You are on page 1of 28

European Journal of Environmental and Civil Engineering

ISSN: 1964-8189 (Print) 2116-7214 (Online) Journal homepage: http://www.tandfonline.com/loi/tece20

Evaluation of one-way shear behaviour of


reinforced concrete slabs: experimental and
numerical analysis

Tan-Trung Bui, Ali Limam, Wendpanga-Serge-Auguste Nana, Emmanuel


Ferrier, Marion Bost & Quoc-Bao Bui

To cite this article: Tan-Trung Bui, Ali Limam, Wendpanga-Serge-Auguste Nana, Emmanuel
Ferrier, Marion Bost & Quoc-Bao Bui (2017): Evaluation of one-way shear behaviour of reinforced
concrete slabs: experimental and numerical analysis, European Journal of Environmental and Civil
Engineering, DOI: 10.1080/19648189.2017.1371646

To link to this article: http://dx.doi.org/10.1080/19648189.2017.1371646

Published online: 03 Sep 2017.

Submit your article to this journal

Article views: 18

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tece20

Download by: [Australian Catholic University] Date: 28 September 2017, At: 20:00
European Journal of Environmental and Civil Engineering, 2017
https://doi.org/10.1080/19648189.2017.1371646

Evaluation of one-way shear behaviour of reinforced concrete


slabs: experimental and numerical analysis
Tan-Trung Buia  , Ali Limamb, Wendpanga-Serge-Auguste Nanaa, Emmanuel Ferrierc,
Marion Bostd and Quoc-Bao Buie
a
SMS-ID, INSA Lyon, University of Lyon, Lyon, France; bUniversity of Lyon, Lyon, France; cLaboratory LMC2, UCBL,
University of Lyon, Villeurbanne Cedex, France; dGERS-RRO, Ifsttar Bron, Bron, France; eSustainable Developments in
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Civil Engineering Research Group, Faculty of Civil Engineering, Ton Duc Thang University, Ho Chi Minh City, Vietnam

ABSTRACT ARTICLE HISTORY


The shear design of concrete slabs is still an unsolved problem. The following Received 14 March 2017
study presents experimental and numerical investigations on the shear Accepted 20 August 2017
behaviour of reinforced concrete slabs (without shear reinforcement)
KEYWORDS
under concentrated loads. The small thick slabs of 10 cm were tested in this Concrete slab; shear;
study. Experimental tests were conducted to quantify the shear strength experiment; finite element;
and the associated failure modes. The influence of several variables was concentrated load
addressed such as the influence of boundary conditions, four supported
side slabs instead of two supported side slabs, the influence of loading plate
length. A series of eight tests on six slabs were presented. The experiments
are firstly used to evaluate the pertinence of Eurocode 2 and Model Code
2010 formulations using the levels of approximation LoA I and LoA II for the
shear design of reinforced concrete slabs without shear reinforcement in
comparison with the French approach, and secondly validated numerical
modelling using the non-linear finite element method. The proposed
numerical model showed good agreement with the experimental results
in terms of slab behaviour.

1. Introduction
In reinforced concrete members, after cracking due to bending, shear can be transferred by a number
of potential actions, thereby leading to failure. A summary of shear-transfer actions can be found in
ASCE-ACI Committee 445 report from 1998 (ASCE-ACI Committee 445 on Shear & Torsion, 1998), with
the actions defined as: shear stresses in uncracked concrete, interface shear transfer (or aggregate
interlock), the dowel action, residual tensile stresses and arch action.
Shear behaviour in slabs can be divided into two cases: one- or two-way shear. The latter reaches
shear failure, for example, in punching under a concentrated load (Belletti, Pimentel, Scolari, & Walraven,
2015; Belletti, Walraven, & Trapani, 2015), which is generally smaller than the flexural failure load calcu-
lated by theory such as yield line theory (Theodorakopoulos & Swamy, 2002). Failure occurs with the
potential diagonal crack following the surface of a truncated cone around the loading area. The problem
with this failure mode is that it is brittle due to the inability of the concrete to support the large tensile
stresses that develop. One-way shear behaviour is defined by the presence of a distinct shear crack on a
single side. Loading was either line loads or concentrated loads (situations such as localised heavy loads

