Вы находитесь на странице: 1из 14

Computers and Structures 160 (2015) 4255

Contents lists available at ScienceDirect

Computers and Structures


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

Numerical limit analysis of steel-reinforced concrete walls and slabs


A.A. Pisano , P. Fuschi, D. De Domenico
Dept. PAU, University Mediterranea of Reggio Calabria, via Melissari, I-89124 Reggio Calabria, Italy

a r t i c l e

i n f o

Article history:
Received 23 January 2014
Accepted 3 August 2015
Available online 22 August 2015
Keywords:
Reinforced concrete walls
Slabs
Limit analysis
Peak load
Collapse mechanism
Large-scale prototypes

a b s t r a c t
A limit analysis based design methodology is hereafter proposed and applied for the peak load evaluation
of steel-reinforced concrete large-scale prototypes of structural walls and slabs. The methodology makes
use of nonstandard limit analysis and predicts the peak load multiplier of the analyzed structures by
detecting an upper and a lower bound to it. Some useful hints on the collapse mechanism the structure
will exhibit at its limit state is also attainable. To check the reliability of the numerically detected peak
loads and failure modes a comparison with experimental laboratory findings, available for the large-scale
specimens considered, is presented.
2015 Elsevier Ltd. All rights reserved.

1. Research motivations and introduction


The present study follows a very recent paper by the authors
[30], dealing with the possibility to predict the peak load and the
failure mechanism of steel-reinforced concrete elements. Such a
possibility, explored in [30] with reference to a standard benchmark on steel-reinforced concrete beams under bending, is investigated here with reference to reinforced concrete (RC) structures of
practical and greater engineering interest, namely: walls and slabs.
The addressed topic belongs to a wider ongoing research programme, started by the authors in [29] in the context of RC structures, but already applied with success to structural elements
made by different constitutive materials, such as laminates of fiber
reinforced polymers, see e.g. Pisano and Fuschi [26,27], Pisano et al.
[28]. The proposed methodology, giving information at an ultimate
(collapse) state of the structure in terms of peak load and collapse
mechanism, can be viewed as a useful preliminary design tool also
for RC-structures of large dimensions. If necessary, more accurate
step-by-step analysis, able to follow the fracturing/damaging processes exhibited by RC-structures in the post-elastic regime, can be
carried out. Such a deeper investigation can however be reserved
only to confined zones or weaker structural elements detected by
a much simpler limit analysis which, as shown hereafter, gives useful hints on the collapse mechanism and predicts, with good accuracy, the ultimate value of the loads acting on the located weaker
members or parts.

Corresponding author.
http://dx.doi.org/10.1016/j.compstruc.2015.08.004
0045-7949/ 2015 Elsevier Ltd. All rights reserved.

The numerical analysis employed here is based, essentially, on


the application of non standard limit analysis theory (Lubliner
[21]). The peak load of a structure made by a non standard material
as concrete (where non associativity is postulated to account for its
dilatancy) can indeed be located between an upper and a lower
bound to it. There are many examples of limit analysis in the realm
of nonstandard materials, from the pioneering works of Drucker
et al. [9] and Radenkovic [33], to studies concerning geotechnical
problems, e.g. Sloan [37], Boulbibane and Ponter [5], or those
specifically dealing with reinforced concrete, see e.g. Liman et al.
[20], Larsen et al. [16]. A comprehensive and updated review of
limit analysis methods can be found in Nielsen and Hoang [25]
in the field of concrete plasticity or, in the wider context of the
so-called Direct Methods, in the very recent book by Spiliopoulos
and Weichert [38]. On the other hand, the application of plasticity
based approaches to reinforced concrete structures, whose ductile
behavior is assured by the presence of reinforcement, is witnessed
by a number of contributions, see e.g. the monographs of Chen
[7,8] and, again, Nielsen and Hoang [25] or the papers by Brisotto
et al. [10], Roh et al. [34], Zhang and Li [41], Benkemoun et al.
[3], Carrazedo et al. [6], Wu and Harvey [40], just to quote few very
recent contributions on this research theme being the list far to be
exhaustive.
There are two finite element (FE) based iterative procedures
promoted by the authors for limit analysis of RC-structures, the linear matching method (LMM), see e.g. Ponter and Carter [31],
Pisano et al. [29], and the elastic compensation method (ECM),
see e.g. Mackenzie and Boyle [22], Pisano et al. [30]. In the above
couples of references, the former paper is the one where (with reference to structures made of von Mises-type materials) the

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

method was conceived and first proposed, the latter that where the
method was rephrased to deal with reinforced concrete structural
elements. In particular, in Pisano et al. [29] the LMM has been
reformulated and adapted to a MentreyWillam-(MW)-type
yield criterion (Mentrey and Willam [24]) focusing all the theoretical aspects, the mathematical and geometrical details with reference to a 3D formulation in the HaighWestergaard coordinates.
A few examples are presented there to show the applicability of
the method to reinforced concrete simple elements. In Pisano
et al. [30], while the effectiveness of the LMM is investigated by
analyzing a standard benchmark on steel-reinforced concrete
beams under bending (Bresler and Scordelis [4], Vecchio and Shim
[39]), a revisited version of the ECM suitable for the MW-type
yield criterion is proposed. In the above paper the use of the two
methods was applied for the first time to simple reinforced concrete structures, namely beams, showing the possibility to locate
the real (experimentally detected) value of the peak load by computing an upper and a lower bound to it.
In the present study, skipping the theoretical details given in
[29,30] to avoid repetition, the above mentioned numerical FE procedures are applied to compute the peak load as well as to predict
the failure mechanisms of large-scale RC-prototypes of walls and
slabs. The relevant experimental data, concerning tests carried
out up to collapse and available for the analyzed specimens, have
been used to validate the numerical predictions so facing real
experimental findings. The following papers/reports have been
considered: Lefas et al. [18], where structural walls were tested
under combined action of a constant axial and a horizontal load
monotonically increasing to failure; El Maaddawy and Soudki
[11], where simply supported one-way RC-slabs were tested to
failure under four-point bending; Sakka and Gilbert [35,36], where
simply, continuous and corner supported square and rectangular
slabs subjected to line or point loads increasing to failure were
tested. Some of the results of these latter tests are also reported

43

in more recent publications of the same Authors to which the


Reader can refer [13,14].
Some information on the followed nonstandard limit analysis
approach as well as on the LMM and ECM are given in the next Section 2 where the key ideas of the iterative FE numerical schemes
are explained with the aid of two geometrical sketches. Details of
the computational steps are given in Appendices A and B for
completeness. Section 3 addresses the geometry, the material data,
laboratory fixtures and the loading conditions of the analyzed
RC-prototypes. The adopted mechanical model, FE meshes,
modeling hypotheses are also expounded in this Section closing
with a comparison between the obtained results and the
experimental findings. Concluding remarks are given at closure
in Section 4 also outlining possible future developments.

