You are on page 1of 13

Construction and Building Materials 43 (2013) 1–13

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


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

Experimental and numerical analysis of bending–buckling mixed failure of


brickwork walls
Ernest Bernat a,⇑, Lluis Gil a, Pere Roca b, Cristián Sandoval c
a
Department of Strength of Materials and Engineering Structures, Technical University of Catalonia UPC, Barcelona Tech, ETSEIAT, Colom 11, 08222 Terrassa, Spain
b
Department of Construction Engineering, Technical University of Catalonia UPC, Barcelona Tech, ETSECCPB, Jordi Girona 1-3, 08034 Barcelona, Spain
c
Department of Structural and Geotechnical Engineering, and School of Architecture, Pontificia Universidad Católica de Chile, Casilla 306, Correo 22, Santiago, Chile

h i g h l i g h t s

" Experimental study of the second order bending failure mode of full-scale unreinforced brickwork walls.
" Experimental analysis of the effect of different slenderness and eccentricities.
" Test setup for testing hinged full-scale masonry walls in laboratory.
" Numerical simulation of walls under eccentric compressive loads taking into account geometric and material non linearity.
" Comparison from different experimental programs, analytic standards solutions and the proposed numerical simulation.

a r t i c l e i n f o a b s t r a c t

Article history: The eccentric in-plane loading of masonry walls involves complex bending performance that includes
Received 19 September 2012 second-order effects. In this work, a bidimensional (2D) simplified micro-model for the analysis of this
Received in revised form 21 January 2013 type of failure is developed. An experimental investigation based on 20 tests of full-scale unreinforced
Accepted 25 January 2013
masonry walls is performed. The tests are characterised by slenderness and load eccentricity. The analyt-
Available online 28 February 2013
ical methods of Eurocode-6 and ACI-530 are compared with experimental data from the present investi-
gation, other experimental results available in the literature and simulation results from the numerical
Keywords:
model.
Buckling failure
Unreinforced brickwork wall
Ó 2013 Elsevier Ltd. All rights reserved.
Experimental tests
Simplified micro-modelling

1. Introduction masonry and its fragile response to tension. Although the tensile
strength of masonry walls is neglected in most codes (see [2,3]),
Although the use of structural masonry has been dramatically some authors (see [4–6]) have reported that this property has a
reduced since the widespread introduction of concrete and steel significant influence, even at small values, on the load-bearing
structures, there is still a large number of existing buildings in capacity of slender or eccentrically loaded walls.
use that are structurally composed of load-bearing brick masonry To understand the structural behaviour of compressed brick-
walls. Assessing the strength of these buildings is a part of the work walls, several experimental research studies have been pub-
maintenance work required to enhance their useful life. Among lished reporting empirical analytical calculation methods for the
the possible failure modes, buckling failure must be addressed to assessment of the wall load-bearing capacity. The application
verify the safety of the walls from sudden collapse. range of such methods is limited to specific support and load con-
The in-plane load-bearing capacity of unreinforced brickwork ditions [4,7–13]. Recently, the development of modern computa-
walls depends on both the material properties and the geometry tional methods (e.g., FEA) has fostered the proposal of a large
of the structure. The typical effects of eccentricity and slenderness number of numerical models to simulate the behaviour of
have been widely analysed and are reported in many works [1]. masonry. Most of the proposed computational methods for the
The behaviour of eccentrically loaded unreinforced masonry walls analysis of masonry structures have been formulated using
is complex because of the non-linear compressive behaviour of the micro-modelling [14–16] or homogenised macro-modelling
approaches [17–20].
⇑ Corresponding author. Tel.: +34 937398728; fax: +34 937398994. In the experimental field, several studies have reported on the
E-mail address: ernest.bernat@upc.edu (E. Bernat). load-bearing capacity of masonry walls. The works of da Porto

0950-0618/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.conbuildmat.2013.01.025
2 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

et al. [21] and Cavaleri et al. [22] are especially significant, with re- Some authors ([6,46,47]) have demonstrated that the most widely utilised
structural masonry codes are too conservative when applied to the analysis of very
spect to the study of the out-of-plane behaviour of load-bearing
slender or very eccentrically compressed unreinforced masonry walls. Comparing
walls. The similarity between full-scale and small-scale tests has the experimental values from the tests with the result of the application of widely
been widely studied ([23]), and full research programmes have used standards is another specific objective of the present work.
been completed considering only small-scale walls ([6,23–25]). Numerical models have been proposed to overcome the limitations of the stan-
Compressive full-scale tests on masonry walls or columns are not dards in calculating the structural response of masonry walls. However, most of the
proposed numerical methods have been implemented in-house, using proprietary
very common. Among the studies on full-scale walls, those of Wat-
software that is not freely available to outside practitioners. Conversely, the present
stein and Allen [26] and Kirtschig and Anstötz [27] are remarkable research is based on a finite element approach that is available in a commercial
because of the use of specific systems to model neatly hinged sup- software environment. Validating and proving the accuracy of this available numer-
ports (i.e., rotationally free) at the wall ends. In these cases, the ical method is another objective of this research. In addition to the experimental re-
sults obtained in the present study, the model is also validated using experimental
clarity and simplicity of the support condition allow the study of
data available in the literature.
the influence of the load eccentricity and the wall slenderness with The available data are extended by using the numerical tool to perform a com-
no uncertainty in the actual boundary conditions. prehensive numerical simulation. The experimental and numerical results are com-
One of the parameters that affect the buckling failure and the pared to two standards (EC-6 [2] and ACI 530 [3]) to assess the accuracy of the
load-bearing capacity of masonry walls is the masonry modulus standards.