CONTACT  Tan-Trung Bui  tan-trung.bui@insa-lyon.fr


© 2017 Informa UK Limited, trading as Taylor & Francis Group
2    T.-T. BUI ET AL.

from equipment, slabs loaded by walls or columns). The slab’s behaviour under line loads can generally
be assumed to be equivalent to a beam’s behaviour. A number of experimental studies conducted on
slabs or on wide beams under line loads, such as Gurutzeaga, Oller, Ribas, Cladera, and Marí (2014), A.
Lubell, E. Bentz, and M. Collins (2009), A. S. Lubell, E. C. Bentz, and M. P. Collins (2009), Sherwood, Bentz,
and Collins (2007), Sherwood, Lubell, Bentz, and Collins (2006), have confirmed this tendency.
The shear behaviour of a concrete beam has long been investigated by many researchers. A review
of the studies published on the past 60 years of research on shear behaviour to the behaviour of beams
can be found in Collins, Bentz, and Sherwood (2008). However, until now, in comparison with shear
behaviour in beams, only a few tests have been conducted to study the shear behaviour of slabs under
concentrated loads. Most of these studies addressed the cases corresponding to the structure used in
slab bridges (Lantsoght, 2013; Lantsoght, van der Veen, & Walraven, 2013; Natário, Fernández Ruiz, &
Muttoni, 2014; Rombach & Kohl, 2013; Reißen & Hegger, 2013a, 2013b; Vaz Rodrigues, Fernández Ruiz,
& Muttoni, 2008), or in thick slab of nuclear buildings (Bui et al., 2016; Nana, Bui, Limam, & Abouri, 2017);
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

However, configurations that correspond to the structure of floor slabs have rarely been addressed.
Therefore, this study responds to a lack of shear behaviour research on floor slabs under a concentrated
load. For traditional floor slabs, a slab thickness of 18 cm was often used in practice. However, in reality,
concrete thickness of 10 cm is large enough in mechanical point of view (to ensure a bearing capacity
in ad equation with classical design loads used for the common residential buildings). This is applied for
the isolated slab specimen. The remaining thickness of concrete member is only for the objective of the
acoustic problem. In the modern construction, the traditional slab thickness of 18 cm can be replaced by
the small thickness slab of 10 cm (for mechanic resistance) and combining with another material layer
(for acoustic problem). Indeed, from the structural resistance view point, considering Eurocode 2, and
classical building floors configurations submitted to dead load and classical service loads, a thickness of
10 cm and steel reinforcements which not exceed the maximum authorised in regulations (Eurocode 2),
permits to sustain the mechanical requirements. The classical adopted thickness for common buildings
which is 15 or 18 cm is mainly needed for acoustic criterions. It is known that the shear capacity of a
concrete member depends on the slab thickness. Therefore, the evaluation of shear behaviour of the
small slab thickness of 10 cm different from the one of the traditional thick slab of 18 cm is necessary.
All the studies mentioned above on the shear behaviour of slab bridges were based on tests con-
ducted on slabs supported on one side (cantilever slab) or two sides to model real deck slabs in bridge
structures. However, in floor concrete structures, the slabs can also be supported on four sides. These
different boundary conditions can produce different shear behaviours due to the influence of the
two supported lateral sides, as in punching shear. The authors in Elstner and Hognestad (1956) tested
similar slabs supported along the entire perimeter or only on two opposite edges. The resulting crack
pattern was different for each slab. The slabs supported along the whole perimeter showed a stiffer
behaviour than the slab supported on two edges and the failure load was also higher. As a result, for
one-way shear, the different boundary conditions when slabs are supported on four sides should also
be compared. Others studies on continuous slabs are also available in Belletti, Cantone, Fernández Ruiz,
and Muttoni (2016), Cantone, Belletti, Manelli, and Muttoni (2016).
For slabs subjected to a concentrated load, the shear resistance should not be calculated over its
entire width bw but over a certain effective width beff. In practice, there are two ways to determine the
effective width (Lantsoght, van der Veen, and Walraven, 2013). The first method was used by the Dutch
by a 45° horizontal load spreading from the centre of the load (beff,1). The second method was used by
the French (Bui et al., 2016; Nana et al., 2017) with a 45° horizontal load spreading from the far edges
of the load (beff,2). The test results in these two studies showed that the shear load determined by the
French practices have a good ability to predict the experimental results. With this method, it could be
seen that beff,2 depends on the length of the loading plates. Only one study, conducted in 1988 (Regan
& Rezai-Jorabi, 1988), and recently in Belletti, Damoni, Hendriks, and de Boer (2014), have investigated
this influence. The conclusion that can be drawn from this research is that there is generally a small
increase in resistance as the loading plate length increases for a constant slab width. Therefore, the
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   3