2. Key ideas of the numerical approach and computational


schemes
The key point of the promoted nonstandard limit analysis
approach (see e.g. [33]) is to encircle the yield surface of a given
nonstandard material with two surfaces, precisely an outer and an
inner surface and to compute, with reference to such surfaces
(referred to two standard materials), an upper and a lower bound
to the real collapse load multiplier pertaining to the nonstandard
material structure under consideration. Concrete is herein modeled
as a nonstandard material obeying to a MW-type yield surface
which can play the double role of inner and outer surface in the
sense specified above. The MW-type yield surface endowed with
cap in compression is shown schematically in Fig. 1. Steel reinforcement, on the other hand, are considered of an infinitely elastic
behavior. Their presence is taken into account only for what concern the confinement effect they exert on concrete. Such effect
injects a ductile behavior on the RC-element that is it confers to

Fig. 1. Adopted MentreyWillam-type yield surface with cap: (a) deviatoric sections at three generic values of hydrostatic pressure; (b) tensile and compressive meridians
in the Rendulic plane at h 0 and h p=3, respectively; and (c) 3D sketch in the principal stress space.

44

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

the element an essential requisite for the present plasticity-based


approach.
The LMM is an iterative procedure belonging to a kinematic
approach of limit analysis and involving one sequence of linear
FE-based analysis. At each iteration say #k  1 (i.e. at each FEi
i , with Pk1 load multiplier and p
analysis under loads P k1 p
assigned reference loads), the elastic moduli as well as some given
initial stresses are modified within the structure assumed by
hypothesis as made of a linear viscous fictitious material. The adjective fictitious stands for the circumstance that the material may
have elastic parameters which assume different values at different
points. The latter being Gauss points (GPs) in a FE discrete model of
the structure. By doing so the fictitious solution computed in terms
of linear strain rate at the generic GP, say e_ k1 , (together with the
k1
related compatible displacement rates u_ i
) can be interpreted as
strain rate (with related displacement rate) at collapse for the real
structure. Basically, the above updating of elastic moduli and initial
stresses allows to build a collapse mechanism for the structure in
terms, in this context, of volumetric and deviatoric strain rate
components and displacement rates, say e_ cv , e_ cdx , e_ cdy , u_ ci , where the
apex c stands for a value at collapse. For a better understanding
of the method it could be very useful to look at its geometrical
interpretation in the principal stress space. In Fig. 2, on the basis
of the formal analogy between the linear viscous problem
and the linear elastic problem, the fictitious linear solution
computed at the generic GP can be located at the stress point
PL on the complementary energy equipotential surface
k1 ; q
 k1
 k1
W k1 n; q; K k1 ; Gk1 ; n
;q
 W k1 const: perx
y
taining to the fictitious material. The surface is indeed a prolate

spheroid (see [29] for details) whose semiaxes, say at iteration


k  1, are 4Gk1 and 6K k1 i.e. they are related to the fictitious
elastic moduli. The spheroid has also its center at 
nk1 ,
k1
k1
 y i.e. located by given fictitious initial stresses. Here
x
q
; q
only a portion of W k1 W k1 is represented for simplicity.
The updating of the fictitious moduli to values K k ; Gk and of
k , q
 k
 k
the initial stresses to values n
x ; q
y is carried out in such
a way that the spheroid is modified in shape and position, or, in
other words, the complementary energy surface of the fictitious
material is modified so that the fictitious linear solution PL is
brought onto the MW-type yield surface, precisely onto the point
PM of normal e_ k1 . It is worth noting that if the yield surface is
strictly convex (as the one here assumed) the point PM is uniquely
determined by the given normal e_ k1 . Such outward normal at PM
can then be viewed as an e_ c k1 , i.e. as a strain rate at collapse, the
related displacement rates u_ c being those pertaining to the collapse
mechanism while stress coordinates of PM ; nY k1 ; qx
; qy

are the associated stresses at yield. All the ingredients to compute


Y k1

Y k1

an upper bound multiplier, say PUB , are known and the following
relation holds true:

R 
k
PUB

nY k1 e_ v

c k1

qx
R

Y k1 c k1
_d
x

@V t

i u_ ci
p

qy

d@V

Y k1 c k1
_d
y

dV

As it appears from the discussed geometrical sketch the


W k W k1 and the MW-type yield surface match at PM , so
the name of the method. Moreover the above stress at yield,
computed through the matching, do not meet the equilibrium

Fig. 2. Geometrical sketch, in the principal stress space, of the matching procedure fulfilled at the generic Gauss point within the current element from iteration
#k  1 to #k.

45

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

i and the procedure is carconditions with the acting loads P k1 p


ried on iteratively until the difference between two subsequent
PUB values is less than a fixed tolerance. Convergence requires that
the W k W k1 matches MW-type yield surface at PM and
otherwise lies outside the yield surface (see [32]).
On the other hand, the ECM is aimed at constructing an admissible stress field suitable for the evaluation of a lower bound to the
peak load multiplier, according to the static approach of limit analysis theory. Also the ECM, like the LMM, acts in an iterative way,
but involving many sequences of linear FE-based analysis. At each
i , with P s load
sequence, defined by a given load value, say Ps p
multiplier of the current sequence, the Young modulus is reduced
within highly loaded regions of the structure, i.e. in those elements
#e where elastic stress is greater than the yield one.
Once again, a geometrical sketch is used to explain the procedure. In Fig. 3, at iteration k  1 of the current sequence, an elastic stress response evaluated within a generic finite element #e (it
is an averaged value of the stresses computed at the GPs of the eleFig. 3. Geometrical sketch, in principal stress space, of the ECM fulfilled within the
generic element in the FE discrete model at iteration #k  1 of the current
sequence.

e k1

Y k1

ment #e) is represented by point P#e . In the same sketch P#e


is the corresponding value at yield represented by the stress point
e k1

lying on the same meridian plane of P#e

, i.e. at h he , located by

Fig. 4. Steel-reinforced concrete walls: (a) mechanical model, geometry, boundary and loading conditions; (b) typical FE-mesh of 3D solid elements for concrete and 1Dembedded-truss elements for re-bars; (c) reinforcement arrangement for walls type 1 square-shaped; and (d) reinforcement arrangement for walls type 2 rectangular.

SW22

SW24

2.27
2.47
2.99

2.92
3.20
3.57

SW16

SW17

2.09

SW23

SW15

PLB
PEXP
PUB

1.16
1.23
1.32

SW21

SW14

3.35
3.55
3.94

4.15

SW13

1.76
1.50

SW12

1.83

SW11

2.23
2.65
2.83

230
355

185
460

182
343

325

1.53

34.4
34.7
31.9
32.3
32.5
34.3
33.6
32.4
34.1
33.5
33.6
32.9
29.2

1.20

2.20
2.23
1.94
1.97
2.00
2.19
2.11
1.99
2.16
2.10
2.11
2.04
1.67

3.21
3.30

F V kN

44.46
45.56
34.51
35.79
36.81
43.95
41.06
36.38
43.01
40.63
41.06
38.25
25.59

1.56
1.80
2.05

Ec GPa

SW11
SW12
SW13
SW14
SW15
SW16
SW17
SW21
SW22
SW23
SW24
SW25
SW26

3.00
3.40
4.03

f t MPa

1.28
1.50
1.71

f c MPa

2.60

Specimen

3.30

Table 1
Steel-reinforced concrete walls: specimen number; compressive and tensile concrete
strengths; concrete Young modulus; constant value of the applied vertical load.