of linear deformation (Young’s modulus). Several recent studies


that focussed on full-scale specimens have provided meaningful 3. Experimental tests
results on the relationship between the compressive strength of
masonry and its modulus of linear deformation ([28]). With To continue the previous work on small-scale masonry models
respect to material properties determination, the work of [6] and to enhance the analysed cases contrasting the results with
Maurenbrecher ([29–32]) is significant. Maurenbrecher ([29–32]) tests on full-scale walls, a complete experimental programme at
studied the effect of building procedures on the properties of the full scale was executed.
resulting masonry, particularly the compressive strength (fc) and Twenty masonry walls were tested under eccentric in-plane
the modulus of linear deformation (E). Maurenbrecher ([29–32]) compression to achieve out-of-plane failure from the combination
also analysed the influence of the load eccentricity on the masonry of second-order bending (which would cause a rapidly increasing
compressive strength (fc). Gazzola et al. [33] and Sinha et al. [34] out-of-plane deformation of the walls) and buckling (which would
have investigated the flexural strength of masonry to understand explain a sudden failure due to the applied axial load associated
out-of-plane behaviour. with a geometric instability). In addition, the components’ (bricks
Most of the experimental studies to date have been focussed on and mortar) mechanical properties and the masonry compound
the derivation of empirical analytical formulations for the estima- properties were characterised. In particular, 131 tests were per-
tion of the load-bearing capacity of walls (see [27]). One of the first formed to determine the compressive strength of the mortar, and
analytical formulations, presented by Yokel ([13,35]), has been the 249 tests were performed to assess the flexural strength of the
basis of several later proposals. mortar. Totals of 19 and 42 tests were performed to characterise
Alternative formulations that cover a wider range of situations the compressive strength and the flexural strength, respectively,
and are meant to be used with the aid of computers are becoming of the bricks. The water absorption of the ceramic was measured
more common in recent times ([10,28]). Because of the growing by testing 29 samples. Also 29 samples were tested to find the
complexity of the analytical formulations, modern computer compressive strength and Young’s modulus of the masonry, and
methods are becoming popular alternatives. Many finite element 32 more were tested to measure the masonry tensile strength.
approaches have been proposed for the analysis of masonry wall In the next subsections, the experimental programme is sum-
structures, including simplified methods, such as one-dimensional marised. Information is included on the preparation of the samples,
(1D) beam descriptions [36], and more realistic and sophisticated the test setup and the results.
applications, such as the micro-models presented by Lourenço
[37], Salerno and de Felice [38] and Martini’s three-dimensional
3.1. Material properties and production of test walls
(3D) models [39]. As a compromise, there is a growing
tendency to work in homogenised 2D models. Homogenisation
Two different types of Portland cement mortar were used to
techniques for masonry have been studied by many researchers
fabricate the walls. Both mortars were acquired as commercial
([15,40–44]).
products, to be mixed with water according to the manufacturer’s
instructions. Standardised tests to determine the compressive and
2. Scope and methods flexural strengths were performed as described in [48]. The results
are summarised in Table 1. Low-compressive-strength mortar was
The greater the slenderness and the load eccentricity are, the greater the effect
used to simulate ancient brickwork.
of the tensile strength and the boundary conditions on the load-bearing capacity
[35]. Typical failure is because of second-order out-of-plane bending caused by
One type of solid-clay brick, with dimensions of 280 
the load eccentricity and possible geometric irregularities. 132  45 mm3, was used. Tests to determine the compressive
Only a few studies on full-scale unreinforced masonry walls under eccentric strength (fcb = 27.93 MPa) and water absorption (from initial dry
compressive load have been performed to investigate this type of failure. Moreover, state, 1.46 mg/(mm2min), and after 1 min wetted, 0.65 mg/(mm2-
the previous researchers did not focus on the effects of the tensile strength on the
min)) were performed as described in [49,50], respectively. A
buckling response of the walls and the corresponding masonry tensile strength was
not measured. In some cases, the support conditions at the wall ends are not accu-
rately reported. These limitations make it difficult to use the data produced by the
previous tests as a basis to calibrate and validate newly developed numerical meth- Table 1
ods. Furthermore, although some extensive research has been successfully under- Characterisation test results of mortar, corresponding to the walls in which each
taken in this area recently ([6,45]), full-scale test results are required to mortar was used. The coefficient of variation of the resulting value is in parentheses.
complement the reduced-scale test data. Therefore, the first objective of the present Mortar Compressive strength Flexural strength Walls
research is to perform experimental tests on full-scale walls to complement the
data available in the literature and to analyse the effects of slenderness and eccen- # Tests Value (MPa) # Tests Value (MPa)
tricity on the load-bearing capacity of masonry walls. This study includes the char- M1 81 3.18 (0.33) 47 1.24 (0.30) 1
acterisation of the material mechanical properties of both the components and the M2 168 3.70 (0.63) 84 1.25 (0.89) 2–20
masonry as a composite material.
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 3

three-point bending test was used to determine the flexural Table 4


strength (ftb = 2.81 MPa). The tensile strength of the bricks was Theoretical geometry of the tested full-scale unreinforced masonry walls.

determined from the flexural values according to the method pro- Series Height, H (mm) Width, b (mm) Thickness, t (mm) Walls #
posed in [24]. H 2700 900 132 1–3, 5–9
The masonry was tested to obtain the compressive strength, fc, M 1650 900 132 10–16
the modulus of linear deformation in the direction perpendicular S 1000 900 132 18–20
to the bed joints, E, and the bonding strength between bricks and F 2700 900 132 4
T 1800 900 270 17
mortar joints, fxt. Specimens consisting of five stacked bricks were
used to obtain values for fc and E, and samples of two stacked
bricks were used for the other tests. The values obtained for each
wall series are summarised in Table 2. The results show that the the vertical direction to allow for free descending displacement on
value of the modulus of linear deformation is low compared with the top of the wall during the test.
the correlations between E and fc provided by the codes (E/fc = A steel beam was installed between the jack and the upper
1000 in [2] and E/fc = 700 in [3]). The measured values are closer hinge to distribute the load along the wall width. The connection
to those found by other researchers (see [31,51–53] and Table 3), between the wall and the test machine was made using two steel
although the measured values are still significantly lower. Using plates at each end. These plates held the wall in contact with the
low-strength mortars and actual brick-laying conditions (less con- hinges and allowed different load eccentricities to be set. Any
trolled than in typical laboratory conditions) may explain the eccentricity between 0 mm and 80 mm, at 5 mm intervals, was
unexpectedly low modulus of deformation. possible. The contact between the wall and the test system con-
Five different geometries of full-scale unreinforced masonry sisted of a 20-mm M2 mortar layer to distribute the load and avoid
walls (H, F, M, T, S) were tested as summarised in Table 4. Twenty points of high stress. The details of the connection system are
walls were tested. All walls were built by professional bricklayers shown in Fig. 2. Eight potentiometers were installed to control
in the vertical position to emulate the actual fabrication process. the rotation of both hinges (two sensors for each hinge) and the
The bricks were wetted before being placed in position to assure descending movement of the upper hinge (four sensors on the steel
good adherence. The alignment of each row was checked during distribution beam). Depending on the test, two or four laser dis-
the construction. The walls were fabricated on an auxiliary steel placement sensors were used to measure the out-of-plane defor-
tool utilised for the transport, elevation and positioning into the mation. The mid-height deformation was always measured with
test setup. Walls W#1–W#9 were fabricated and air-cured under the deformation at 1=4 and 3=4 height in the case of walls W#10–
indoor atmospheric conditions, whereas the rest of the walls were W#20.
fabricated and air-cured under outdoor atmospheric conditions. The tests were force-controlled. All data were recorded at 50 Hz.
An ordinary video camera was used to record all tests, and a high-
velocity camera was used to record the W#10–W#20 tests to cap-
3.2. Wall test setup and procedure ture the details of the failure mode.
The testing was executed as follows: (a) The wall was placed in
The test setup is described in Fig. 1. To produce accurate and the test position using an overhead travelling crane and a pallet
clear boundary conditions, the walls were prepared with a pair of truck, affixing the wall to the loading system in the theoretically
steel hinges as part of the test setup. The lower hinge was laid desired position; (b) The actual geometry was measured with a
on the floor (H and F series) or placed over a structural steel beam laser sensor, including local and global geometric deviations. Two
(M, S and T series) to increase the height of the wall’s base, whereas out-of-plane measurements, one at each border of the wall, were
the upper hinge was connected to a hydraulic jack with a compres- obtained every three masonry rows; (c) The sensors were put in
sion load capacity of 1000 kN. The upper hinge could move only in place; (d) The surrounding area was protected from the projection
of rubble; and (e) The test proceeded until the wall collapsed. The
Table 2
load increased continuously at a uniform rate until the sudden col-
Characterisation test results of masonry. lapse point.
Wall # Mortar Modulus of linear Compressive Bonding
deformation, E strength fc strength fxt 3.3. Experimental results
(MPa) (MPa) (MPa)
1 M1 780 18.2 0.23 Table 5 summarises the test conditions and results for the wall
2–5 M2 12.9 0.36 series. The geometry type (as outlined in Table 4) and slenderness
6–9 13.7
(Eq. (1)) are indicated for each wall. The mid-height eccentricity ra-
10–20 10.8
tio (em/t), calculated as the ratio of the initial mid-height eccentric-

