Sie sind auf Seite 1von 18

SIMPLIFIED METHODS FOR ASSESSMENT OF MASONRY

SHEAR-WALLS

Pere ROCA
Professor
Universitat Politcnica de Catalunya
Barcelona
Spain

ABSTRACT

The paper presents a discussion on the possibility of using simple load-path or strut and tie
models, based on equilibrium, to estimate the ultimate capacity of masonry shear-walls.
Tentative rules and specific models are presented for elementary solid walls subjected to
different load conditions. The performance of the suggested models is analysed by comparing
their predictions with available experimental results. Tests carried-out on dry-joint and mortar-
joint, true-scale or scaled laboratory walls are considered for this purpose. Some remarks are
presented on the applicability of the simplified models.

1. INTRODUCTION

Modelling the ultimate condition of reinforced concrete components by means of strut and tie
models, where the struts describe the compression stress fields and ties represent the
reinforcing bars, constitutes a well acknowledged approach accepted by modern concrete codes
and commonly used for practical design and assessment. A comprehensive description of the
technique and possibilities, in the case of reinforced concrete, can be found in [1].

The possibility of using a similar approach for the study of the ultimate response of plain
masonry walls has not yet deserved much attention. Certainly, significant difficulties or
limitations can be envisaged due to the fact that the principles which permit the use of the strut
102 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

and tie models in reinforced concrete may not be entirely applicable to a material of brittle
nature such as un-reinforced clay brick or stone masonry. A specific difficulty comes from the
lack, in this case, of identifiable, discrete tension-carrying components showing plastic
behaviour, such as the reinforcing bars.

However, the use of simple equilibrium methods is not completely unexplored. The working
condition of a shear wall at maximum loading has been, in some occasions, described with the
help of simple schemes involving a set or a continuum of diagonal struts. Ganz and Thrlimann
[2] envisaged parallel or fan compression field models to estimate the ultimate load resisted by
confined shear walls. De Tomasi et al. [3] have recently proposed an approach which considers
the need for ties, in combination with the compression struts, to account for the deviation of the
compression fields in masonry faades.

A more specific proposal, oriented to the analysis of elementary shear walls subjected to a
combination of vertical and horizontal forces, is described in this paper (sections 2 and 3). The
ability of the proposed approach to estimate the ultimate capacity of masonry shear-walls has
been analysed by comparing their predictions with the experimental results obtained for
different walls tested in the laboratory, as described in section 4. These series include full scale
dry-joint walls and one-forth-scaled cohesive walls subjected to different loading conditions.

It must be remarked that the approach presented here constitutes only a first proposal still
requiring further assessment. A more comprehensive calibration, based on both additional
experimental evidence and numerical simulation by means of up-to-date computer methods, is
currently being undertaken.

Even if further calibration is carried out, the simple nature of the models prevent them from
describing significant phenomena observed in plain stone or brick masonry, such as
anisotropy, dilantancy, contact effects, or the acknowledged influence of the size and geometry
of the blocks. Because of these possible limitations, the models proposed are only to provide an
auxiliary tool intended for first-approach calculations. They can not be used for problems
where the deformation of the wall is significant for the determination of the response; more
specifically, the models can not be used to determine the strength contribution of wall panels
confined in concrete or steel frames and subjected to imposed lateral displacements.

2. FEATURES OF ELEMENTARY EQUILIBRIUM MODELS

2.1 Proposed features

Given the very limited tensile strength of the material, the ultimate capacity of shear walls can
only be explained by the generation of diagonal fields of compression stresses in equilibrium
with the external loads. It must be recognised, however, that the geometry of the wall and the
particular loading conditions may not cause pure uniform diagonal fields but more disturbed
ones experiencing significant deviations within the wall (Fig. 1,a); this occurs, in particular, in
the case of a elementary solid wall subjected to vertical and horizontal loading. The deviation
of the main compression stress fields produce inward or outward resulting thrusts which, in
Pere ROCA 103

turn, are balanced with complementary tension or compression internal forces. Where no
confinement exists, the deviation of a compression stress field is only possible if a horizontal
tensile force can be developed within the fabric. Observed damage in walls tested to failure
under a combination of vertical and horizontal load seems compatible with such understanding
of the internal distribution of forces (see section 4).