3.81

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

1.21
1.27
1.48

46

!e k1
the intersection of the stress vector OP #e
with the MW-type
Y k1

yield surface. By hypothesis, in the sketch of Fig. 2, P#e

is low-

e k1
P#e .

By the ECM, at the current iteration k  1 of the


er than
current sequence s, the elastic modulus of element #e has then to
be reduced according to the formula:

E#e

2
32

 !Y k1
OP #e 

7
k1 6
: E#e 4
5 ;

 !e k1
 OP #e 

(a)

20

#SW16

3. Analyzed large scale RC-prototypes: walls and slabs


As pointed out in Section 1, the main goal of the present work
is to verify the effectiveness and reliability of the promoted

PEXP
PLB

12

3.94

3.55

3.35
0

10

12

14

16

iteration number

(b)

2.5

PUB

#SW26

PEXP

2.0

load multiplier P

where referring again to Fig. 3, PR has to be intended as the point


Pe#e farthest away from the MW-type yield surface, PYR being the
corresponding stress point at yield.
The possibility for applying the discussed limit analysis procedures either to concrete, obeying to a MW-type yield criterion,
and to steel-rebars, obeying to a von Mises-type yield condition,
within a layered FE-based limit analysis approach is actually the
object of an ongoing research. Owing to the different FE types
adopted for the two materials within the proposed FE formulation,
two simultaneous yield criteria could be considered in the iterative
procedure, one for concrete elements and one for steel elements.
This would imply a simultaneous updating of the material properties of both concrete and steel elements in the iterative procedure,
so enabling the prediction of possible steel bar yielding at incipient
collapse. The latter is indeed oriented to include those collapse
mechanisms characterized by reinforcement yielding not treated
in the following.

PUB

16

load multiplier P

where
is the modulus to be used at next iteration and the square
of the reducing factor accelerates the procedure (see e.g. [27]). The
described operation redistributes the stresses within the structure and allows to define a maximum admissible stress value in
the whole structure for the given load. Increased values of loads
are then considered in the subsequent sequences of analysis until
further load increase does not allow the maximum stress to be
brought below yield by the reduction (or redistribution) procedure.
A lower bound multiplier can be eventually evaluated at last
admissible stress value attained for a maximum acting load, say
s
i , in the shape:
PD p

PLB

SW26

Fig. 5. Steel-reinforced concrete walls: values of the upper P UB and lower P LB


bounds to the peak load multiplier against the experimentally detected one P EXP .

k
E#e

s
 ! k1
PD


 OP YR 
 ! k1 ;


 OP R 

SW25

PLB
1.5

1.32

1.23

1.16

1.0

0.5

0.0
1

10

12

14

16

18

iteration number
Fig. 6. Steel-reinforced concrete walls. Values of the upper P UB and lower P LB
bounds to the peak load multiplier versus iteration number: LMM prediction, solid
lines with square markers; ECM prediction, solid lines with triangular markers;
collapse experimental threshold (after Lefas et al. [18]) dashed lines. (a) Specimen
SW16 and (b) specimen SW26.

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

procedure tackling two types of steel-reinforced concrete structural elements of engineering interest, specifically: walls and slabs.
Being aware of the limitations congenital to a plasticity-based
approach it is worth noting that the ductile behavior, which is an
essential requisite for the adoption of limit analysis, is actually
assured by the presence of steel bars which mitigate, or even
nullify, many complex post-elastic phenomena exhibited by plain
concrete at incipient failure such as localization and/or fracturing/
damaging mechanisms. These phenomena, due to a mainly brittle
behavior, cannot be treated with the present procedure which has
to be confined to ductile steel-reinforced concrete structures.
Indeed, the RC-structures of common use in civil engineering
applications addressed here luckily belong to the above category
as the selected experimental tests analyzed below.
Reference is made to experimental findings taken from the relevant literature; in particular thirteen large-scale walls, tested up to
failure by Lefas et al. [18], and seven large-scale slabs, tested up to
failure by Sakka and Gilbert [35,36], have been considered. The
experimental tests, in practice, have been numerically reproduced
to predict the peak load as well as the failure mechanism of each
structure and the obtained numerical results compared with those
given by the laboratory tests. It is worth noting that the prediction
of the failure mechanism, which for the examined cases is quite

47

obviousa flexural mode of failure is mainly to be expectedhas


to be intended rather as the capacity of the methodology to localize
numerically the zones where mechanism starts and spreads at
collapse.
The elastic analysis performed within the two summarized iterative procedures have been carried out using the FE code ADINA,
(ADINA [1]), with meshes of 3D-solid 8-nodes elements with
2  2  2 GPs per element for modeling concrete and embedded
2-nodes, 1-GP, truss elements utilized for modeling re-bars and
stirrups. Such embedded truss elements (refer to ADINA for
details) are 1D FEs connecting the intersections of the rebar axes
with the faces of the 3D solids concrete elements. Such intersections are generated nodes on the 3D solid FEs faces constrained
to the three closest corner nodes of the 3D element itself. A perfect
bond between concrete and steel reinforcement, assumed as said of
an indefinitely elastic behavior, has been considered. The number
of finite elements is different for each specimen type and is chosen
after a preliminary mesh sensitivity study to assure an accurate FE
elastic solution. Nodal loads, equivalent to the distributed load
exerted by rigid plates in the laboratory fixture, are considered.
Boundary conditions (constraints) consistent with those of the
experimental tests will be specified next for each specimen
type. For what concerns the choices related to the assumed 3D

(a)

(b)

y
x

(c)

(d)

y
z

y
x

Fig. 7. Steel-reinforced concrete walls. Band plots of the Cartesian strain rate component e_ cy in the deformed configurations at the ultimate value of the horizontal load P UB F H
for specimens SW16, (a) and (b), and SW26, (c) and (d). In particular: (a) and (c) show the results obtained at last converged solution of the LMM on the fictitious structure
localizing the plastic zone and/or the collapse mechanism; (b) and (d) show the results pertaining to an elastic solution of the real structure, i.e. with the real elastic parameters.