present study was conducted to contribute to gauging the influence of the loading plate length on
the shear behaviour of slabs.
Sherwood et al. (2006) and Gurutzeaga et al. (2014) confirm that transverse flexural reinforcement
does not influence the shear stresses at failure of one-way slabs under line loads. Note that the transverse
reinforcement is the flexural reinforcement in the transverse direction (it’s not the shear reinforcement).
The behaviour of one-way slabs under line loads can be assumed to be the same as beams. However,
in one-way slabs under concentrated loads, moments occur in the span direction and the transverse
direction, so a one-way slab under concentrated loads does not behave like a beam. Most of the inter-
national codes such as ACI-318-08 (318, 2008) and Model code 2010 (Fib, 2013) for design provisions do
not take into account the influence of the transverse reinforcement. Only in the French National Annexe √
for Eurocode 2 (European Committee for Standardisation, 2005), a different approach vmin = 0.34 𝛾
fck
c

was used to evaluate the shear strength for slabs with transverse redistribution of loads. Nonetheless,
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

this is not made clear that how the transverse reinforcement influence on the shear strength.
The non-linear finite element analysis in modelling the engineering structures is currently an
advanced numerical tool which can provide a very good prediction. However, the modelling of the
non-linear behaviour for reinforced concrete structures under shear loading is generally difficult, espe-
cially for thick slabs when the 3D effect in the thickness increases. Furthermore, the failure mode in
shear of the concrete structures is a quasi-brittle mode, which often causes a divergence problem in
the model’s algorithm. There are many non-linear concrete models from different theories (Grassl &
Jirásek, 2006; Hansen & Schreyer, 1994; Kachanov, 1986; Massonnet, Olszak, & Phillips, 1979). However,
no model can reproduce all the failure mechanisms for reinforced concrete structures. Therefore, the
verification of the robustness of the numerical model applied for the non-linear behaviour of the rein-
forced concrete is still necessary, in particular for the case of shear loading.

2.  Experimental programme


2.1.  Test configuration
The experimental study of the behaviour of concrete slabs was carried out on a total of six slabs,
numbered from S1 to S6. The concrete and reinforcement properties are reported in Table 1. All slabs
were without shear reinforcement. The slabs were designed to study the shear resistance subjected to
a concentrated load near the support. Therefore, the shear span av/d ratio is taken to be bigger than

Table 1. Specimen properties.

fcm fctm Longitudinal rein- Transverse Rein-


(MPa) (MPa) forcement forcement
Line
load Mean Mean Bars/spac- Bars/spac-
Specimen bw (mm) d (mm) av /d L (cm) values values ing (mm) ρl (%) ing (mm) ρt (%)
S1 2500 85 2 100 25.8 2.9 21Φ10/120 0.77 10Φ10/240 1.039
(4 supports) 11Φ12/240
S2 2500 85 2 100 30.4 3.1 31Φ10/80 1.16 21Φ6/120 0.303
(2 supports)
S3 2500 85 2 20 30.4 3.1 31Φ10/80 1.16 13Φ6/200 0.182
(2 supports)
S4 400 85 2 40 25.8 2.9 7Φ10/50 1.86 – 0
(2 supports)
S5 2500 85 2 100 30.2 3.4 8Φ8/120 1.00 16Φ8/160 0.406
S5bis 14Φ6/40
(2 supports) 24Φ6/35
20Φ10/40
S6 2500 85 2 40 19.2 2.3 6Φ8/120 1.01 16Φ8/160 0.406
S6bis 10Φ6/40
(2 supports) 14Φ6/35
16Φ10/50
4    T.-T. BUI ET AL.