Table 3
E/fk ratio from literature and experimental tests.

Source of information Modulus of linear deformation, E (MPa) Compressive strength, fc (MPa) E/fc
[31] Clay brickwork series 2 21,700.00 26.04 833.33
[31] Clay brickwork series 1 8675.00 13.82 627.50
[51] Concentric loading 2172.50 12.92 168.19
[52] Concentric brickwork type 1 1867.00 13.50 138.30
[52] Concentric brickwork type 2 1700.00 13.25 128.30
[52] Concentric brickwork type 3 2313.00 7.53 307.04
[51] Concentric load 1440.00 11.70 123.08
Mortar M1. Wall 1 780.00 18.20 42.86
Mortar M2. Walls 2–5 780.00 12.90 60.47
Mortar M2. Walls 6–9 780.00 13.70 56.93
Mortar M2. Walls 10–20 780.00 10.80 72.22
4 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

Fig. 1. Test setup (a) for walls from series S, (b) for walls from series M and (c) for walls from series H.

ity (em) to the thickness (t) of the wall, is also listed. A dimension- displacement of the upper hinge at the maximum load, vmax, both
less measure of the hinge vertical alignment (d/t), describing the measured from the effective height (Hef), are shown in the last col-
overall initial rotation of the wall, is also calculated and shown in umns of Table 5. Hef is the real distance between the axes of the
Table 5. Fig. 3a shows the geometric configuration of a typical wall hinges.
of the current campaign. With respect to the experimental results,
k ¼ Hef =t ð1Þ
the load-bearing capacity of the wall, Pmax, and the ratio (Eq. (2)) of
the maximum load to the maximum theoretical uniformly distrib- / ¼ Pmax =Pu ð2Þ
uted load (Eq. (3)), are included. The mid-height out-of-plane
deformation at the maximum load, hmax, and the descending Pu ¼ b  t  fc ð3Þ

Fig. 2. Detail of the connection between the wall and the test system at the lower end (left) and the upper end (right).
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 5

Table 5
The results of the tests on unreinforced masonry walls under eccentric compressive loads.

Wall Geometry k e/t (%) d/t (%) Pmax (kN) U (%) hmax/Hef  103 vmax/Hef  103
W#1 H 22.3 4.2 0.0 169.3 7.8 4.5 2.5
W#2 H 22.2 14.9 0.0 65.7 4.3 5.9 1.4
W#3 H 22.1 8.5 7.6 133.8 8.7 9.2 2.4
W#4 F 15.0 3.9 4.9 578.6 37.8 12.7 10.8
W#5 H 22.1 10.6 5.3 239.8 15.6 5.1 2.2
W#6 H 21.7 16.1 9.5 30.0 1.8 4.9 0.5
W#7 H 21.8 5.4 4.5 134.7 8.3 6.7 1.9
W#8 H 22.3 0.3 6.1 129.4 8.0 5.6 1.5
W#9 H 21.9 9.8 6.4 109.8 6.7 4.5 1.3
W#10 M 14.1 8.1 8.0 423.9 33.0 8.2 6.0
W#11 M 14.3 18.1 7.6 371.2 28.9 9.2 5.5
W#12 M 13.9 5.0 3.8 471.1 36.7 8.9 6.0
W#13 M 14.1 25.6 1.5 83.8 6.5 9.4 2.2
W#14 M 14.1 16.0 0.0 518.5 40.4 9.0 5.4
W#15 M 14.0 24.3 2.3 236.7 18.4 10.1 3.2
W#16 M 14.1 12.5 2.3 408.2 31.8 8.5 4.5
W#17 T 6.8 66.9 9.1 491.1 38.3 8.8 4.8
W#18 S 9.1 15.8 2.3 803.4 62.6 2.9 5.4
W#19 S 9.2 22.9 0.0 686.1 53.5 4.5 8.7
W#20 S 9.3 25.0 2.3 152.6 11.9 7.6 4.3