48

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255


0

concrete constitutive model, the uniaxial compressive strength, f c ,


has been taken from the quoted references; the uniaxial tensile
0
strength, f t , when not available, has been assumed as
q
0
0
f t 0:33 f c as suggested by Bresler and Scordelis [4]; the value
of the eccentricity parameter e of the MW-type yield surface
0
0
0
0
has been evaluated by the expression e 2 f t =f c =4  f t =f c  as
0
0
suggested in Balan et al. [2], the f t =f c ratio being assumed as a measure of the material brittleness. Other three values have been
finally fixed to locate the MW-type yield surface in the principal
p 0
stress space, namely: nv assumed as nv 3f c =m with m given by
02

02

m : 3 f c  f t e=f c f t e 1, see Pisano et al. [29]; na 0:7923 f c


0
and nb 1:8964 f c as suggested by Li and Crouch [19].
Finally, a Fortran main program has been used to drive both the
iterative procedures here proposed updating, at each GP of each
element, the fictitious elastic parameters when performing the
LMM or realizing the redistribution procedure within the ECM.
3.1. Walls
The large-scale wall models, addressed by Lefas et al. [18], were
tested under the combined action of a constant axial (vertical) and
a horizontal load monotonically increasing to failure. The experiments were oriented to investigate the effect on wall behavior of
the height-to-width ratio, the axial load, the concrete strength
and the amount of web horizontal reinforcement also performing
a critical examination of some concepts underlying the ACI Building Code provisions for the design of reinforced concrete structural
walls in force in the early eighties. The attention, for obvious reasons, is hereafter focused exclusively on the experimental data
and results given in the above quoted paper where details on the
test set up are given.
The assumed mechanical model reporting geometry, loading
and boundary conditions of the analyzed walls is sketched in
Fig. 4(a) where, besides the geometrical dimensions L; t and H

specified for the two types of walls shown in Fig. 4(c) and (d), F H
denotes the horizontal reference load, assumed equal to 100 kN
and amplified by the load multiplier P; F V is the vertical constant
(fixed) load whose value is given, for each specimen, in Table 1
(f V being the equivalent distributed load). As specified in the referenced experimental paper all the walls were monolithically connected to an upper and a lower reinforced concrete beam both
working as cages for the anchorage of the vertical bars. The upper
beam was also the rigid element through which axial and horizontal loads were applied to the wall specimens, the lower beam, simulating a rigid foundation, was used to fix the specimen to the
floor.
The model of Fig. 4(a) takes into account such experimental fixture assigning zero displacements to the points belonging to the
shaded bottom cross-section of the wall and zero relative horizontal displacements (x and z directions) to the points belonging to the
shaded top cross-section. The horizontal load P F H is then applicable to any point (or FE node) of such horizontal (rigid) top crosssection. Fig. 4(c) and (d) specify, as said, specimens type 1 and type
2, respectively and, besides their geometrical dimensions, give the
reinforcement arrangement made of vertical and horizontal
high-tensile steel bars of diameters equal to 8 and 6.25 mm,
respectively. Stirrups of mild steel bars of 4 mm diameter were
also present as additional horizontal reinforcement to confine the
wall edges. Table 1 completes the needed data giving the concrete
material parameters for each specimen labeled borrowing from
Lefas et al. [18]. In particular, for concrete the Young modulus, Ec ,
0:3

has been evaluated as Ec 22; 000 f c =10 MPa, following Eurocode 2 [12] while a Poisson ratio of m 0:2 has been assumed for
all the specimens. E 200 GPa has also been assumed for the steel
reinforcement. Finally, in Fig. 4(b) the FE mesh scheme adopted in
the analysis is shown with 3D-solids for concrete (616 and 624 for
specimens type 1 and type 2, respectively) and 1D-embedded
trusses (686 and 718 for specimens type 1 and type 2, respectively). The number of truss elements reduced to 446 and 502 for

P p1

(a)
P p2

P p1

z
x

L2

Ly

L3

L1
L1

L3

Lx

L2

(b)

(c)
z

sy
15 mm

t
15 mm

barx

sx

15 mm

bary

15 mm

Ly
Fig. 8. Steel-reinforced concrete simply-supported slab #1 (which is the control specimen of El Maaddawy and Soudki [11]) and slab #2 (coincident with specimen SS4 of
Sakka and Gilbert [35]): (a) mechanical model, geometry, boundary and loading conditions; (b) reinforcement arrangement along x direction; and (c) reinforcement
arrangement at cross-section.

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

specimens SW17 and SW26, respectively, where a lower percentage of horizontal bars was used.
The obtained results are summarized in Fig. 5 where for each
analyzed wall specimen, are given: the peak load multiplier values
detected experimentally (after Lefas et al. [18]) and the predictions,
in terms of upper and lower bound values to such peak, furnished
by the present approach.
By inspection of the numerical findings, in three over the
thirteen examined cases (precisely walls SW11, SW24, SW25)
the experimental peak load value lies outside the range numerically predicted, such cases are wrong predictions of the adopted
methodology. However, it is worth noting that the error on wall
SW25 could indeed be due to a trouble which occurred during
the laboratory test and which had been reported by the Experimenters, that might have invalidated the experimentally detected
peak load value. To be precise, an unintended eccentricity of the
vertical and horizontal loads occurred due to an unexpected shift
of the fixture transferring the axial load on the wall. By excluding
specimen SW25 the percentage of wrong predictions on the peak
load was reduced from about 23.08% to about 16.67% which is
an acceptable value from an engineering point of view when
dealing with practical (real) structures. On the other hand, the
percentages of error between the experimental peak load value
and the predicted lower bound to it for walls SW11, SW24 and
SW25 range from a maximum of 27.5% to a minimum of 17.3%.
If it is taken into account that the promoted procedure can be
viewed as a preliminary design tool at ultimate limit states, the

49

above percentages are fully compensated by the safety factors


on acting loads and material strengths required by design rules
(see e.g. Eurocode 2). In all other cases the predicted interval,
within which the experimentally detected value falls, is quite narrow showing a good performance of the proposed methodology.
Fig. 6 shows, for two of the analyzed specimens (i.e. one for each
type, precisely: SW16 and SW26), the plots of the upper and the
lower bounds to the peak load multiplier versus the iterations
number.
Analogous results are obtained for all the other cases but are
omitted for sake of brevity. In all cases a monotonic and rapid convergence of both methods is exhibited.
A deeper comprehension of the mechanical behavior of the
walls at collapse can be gained by the prediction of the walls failure modes. As said, the LMM builds the collapse mechanism the
structure exhibits when the loads attain their peak value or, more
exactly, they reach the evaluated upper bound value to such peak.
Moreover, such a mechanism is built on a fictitious structure i.e. it is
located within the analyzed structure made, by hypothesis, of a
material endowed with a fictitious spatially varying distribution
of elastic parameters and initial stresses.
A possibility to predict the walls failure mechanism is then
given by the possibility to point out the plastic zone (collapse
mechanism) at last converged solution of the LMM. To this aim
the plots of the displacement rates (i.e. the final deformed configuration), as well as those of the Cartesian strain rate component e_ cy
have been considered on the walls loaded by P UB F H (plus the

Fig. 9. Steel-reinforced concrete continuous-supported slab #3 (coincident with slab CS5 of Sakka and Gilbert [35]): (a) mechanical model, geometry (all dimension in mm),
boundary and loading conditions; (b) reinforcement arrangement at roller support and at mid-span along longitudinal axis; (c) reinforcement arrangement at roller support
and at mid-span at cross-section; (d) reinforcement arrangement at interior support along longitudinal axis (top bars along x have a length of 1800 mm and are centered at
midspan); and (e) reinforcement arrangement at interior support in the cross-section.