or equal to 2.0 (Kani, 1964) (av is the horizontal distance of the section from the face of support, d is
effective depth of the slab). To avoid the direct transmission of the loads to the supports, the av/d ratio
was chosen to be sufficiently large, so a constant value av/d = 2 (so a/d = 3.18) was chosen for all slabs.
This has been proved experimentally in the study of (Kani, 1964). An experimental programme of two
series of slabs without shear reinforcement was developed to evaluate the shear strength. The tests
were carried out with a structure like a floor system in which a concrete floor slab was supported by
the reinforced concrete beams.
The first series consists of slabs S1–S4 in which the RC beams are directly supported by the laboratory
floor (Figure 1). A polyene layer was placed between the surface of the reinforced concrete beam and
the laboratory floor. All slabs were identical in thickness: 100 mm. Slab S1 was a square slab supported
on the RC beams on all four sides measuring 2.9 m × 2.9 m and 2.5 m × 2.5 m between the faces of
supports. The loading device included a 1 m × 0.2 m rectangular loading plate (Figure 1(a)). To evaluate
the influence of the boundary conditions, slab S2 was supported on two sides (Figure 1(b)). The same
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

loading surface as for S1 was applied to slab S2. Slab S3 had the same geometry properties as slab S2.
However, the loading surface was decreased by reducing the size of the loading plate to 0.1 m × 0.1 m
(Figure 1(c)). The second test on the same slab S3 was conducted, but the load was applied to the right
side of the slab, called the S3bis test. The av/d = 2 value remained constant. Slab S4 was a beam measur-
ing 2.9 m × 0.4 m. The load was applied uniformly over its full width with the loading plate measuring

(a) (b) (c)

(d)

Figure 1. Slabs S1 to S4.


EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   5

0.4 m × 0.2 m. This test was conducted with the aim of showing a different shear behaviour for slabs
under concentrated loads than for beams.
A second series consisted of tests on slabs S5, S5bis, S6 and S6bis. These slabs had the same geomet-
rical properties and boundary conditions (two supports) as the slab S2 test but with different loading
plate sizes. Slabs S5 and S5bis had a loading plate length of L = 1 m and slabs S6 and S6bis were smaller:
L = 0.4 m (Figure 2). The S5bis and S6bis tests have the same geometry as the S5 and S6 tests, but the
load was applied in the other side (right side) of the slabs. The loading plate width was the same: 0.2 m.
A mortar layer was placed between the surface of the reinforced concrete beam and the laboratory
floor. This mortar layer is to prevent the small sliding of the RC beams supported on the laboratory
floor. The first objective of this second series is to study through the slabs S5 and S6, the influence of
the loading plate length on the shear behaviour of slabs. The second objective of this series through
the slabs S5bis and S6bis is to study the effect of a type of boundary condition which is intermediate
between the simple supported and the embedded. Therefore, for the S5bis and S6bis tests, the rotation
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

of two RC beams was rigidified by a system consisting of longitudinal metal bars welded with the pre-
stressed steel bars anchored on the laboratory floor (Figure 2).
The shear resistance of a slab subjected to a concentrated load should not be calculated over its
entire width bw, as is done for beams, but over a certain effective width beff used in the French code
(FD P 18-717, 2013). Horizontal load spreading is assumed to be 45° diffused from the far corners of
the load (Figure 3). As a result, the effective width calculated is beff = L + 0.74 (units in m), with L the
length plate of the load.
An overview over the properties of the six slabs tested is given in Table 1. The specimens were cast
using ready-mixed concrete, with specified 28-day strength of 25 MPa. A maximum aggregate size of
20 mm was used. The reinforcement had the yield strength fy = 567 MPa and the ultimate strength
fu = 672 MPa. The concrete cover was 10 mm, resulting in an effective depth d = 85 mm. The amount,
the spacing and the diameter of longitudinal and transverse reinforcement are given in Table 1. Figure
4 illustrates the reinforcement layout of S5 and S5bis slabs.

