
bricks [8,9].
Fig. 10 shows the lateral deflection curves of the walls tested, where Fig. 9. Average load-axial deformation curve of hollow wall.
the results showed a minimal lateral deformation (1.56 mm). This lateral
deflection is due to the effect of Poisson’s ratio of the lateral strain to the
corresponding axial strain reported by Nehdi and Khan [54]. It means,
therefore, that there was no buckling as the test walls were stable.

4.2.2. Grouted walls


The compression results of grouted wall are shown in Table 6. It
shows an average mean compressive strength (f ’mw ) of 5.75 MPa corre­
sponding to 32% of the average compressive strength of each brick unit
(f ’b ) which is similar to that found by Jaafar et al. [2]. Strength reduction
can be attributed to the premature failure of the face shell due to the
lateral expansion of the grout inside the central cores, which caused high
tension in the masonry bricks. Therefore, together with the surrounding
brick plates, the grout resists the initial stage loads which are bonded to
Fig. 10. Load-lateral displacement curve of hollow wall.

Table 6
Compression test results of hollow and grouted walls.
Wall Specimen Load bearing Ultimate Mean strength, Max. axial Max. lateral Web splitting Web splitting Efficiency, %, Stiffness,
area, mm2 load, kN (f ’mw ), MPa displ. (mm) display. (mm) load, kN stress, σw, MPa f ’mw /f ’b kN/mm

Hollow HW1 46875 174.5 3.73 7.73 1.68 120.2 2.56 20.63 28
HW2 46875 214.9 4.59 7.08 1.64 140.1 2.99 25.50 33
HW3 46875 153.6 3.28 6.13 1.37 90.25 1.93 18.17 21
Mean 46875 181.0 3.87 6.98 1.56 116.8 2.49 21.43 27
COV, % 14.1 9.41 8.81 17.5 14.21 18
Grouted GW1 72107 424.8 5.89 5.29 0.66 140.1 1.95 0.33 104.7
GW3 72107 403.9 5.60 5.39 0.59 138.6 1.92 0.31 108.5
GW2 72107 416.0 5.76 5.76 0.81 160.5 2.23 0.32 96.5
Mean 72107 414.9 5.75 5.48 0.69 146.4 2.03 0.32 103.2
COV, % 2.07 3.69 13.4 6.85 2.55 4.85

7
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Fig. 11. Average Load-axial displacement curve of grouted wall.

4.3. Failure mode and cracking pattern of hollow and grouted walls

Typical failure modes observed in hollow walls subject to vertical


Fig. 13. Cracking and failure modes of hollow wall.
compression are shown in Fig. 13. It was observed that the development
of micro cracks, at about 20–30% (45 kN) of the final load, was triggered
by the failure of the middle units on the face shells. As the load
increased, the micro cracks developed vertical and inclined cracks on
the sides of the middle units, as shown in Fig. 13a. In addition, localized
shear failure near the interlocking joints was observed. This can be
attributed to the uneven distribution of bed stress due to the inconsis­
tency of the interlocking surface caused by roughness and irregularity of
the dry bed joints as well as the variations in height of the units. The
spalling of the brick face shells was also observed in the hollow walls,
but it was not as intense as in the case of the grouted walls. Web splitting
was another typical failure mode observed in the hollow wall specimens
tested, as shown in Fig. 13b. The average stress that web-initiated cracks
were 2.49 MPa, as shown in Table 6. These types of failure modes are
common in the mortarless brick system [2,26] and conventional hollow
brick system [55,56].
The cracking and failure modes of the cracked walls are shown
Fig. 14. The vertical micro-crack on the face-shells of the grouted walls
was observed at about 25–40% (90–140 kN) of the ultimate failure load,
while the start of the web splitting occurred at an average web splitting
stress of 2.03 MPa. Failure modes were similar to hollow walls except
that the vertical openings on the face-shells and webs were larger and
the local crushing near the interlocking grooves was more pronounced.
Fig. 14. Cracking and failure modes of grouted wall.
Similar failure patterns have been reported for cracked interlocking
mortarless brick walls [2,3,26,57].
dividing the ultimate axial load by the load bearing area while the strain
was calculated by dividing the average longitudinal deformation, ob­
4.4. Axial stress-strain relationship tained from LVDTs, by the original height of the test specimen.
Furthermore, due to the initial seating deformation at the dry bed joints,
For both hollow and grouted wall panels, stress was calculated by the modulus of elasticity (Emw ) was determined by calculating the
ascending slope of the stabilized portion of the curves which lies
approximately between 20 and 50% of ultimate stress.