50

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

Fig. 10. Steel-reinforced concrete two-way corner-supported slabs #4; #5; #6 and #7, coincident with specimens S2S-5, S2S-6, S2R-4 and S2R-5 of Sakka and Gilbert [36]:
(a) mechanical model, geometry (all dimensions in mm), boundary and loading conditions; (b) reinforcement arrangement along x direction; and (c) reinforcement
arrangement along y direction.

pertinent constant F V ) and at the last distribution of fictitious


parameters and initial stresses.
Fig. 7(a) and (c) show, again only for specimens SW16 and
SW26, the Cartesian strain rate component e_ cy distribution in the
deformed (final) configuration attained by LMM at convergence.
The plasticized zones so located appear sufficiently confined and
reasonably close to the damaged zones experimentally detected.
Also the deformed shapes, particularly that of the slender (type
2) specimen (see again Fig. 7(c)), show clearly how around such
plasticized zones the remainder structure rotates rigidly exhibiting
a collapse/failure mechanism. The distribution of e_ cy on the
deformed configuration located on the fictitious structure of the
LMM, as the ones given by Fig. 7(a) and (c), that is the collapse
mechanism the method builds seems indeed to give some reliable
hints on the expectable failure mode of the structure but, obviously, only from a qualitative point of view. The level of detail in
describing the state of incipient collapse even if not exhaustive
can however be useful to localize critical zones or weaker members
for example within reinforced concrete structures of larger dimensions. A prediction, even if qualitative, of the plasticized zones
locating the collapse mechanism cannot obviously be obtained if
a simple elastic solution of the real wall (i.e. with the real material
parameters) loaded by P UB F H is carried out, as shown by the plots
given in Fig. 7(b) and (d).
3.2. Slabs
Three types of large-scale steel-reinforced concrete slab specimens have been analyzed for an overall number of seven specimens: two simply-supported, one continuous-supported and four
corner-supported. The analyzed specimens have been selected from

a wider survey of experimental tests. The simply-supported slabs,


hereafter named slab #1 and slab #2 as well as the continuous
supported, slab #3, are those tested by El Maaddawy and Soudki
[11] and Sakka and Gilbert [35] (see also [13,14]). The first paper
was oriented to verify the use of mechanically-anchored unbonded
fiber reinforced polymer system to upgrade reinforced concrete
slabs. This goal is out of the present study but, among the six slabs
there tested to failure under four-point bending, one was used as a
control specimen, i.e. without any fiber reinforced polymer
strengthening system and it has been here chosen as slab #1.
The second quoted paper/report analyzes eleven slabs, four slabs
were simply supported and seven were continuous over two equal
spans; two classes of steel reinforcement were also considered:
low ductility and normal ductility. The experimental work was carried on to investigate on the effect of reinforcement ductility on the
strength and failure modes of such eleven one-way reinforced concrete slabs. The slabs with low ductility reinforcement showed to
fail by brittle fracture of the reinforcement (see e.g. [35]). Such
non conventional flexural mode of failure is out of the predictive
capacities of the present approach.
The only two samples, among the eleven, reinforced with
normal-ductility steel bars (exhibiting a ductile mode of failure)
have then been chosen: the simply supported SS4, named slab
#2 hereafter, and the continuous supported CS5, named slab #3
in the following. Finally the four corner-supported specimens
hereafter analyzed are taken from the paper/report by Sakka and
Gilbert [36], where the strength and ductility of two-way cornersupported reinforced concrete slab panels containing low- and
normal-ductility bars were investigated by laboratory tests up to
failure. Once again, among eleven slabs, six square and five rectangular, the four containing normal-ductility bars have been chosen,
namely the square shaped S2S-5 and S2S-6, here denoted as slab

51

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

Table 2
Steel-reinforced concrete simply-supported slabs #1 and #2 sketched in Fig. 8: specimen number; geometrical data, bars diameters and spacing, value of the applied reference
load.
Slab specimen

L1 (mm)

L2 (mm)

L3 (mm)

Lx (mm)

Ly (mm)

t (mm)

#1
#2

250
500

500
500

150
250

1800
2500

500
850

100
106

bar x (mm)

sx (mm)

bar y (mm)

sy (mm)

1 (N/mm)
p

2 (N/mm)
p

11.3
12

235
410

100

117.65

#1
#2

12

200

Table 3
Steel-reinforced concrete corner-supported slabs #4, #5, #6, and #7 sketched in Fig. 10: specimen number; geometrical data, bars spacing and diameters, value of the applied
reference load.
Lx (mm)

L1x (mm)

L2x (mm)

Ly (mm)

L1y (mm)

L2y (mm)

t (mm)

#4
#5
#6
#7

2400
2400
2400
2400

790
790
520
790

250
250
520
250

2400
2400
3600
3600

790
790
820
1390

250
250
820
250

106.1
100
100
101.6

sx (mm)

sy (mm)

dx (mm)

dy (mm)

bar x (mm)

bar y (mm)

1 (N/mm2)
p

2 (N/mm2)
p

200
300
300
200

200
300
300
200

80
130
130
80

0.278
0.278

0.278

4.444

Ec (GPa)

#1
#2
#3
#4
#5
#6
#7

25.00a
38.00c
37.80c
32.20f
26.70f
58.00f
44.00f

1.65b
3.68d
3.17d
3.09f
2.88f
4.68f
3.56f

28.96e
27.47c
26.47c
27.97f
25.59f
34.19f
28.29f

0.2
0.2
0.2
0.2
0.2
0.2
0.2

PEXP
PUB

1.06

f t (MPa)

0.58
0.67
0.73

f c (MPa)

0.35
0.41
0.46

Slab specimen

PLB

0.77
0.84

Table 4
Steel-reinforced concrete slabs: specimen number; compressive and tensile concrete
strengths; elastic concrete properties.

10
12
12
10

1.30
1.33
1.45

10
12
12
10

0.44
0.45
0.51

80
130
130
80

0.30
0.34
0.39

#4
#5
#6
#7

1.05
1.15
1.31

Slab specimen

After El Maaddawy and Soudki [9].


q
0
0
Computed as f t 0:33 f c as suggested by Bresler and Scordelis [3].
c
After Sakka and Gilbert [35].
0
0
d
Derived by the flexural strength as f t f cf =1:2 according to Eurocode 2 [10].
0:3
0
e
Derived by the compressive strength as E 22f c =10
GPa according to
Eurocode 2 [10].
f
After Sakka and Gilbert [36].
b

Table 5
Steel-reinforced concrete slabs: specimen number; number of 3D-Solid elements and
1D-Embedded Truss elements used for the FE analyses.
Slab specimen

Number of 3D-solid

Number of 1D-Emb. trusses

#1
#2
#3
#4
#5
#6
#7

660
960
912
768
768
576
1056

60
162
290
336
336
232
492

#4 and slab #5; the rectangular S2R-4 and S2R-5 here named slab
#6 and slab #7, respectively. The tests set up, the material data
and many other details are given in the above quoted papers to
which reference is made; in the following the essential information
are summarized focusing the considered slab specimens.
The mechanical model, the geometry, loading and boundary
conditions, as well as the reinforcement arrangement of the seven
slabs analyzed are sketched in Fig. 8(a)(c) for slab #1 and #2,

Slab1

Slab2

Slab3

Slab4

Slab5

Slab6

Slab7

Fig. 11. Steel-reinforced concrete slabs: values of the upper P UB and lower P LB
bounds to the peak load multiplier against the experimentally detected one P EXP .