2.2.  Test set-up


A quasi-static loading was applied by a hydraulic jack with a maximum load capacity of 400 kN using
displacement control. The stiff beams attached to the loading machine were used to transversely dis-
tribute the load, with a different configuration depending on the loading area (Figure 5). During the test,
the force and the displacement at different points on the slab were measured. Force was measured by a
sensor placed between the head of the jack and the stiff beams. The slab displacements were measured

System to rigidify the rotation

Figure 2. Slabs S5, S5bis; S6, S6bis.


6    T.-T. BUI ET AL.

Figure 3. Assumed effective width for slabs in French practice (FD P 18-717, 2013).
A-A

100 A

B B
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Load Left Load Right

31 Frames
Ø6/40 Ø6/80

10 10
8 Ø8/120
Ø8/120 285
2500
14 Ø6/40
250
Ø6/35 2*12 Ø6/35
250

Ø6/80

10 10
2*10 Ø10/40
Ø10/40 285

400 16Ø8/160
249
Ø6/80 Ø8/160
200
100
B-B 400
Ø10/40 + Ø6/35 + Ø6/40 + Ø8/120

2900

Linear loading
200 170 200 on 1000*200
av

d=85
400
300
2*31 Frame
Spacing 60
Ø6/120 60
245
38 38
Ø4/100 18

Figure 4. Reinforcement layout of slabs 5 and 5bis. Units: [mm].

using seven displacement sensors (LVDT) positioned on the slab (S1 to S3; S5 and S6) as shown in Figure
6(a). For slab S4 (beam), only four LVDTs were used, as shown in Figure 6(b).

3.  Experimental results and discussion


The values of the peak loads of the slabs and the corresponding failure modes are summarised in Table
2. The shear strength Vfailure was calculated from the peak load Pfailure by determining the reaction force
in the support, similar to a calculation for beam shear:
( )
Vfailure = Lspan − dm ∕Lspan × Pfailure
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   7
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 5. Load system and the dimensions of the stiff beams: (a) L = 100 cm; (b) L = 40 cm.

575 825

LVDT 3
375
LVDT 7

625
500 160 220 870 1150 100

LVDT 2
LVDT 4 500
LVDT 5 625

375 LVDT 6
160 220 870 1150 100
LVDT 1 LVDT 4
575 825
LVDT 3
LVDT 1
LVDT 2

(a) (b)

Figure 6. Transducers positioned on the slabs of: (a) Specimen 1; (b) Specimen 4.

where Lspan is free span (equal to 2.5 m) and dm is distance from the middle point of the loading plate
to the support (equal to 0.27 m).
However, this method is not used for the four sides supported slabs. For the four sides supported slab
(S1), the shear load was calculated from the peak load by a Navier solution consisting of a Fourier serial
development for rectangular plates simply supported at on all four sides (Timoshenko & Woinowsky-
Krieger, 1959). The geometry dimension and the peak load were implanted in a computer programme
and then using a Navier solution.
Based on these results, we evaluated the influence of the boundary conditions, the length of the
loading surface and the transverse reinforcement.
8    T.-T. BUI ET AL.

Table 2. Experimental results.