(a) (b)

tan=tan+c/

(c) (d)

Figure 1: Features of models: (a) deviation of compression stress fields by horizontal tensile
forces; (b) parallel distributed struts; (c) reverse bottle neck struts combined with ties; (d)
limitation of the maximum angle of the strut with respect to the vertical.

The above considerations have been taken into account to envisage possible rules for the
construction of the models and to propose specific simple models. In some cases, it is possible
to conceive models which consist only of struts. In other cases, ties are necessary to explain the
equilibrium or to improve the consistency of the model with the experimental evidence. Based
on these ideas, the following rules are tentatively proposed:
104 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

(1) The models must be as simple as possible to provide practical and efficient approaches.
The number of elements (struts, ties) is to be limited to the minimum amount required to
obtain an acceptable description of the ultimate mechanism.

(2) The struts used in a model describe compression fields covering a certain volume. Because
of that, the effect of a distributed force on a wall is to be modelled as a distributed strut
(Fig. 1, b) or, alternatively, as a set of discrete struts. A minimum amount of two struts
should be used, in any case, to represent the effect of a distributed load on a wall.

(3) Concentrated or partial loads as well as reactions- will cause compression fields to
experience a reverse bottle-neck effect which must be described by means of a
mechanism combining a minimum of two opening struts with a balancing tie (Fig. 1,c).

(4) The maximum slope of a strut with respect to the vertical is limited by the frictional
response of the joints. If the Mohr-Coulomb criteria is adopted to describe the maximum
shear force that can be transferred by the horizontal joints, the slope of the struts with
respect to the vertical is limited to tan, where is the friction angle of the unit-mortar
interface, in the case of a dry-joint wall. In a cohesive wall, the slope of struts is limited to
tan,
c
tan = tan + (1)
n
where c is the cohesion and n is the average vertical compression (Fig. 1,c).

(5) Ties can only be admitted in the horizontal direction given the very low tensile strength of
the unit-mortar interfaces. The tensile force experienced by a tie is resisted thanks to the
combined contribution of friction in joints and tensile strength of the units. The maximum
tensile force T carried by a tie is estimated through the following two conditions:

T Vi tan (2)

T Ab bt (3)

where Vi can be taken as the minimum of the vertical forces carried by the two struts
linked to the tie, Ab is the sectional area of unit courses which are contributing to resist
the tensional force and bt is the average increment of tensile stress which can be resisted
by the units.

(6) Two different types of nodes (or connections between linear elements) can be identified.
The first one is consists of the connection between two struts and a tie, in which a tensile
internal force (T) is anchored by a deviating compression stress field. In that case, the
following condition, also related to the transference of shear forces through the joints,
must be satisfied:
Pere ROCA 105

T V1 tan (4)

where V1 is the vertical force carried by the strut.

(7) The second type of node corresponds to the region where one or more compression forces
converge with a reaction. This type of node is represented as a finite region whose
minimum dimensions are determined by the compression strength of the fabric.

(8) Last, but not least, the mechanisms should be consistent with the evidence obtained from
experiments and micro-modelling, in terms of distribution of stresses, cracking and other
observable or measurable aspects. However, the way mechanisms are to be related to the
observed cracking and distribution of stresses is not obvious.

Section 3 introduces a set of possible models stemming from the above rules. In these models,
both the vertical and the horizontal load are assumed to be applied at the upper edge of the wall
while the self-weight is considered negligible.

2.2 Discussion

The afore mentioned features should not be regarded as a definite set of rules for the
construction of mechanisms, but as a set of initial criteria still requiring refinement based on
additional research. Significant difficulties may be foreseen in any attempt to apply them in a
general manner.

First, rule (4) does not yield complete objectivity since the slope of the struts may depend on
the refinement of the models. Rule (4) is intended for struts representing the main compression
fields. Note that rules (4), (5) -equation(2)- and (6) are in fact equivalent and simultaneously
satisfied in the case of a strut with slope equal to tan connected to a vertical strut and a
horizontal tie.

Equation (3) requires the determination of the area of the transverse section Ab, that is to say,
of the number of unit courses contributing to resist the tensile force corresponding to the tie. It
is suggested, as a first approach, to define Ab as the largest available area centred in the tie, of
which, due to the vertical joints, only one half of the unit courses contribute simultaneously to
carry the tensile force. Note that equation (2) is intended to ensure the friction needed to tie the
courses transversally.