Fig. 9(a)(e) for slab #3, Fig. 10(a)(c) for slabs #4#7. Tables 2
and 3, which refer to Figs. 8 and 10 respectively, complete the
needed data specifying also the values of the applied reference
loads. Finally, Table 4 gives, for all specimens, the concrete material properties while the steel Young modulus has been assumed
equal to 205 GPa for all samples. No FE-meshes are shown for
the slabs being worth the general remarks given at the beginning
of Section 3 for what concerns the adopted finite elements,
namely: 3D-solids are used to model concrete and 1D-embedded
trusses to model re-bars. In Table 5, for completeness, are given
the numbers of elements used for the FE analyses of each
specimen.
Fig. 11 shows the results obtained for the seven slabs analyzed
and given, as before, in terms of experimentally detected peak load
multipliers (after the quoted paper) and upper and lower bound
predictions to them. In all cases the predicted interval of bounding
multipliers embraces the real peak load multiplier value and is also
quite narrow confirming a good performance of the proposed
methodology also for this type of steel-reinforced concrete element. Also in this case a monotonic and rapid convergence is
exhibited (less than ten iterations are required to stop both procedures). Plots of the type shown in Fig. 6(a) and (b) for the walls are
obtained for the slabs and are here omitted for brevity.

52

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

(b)

(a)
z

(d)

(c)

Fig. 12. Steel-reinforced concrete simply-supported slab #2. Band plots of the Cartesian strain rate component e_ cx in the deformed configurations at the ultimate value of the
acting loads: (a) results obtained at last converged solution of the LMM on the fictitious structure localizing the plastic zone and/or the collapse mechanism; (b) results pertaining
to an elastic solution of the real structure, i.e. with the real elastic parameters; (c) comparison of the deformed shapes given in (a) and (b) in the plane xz; and (d) photograph
of slab #2 at failure after Sakka and Gilbert [35].

(b)

(a)

(d)

(c)

Fig. 13. Steel-reinforced concrete corner-supported slab #7. Band plots of the Cartesian strain rate component e_ cy in the deformed configurations at the ultimate value of the
acting loads: (a) results obtained at last converged solution of the LMM on the fictitious structure localizing the plastic zone and/or the collapse mechanism; (b) results pertaining
to an elastic solution of the real structure, i.e. with the real elastic parameters; (c) comparison of the deformed shapes given in (a) and (b) in the plane yz; and (d) photograph
of slab #7 at failure after Sakka and Gilbert [36].

For what concerns the slabs failure mode prediction, all the
remarks made for the walls in the previous subsection hold true.
In Figs. 12(a)(d) and 13(a)(d) the predicted failure modes are
shown, for sake of brevity, only for slabs #2 (simply-supported)
and #7 (corner-supported), respectively. Once again the failure
modes predicted by the LMM on the fictitious structure at convergence are compared with the elastic solutions obtainable on the
real structure at ultimate load level. It is clear how the LMM locates
a reliable collapse mechanism, a sort of plastic hinge arises at
mid-span spreading to the whole slab width while the remainder

portions of the slab (till the supports) rotate rigidly around such
hinge. Quite impressive is the comparison between the numerically predicted mechanisms and the ones documented by the photographs at failure of the experimental tests. Peculiar are the e_ cy
concentrations at the corner supports as well as the prominent
flexural deformations, in the xz plane, showed by the meaningless elastic solution on the real slab at ultimate load for specimen
#7, see Fig. 13(b). Such effects are far from the real state of the slab
at incipient collapse and, in facts, they do not appear in the predicted mechanism shown in Fig. 13(a). The latter exhibits rigid

53

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

rotations around a narrow plasticized zone at mid span and almost


no flexural deformation in the xz plane, despite the punctual supports, that is it predicts almost exactly the collapse mechanism
detected experimentally, see Fig. 13(d). It is worth noting that
the predictions on the failure mechanisms are qualitative rather
than quantitative. Once again, the reliable information lie indeed
either on the e_ cy distribution and concentration evidenced by the colored contour maps which locate the plasticized zones or on the
shape of the deformed configuration, both predicting the collapse
mechanism the slab will exhibit.

Appendix A. Linear matching method


The sequence starts assigning to all FEs of the discrete structural
model an initial set of fictitious elastic parameters, say E0 ; m0 , and
 0
 0
initial stresses 
n0 ; q
x ; q
y . To initialize the iterative procedure,
k1

for upper bound value and where, for k 1; P UB can be any arbitrary value to start the sequence (it can indeed be assumed equal
to unity as suggested above). At the beginning compute also the
bulk

4. Concluding remarks and future developments


Large-scale prototypes of steel-reinforced concrete walls and
slabs have been analyzed to validate a nonstandard limit analysis
numerical approach recently proposed by the authors. With this
aim detailed experimental laboratory tests carried out up to failure
and available in the literature have been considered as a benchmark. The combination of two FE-based procedures, the former
aimed to search for an upper bound the latter able to give a lower
bound to the peak load multiplier, is the key-idea of the promoted
plasticity-based methodology oriented to practical reinforced concrete structures. Indeed, the confining effect of the steel bars and
the ductile behavior injected by their presence in such structures,
a circumstance shared by many RC-structures of common use in
the civil engineering applications, allows to apply a nonstandard
limit analysis approach as the one pursued here.
With all the limitations of a plasticity based approach for concrete, the obtained numerical results, as witnessed by comparison
with the experimental findings, appear to give useful and quite
reliable information on the peak load, failure modes and critical
zones of the analyzed reinforced concrete structural elements. A
fully 3D formulation has been employed that is at structural level,
by 3D-solid FEs, as well as at constitutive level, with a MW-type
concrete model implemented in the principal stress space. To this
concern it is worth remarking that a formulation based on principal stress components instead of generalized stress variables, such
as membrane forces, shear forces and bending moments, often
employed for limit analysis of RC-structures (see e.g. [15,17,23])
and requiring section-forces-based yield criteria strictly related
to geometry and loading conditions of the mechanical problem
under study, can be easily applied to structural RC elements of general shapes suffering general loading conditions. It is also worth
noting that the utilized FE numerical procedures are both based
on sequences of elastic analysis so resulting easy to apply with
any commercial FE code. Moreover, the whole methodology can
be easily rephrased with reference to DruckerPrager or Mohr
Coulomb or other criteria specifically oriented to concrete. The
choice of MW-type model being justified by the strict convexity
of such yield surface which is an essential requisite to achieve
the matching within the LMM. Straight meridians, if present as
in the above mentioned widely used criteria, it should be approximated by smooth curves, a drawback which can be overcome
without much effort (see e.g. [32]).
Finally, two possible future steps can be envisaged: (i) the application of the presented limit analysis approach to existing RCframed structures for the evaluation of their actual load bearing
capacity. To this concern the material parameters and strengths,
detected in situ, could be used to calibrate the adopted MWtype concrete model and (ii) the simultaneous application of the
present FE-based limit analysis either to concrete, governed by
the MW-type criterion, or to steel bars, handled by a von Misestype criterion. Such improvement would allow to take into account
the bars yielding often exhibited by RC-structures at a state of
incipient collapse. Both goals are the object of an ongoing research.