Line Longi. Tranv.


load rein. rein.
L
Number of Failure Pfailure √ Vfailure
Specimens supports (mm) ρl (%) ρt (%) mechanism (kN) Pfailure/ fcm (kN) Remarks
S1 4 1000 0.77 1.039 Shear 294 57.9 211.9 Reference
slab
S2 2 1000 1.16 0.303 Shear 308 55.9 274.7 Boundary
conditions
effect
compared
to slab S1
S3bis 2 400 1.16 0.182 Shear 196 35.5 174.8 Loading
plate
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

length
effect
compared
to slab S2
S4 (beam) 2 400 1.86 0 Shear 81 15.9 72.3
S5 2 1000 1.04 0.406 Shear 342 62.2 305.1 Reference
slab
S5bis Shear 351 63.9 313.1 Boundary
constraint
effect
compared
to slab S5
S6 2 400 1.01 0.406 Shear 185 42.2 165.0 Loading
plate
length
effect
compared
to slab S5
S6bis Shear 166 37.9 148.1 Boundary
constraint
effect
compared
to slab S6

3.1.  Influence of the boundary conditions (supported on four sides vs. two sides)
The overall responses of slabs S1 and S2 are shown in Figure 6 representing the load/displacement
curves measured at the point of LVDT4. The peak load is determined by a loading drop from the load/
displacement relation measured at the jack during testing. After this drop, in the post-peak behaviour,
a very small amount of ductility was observed which presented the typical brittle failures of concrete
in shear. Comparing the load/deflection curves in slabs S1 and S2 shown in Figure 7, it can be seen that
the slab supported on four sides (slab S1) was a bit stiffer than the slab supported on two sides (S2). The
failure load of slab S1 seemed to be a bit smaller than that of slab S2 (Pfailure-S1 = 289 kN; Pfailure-S2 = 306
kN). However, if the effect “concrete variability” is removed using the failure load normalised by the
square-root of the compressive strength fcm (Table √ 2), we found that√ the failure load of slab S1 seemed
to be a bit bigger than that of slab S2 (Pfailure-S1/ fcm  = 57.9; Pfailure-S2/ fcm  = 55.9 kN).
The cracking patterns of all slabs were obtained after loading by observing and measuring directly
using magnifying glass. The crack patterns of two slabs are shown in Figures 8 and 9. Shear failure was
observed in both slabs. On the top surface, the progressive crack growth developed in a different way
for the two slabs. For slab S1, the cracks appeared and spread over four sides on the top surface (Figure
8(a) and (b)). At the four corners, a crack line was formed with an approximate angle of 45° to each of
the two adjacent sides. In other words, these boundary conditions allowed the cracks to develop on
each side on the total top surface. However, for slab S2, the cracks concentrated mainly on the top side
and around the loading plate (Figure 9(a)).
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   9

Comparison of Slab S1 and S2


350

300

250

200
LVDT 4

Load (kN)
150

100
LVDT 4 - 4 supports
50
LVDT 4 - 2 supports
0
0 2 4 6 8 10 12 14 16 18
Displacement (mm)
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 7. Comparison of load/displacement curve between Slab S1 and S2.

(a) (b) (c)

Figure 8. Crack pattern of slab S1 on: (a), (b) top surface; (c) bottom surface.

(a) (b)
Figure 9. Crack pattern of slab S2 on: (a) top surface; (b) bottom surface.
10    T.-T. BUI ET AL.

For slab S1, in contrast to the wide development of the cracks on its top surface, due to the two lat-
eral sides supported, the cracks on its bottom surface seemed to be limited in the transverse direction
and developed mainly in the longitudinal direction (Figure 8(c)). Nevertheless, for slab S2, the cracks
tended to develop in two directions – the longitudinal and transverse directions (Figure 9(b)). These
results indicate that the crack growth of slabs under concentrated loads near the support depends on
the boundary conditions related to the numbers of sides supported. However, the difference in shear
capacity was not clear. However, between slab S1 and S2, there is also a change in the horizontal and
transversal reinforcements. This change could be also responsible for the change in the failure force.