4.4.1. Hollow walls


Average stress-strain curves of hollow walls and their parameters are
shown in Fig. 15 and Table 7 respectively. As can be seen in the figure,
the first portion of the curve, which extended to approximately 25%–
30% of the ultimate stress level, was nonlinear with low stiffness
resulting from the initial settlement of uneven and irregularities surfaces
and height variations of adjacent bricks. The second portion of the curve
was linear and more dependent on the deformation of the bricks,
becoming reasonably stabilized and showing similar behavior until
failure. The average ultimate strain was 0.0055 mm/mm length which is
in line with the concrete containing crumb rubber [44,58,59] and with
the strain of rubberized long hollow bricks [60]. The average elastic
modulus, the slope of the stabilized portion of the curves which is
Fig. 12. Load-lateral displacement curve of grouted wall.

8
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Fig. 15. Average axial stress-strain curves of hollow walls. Fig. 16. Average axial stress-strain curves of grouted walls.

continued to gain strength with reduced stiffness until the final strength
Table 7 was reached. Unlike conventional masonry walls, a sudden drop in the
Stress-strain properties of rubberized interlocking hollow and grouted walls. stress was not observed and a descending portion was then observed
Wall ID Peak stress, Ultimate Modulus Emw =f ’mw with a mild slope.
f ’mw MPa strain, ε, mm/ elasticity, Emw , % The stress–strain curve for both hollow and grouted walls was further
mm MPa
analyzed to determine the energy absorption capacity and it was defined
Hollow HW1 3.73 0.0061 714.13 191.46 as the area under the stress-strain curve [58]. It is found that the total
HW2 4.59 0.0056 864.00 188.24 compressive toughness (energy absorption capacity) of the hollow and
HW3 3.28 0.0049 604.12 184.18 grouted rubberized interlocking wall is 0.025 MPa and 0.032 MPa and,
Mean 3.87 0.0055 727.42 188
respectively. Similar findings have been reported by Topcu and Ozceli­
COV, % 14.1 8.9 14.6
Grouted GW1 5.89 0.0042 1830.03 310.7 kors [63] and Sadek and El-Attar [45] that the addition of a 10% rubber
GW2 5.77 0.0046 1687.02 292.4 chip has increased the toughness (energy absorption) of concrete by
GW3 5.601 0.0043 1817.98 324.6 more than 20%. The results confirmed that the use of rubber signifi­
Mean 5.75 0.00436 1778.34 309.2
cantly increased the toughness of the masonry walls, regardless of the
COV, % 2.06 3.89 3.6
size of the rubber. This means that rubberized interlocking walls have a
significant ability to withstand post-failure loads and absorb more en­
approximately between 20 and 50% of the ultimate stress, was found to ergy for fracture than unrubbed walls.
be 727.42 MPa. This value is equal to 188 times the ultimate stress,
(188f ’mw ), which is much lower than the concrete masonry. The highly 5. Conclusion
significant decrease in elastic modulus may be contributed to the
weakness of crumb rubber existed in the test walls which accompanied The conclusions drawn from this study are as follows:
by a large strain and large deformation in addition to the gaps inherent
in interlocking masonry. 1 Rubberized interlocking brick can be categorized as load-bearing
masonry brick because its strength is 18.4 MPa, which is greater
4.4.2. Grouted walls than the required strength (13.7 MPa)..
Stress-strain curves of each grouted wall and significant parameters 2 The use of crumb rubber as a fine aggregate in interlocking bricks
are presented in Fig. 16 and Table 7, respectively. The figure shows an reduces the maximum stress and stiffness while increasing the strain
initial non-linear curve at low stress, approximately up to 20% of the and energy absorption capacity before failure (ductility).
final stress due to the existence of crumb rubber leaving gaps between 3 Rubberized interlocking walls increase ductility and undergo
the brick constituents and resulting in a decrease in the ductility of the measurable post-failure loads with significant displacement due to
strain and a less linear slope of the strain [46]. This behavior is referred the ability of crumb rubber to withstand large elastic deformation
to as non-linear progressive stiffening at the beginning of the compres­ prior to failure.
sion loading regime for the dry surface interface, which was observed 4 The compressive strength of the rubberized interlocking grooves
significantly by Jaafar et al. [3], Zahra [61] and Andreev et al. [62]. The increased by almost 50% compared to the rubberized interlocking
linear portion was observed up to 90% of the ultimate stress in the hollow walls. Designers shall therefore consider an increase of at
specimens. Subsequently, the results recorded a mean final stress of least 45% in the design of rubberized interlocking grouted wall
5.75 MPa with a corresponding axial strain of 0.00436, which in turn is strength compared to the hollow wall strength.
in good agreement with that reported by Jaafar et al. [3] for interlocking 5 Rubberized interlocking walls tend to have large initial deflections as
masonry wall. The elastic modulus (Emw ) has been found to be 1778.34 the bricks settle and the gaps caused by dry joints close. After major
MPa on average which is 2.44 higher than the modulus of elasticity for early displacements, stiffness tends to increase.
hollow wall. 6 The cracks on the sides of the shells were a combination of vertical
Generally, the stress-strain response of the grouted walls was slightly flexural cracks and shear cracks. The common failure modes of
different from the hollow walls, especially at higher loads, which were hollow and grouted rubberized interlocking masonry walls were web
attributed to the local crushing of the groove core on the grouted splitting and face spalling.
specimens. However, at levels approaching peak stress, the behavior of 7 The conduct of rubberized interlocking masonry walls subject to
grouted and hollow walls tends to be non-linear. It was observed that, eccentric loads, lateral loads and out-of-plane loads must be further
following the occurrence of the yielding stress, the grouted specimens investigated.