say for k 1, set P UB PUB 1; where capital P denotes the load


i , the subscript UB stands
multiplier of assigned reference loads p

and

shear

modulus,

K 0 E0 =31  2m0

and

G E =21 m , respectively. The following operative steps


can then be envisaged:
0

Step #1: Perform a fictitious elastic analysis with elastic parame k1
 k1
ters K k1 ; Gk1 ; initial stresses 
nk1 ; q
; q
;
x
y
k1

loads P UB

i . Compute a fictitious kinematic linear solup

,
tion (at each Gauss point on each FE), namely: e_ v
_ed k1 , e_ dk1 , u_ i k1 together with the corresponding
k1

stress values n k1 ; qx


; qy
(the apex standing
for a value computed by the fictitious linear analysis).
Compute also the (constant) value of the complementary energy potential pertaining to the fictitious solution, namely:
k1

W k1

k1


1  k1 k1
k1
k1
:
n
e_ v
qx k1 e_ dx
qy k1 e_ dy
2

It is worth noting that n k1 ; qx


; qy
, identify a
stress point, say PL , on the equipotential surface
h
i
 k1
 k1
W k1 whose
W n; q; K k1 ; Gk1 ; 
nk1 ; q
;q
x
y
k1

k1

k1
k1
k1
outward normal has components e_ v
, e_ dx
, e_ dy
.
Step #2: Locate on the MW-type yield surface the stress point
at yield, say PM nM ; qM ; hM , whose outward normal has
k1
k1
k1
components e_ v
, e_ d
, e_ d
. The uniqueness of such
x

point is assured by the strict convexity of the MW-type


yield surface.
Step #3: Impose the so-called matching conditions, i.e. find
those values of the fictitious moduli and initial stresses,
k ; q
 k
 k
say: Gk ; K k ; n
x ; q
y , (to be used, if necessary, at
next iteration) in such a way that the equipotential surh
i
 k
 k
W k1 matches the
face W n; q; K k ; Gk ; 
nk ; q
x ;q
y
MW-type yield surface at PL  PM . The components

e_ vk1 , e_ dxk1 and e_ dyk1 can in this way be interpreted


(together with the related compatible displacement
rates) as a collapse mechanism.
c k1
k1
c k1
k1
c k1
k1
Step #4: Set e_ v
e_ v
, e_ dx
e_ dx
, e_ dy
e_ dy
,
k1
u_ ci u_ i
and evaluate the upper bound multiplier
with the standard format of the kinematic approach of
limit analysis, namely:

R  Y k1 c k1
Y k1 c k1
n
e_ v
qx
e_ dx qYy k1 e_ cdyk1 dV
V
k
R
P UB
;
 u_ c d@V
p
@V t i i

where nY k1 ; qx
; qy
are the stresses at yield given
by the detected PM on the MW-type surface and where
the integrals have to be performed numerically on the FE
mesh. At current iteration, the work done by constant loads
(i.e. not amplified by the load multiplier), if any, is subtracted
from the numerator.
Y k1

Y k1

54

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255

Step #5: The above stresses at yield nY k1 ; qx

Y k1

; qy

Y k1

do not

k1
i
satisfy the equilibrium conditions with the loads P UB p
and a new fictitious analysis is performed with the
k
i . The iterations
updated values k and loads P UB p
(analyses) start and they carry on till the equilibrium
is fulfilled. This latter circumstance is verified by evaluating the difference between two subsequent PUB values.
k

k1

Namely: if jPUB  P UB j is less than or equal to a fixed


tolerance stop the iterations, else, set k k  1 and go
to step #1.
Appendix B. Elastic compensation method
The procedure starts with a first sequence, say sequence s 1,
s

i P1

carried out for given loads PD p
D pi , where: pi denote, as
s

before, assigned reference load; P D is the design load multiplier


1

pertaining to the sequence s (at start, for s 1, PD can be assumed


k

equal to unity). Denoting with  a quantity  pertaining to the


k-th iteration, or k-th elastic analysis, within the current sequence
s, the following steps can be envisaged:
Step #1: Set k 1 and assign to all FEs the known (real) mateelastic
parameters,
namely:
Ek1 E0 ,
0
m
m (the latter is kept constant during the analysis) or, if necessary, assign the bulk and shear modurial

k1

lus, K 0 and G0 , respectively.


Step #2: Perform an elastic analysis with material parameters
s
i computing the principal
Ek1 ; mk1 and loads PD p
stresses at each Gauss point of each FE. Let refer to
the element #e for simplicity. Average also such stresses inside the element so obtaining an averaged elastic solution within the e-th element. The latter, after
transformation in the HW coordinates, locates in
the principal stress space a stress point, say
e k1

P#e

ne ; qe ; he .

Step #3: Compute, in the Rendulic plane at h he , the HW


coordinates of the corresponding stress point at yield
Y k1

within the element #e, say P#e


Y

nY ; qY ; hY , with

h  h . That is locate on the MW-type yield surface


!e k1 . !e k1
the point on the direction OP #e
.
 OP #e 
Step #4: Update the Young moduli within the current element
#e
according
to
the
formula:

k1  ! k1 2


k
k1  !Y 
s
. For P D 1,
E#e : E#e  OP #e 
 OP e#e 
i.e. after the first sequence, the above updating formula
is applied only in the highly loaded elements where
 ! k1  ! k1




>  OP Y#e 
.
 OP e#e 
Step #5: Detect the maximum stress in the whole FE mesh, or,
k1

equivalently, locate the stress point, say PR ,


farthest away from the MW-type yield surface
and evaluate the pertinent stress point at yield,
Yk1

PR
. (The latter being the intersection between the
MW-type yield surface and the straight line having

!k1 . ! k1


:
direction OP R
 OP R 
k1

just reaches (is below) its


Step #6: Check if the stress point PR
corresponding yield value:
 ! k1  ! k1




(i) if  OP R 
<  OP YR 
then compute a lower bound
multiplier as:

k1

PLB

s
 ! k1
PD


:  OP YR 
 ! k1 ;


 OP R 

s1

k1

set PD
> P LB , i.e. increase the intensity of the acting
loads, set also s s 1 and go to step #1 to perform a
new sequence of elastic analyses.
 ! k1
 ! k1




(ii) if  OP R 
P  OP YR 
try to redistribute, that is set
k k  1 and go to step #2 to start a new analysis with
s

the current P D , i.e. within the current sequence but with


the updated Young moduli of step #4.
When inside the current sequence s the maximum
stress detected at step #5 of analysis k-th is such that
 ! k
 ! k1




it means that the stress redistribu OP R  P  OP R 
tion (or compensation) failed and the searched P LB value
coincides with the load multiplier of the previous
s1

sequence, i.e. PLB  PD

; the procedure ends.