3.2.  Influence of loading plate length


To evaluate the influence of loading plate length, for slab S3 the load area was reduced to a short length
of L = 20 cm. However, this test led to failure that did not correspond to the expected failure mode.
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

The failure occurred by punching instead of shearing. To avoid this local damage with punching shear
failure created around the small loading area, an additional test was conducted on the same slab (S3)
with the longer loading plate length L = 40 cm. The force was applied to the other side of the support
line (S3bis test). The shear span av/d = 2 ratio remained constant in these experiments. It had only the
measurements of force and the displacement by the transducer included in the hydraulic jack. The
results obtained for the S3bis slab was shear failure.
From the observations of the tests above, to avoid punching failure, the length L ≥ 40 cm of the
loading plate was used for the other slabs. Slab S5 with L = 100 cm and slab S6 with L = 40 cm was tested
to evaluate the influence of the length of the loading plate. The two slabs had the same reinforcement
ratio. The comparison of the load/displacement responses for S5 and S6 are shown in Figure 10. These
experiments demonstrate that shear capacity increases as the length of the loading plate increases. This
can be explained by the fact that the shear capacity of slabs can be calculated using the concept of the
effective width (beff ) which depends on the length plate of the load (described in the previous Section
2.1). Therefore, if the concept of an effective width can be applied to concrete slabs loaded in shear,
then larger effective widths lead to larger capacities. The concept of the effective width is accepted and
included in the recent version of the French code (FD P 18-717, 2013). The same phenomenon of the
influence of the length plate load on the shear strength can be observed also in the recent studies of
Lantsoght, de Boer, van der Veen, and Walraven (2015), Lantsoght, van der Veen, and Walraven (2012),
Nana et al. (2017). Furthermore, the studies in literature (experiments and simulations) showed that the
observed increase in shear capacity is almost linear with the increase in length plate.
Both slabs S5 and S6 failed in shear (Figure 11). In both cases, the crack patterns developed in two
directions: longitudinal and transverse. The shear behaviour of these slabs loaded with a concentrated

Comparison of Slab S5 and Slab S6


400

350 L=100cm
L=100cm
300

250
L=40cm
Load (kN)

L=40cm
200

150

100 LVDT 2 - Slab S6


LVDT 4 - Slab S6
50 LVDT 2 - Slab S5
LVDT 4 - Slab S5
0
0 2 4 6 8 10 12 14
Displacement (mm)

Figure 10. Influence of loading plate length (L = 100 cm and 40 cm).


EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   11

Loading area

(a) (b) (c)

Figure 11. Crack patterns of: (a) slab S5, top face; (b) slab S5, bottom face; (c) slab S5bis, bottom face.
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

load was found to be different from the shear behaviour of the two-dimensional beam, as in the analyses
given in Lantsoght, Van der Veen, Walraven, and De Boer (2013). The S4 beam failed in shear (Figure 12),
and cracks developed only perpendicular to the span direction, while in the slabs the cracks developed
perpendicular and parallel to the span direction. This indicates the ability of the slab to distribute the
load over the longitudinal and transverse directions.

3.3.  Influence of the boundary constraint


In slabs S5bis and S6bis, the rotation of the two RC beams was restrained. The objective of this series
through the slabs S5bis and S6bis is to study the effect of a boundary condition type which is inter-
mediate between the simple supported and the embedded. For slabs with L = 1 m (S5 and S5bis), the
failure load was 2% higher when the two RC beams supporting the slab were blocked in horizontal
displacement (Figure 13). However, this horizontal displacement restraint reduced the shear strength
by 8.8% for slabs with L = 0.4 m (Figure 13). Basing on our experimental data available, the influence of
this rotation restraint could not be quantified. Therefore, in the numerical part, two tests of S5bis and
S6bis were chosen to simulate.