The units experience complex stress states caused by the shear and vertical compression forces
applied to them, to which their role as components of a tie contributes to additional tensile
stress. If the Rankine criterion is adopted to characterize the strength response of the material
of the unit (meaning that the first principal stress is limited to the tensile strength of the
material fbt) the maximum acceptable stress bt can be estimated from:

f bt =
1
2
(
bt n ( bt + n ) 2 + 4 2 ) (5)
106 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

where n and , are the average compression and shear acting on their faces. n, and can be
roughly estimated from the forces supported by the struts which are anchoring the tie.

The difficulty in applying rule (5) lies in the determination of the compressive strength to be
considered in the analysis of the compression node. The effective compression strength may be
strongly influenced by the angle of the resulting compression forces with respect to the
direction of the courses, due to anisotropy, and by the biaxial stress states produced by the
converging struts. Such effects may be particularly significant in walls with small h/b ratio, or
walls made with perforated or hollow bricks.

2.3 Theoretical background

The proposal finds its theoretical justification in the lower bound theorem of plasticity. This
implies that, if an equilibrated mechanism if found, the estimated ultimate load is a lower bound
of the real load causing failure. The possibility of applying the bound theorems of plasticity stems
from the fact that two physical mechanisms determining the ultimate condition in the models,
namely, friction in joints and yielding of the fabric in compression, can be considered plastic.

However, this not applicable in the case of walls for which the ultimate load is determined by
cracking in tension. Nor is it applicable to walls for which crushing of the fabric in
compression appears prior to the full development of the mechanism. In other words, the
mechanism should be chosen in a way that (a) the deformation limit of the fabric in
compression is not exceeded at any point and (b) the tensile forces due to the ties can be
resisted by the fabric. An acceptable ultimate mechanism must be determined by the condition
of maximum friction in joints (i.e., maximum slope of struts) or maximum compression in
nodes. Rule (3), related to cracking of units, consists, in fact, of a requirement for the
applicability of the mechanism. If rule (3) is not satisfied, cracking of units may prevent the
full formation of the envisaged mechanism, meaning that a different mechanism such as a
residual one, see section 3.4- is to be considered. However, partial cracking may be compatible
with this requirement. In any case, severe diagonal cracking is always expectable after the
attainment of the peak loading due to secondary effects caused by the sliding of the units or by
intense compression stresses close to loads and reactions.

According to the lower-bound theorem, if two acceptable models are found, the more realistic
of the two is the one providing the upper ultimate load. When different mechanisms can be
found, all based on the aforementioned rules, the one providing the upper estimation is likely to
be the more realistic one.

3. SUGGESTED MODELS FOR ELEMENTARY SOLID SHEAR WALLS

3.1 Walls subjected to partial or concentrate loading

Several models, all representing reverse bottle-necked compression fields, and corresponding
to different degrees of refinement, are envisaged for the case of walls subjected to partial or
concentrated vertical load combined with horizontal load (Fig. 3).
Pere ROCA 107

The model indicated in Fig. 3,d is preferred because of its simplicity; compared with the, also
simple, model of Fig. 3,c, it yields an upper estimation of the ultimate load.

Mechanism of Fig. (3,b) leads to the following calculation of the ultimate horizontal load H:

H = min{H 1 , H 2 } (6)

2
H 1 = V tan (7)
3

bd m V
H2 = V tan m = (8)
h tf c
where V is the total applied vertical load, b,h and t are the width, height and thickness of the
wall and fc is the compression strength of the masonry; d is the exceeding dimension defined
in Fig 2,c; m is de width of the minimum contact surface required to transfer the vertical load
in plastic regime; tan is defined in equation (1). Equation (6) refers the case of a failure
initiated by the sliding of the units across the horizontal joints, while equation (7) corresponds
to a failure caused by the yielding of the material in compression; in the first case, a stepped
diagonal crack associated to the sliding of units along the joints is to be expected; in the second
case, cracking of units due the tie action may be expected to initiate in the centre of the wall
(where tension stresses are maximum) and then to develop toward the corners of the wall, to
encounter additional cracks caused by the failure of the material in compression.

V e V g g V
H H H

a m=V/(tfc) g g d
a a g=(b-d)/4 a
b

(a) (b) (c)

Figure 2: Models proposed for walls subjected to concentrated or partial vertical loads.
108 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

As mentioned in section 1, the models can not be used to determine the strength contribution of
wall panels confined in concrete or steel frames subjected imposed lateral displacements,
meaning that no possibility exists, based only in equilibrium, to determine the mobilised couple
of V,H forces. However, similar models to those above described may be used to determine the
relationship between the two forces. Models of Fig. 3, producing the same equations than those
previously described, are proposed to determine this relationship in walls confined in the
vertical direction. Model (b) shows satisfactory correspondence with the cracking scheme
observed in real experiments (Fig. 11).