9
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

Author statement [27] T. Uyguno� glu, I.B. Topcu, O. Gencel, W. Brostow, The effect of fly ash content and
types of aggregates on the properties of pre-fabricated concrete interlocking blocks
(PCIBs), Constr. Build. Mater. 30 (2012) 180–187.
Amin Al-Fakih: Conceptualization, Methodology, Formal analysis, [28] ASTM C150/C150M-17, Standard Specification for Portland Cement, ASTM
Investigation, Writing - Original draft preparation. M. M. A. Wahab: International, West Conshohocken, PA, 2017, pp. 1–7.
Writing - Review & Editing, Funding acquisition. Bashar S. Moham­ [29] ASTM C618-17a, Standard Specification for Coal Fly Ash and Raw or Calcined
Natural Pozzolan for Use in Concrete, ASTM International, West Conshohocken,
med: Conceptualization, Resources, Supervision, Funding acquisition. PA, 2017.
M.S. Liew: Resources. Noor Amila Wan Abdullah Zawawi: Resources. [30] A. Al-Fakih, B.S. Mohammed, M.S. Liew, W.S. Alaloul, Physical properties of the
Sholihin As’ad: Writing - Review & Editing. rubberized interlocking masonry brick, Int. J. Civ. Eng. Technol. 9 (6) (2018)
656–664.
[31] A. Al-Fakih, B.S. Mohammed, M.S. Liew, W.S. Alaloul, M. Adamu, V.C. Khed, M.
Acknowledgment A. Dahim, H. Al-Mattarneh, Mechanical behavior of rubberized interlocking bricks
for masonry structural applications, Int. J. Civ. Eng. Technol. 9 (9) (2018)
185–193.
The authors would like to thank Universiti Teknologi Petronas (UTP) [32] ASTM C136/C136M-14, Standard Test Method for Sieve Analysis of Fine and
Malaysia for granting the project under Yayasan UTP with grant code Coarse Aggregates, ASTM International, West Conshohocken, PA, 2014, pp. 1–5.
YUTP 0153AA–H30. [33] ASTM C33/C33M-16e1, Standard Specification for Concrete Aggregates, ASTM
International, West Conshohocken, PA, 2016.
[34] ASTM C140/C140M-17b, Standard Test Methods for Sampling and Testing
References Concrete Masonry Units and Related Units, ASTM International, West
Conshohocken, PA, 2017.
[1] S.I. Fundi, J.W. Kaluli, J. Kinuthia, Performance of interlocking laterite soil block [35] ASTM C90-16a, Standard Specification for Load-Bearing Concrete Masonry Units,
walls under static loading, Constr. Build. Mater. 171 (2018) 75–82. ASTM International, West Conshohocken, PA, 2016.
[2] M.S. Jaafar, W.A. Thanoon, A.M. Najm, M.R. Abdulkadir, A.A.A. Ali, Strength [36] ASTM C1019-16, Standard Test Method for Sampling and Testing Grout, STM
correlation between individual block, prism and basic wall panel for load bearing International, West Conshohocken, PA, 2016.
interlocking mortarless hollow block masonry, Constr. Build. Mater. 20 (7) (2006) [37] I. Satyarno, A.P. Solehudin, C. Meyarto, D. Hadiyatmoko, P. Muhammad, R. Afnan,
492–498. Practical method for mix design of cement-based grout, Proc. Eng. 95 (2014)
[3] M.S. Jaafar, A.H. Alwathaf, W.A. Thanoon, J. Noorzaei, M.R. Abdulkadir, 356–365.
Behaviour of interlocking mortarless block masonry, Constr. Mater. 159 (3) (2006) [38] ASTM E111-17, Standard Test Method for Young’s Modulus, Tangent Modulus, and
111–117. Chord Modulus, ASTM International, West Conshohocken, PA, 2017.
[4] Y.H. Lee, P.N. Shek, S. Mohammad, Structural performance of reinforced [39] BS 5628-1, Code of Practice for the Use of Masonry Part 1: Structural Use of
interlocking blocks column, Constr. Build. Mater. 142 (2017) 469–481. Unreinforced Masonry, BSI, 2005, p. 80.
[5] P. VanderWerf, Mortarless block systems, Masonry Construction 12 (2) (1999) [40] ASTM E72-15, Standard Test Methods of Conducting Strength Tests of Panels for
20–24. Building Construction, ASTM International, West Conshohocken, PA, 2015.