References
[1] ADINA R&D Inc. Theory and modeling guide. Watertown (MA, USA): ADINA
R&D; 2002.
[2] Balan TA, Spacone E, Kwon M. A 3D hypoplastic model for cyclic analysis of
concrete structures. Eng Struct 2001;23:33342.
[3] Benkemoun N, Ibrahimbegovic A, Colliat JB. Anisotropic constitutive model of
plasticity capable of accounting for details of meso-structure of two-phase
composite material. Comput Struct 2012;9091:15362.
[4] Bresler B, Scordelis AC. Shear strength of reinforced concrete beams. J Am
Concr Inst 1963;60(1):5172.
[5] Boulbibane M, Ponter ARS. Extension of the linear matching method
to geotechnical problems. Comput Methods Appl Mech Eng 2005;194:
463350.
[6] Carrazedo R, Mirmiran A, Bento de Hanai J. Plasticity based stressstrain model
for concrete confinement. Eng Struct 2013;48:64557.
[7] Chen WF. Plasticity in reinforced concrete. USA: McGraw-Hill; 1982.
[8] Chen WF, Han DJ. Plasticity for structural engineering. New York
(USA): Springer-Verlag; 1988.
[9] Drucker DC, Prager W, Greenberg HJ. Extended limit design theorems for
continuous media. Quart Appl Math 1952;9:3819.
[10] Brisotto D de S, Bittencourt E, Bessa VMR. Simulating bond failure in reinforced
concrete by a plasticity model. Comput Struct 2012;106107:8190.
[11] El Maaddawy T, Soudki K. Strengthening of reinforced concrete slabs with
mechanically-anchored unbounded FRP system. Constr Build Mater
2008;22:44455.
[12] Eurocode 2. Design of concrete structures Part 11: General rules and rules
for buildings. Final draft prEN 1992-1-1, 2003.
[13] Gilbert RI, Sakka ZI. The effect of reinforcement type on the ductility of
suspended reinforced concrete slabs. J Struct Eng 2007;133(6):83443.
[14] Gilbert RI, Sakka ZI. Strength and ductility of corner supported two-way
concrete slabs containing welded wire fabric. In: CONCRETE 09, 24th Biennial
Conference of the Concrete Institute of Australia, Paper 5b-1, 1719 September
2009, Sydney.
[15] Krabbenhft K, Damkilde L. Lower bound limit analysis of slabs with nonlinear
yield criteria. Comput Struct 2002;80:204357.
[16] Larsen KP, Poulsen PN, Nielsen LO. Limit analysis of 3D reinforced concrete
beam elements. J Eng Mech 2012;138(3):28696.
[17] Le CV, Gilbert M, Askes H. Limit analysis of plates and slabs using a
meshless equilibrium formulation. Int J Numer Methods Eng 2010;83:
173958.
[18] Lefas ID, Kotsovos MD, Ambraseys NN. Behavior of reinforced concrete
structural walls: strength, deformation characteristics, and failure
mechanism. ACI Struct J 1990;87:2331.
[19] Li T, Crouch R. A C 2 plasticity model for structural concrete. Comput Struct
2010;88:132232.
[20] Liman O, Foret G, Ehrlacher A. RC two-way slabs strengthened with CFRP
strips: experimental study and a limit analysis approach. Compos Struct
2003;60:46771.
[21] Lubliner J. Plasticity theory. New York: Macmillan Pub. Co.; 1990.
[22] Mackenzie D, Boyle JT. A method of estimating limit loads by iterative elastic
analysis. Parts IIII. Int J Pressure Vessels Pip 1993;53:77142.
[23] Maunder EAW, Ramsay ACA. Equilibrium models for lower bound limit
analyses of reinforced concrete slabs. Comput Struct 2012;108109:1009.
[24] Mentrey P, Willam KJ. A triaxial failure criterion for concrete and its
generalization. ACI Struct J 1995;92:3118.
[25] P Nielsen M, Hoang LC. Limit analysis and concrete plasticity. 3rd ed. CRC
Press, Taylor & Francis Group; 2011.
[26] Pisano AA, Fuschi P. A numerical approach for limit analysis of orthotropic
composite laminates. Int J Numer Methods Eng 2007;70:7193.

A.A. Pisano et al. / Computers and Structures 160 (2015) 4255


[27] Pisano AA, Fuschi P. Mechanically fastened joints in composite laminates:
evaluation of load bearing capacity. Compos: Part B 2011;42:94961.
[28] Pisano AA, Fuschi P, De Domenico D. A layered limit analysis of pinned-joints
composite laminates: numerical versus experimental findings. Compos: Part B
2012;43:94052.
[29] Pisano AA, Fuschi P, De Domenico D. A kinematic approach for peak load
evaluation of concrete elements. Comput Struct 2013;119:12539.
[30] Pisano AA, Fuschi P, De Domenico D. Peak loads and failure modes of steelreinforced concrete beams: predictions by limit analysis. Eng Struct
2013;56:47788.
[31] Ponter ARS, Carter KF. Limit state solutions, based upon linear elastic solutions
with spatially varying elastic modulus. Comput Methods Appl Mech Eng
1997;140:23758.
[32] Ponter ARS, Fuschi P, Engelhardt M. Limit analysis for a general class of yield
conditions. Eur J Mech/A Solids 2000;19:40121.
[33] Radenkovic D. Thormes limites pour un materiau de Coulomb dilatation
non standardise. CR Acad Sci Paris 1961;252:41034.
[34] Roh H, Reinhorn AM, Lee JS. Power spread plasticity model for inelastic
analysis of reinforced concrete structures. Eng Struct 2012;39:14861.

55

[35] Sakka ZI, Gilbert RI. Effect of reinforcement ductility on the strength
and failure modes of one-way reinforced concrete slabs. UNICIV REPORT No.
R450, Sydney (Australia): The University of New South Wales; 2008,
ISBN:85841 417 1.
[36] Sakka ZI, Gilbert RI. Strength and ductility of corner supported two-way
concrete slabs containing welded wire fabric. UNICIV REPORT No. R453,
Sydney (Australia): The University of New South Wales; 2008, ISBN: 85841
420 1.
[37] Sloan SW. Lower bound limit analysis using finite elements and linear
programming. Int J Numer Anal Methods Geomech 1988;12:6177.
[38] Spiliopoulos K, Weichert D. In: Spiliopoulos K, Weichert D, editors. Direct
methods for limit states in structures and materials. Dordrecht: Springer
Science+Business Media B.V.; 2014. ISBN:978-94-007-6826-0.
[39] Vecchio FJ, Shim W. Experimental and analytical reexamination of classic
concrete beam tests. J Struct Eng 2004;130(3):4609.
[40] Wu R, Harvey JT. A J2 -plasticity model based on bounding surface concept. Int J
Numer Anal Methods Geomech 2013;37:74457.
[41] Zhang J, Li J. Investigation into Lubliner yield criterion of concrete for 3D
simulation. Eng Struct 2012;44:1227.

Вам также может понравиться