3.2 Walls subjected to vertical uniform loading

The case of walls resisting vertical uniformly distributed loads requires the consideration of
smeared struts arranged according to a parallel or fan distribution. Different envisaged models
are described in Fig. 3. In this figure, thick solid lines are used to represent the struts,
horizontal thin lines are used to represent the ties, and additional thin, discontinuous lines are
included to indicate the stress fields associated to the struts.

Note that in all the models the horizontal load can not be uniformly distributed along the upper
edge of the wall. Model (a) of Fig. 3 -the fan model- recognizes the fact that the slope of the
load paths developed along the wall must vary gradually in order to become compatible with
the geometry of the wall. The width of the bottom distributed node is estimated as m= V/(t fc),
where V is the total vertical load applied. Model (a) does not include a limitation on the
maximum slope of the struts.

Model (b) consists of a modified fan model which complies with the limitation of the
maximum angle of the struts. In the case of narrow walls, or very compressed walls, with

V
m b h tan( ) m= (9)
tf c

it becomes equivalent to the fan model.

Models (c) and (d), Fig. 3, are simple variations of model (b), although more consistent with
the type of failures observed in the experiments; they are acceptable if the fabric is resisting the
tensile force carried by the tie at maximum loading. Note that models (c) and (d) predict a
concentration of compression stresses close to the lateral edge (region A in Fig. 3, see also
Figs. 7 and 9). Alternative and more continuous models, not causing such concentrations of
compression stresses have been also considered; however, they produce lower estimations of
the ultimate load.

Using models (b), (c) or (d), the ultimate horizontal load is estimated as

bm
H =V if m b h tan( ) (10)
2h
Pere ROCA 109

h 1
H = V tan 1 tan if m < b h tan( ) (11)
2b 1 v
where
m V
v= = (12)
b tbf c

fc
fc

(a) (b)

a b-2a a

h/2

fc fc

(c) (d)

Figure 3: Models proposed for walls subjected to uniform vertical load.


110 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

3.3 Walls with openings

Similar mechanisms can be considered for more complex walls or faades with openings or
other geometric alterations. Fig. 4 shows an example of a mechanism describing the internal
forces experienced by a wall with a central opening; in this example, almost no vertical
compression exists close to the top and bottom edges of the opening, meaning that such a
mechanism is only expectable in very cohesive walls; if the ties can not be mobilised, a more
weak, marginal mechanisms similar to those described in section 3.4 is to be considered.

Figure 4: Primary mechanism for a Wall with opening.

3.4 Residual mechanisms

As mentioned in section 2.3, cracking of units may prevent the full development of
mechanisms involving horizontal ties. Whenever cracking of units is predicted (for instance,
using equation (5) or another similar criteria), an alternative, secondary or residual mechanism,
is to be considered. This is illustrated by the example of Fig. 5, consisting of a wall subjected
to concentrated vertical load; if the main tie can no be mobilised meaning that a diagonal
crack C (Fig.5,a), developed through the units, is expected- alternative combinations of
mechanisms such as those shown in Fig. 5,b-c must be considered. In case of direct loading,
the combination of mechanisms A and B may be foreseen to describe the possible response of
the cracked wall; however, mechanism B is not likely to be mobilised due to the deterioration
of the region close to the wall, and only mechanism A may contribute to resist part of the load.
In the case of a wall confined in the vertical direction, subjected to an imposed horizontal
displacement, a secondary resisting scheme, consisting of the combination of mechanisms A
and B (Fig. 5,c) may be envisaged. Mechanisms A and B maintain in this case the central
symmetry of the problem.
Pere ROCA 111

B
B
A
C
A

(a) (b) (c)

Figure 5: Primary (a) and residual mechanisms of vertically free (b)


and confined (c) walls.

4. COMPARISON WITH EXPERIMENTAL RESULTS

4.1 Summary of considered experimental series

The predictions of the above models are compared with experimental results obtained for three
different series of walls tested in the Laboratory of Technology of Structures of the Technical
University of Catalonia. The geometry of the walls, load conditions and properties significant
for the presented models are summarized in table 1.