[6] T. Zahra, M. Dhanasekar, Characterisation and strategies for mitigation of the [41] H.O. K€ oksal, C. Karakoç, H. Yildirim, Compression behavior and failure
contact surface unevenness in dry-stack masonry, Constr. Build. Mater. 169 (2018) mechanisms of concrete masonry prisms, J. Mater. Civ. Eng. 17 (1) (2005)
612–628. 107–115.
[7] S.A. Osman, Z.S. Mohamed, A. Sulaiman, M.F. Ismail, Experimental analysis of [42] R.O.G. Martins, G.H. Nalon, R.d.C.S.S.A. Alvarenga, L.G. Pedroti, J.C.L. Ribeiro,
Interlocking load bearing wall brickool system, Key Eng. Mater. 594–595 (2014) Influence of blocks and grout on compressive strength and stiffness of concrete
439–443. masonry prisms, Constr. Build. Mater. 182 (2018) 233–241.
[8] S. Ahmad, S. Hussain, M. Awais, M. Asif, H. Muzamil, R. Ahmad, S. Ahmad, To [43] M. Danko, E. Cano, J. Pena, Use of Recycled Tires as Partial Replacement of Coarse
study the behavior of interlocking of masonry units/blocks, IOSR J. Eng. 4 (3) Aggregate in the Production of Concrete, Purdue University Calumet, 2006.
(2014) 39–47. [44] S.D. Parveen, A. Sharma, Rubberized concrete: needs of good environment
[9] Z. Ahmad, S. Othman, B. Md Yunus, A. Mohamed, Behaviour of masonry wall (overview), Int. J. Emerg. Technol. Adv. Eng. 3 (2013) 192–196.
constructed using interlocking soil cement bricks, World Acad. Sci. Eng. Technol. [45] D.M. Sadek, M.M. El-Attar, Structural behavior of rubberized masonry walls,
(2011) 1263–1269. J. Clean. Prod. 89 (2015) 174–186.
[10] G.D. Robert, A.H. Ahmad, Behavior of concrete block masonry under axial [46] N. Eldin Neil, B. Senouci Ahmed, Rubber-Tire particles as concrete aggregate,
compression, ACI Journal 76 (6) (1979) 707–722. J. Mater. Civ. Eng. 5 (4) (1993) 478–496.
[11] M. Martínez, S. Atamturktur, Experimental and numerical evaluation of reinforced [47] J. Lv, T. Zhou, Q. Du, H. Wu, Effects of rubber particles on mechanical properties of
dry-stacked concrete masonry walls, J. Build. Eng. 22 (2019) 181–191. lightweight aggregate concrete, Constr. Build. Mater. 91 (2015) 145–149.
[12] E. Carrasco, J. Mantilla, T. Esp�
osito, L. Moreira, Compression performance of walls [48] R.A. Silva, E. Soares, D.V. Oliveira, T. Miranda, N.M. Cristelo, D. Leit~ ao,
of interlocking bricks made of iron ore by-products and cement, Int. J. Civ. Eng. Mechanical characterisation of dry-stack masonry made of CEBs stabilised with
Technol. 13 (3) (2013) 56–62. alkaline activation, Constr. Build. Mater. 75 (2015) 349–358.
[13] M. Ali, R.J. Gultom, N. Chouw, Capacity of innovative interlocking blocks under [49] ASTM C469/C469M-14, Standard Test Method for Static Modulus of Elasticity and
monotonic loading, Constr. Build. Mater. 37 (2012) 812–821. Poisson’s Ratio of Concrete in Compression, ASTM International, West
[14] G. Marzahn, Investigation on the initial settlement of dry-stacked masonry under Conshohocken, PA, 2014.
compression, Leipzig Ann. Civ. Eng. Rep. 3 (1999) 247–261. [50] B.S. Institution, Part 1: Code of Practice for Dead and Imposed Loads, 1996.
[15] H.C. Uzoegbo, R. Senthivel, J.V. Ngowi, Loading capacity of dry-stack masonry [51] B.S. Institution, Structural Use of Concrete, Part 1: Code of Practice for Design and
walls, Masonry Soc. J. 25 (1) (2007) 41–52. Construction, British Standards Institution, UK, 1997.
[16] K. Anand, K. Ramamurthy, Development and evaluation of hollow concrete [52] T. Miranda, R.A. Silva, D.V. Oliveira, D. Leit~ao, N. Cristelo, J. Oliveira, E. Soares,
interlocking block masonry system, Masonry Soc. J. 23 (1) (2005) 11–19. ICEBs stabilised with alkali-activated fly ash as a renewed approach for green