The walls were tested so that the eccentricities at the lower and ing capacity. As the results in Fig. 8 show, the descending displace-
upper ends were equal. Although a constant eccentricity along the ment of the wall top section at the instant of failure depends on the
wall height was expected under these conditions, additional acci- initial eccentricity at mid height (which, if the alignment of the
dental eccentricities appeared because of construction imperfec- hinge is not taken into account, is equal to e/t  d/2t). Practically,
tions. The actual wall shape was measured after the placement of the experiments show that, for a fixed slenderness, the vertical dis-
each wall in the test position. The actual mid-height eccentricity, placement required to cause the collapse of the wall decreases
em, was calculated based on the measurements. Construction with an increase in eccentricity. Slenderness also affects this rela-
imperfections also caused the hinges to be imperfectly aligned, tionship. The results differ between walls from series H, M and S, as
requiring the measurement of the initial horizontal distance shown in Fig. 8. For example, an increase of the normalised initial
between the upper and bottom axes (d). In Tables 5 and 7–9, if d eccentricity (without the hinge-misalignment effects) from 4.2% to
is positive the upper hinge was displaced in the out-of-plane direc- 20.8% causes a reduction in the vertical displacement needed to
tion (the right side in the configuration of Fig. 1) with respect to the reach failure from 2.5% of the effective height to 0.5% in H-series
lower hinge. The contribution of the initial alignment (d) to the walls (i.e., a reduction of 80%). Similar behaviour is observed for
eccentricity at mid-height was systematically estimated as d/2 for the M-series walls. As shown in Fig. 8, an increase in the norma-
the present experimental campaign. lised initial eccentricity from 3.1% to 24.8% causes a 63% reduction
The walls were compressed until failure. Three different col- in the vertical descending movement needed to cause the failure of
lapse mechanisms were observed. The most common failure mech- an M-series wall. Similar to the results in Fig. 7, these results are
anism observed was the formation of a single horizontal crack at consistent with the fact that most of the walls, as mentioned, failed
approximately mid-height with a sudden out-of-plane displace- because of geometric instability.
ment and collapse of the entire wall. Tensile failure caused the for- The experimental results show that the initial shape (eccentric-
mation of a natural joint and was the mechanism leading to the ity) affects both the load-bearing capacity and the vertical defor-
corresponding wall collapse. This mechanism was observed for mation at the instant of failure. This effect increases as the
all walls (Fig. 4) except W#4, W#14 and W#17. Walls W#14 (M slenderness of the wall increases.
series) and W#17 (T series) failed in a mixed mode, i.e., a combina- The results are also scattered. The scatter could be attributed to
tion of joint opening at the tension side with masonry crushing at the lack of homogeneity in the component material properties, the
the compression side (Fig. 5). Wall W#4 failed by out-of-plane manual procedure used to build the walls and/or the variation in
shear near the upper edge of the wall. The results of Wall W#5 the curing processes. Tests performed on the material components
were not included in the analyses because they show an anoma- (summarised in Table 1) resulted in coefficients of variation
lous response that is inconsistent with the rest of the H-series slightly higher than those found in literature. The vertical position
walls. of the walls during the construction and the manual brick laying
The experimental results are summarised in Fig. 6. As expected, may also explain some of the scatter compared with other test
the results show that the load-bearing capacity decreases signifi- studies (e.g. [6]). In addition, imperfections in the wall geometry
cantly with increases in the load eccentricity and the slenderness. were measured in the present study and may partially explain
For an actual eccentricity of approximately 16% of the wall thick- the scatter in the results. These imperfections are acceptable
ness, changing the slenderness from 9.1 to 14.1 or 21.7 causes a according to the current manufacturing standards. The air curing
loss of the load-bearing capacity of 35.5% or 96.3%, respectively. conditions may also explain part of the scatter. Some of the walls
Similarly, the load-bearing capacity of the walls in the H series were built and cured in summer outdoor conditions, whereas oth-
(with a theoretical slenderness of approximately 21) decreases ers were produced in winter conditions, both indoors and
by approximately 75% if the eccentricity is increased from 0.3% outdoors.
to 16.1% of the wall thickness. The load-bearing capacity decreases The scatter in the results is comparable to the scatter typically
by approximately 80% if the eccentricity is reduced from 5% to observed in similar studies [26,27]. Although the scatter is signifi-
25.6% of the wall thickness for M-series walls. cant, these tests more closely represent actual masonry structures
The results in Fig. 7 exhibit a linear relationship between the than strict laboratory-controlled tests. The data obtained in this
vertical displacement of the walls and their maximum load-bear- study are representative of real structures. Nonetheless, scatter
6 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

should be taken into account when analysing and comparing the


results in Section 5.

4. Numerical tool and standard codes

4.1. Features of the finite element analysis method

Although there are widely accepted models to represent the


behaviour of masonry (see [14–20,54–58]), a different approach,
intended to be easily implemented in commercial software
(ANSYSÓ 12.1), is preferred because of its wider availability to
practitioners. The model used in this study is a 2D plane-strain
simplified micro-model that combines solid elements to represent
the units with interface joint elements to represent the bed joints.
As it is normally done when using simplified micro-models, the
unit volume is increased to represent the surrounding mortar
joints. Hence, the model is composed of extended units that model
both bricks and mortar and zero-thickness interface elements that
simulate the discontinuities. The interface elements do not affect
the response to compression but can open and rotate under ten-
sion. In compression, the units are assumed to behave elastically
up to the compression limit; subsequently, perfect plasticity is
applied. Under tension, the response is controlled by the interface
joint elements, which have a perfectly brittle response. Hence, the
non-linear inelastic behaviour of the material in tension is concen-
trated in the joint elements.
Because the interface elements do not deform in compression,
the masonry deformation modulus is determined by the deforma-
bility of the extended units. The Young’s modulus used for the
extended units is equal to the value determined experimentally
on masonry specimens.
The contact elements are characterised by a cohesive zone mod-
el (CZM) that was originally proposed by [59] to describe tensile
behaviour. The CZM defines elastic behaviour (using the same E
as in the masonry homogenised elements) up to the fragile failure
of the joint. When the tensile strength (brick–mortar interface) is
reached, the contact opens while the normal tensile stress
decreases linearly. The parameters required to model the joints
in the CZM are the ultimate tensile stress (fxt) of the brick–mortar
interface and the fracture energy of the first fracture mode (GIf ) of
the brick–mortar interface. The model includes large deformations
to allow second-order bending effects and to allow the expected
buckling failure. The load is indirectly applied as an imposed verti-
cal descending displacement of the upper end of each wall. The
analyses are performed by increasing the displacement step-by-
step and determining the post-critical behaviour.
All simulated cases are defined with a hinge test setup. In Fig. 3,
the test configuration and model geometry used for each case is
presented. Images a–c of this figure show the real test configura-
tion, with the actual shape of the hinges, for the walls of the cur-
rent experimental campaign (a), the experimental campaign
presented in [27] (b) and the cases in [26] (c). The second column
of images in Fig. 3d–f, shows the geometric definition of the model
used for the same experimental campaigns. Finally, illustrations
g–i of the figure in Fig. 3 (cases g–i) represent the model used for
each case including the boundary conditions, the masonry rows
definition and the loading process.
As can be seen in Fig. 3d–f, the hinged condition of the wall has
been simulated in a simplified way by means of stiff triangular
objects placed at each end of the walls. These triangular objects
Fig. 3. Test configuration for the presented campaign (a), the campaign from [27]
have been defined with a much larger stiffness than that of ma-
(b) and the campaign from [26] (c); geometric definition for the presented
campaign (d), the campaign from [27] (e) and the campaign from [26] (f); model sonry. This simplified way of modelling the hinges has been suc-
used for the simulation of the walls from for the presented campaign (g), the cessfully used in similar applications (see [6]). These objects
campaign from [27] (h) and the campaign from [26] (i). provide a clear position for the load so that the eccentricity can
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 7

Table 6
Geometric values used in the numerical simulation of the experimental campaign from [26].

Case et (mm) eb (mm) em (mm) d (mm) Hef (m) t (mm) hrow aprox. (mm)
HB-10-0 0 0 0 0 1.15 93 73
HB-20-0 2.12
HB-30-0 2.98
HB-40-0 3.95
HB-10-1/6 15.5 15.5 1.15
HB-20-1/6 2.12
HB-30-1/6 2.98
HB-40-1/6 3.95
HB-10-1/3 31.0 31.0 1.15
HB-20-1/3 2.12
HB-30-1/3 2.98
HB-40-1/3 3.95

Table 7
Geometric values used in the numerical simulation of the experimental campaign from [27].