Table 1: Summary of features of experimental wall series

Series Joint Unit b h t fc tan c Vertical


(n) cm cm cm N/mm2 N/mm2 loading
1(7) dry Sandstone 100 100 20 30 0.66 0 uniform
2(15) mortar Clay brick 30 25 3,5 15 0.8 0.22 uniform
3(20) mortar Clay brick 30 25 3,5 12 0.8 0.20 partial
(n) = number of walls included in the series

The dry-joint walls of series 1 were built with sandstone blocks with dimensions 202010 cm
(Fig. 6). The values considered for tan and fc (see in table 1) where determined in the
laboratory with complementary experiments on units of the same type of sandstone and having
the same dimensions than those used to build the walls; in particular, the value taken for fc was
obtained by testing a specimen consisting of 4 staked blocs. Additional experiments showed
that, in the case of the dry-joint walls tested, the value of fc, thus obtained, was significantly
dependent upon the size of the blocks (Roca et al. [4]).
112 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

Walls of series 2 and 3 were made of on-purpose manufactured small solid clay bricks with
dimensions 72,53512.5 mm and 2.5 mm thick micro-mortar bed joints with measured
uniaxial compression strength of 13.2 N/mm2 for series 2 and 8.6 N/mm2 for series 3. In all
cases, the average compressive strength of the masonry fc, the cohesion c and the angle of
friction where measured by testing specimens made of bricks with the same dimensions than
those composing of the walls. The angle of friction was measured using the simple method
described in [5]; it is intended to improve the measurement of the same parameter by using
Van der Pluijms [6] more accurate procedure.

4.2 Comparison with experiments on dry joint walls with uniform vertical loading

The experiments carried out on dry-joint sandstone unit walls by Oliveira [7] (series 1 of
table 1, see also Roca et al. [4]) have been considered to carry out a first comparison with the
predictions obtained by the proposed simple models introduced in section 3. These experiments
have been already considered by Ordua and Loureno [8] to validate a proposed cap model
for limit analysis of masonry constructions.

Vertical
Load

Horizontal
Reinforced Concrete Beam Load

10
10

100

20
10

10

[cm]
Reaction Slab
20
100

Figure 6: Geometry of walls and loading system (series 1).

The main steps of the testing procedure consisted of the application of a vertical load by means
of a hydraulic actuator kept under force control, resulting in a constant vertical load; and the
gradual application of the horizontal load by small displacement increments by means of a
second hydraulic actuator. The load was applied against the reinforced concrete beam built
over the stone walls. The vertical actuator was provided with a triaxial hinge at each end. The
damaged condition of two of the walls after the attainment of maximum horizontal load is
presented in Fig. 7.
Pere ROCA 113

Figure 7: Walls from series 1 after the experiment.

Fig. 8 shows the comparison between the experimental results and numerical predictions
obtained from equations (10-11). Al the walls tested verified the condition corresponding to
equation (11), meaning that in all cases the failure was due to the sliding of the joints.

120

100
Horizontal load (kN)

80

60

40 experimental
simple model
cap model [8]
20

0
0 50 100 150 200 250 300
Vertical load (kN)

Figure 8: Series 1. Comparison between experimental and predicted ultimate loading.

4.3 Comparison with experimental cohesive walls subjected to uniform vertical loading

The experiments considered here where tested using the same procedure above described for
the dry-joint walls. In the case of series 2, consisting of smaller, one-forth scaled walls, the
tests were carried out in a press machine combined with a horizontal actuator. The damaged
condition of two walls subjected to different vertical loads, after the attainment of maximum
horizontal loading, is shown in Fig. 9.
114 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

Figure 9: Walls from series 2 after the experiment.

Fig. 10 shows the comparison between the experimental load combinations leading to failure
and the corresponding predictions of the simple models. Equations (10-11) corresponding to
any of the models (b-c-d) of Fig. 4 are used.

25

20
Horizontal Load (kN)

15

10
Experimental
Simple model
5
Serie1

0
0 20 40 60 80 100 120 140
Vertical load (kN)

Figure 10: Series 2. Comparison between experimental and predicted ultimate loading.

4.4 Comparison with experimental cohesive walls subjected to partial vertical loading

In the case of series 3 the test was similar to those described in the above sections except for
constraining the rotation of the top stiff beam. Consequently, applying horizontal load caused
the vertical load to acquire a certain eccentricity which, in the ultimate condition, can be
assumed to be the maximum possible one.
Pere ROCA 115

The damaged condition of two walls subjected to a different vertical load, after the attainment
of maximum horizontal loading is presented in Fig. 12.