[17] M. Hatzinikolas, Structural behavior of an interlocking masonry block, in: building: exploitation of the masonry mechanical performance, Constr. Build.
Proceedings of the 4th Canadian Masonry Symposium, June 1986, 1986. Mater. 155 (2017) 65–78.
[18] A.V. Dyskin, E. Pasternak, Y. Estrin, Mortarless structures based on topological [53] N.J. Azmi, B.S. Mohammed, H.M.A. Al-Mattarneh, Engineering properties of
interlocking, Front. Struct. Civ. Eng. 6 (2) (2012) 188–197. concrete containing recycled tire rubber, ICCB B 34 (34) (2008) 373–382.
[19] K.B. Najim, M.R. Hall, Mechanical and dynamic properties of self-compacting [54] M. Nehdi, A. Khan, Cementitious composites containing recycled tire rubber: an
crumb rubber modified concrete, Constr. Build. Mater. 27 (1) (2012) 521–530. overview of engineering properties and potential applications, Cem. Concr.
[20] P. Turgut, B. Yesilata, Physico-mechanical and thermal performances of newly Aggregates 23 (1) (2001) 3–10.
developed rubber-added bricks, Energy Build. 40 (5) (2008) 679–688. [55] N.G. Shrive, The Failure Mechanism of Face-Shell Bedded (Ungrouted and
[21] S. Naganathan, A.Y.O. Mohamed, K.N. Mustapha, Performance of bricks made Unreinforced) Masonry Subject to Compressive Loading, University of Calgary.
using fly ash and bottom ash, Constr. Build. Mater. 96 (2015) 576–580. Department of Civil Engineering1981.
[22] J. Feng, J. Sun, P. Yan, The influence of ground fly ash on cement hydration and [56] B. Jonaitis, R. Zavalis, Experimental research of hollow concrete block masonry
mechanical property of mortar, Adv. Civ. Eng. 2018 (2018) 7. stress deformations, Proc. Eng. 57 (2013) 473–478.
[23] B.S. Mohammed, M.S. Liew, W. S Alaloul, A. Al-Fakih, W. Ibrahim, M. Adamu, [57] K.-H. Oh, H.G. Harris, A.A. Hamid, Behavior of Interlocking Mortarless Masonry
Development of rubberized geopolymer interlocking bricks, Case Stud. Constr. under Compressive Loads, Seventh Canadian Masonry Symposium.
Mater. 8 (2018) 401–408. [58] A.R. Khaloo, M. Dehestani, P. Rahmatabadi, Mechanical properties of concrete
[24] M.E. Nazar, S.N. Sinha, Fatigue behaviour of interlocking grouted stabilised mud- containing a high volume of tire–rubber particles, Waste Manag. 28 (12) (2008)
fly ash brick masonry, Int. J. Fatigue 29 (5) (2007) 953–961. 2472–2482.
[25] B. Mohammed, M. Aswin, Properties and structural behavior of sawdust [59] I. Alam, U.A. Mahmood, N. Khattak, Use of rubber as aggregate in concrete: a
interlocking bricks, in: Proceedings of the 3rd International Conference on Civil, Review, Int. J. Adv. Struct. Geotech. Eng. 4 (2) (2015) 92–96.
Offshore and Environmental Engineering, 2016, pp. 437–442. Kuala Lumpur- [60] E. Sodupe-Ortega, E. Fraile-Garcia, J. Ferreiro-Cabello, A. Sanz-Garcia, Evaluation
Malaysia. of crumb rubber as aggregate for automated manufacturing of rubberized long
[26] T. Sturm, L.F. Ramos, P.B. Lourenço, Characterization of dry-stack interlocking hollow blocks and bricks, Constr. Build. Mater. 106 (2016) 305–316.
compressed earth blocks, Mater. Struct. 48 (9) (2015) 3059–3074.

10
A. Al-Fakih et al. Journal of Building Engineering 29 (2020) 101107

[61] T. Zahra, M. Dhanasekar, A generalised damage model for masonry under [63] I.B. Topçu, N. Avcular, Collision behaviours of rubberized concrete, Cement Concr.
compression, Int. J. Damage Mech. 25 (5) (2016) 629–660. Res. 27 (12) (1997) 1893–1898.
[62] K. Andreev, S. Sinnema, A. Rekik, S. Allaoui, E. Blond, A. Gasser, Compressive
behaviour of dry joints in refractory ceramic masonry, Constr. Build. Mater. 34
(2012) 402–408.

11

You might also like