Case et (mm) eb (mm) em (mm) d (mm) Hef (m) t (mm) hrow aprox. (mm)
5.6-0 0 0 0 0 0.635 115 105
11.1-0 1.25
18.8-0 2.12
27.7-0 3.12
5.6-1/8 14.4 14.4 14.4 0.635
11.1-1/8 1.25
18.8-1/8 2.12
27.7-1/8 3.12
5.6-1/4 28.8 28.8 28.8 0.635
11.1-1/4 1.25
18.8-1/4 2.12
27.7-1/4 3.12
5.6-1/3 38.3 38.3 38.3 0.635
11.1-1/3 1.25
18.8-1/3 2.12
27.7-1/3 3.12

be modelled accurately in all cases. A minimum 1-mm eccentricity experimental load-bearing capacity of the walls presented in
is always used to provide an initial imperfection to activate the [26,27] and the numerical results for different slenderness and
geometric non-linear response. All the models (see Fig. 3g–i) have eccentricities. Table 8 presents the geometric data for the simula-
been defined with the actual geometry, measured as discussed in tion of the walls tested in the current research. It must be noted
Section 3.2 for the present experimental campaign and available that previous researches do not provide comprehensive informa-
in [26,27] for the comparison cases tested by other authors. Tables tion about the lateral or vertical displacement of the walls at the
6–8 summarise all geometric data used in numeric models. The collapse load. In contrast, the work herein presented includes a
description of the geometry (see Fig. 3d–f) takes into consideration comparison of the numerical predictions of the lateral deformation
the eccentricity at each end of the wall (eb and et which are the of the walls at mid-height (Fig. 8) and the vertical descending
eccentricity at the bottom and top of the wall respectively), the movement of the top of the specimen with the corresponding
hinge alignment (d) and the actual out-of-plane geometric imper- experimental values (Fig. 7).
fections which determine the mid-height eccentricity, em. The
effective height (Hef), wall thickness (t) and average masonry row
4.2. Application of the standard codes EC-6 and ACI-530
height (hrow) are also presented in Tables 6–8.
A mapped mesh of 8-node quadrangular elements is used in
The load-bearing capacity of the walls is calculated using two
combination with the interface elements. The parameters required
standard codes, EC-6 and ACI-530 ([2,3]), which have been applied
by the numerical model are the modulus of elasticity (E), the Pois-
to calculate the load-bearing capacity of walls in the literature
son’s coefficient (t) and the compressive strength (fc) of the ma-
([26,27]). The comparison of the standards’ analytical results with
sonry. The model requires also the bonding strength (fxt) of the
the experimental results is used to determine the accuracy of these
brick–mortar interface and the fracture energy of the brick–mortar
two widely used codes.
interface of the first fracture mode (GIf ) as input data. In the lack of
The codes are compared using the experimental values of the
direct experimental information, the fracture energy has been esti-
modulus of linear deformation, E, and the masonry compressive
mated based on a linear fitting, based on experimental values from
strength, fc (Table 9). Safety factors are not considered because
[60], relating the fracture energy of the brick–mortar interface, GIf
the purpose of the calculation is the direct comparison with exper-
with the corresponding bonding strength fxt. The resulting correla-
imental data.
tion allows the estimation of the fracture energy as GIf ¼ 36:65  fxt ,
where GIf is given in N/m and fxt in MPa. Table 9 summarises the
values of all variables used for each simulation case. Tables 6 and 5. Comparison of results and discussion
7 show the comparison of the numerical predictions with all the
experimental results for the cases taken from the literature In this section, the experimental results obtained by the
[26,27]. Figs. 9 and 10 show a graphical comparison between the present study and by other researchers are compared with the
8 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

Fig. 4. Failure mode of wall W#10. Opening of a joint and mechanism formation.

Fig. 5. Failure mode of wall W#14. Crushing of the compressed side mixed with the tensile opening of two joints to form a mechanism.

predictions of the numerical model and with codes EC-6 [2] and ends. The same numerical approach is compared with the experi-
ACI-530 [3]. mental results of the tests described in Section 3. The goal of this
Experimental results from other researchers [26,27] are ana- comparison is to evaluate the accuracy of the simplified micro-
lysed, and the numerical model is validated by comparison to these model in predicting the load-bearing capacity of walls.
tests. These studies involve 28 different combinations of slender- The methods proposed by EC-6 [2] and ACI-530 [3] for the
ness and eccentricity (see Tables 6 and 7). All of them are compres- determination of the wall capacity are also applied and compared
sion tests on full-scale unreinforced masonry walls hinged at both with the experimental and numerical results.
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 9

Fig. 6. Experimental relationship between eccentricity and load-bearing capacity Fig. 8. Experimental relationship between actual eccentricity (not taking into
for three different theoretical values of slenderness corresponding to wall geom- account hinge alignment) and the descending vertical displacement at maximum
etries H, M and S. load for three different theoretical values of slenderness corresponding to wall
geometries H, M and S.

e = 0, the error was 6.4%; for e = t/3, the error was 19.5%. The slen-
derness did not seem to affect the accuracy of the model, as good
results were obtained regardless of the value of the slenderness.
On the whole, the model slightly overestimates the load-bearing
capacity of the walls. By contrast, the EC-6 code produces conser-
vative results, underestimating the capacity of the walls for all
cases of [26] (with an average relative error of 62.1%), while the
ACI-530 code shows greater accuracy, with an error of 45.3%, and
a tendency to underestimate the load-bearing capacity of the most
eccentrically loaded walls. Conversely, the ACI-530 code overesti-
mates the load-bearing capacity of the concentrically loaded walls.
In Fig. 10, the results from tests on walls with calcium silicate
units reported in [27], are compared with the results obtained with
the numerical model. Good agreement between experimental and
numerical results is shown, with an average relative error of 28.4%.
This error is within the range of the scatter in the experimental
Fig. 7. Experimental relationship between downward displacement at maximum results in [27]. The agreement is particularly good in the cases with
load and load-bearing capacity for three different theoretical values of slenderness lowest eccentricities and/or slenderness (4.2% for e = 0 and k ¼ 5:6)
corresponding to wall geometries H, M and S.
in contrast with the trend obtained for the results reported in [26].
The numerical model seems to overestimate the load-bearing
Experimental data from two sources, namely Watstein and capacity of the walls tested by Kirtschig and Asntötz [27]. The
Allen [26] and Kirstchig and Anstötz [27] were chosen for the com- EC-6 standard code conservatively underestimates the strength
parison because of the clear boundary conditions of the tests (the of almost all walls compared with the data in [27] (Fig. 10). In all
walls were hinged at both ends), the detailed information regard- the comparisons, EC-6 overestimates the load-bearing capacity in
ing the material properties and the fact that all walls were at full only two cases, those with the highest eccentricity. The ACI-530
scale. standard code overestimates most of the ultimate loads. It shows
Table 9 summarises the parameters used in each simulation. For more accurate results for the cases with low load eccentricity
Poisson’s coefficient, which is of little relevance for the type of 2D and overestimates the load-bearing capacity for the cases with
modelling utilised, a common value taken from the literature the greatest load eccentricity. Compared with these data, the aver-
[61,62] was used. age relative error of the EC-6 standard code is 32.3% and that of the
The brick–mortar bonding strength (fxt) for the walls tested in ACI-350 standard code is 28.9%.
[27] was estimated because it was not provided by the authors. The comparison of the numerical model with the results from
In this case, the value of fxt was chosen to satisfy the ratio fc/fxt [26,27] shows that the largest influence of the Young’s modulus
equal to the experimentally obtained value in the present study (E) over the axial load-bearing strength of the walls occurs when
(fc/fxt = 40). The rest of the parameters in Table 9 were obtained the load is eccentrically applied and second-order bending appears.
from the experimental tests. As the Young’s modulus increases, the deformation of the wall
The series of walls labelled as ‘‘HB’’ in [26] was chosen for com- from second-order effects decreases because of the increased stiff-
parison. The numerical model provided a very satisfactory predic- ness of the wall. A larger axial load is thus needed to develop sec-
tion of the corresponding ultimate wall load-bearing capacities. ond-order bending failure, and buckling becomes more likely.
The average relative error was 10.5%. Some calculated results were Similarly, larger flexural brick–mortar bonding strength (fxt) con-
higher than the experimental ones and others were lower, so the tributes by providing more strength against second-order bending
model appears to be balanced (see Fig. 9). It predicts without either failure. The load-bearing capacity of the walls can also be improved
a conservative or an unsafe bias. However, the results with the by increasing the masonry compressive strength (fc). However, the
least error were in the cases with the lowest eccentricities. For effect of the compressive strength is more apparent when the pre-
10 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