Due to the expectable partialization of the contact length between the wall and the top and
bottom elements (plate, top beam) which confine it the walls in this series are treated as walls
subjected to concentrate or partial vertical load, according to section 3.1, with maximum
compatible eccentricity. The ultimate horizontal force H is estimated with equations (6-8).

Fig. 13 shows the comparison between the experimental vertical-horizontal load combinations
leading to failure and the corresponding predictions of the simple models. The obtained
predictions agree satisfactory both in the region determined by the sliding of joints (equation 7)
and that governed by the yielding of the fabric in compression (equation 8).

Figure 11: Walls from series 3 after the experiment.

40
35
Horizontal load (kN)

30
25
20
15
10 Simple model
Experimental
5
0
0 20 40 60 80 100 120 140
Vertical load (kN)

Figure 12: Series 3. Comparison between experimental and predicted ultimate loading.
116 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

5. FINAL REMARKS

The possibility of using simple equilibrium models for the assessment of the in-plane ultimate
response of shear masonry walls, such as those resulting from load-path or strut and tie
combinations, has not yet deserved significantly attention. Certainly, the principles which
encourage their application to other structural materials such as reinforced concrete- are not
so neatly present in the case of plain masonry shear walls.

Aiming to contribute to the possible use of simple equilibrium models for the assessment of
shear walls, a tentative set of rules for the construction of ultimate mechanisms has been
proposed here. Additionally, some particular models have been suggested as a first solution for
elementary cases of solid walls subjected to combined vertical and horizontal external loads.
The predictions obtained using these models agreed acceptably with some available
experimental measurements. However, such proposals are not the only possible, and the
opportunity to envisage more realistic and accurate solutions remains fully open.

The study carried-out suggests that the use simple models for first-approach calculations may
be feasible. Such models may show a certain ability to predict ultimate loading conditions, in
spite of neglecting important phenomena related to the real observed in-plane mechanical
response of plain masonry walls.

A further development of the concepts discussed should be based on significant additional


experimental evidence combined with systematic numerical simulation. In particular, detailed
micro-modelling is regarded as a very promising means to progress in the development of a
more consistent technique of analysis based on simple equilibrium models.

6. ACKNOWLEDGEMENTS

The experimental studies presented here were developed within the research project ARQ2002-
04659, funded by DGE of the Spanish Ministry of Science and Technology, whose assistance
is gratefully acknowledged.

7. REFERENCES

[1] Schlaich, J., Schfer, K., Jennewein M. - Toward a consistent design of structural
concrete. PCI Journal, Vol 32, N 3, 1987, p.72-150.
[2] Ganz, H. R., Thrlimann, B. Strength of brick walls under normal force and shear. Proc.
8th Int. Symposium on load bearing brickwork, London, 1983, p 27-29.
[3] De Tommasi, G., Monaco, P., Vitone, C. - A first approach to the load-path method on
masonry structure behaviour. Structural studies, repair and maintenance of heritage
architecture VIII, WITpress, Southampton, 2003, p. 287-296.
Pere ROCA 117

[4] Roca, P., Oliveira, D., Loureno, P., Carol, I. - Mechanical response of dry joint
masonry. Studies in Ancient Structures, Yildiz Teknik Universitesi, Istanbul, 2001,
p. 291-300.
[5] Ghazali, M. Z., Riddington, J. R. Simple test method for masonry shear strength. Proc.
Instn. Civ. Engrs, Part2, N 85, 1988, p. 567-574.
[6] Pluijm, R., Van Der Shear behaviour of bed joints. Proc. 6th North American
Masonry Conf, Drexel University, Philadelphia, Pennsylvania, USA, 1993, p. 236-136.
[7] Oliveira, D. V. Mechanical characterization of stone and brick masonry. Rep. No. 00-
DEC/E-4, Univ. do Minho, Guimaraes, Portugal, 2000.
[8] Ordua, A., Loureno, P. Cap model for limit analysis and strengthening of masonry
structures. Journal of Structural Engineering, V. 129 N 10, 2003, p.1367-1375.
118 SSMICA 2004 - 6 Congresso Nacional de Sismologia e Engenharia Ssmica

Das könnte Ihnen auch gefallen