Table 8
Geometric values used in the numerical simulation of the current experimental campaign.

Case et (mm) eb (mm) em (mm) d (mm) Hef (m) t (mm) hrow aprox. (mm)
W#1 20 20 5.6 0.0 2.95 132 60
W#2 20 20 19.6 0.0 2.93
W#3 10 10 1.2 10.0 2.92
W#4 0 0 0 6.5 1.98
W#5 5 5 7.0 7.0 2.92
W#6 25 25 33.7 12.5 2.86
W#7 10 10 13.1 6.0 2.87
W#8 10 10 7.6 8.0 2.94
W#9 5 5 21.5 8.5 2.89
W#10 10 10 0.2 10.5 1.87
W#11 20 20 13.9 10.0 1.89
W#12 10 10 1.6 5.0 1.84
W#13 30 30 31.8 2.0 1.86
W#14 20 20 21.2 0.0 1.86
W#15 30 30 29.0 3.0 1.85
W#16 20 20 19.5 3.0 1.86
W#17 80 80 88.0 12.0 1.91 282
W#18 20 20 20.8 3.0 1.20 132
W#19 30 30 30.3 0.0 1.22
W#20 30 30 33.0 3.0 1.22

Table 9
Values of the parameters for the numerical model.

Experimental reference E (MPa) fc (MPa) t fxt (MPa) GIf (N/m)

[26] 25,400 44.81 0.3 2.98 110


[27] 8540 12.2 0.3 7
W#1 780 18.2 0.35 0.23 8
W#2, W#3 and W#5 12.9 0.36 13
W#6–W#9 13.7
W#10–W#16 and W#18–W#20 10.8

Fig. 9. Comparison of the experimental results of [26] with the numerical


simulation and application of the standard codes. Fig. 10. Comparison of the experimental results of [27] with the numerical
simulation and application of the standard codes.

dominant failure mode is material crushing, as in the case of walls


with low load eccentricity (e) and a low slenderness ratio these results are not as good as those obtained for the experimen-
(k ¼ Hef =t). tal results of previous studies ([26,27]), the accuracy is acceptable
Once validated, the numerical model is applied to predict the because it is within the range of the common scatter obtained in
response of the H-, M- and S-series experimental walls presented the experimental tests. As mentioned previously, experimental re-
in Section 3.3 (except for wall W#5, whose inconsistent response sults of tests on masonry walls that rely on the flexural strength of
has been previously discussed). Table 9 summarises the parame- the material normally show highly scattered results, e.g. [33]. In
ters used in the simulations. Comparisons between calculated fact, the scatter in the experimental results in this study is larger
load-bearing capacity and experimental results are shown in Figs. than the average relative error of the numerical model. For H-series
11–13. The best agreement is obtained for the most slender walls walls, the results show no general over- or underestimation. The
(H series, Fig. 11), with an average relative error of 38.4%. Although best results are obtained for the most slender walls. This is the
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 11

increases. This result indicates that the numerical model is more


applicable for slender and eccentrically loaded walls. The average
relative errors are 43.2% and 61% for M- and S-series walls, respec-
tively. The errors in the most eccentric tests are 5.1% and 52.8% for
M- and S-series walls, respectively. The ultimate load calculated
with the numerical model follows the trend of the experimental
load measurement (see the linear fit in Fig. 12 and the results in
Fig. 13) and is accurate within the limits of the experimental
scatter. The behaviour of the model could be improved by imple-
menting a constitutive equation to account for the over-strength-
ening of masonry in compression, e.g., the increase of the
compressive strength for eccentric loading noticed by other
authors [52]. The inclusion of the over-strengthening in compres-
sion would cause higher load-bearing capacities when the crushing
failure process is significant, as in the case of less slender walls
subjected to concentric or moderately eccentric loads.
Fig. 11. Comparison of the experimental results of H-series walls with the In general, acceptable agreement is obtained between the
numerical simulation and application of the standard codes. experimental results and the numerical predictions. With respect
to the application of standards [2,3], in the case of the experimen-
tal programmes considered, EC-6 always underestimates the load-
bearing capacity of the walls, with an average relative error of
96.0%. The best agreement is observed for the less slender walls
(S series), with an error of 88.6%. The results from ACI-530 produce
an average relative error of 48.3%, producing conservative results
for almost all cases. ACI-530 yields the best results for walls of
the H series with an average relative error of 34.6%.

6. Conclusions

Experimental results on full-scale brickwork walls, tested with


in-plane eccentric compression loads, are presented. The experi-
ments comprise tests on masonry walls, changing the parameters
of slenderness and load eccentricity. All walls are hinged at both
ends. This study completes a previous work on small-scale models.
The experimental results show considerable scatter. Nonethe-
Fig. 12. Comparison of the experimental results of M-series walls with the
less, the relationships with eccentricity and slenderness exhibit
numerical simulation and application of the standard codes.
the same trends reported in previous research.
The collapses are caused by geometric instability, except for
walls W#4, W#14 and W#17, as discussed in Section 3.3. This type
of failure is observed in 17 of the 20 tests. As expected, this col-
lapse mode is more evident in the case of the most slender walls
or the most eccentrically loaded walls.
Using two mechanical hinges in the experimental design is
advantageous for the purpose of validating the numerical tools
because the boundary conditions are more precisely known and
can be more accurately modelled.
A numerical model for the prediction of the ultimate load-bear-
ing capacity of masonry walls is chosen and validated by compar-
ison with the experimental results. The experimental results are
also compared with the methods proposed by standards ACI-350
and EC-6 for the calculation of the load-bearing capacity of
masonry walls.
The numerical model yields better results than the two stan-
dard codes for all the wall series tested. Moreover, it is the most
Fig. 13. Comparison of the experimental results of S-series walls with the balanced method, with predictions of the load-bearing capacity
numerical simulation and application of the standard codes. both over- and underestimating the experimental results. The
EC-6 predictions are conservative for all cases except for those
shown in Fig. 11. Generally, ACI-530 is more accurate for cases
same trend that is observed in the modelling tests from [26]. with low eccentricity and low slenderness. In other words, ACI-
The comparison of a linear fit of the experimental results to a linear 530 is more applicable to cases when the failure is directly associ-
fit of the numerical model results (see Fig. 11) show that the model ated with masonry crushing. The numerical model is more accu-
correctly predicts the behaviour of the more slender walls (the rate in the cases with high eccentricity and more slenderness, as
H-series walls). observed by comparing Fig. 11 with Fig. 13. The model is less accu-
For the M- and S-series walls, the numerical results are consis- rate in cases with low eccentricity and/or less slenderness. This
tently conservative. The error decreases as the eccentricity reduction in accuracy may be caused by the simplicity of the com-
12 E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13

pressive behaviour model (linearly elastic and perfectly plastic). [17] Ainsworth M, Mihai LA. An adaptive multi-scale approach to the modelling of
masonry structures. Int J Numer Method Eng 2009;78:1135–63.
Future work to include an improved compressive stress–strain
[18] Chen S, Moon F, Yi T. A macroelement for the nonlinear analysis of in-plane
relationship in the model is recommended. Further future works unreinforced masonry piers. Eng Struct 2008;30:2242–52.
could also be oriented toward incorporating the out-of-plane shear [19] Mistler M, Anthoine a, Butenweg C. In-plane and out-of-plane homogenisation
in the analysis of unreinforced masonry walls. of masonry. Comput Struct 2007;85:1321–30.
[20] Ramalho M, Taliercio A, Anzani A, Binda L, Papa E. A numerical model for the
To obtain acceptable results with a finite element analysis, a description of the nonlinear behaviour of multi-leaf masonry walls. Adv Eng
realistic Young’s modulus must be used because it determines Softw 2008;39:249–57.
the deformation leading to the geometric instability associated [21] da Porto F, Mosele F, Modena C. Experimental testing of tall reinforced
masonry walls under out-of-plane actions. Constr Build Mater 2010;24:
with buckling failure. The formulations included in the standard 2559–71.
codes ([2,3]) are highly dependent on the value of Young’s modu- [22] Cavaleri L, Failla A, La Mendola L, Papia M. Experimental and analytical
lus. Given the variability and limited information on the value of response of masonry elements under eccentric vertical loads. Eng Struct
2005;27:1175–84.
Young’s modulus in masonry, it is necessary to do further work [23] Roca P. Simplified methods for assessment of masonry shear-walls. 6°
on the study of this essential parameter and its influence of the Congresso Nacional de Sismologia e Engenharia Sísmica; 2004. p. 101–18.
behaviour of slender and eccentrically loaded walls. [24] Charry J. Estudio experimental del comportamiento de paredes de obra de
fábrica de ladrillo ante la acción de cargas laerales; 2010.
These results emphasise that a realistic description of the joint [25] Murthy C, Hendry a. Model experiments in load bearing brickwork. Build Sci
behaviour is essential to accurately predict the load-bearing capac- 1966;1:289–98.
ity of unreinforced masonry walls, especially when the failure [26] Watstein D, Allen MH. Structural performance of clay masonry assemblages
built with high-bond organic-modified mortars. In: Second international brick
mode is primarily second-order bending.
masonry conference: structural clay products institute; 1970. p. 99–112.
[27] Kirtschig K, Asntötz W. Buckling tests on masonry. In: Proceedings of the 9th
international brick/block masonry conference. Berlin: IBMAC; 1991. p. 202–9.
Acknowledgements [28] Knutsson H. Vertical load bearing masonry – the danish approach. Mason Int
1991;5:23–6.
The authors wish to acknowledge Dr. Marco Antonio Pérez, Mr. [29] Maurenbrecher AHP. Effect of the test procedures on compressive strength of
masonry prisms. In: 2nd Canadian masonry symposium. Ottawa: National
Juan José Cruz, Mr. Francesc Puigvert and Mr. Christian Escrig for Research Council Canada; 1980. p. 119–32.
providing essential support in the experimental programme and [30] Maurenbrecher AHP. Compressive strength of eccentrally loaded
Dr. Jordi Marcé for his assistance in the development of the numer- masonry prisms. In: NRCC, editor. Third Canadian masonry
symposium. Edmonton: National Research Council Canada; 1983.
ical model.
[31] Maurenbrecher AHP. Axial compression tests on masonry walls and prisms. In:
This study has been partially performed within the project BIA Third North American masonry conference. Texas: National Research Council
2006-04127, funded by the Spanish Ministry of Science and Inno- Canada; 1985. p. 17.
[32] Maurenbrecher AHP. Compressive strength of hollow concrete blockwork. In:
vation, and the project MURETEX, funded by CDTI – Spanish Min-
Proceedings of the fourth Canadian masonry symposium. New
istry of Economics and Competitiveness, whose assistance is Brunswick: Institute for Research in Construction of National Research
gratefully acknowledged. Council Canada; 1986. p. 997–1009.
[33] Gazzola EA, Drysdale RG, Essawy AS. Bending of concrete masonry wallettes at
different angles to the bed joints. In: Proceedings of the third north american
References masonry conference. Arlington, Texas: Construction Research Center Civile
Engineering Department University of Texas; 1985. p. 15.
[34] Sinha B, Ng C, Pedreschi R. Failure criterion and behavior of brickwork in
[1] Mann W. Basics of design of masonry walls subjected to vertical loads and
biaxial bending. J Mater Civil Eng 1997;9:70–5.
buckling according to the present draft of Eurocode EC6 and comparison with
[35] Yokel FY. Stability and load capacity of members with no tensile strength. J
experimental data. In: Proceedings of the 9th international brick/block
Struct Div 1971;97:1913–26.
masonry conference. Berlin: Deutsche Gesellschaft fur Mauerwerksbau eV;
[36] Dawe JL, Liu Y. Analytical modeling of masonry load-bearing walls. Can J Civil
1991. p. 1281–91.
Eng 2003;30:795–806.
[2] Standardization ECf. Eurocode 6: design of masonry structures. Part 1-1:
[37] Lourenço PB. Analysis of masonry structures with interface elements. Theory
General rules for buildings. Rules for reinforced and unreinforced masonry;
and applications. TU-DELFT report no 03-21-22-0-011994. p. 34.
1997. p. 150.
[38] Salerno G, de Felice G. Continuum modeling of periodic brickwork. Int J Solids
[3] Committee MSJ. Building code requirements for masonry structures. ACI 530-
Struct 2009;46:1251–67.
05. Farmington Hills: American Concrete Institute; 2005.
[39] Martini K. Finite element studies in the two-way out-of-plane failure of
[4] Lu M, Schultz A, Stolarski H. Analysis of the influence of tensile strength on the
unreinforced masonry. In: Institute EER, editor. Proceedings of the 6th national
stability of eccentrically compressed slender unreinforced masonry walls
conference on earthquake engineering. Seattle; 1998. p. 1–12.
under lateral loads. J Struct Eng 2004;130:921.
[40] Cecchi A, Sab K. A homogenized Reissner–Mindlin model for orthotropic periodic
[5] Brencich A, Corradi C, Gambarotta L, Mantegazza G, Sterpi E. Compressive
plates: application to brickwork panels. Int J Solids Struct 2007;44:6055–79.
strength of solid clay brick masonry under eccentric loading. In: Proceedings
[41] Mercatoris BCN, Bouillard P, Massart TJ. Multi-scale detection of failure in
British Masonry Society; 2002. p. 37–46.
planar masonry thin shells using computational homogenisation. Eng Fract
[6] Sandoval C, Roca P, Bernat E, Gil L. Testing and numerical modelling of buckling
Mech 2009;76:479–99.
failure of masonry walls. Constr Build Mater 2011;25:4394–402.
[42] Milani G, Lourenço PB, Tralli A. Homogenised limit analysis of masonry walls,
[7] Dajun D. Studies on brick masonry under compression. Mater Struct
Part II: Structural examples. Comput Struct 2006;84:181–95.
1997;30:247–52.
[43] Milani G, Lourenço PB, Tralli A. Homogenised limit analysis of masonry walls,
[8] Defalco A, Lucchesi M. No tension beam-columns with bounded compressive
Part I: Failure surfaces. Comput Struct 2006;84:166–80.
strength and deformability undergoing eccentric vertical loads. IJMS
[44] Zucchini A, Lourenço PB. Homogenization of masonry using a micro-
2007;49:54–74.
mechanical model: Compressive behaviour. In: Soares CAM, editor. III
[9] Hansen L, Nielsen MP. Stability of masonry columns. orbitdtudk. Lyngby:
European conference on computational mechanics solids, structures and
Technical University of Denmark, Department of Civil Engineering, Section for
coupled problems in engineering. Lisbon, Portugal: ECCOMAS; 2006.
Structural, Engineering; 2003. p. 147.
[45] Reyes E, Casati M, Gálvez J. Experimental scale model study of cracking in brick
[10] Mura I. Stability of nonlinear masonry members under combined load. Comput
masonry under tensile and shear stress. Mater Construcc 2008;58:69–83.
Struct 2008;86:1579–93.
[46] Jager W, Bergander H. Comparison of buckling safety of masonry walls
[11] Romano F, Ganduscio S, Zingone G. Cracked nonlinear masonry stability under
according to EC6 and German standards. Proc Br Mason Soc 1998:279–83.
vertical and lateral loads. J Struct Eng 1993;119:69–87.
[47] Lijdens A, Villegas L. Aspectos del diseño de muros portantes de fábrica simple
[12] Schultz A, Mueffelman J, Ojard N. Critical axial loads for transversely loaded
según diferentes normas: excentricidades, reducción de la capacidad portante,
masonry walls. In: 12th int brick/block masonry conf proc; 2000. p. 1633–46.
e hipótesis de cargas a considerar. Rev Obras Pub 1994;141:43–53.
[13] Yokel FY. Strength of load bearing masonry walls. J Struct Div
[48] 83 CAC. UNE-EN 1015-11:2000/A1:2007. Métodos de ensayo de los morteros
1971;97:1593–609.
para albañilería. Parte 11: Determinación de la resistencia a flexión y a
[14] Orduña A. Seismic assessment of ancient masonry structures by rigid blocks
compresión del mortero endurecido. Madrid: AENOR; 2007.
limit analysis. PhD thesis 2003. University of Minho, Portugal. <www.civil.
[49] 41 CAC. UNE-EN 772-1:2002. Métodos de ensayo de piezas para fábrica de
uminho.pt/masonry> 2003. p. 171.
albañilería. Parte 1: Determinación de la resistencia a compresión. Madrid:
[15] Sacco E. A nonlinear homogenization procedure for periodic masonry. Eur J
AENOR; 2002.
Mech – A/Sol 2009;28:209–22.
[50] 41 CAC. UNE-EN 772-11:2001/A1:2006. Métodos de ensayo de piezas para
[16] Zucchini A, Lourenço PB. Mechanics of masonry in compression: results from a
fábrica de albañilería. Parte 11: Determinación de la absorción de agua por
homogenisation approach. Comput Struct 2007;85:193–204.
E. Bernat et al. / Construction and Building Materials 43 (2013) 1–13 13

capilaridad de piezas para fábrica de albañilería, en hormigón, piedra natural y [57] Lourenço PB, Rots JG. Multisurface interface model for analysis of masonry
artificial, y de la tas. Madrid: AENOR; 2006. structures. J Eng Mech 1997;123:660–8.
[51] Brencich A, Corradi C, Gambarotta L. Eccentrically loaded brickwork: [58] Milani G, Lourenço PB, Tralli A. 3D homogenized limit analysis of masonry
theoretical and experimental results. Eng Struct 2008;30:3629–43. buildings under horizontal loads. Eng Struct 2007;29:3134–48.
[52] Brencich A, Felice GD. Brickwork under eccentric compression: experimental [59] Alfano G, Crisfield MA. Finite element interface models for the delamination
results and macroscopic models. Constr Build Mater 2009;23:1935–46. anaylsis of laminated composites: mechanical and computational issues. Int J
[53] Brencich A, Gambarotta L. Mechanical response of solid clay brickwork under Numer Method Eng 2001;50:1701–36.
eccentric loading. Part I: Unreinforced masonry. Mater Struct 2005;38:257–66. [60] Pluijm R. Material properties of masonry and its components under tension
[54] Attard M, Nappi A, Tin-Loi F. Modeling fracture in masonry. J Struct Eng and shear. In: Neis VV, editor. 6th Canadian masonry symposium. Saskatoon,
2007;133:1385. Saskatchewan, Canada; 1992. p. 675–86.
[55] Brasile S, Casciaro R, Formica G. Multilevel approach for brick masonry walls – [61] Oliveira D. Mechanical characterization of stone and brick masonry. report 00-
Part I: A numerical strategy for the nonlinear analysis. Comput Method Appl M Dec/E-4, Universidade do Minho, Departamento de Engenharia Civil,
2007;196:4934–51. Guimarães, Portugal. Guimarães, Portugal; 2000. p. 1–77.
[56] Brasile S, Casciaro R, Formica G. Multilevel approach for brick masonry walls – [62] Rosas Rodríguez JG, Villegas L, Lorenzo Esperante D. The numerical models
Part II: On the use of equivalent continua. Comput Method Appl M front to the behaviour of masonry elements tested in laboratory. Informes
2007;196:4801–10. Construcc 2008:53.

You might also like