4.  Eurocode 2 and Model Code 2010 models versus experimental results
4.1.  Eurocode 2 model
In this section, the experimental results are compared to EN 1992-1-1:2005 with recommended values,
and EN 1992-1-1:2005 with values from the French National Annexe are compared with the experimental
results from the tests. In EC2, the design value for the shear resistance VRd,c of members without shear
reinforcement was evaluated empirically out of experimental data collected worldwide (Feenstra, de
Borst, & TU Delft: Civil Engineering & Geosciences, & TU Delft, Delft University of Technology, 1993).
According to EN 1992-1-1:2005, Section 6.2.2 (European Committee for Standardisation, 2005), the shear
resistance for a structural member without shear reinforcement is calculated as follows:
� � √ �
0.18
⋅ k ⋅ 3 100 ⋅ 𝜌1 ⋅ fck + k1 ⋅ 𝜎cp ⋅ bw ⋅ d = VRd,c1
VRd,c = Max � c
𝛾
� (1)
𝜈min + k1 ⋅ 𝜎cp ⋅ bw ⋅ d = VRd,c2

where, √
k =1+ 200
d
≤ 2 in mm
ρl: Geometric amount of longitudinal tensile reinforcement,
Asl
𝜌l = < 2%
bw .d
12    T.-T. BUI ET AL.

(a) (b)

Figure 12. Crack patterns of beam S4 on the: (a) top face; (b) bottom face.
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Slab S5 and Slab S5bis Slab S6 and Slab S6bis


400 200

350 180
160
300
140
250
120
Load (kN)

Load (kN)
200 100
150 80
LVDT 2 - Slab S5 60
100 LVDT 2 - Slab S6
LVDT 4 - Slab S5
LVDT 2 - Slab S5bis
40 LVDT 4 - Slab S6
50 LVDT 2 - Slab S6bis
LVDT 4 - Slab S5bis 20
LVDT 4 - Slab S6bis
0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18
Displacement (mm) Displacement (mm)

Figure 13. Influence of the restraint in the horizontal displacement of the two RC beams supported.

bw: Smallest width of the cross-section in the tensile area, d: Effective depth of the cross-section, fck:
Compressive strength (MPa) and fck ≤ 90 MPa, Asl: Tensile reinforcement (mm2), σcp: The average normal
concrete stress over the cross-section, positive in compression. √
The recommended value for νmin is given by (6.3 N): 𝜈min = 0.035 ⋅ k 3∕2 ⋅ fck in general.
In the French National Annexe (FNA) requirements, a different approach is used with regard to νmin
using the following value of νmin:
0.34

⎧ 𝛾 fck (2)
⎪ 0.053
c
3∕2

𝜈min = ⎨ 𝛾 k ⋅ fck (3)
⎪ 0.35 √f
c

⎩ 𝛾c ck (4)

Equation (2) for slabs with transverse redistribution of loads,


Equation (3) for beams and for other types of slab,
Equation (4) for walls. √ √
√ FNA, the value of υmin is 𝛾c
In our tests, following the 0.34
fck (𝜐min = 0.23 fck for characteristic value
fck and γc = 1.5; 𝜐min = 0.34 fck for mean value fcm and√γc = 1). This approach can be compared with the
approach recommended by EC2 𝜈min = 0.035 ⋅ k 3∕2 ⋅ fck .

4.2.  Model Code 2010 (LoA I and LoA II)


This code includes a mechanical basic set in the shear design process, which is intended to offer more
flexibility to the engineer in his calculations, while maintaining a balance between complexity and
precision, for the design of new structures and for the evaluation or verification of already existing struc-
tures. The new code provisions (Fib, 2013) contains four Levels of Approximation (LoA). This approach
consists in approaching a calculation by increasing level of complexity leading to an increasing precision
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   13

of the results. Only the first two Levels of Approximation are examined in this study. Level I (LoA I)
provides a simple calculation, it is the most classical method among others. Level II (LoA II) is a bal-
anced model between complexity and precision. The shear resistance for a structural member without
shear reinforcement is calculated as follows, where kv is a parameter defined for the different levels of
approximation (LoA):

fck
VRd,c = kv × × bw × d
𝛾c

4.2.1.  Levels of Approximation (LoA I)


The LoA I is suitable for pre-dimensioning of structural elements, where a conservative calculation
method is acceptable. In the case that the concrete compressive strength fck ≤ 70 MPa, the yield stress
of the longitudinal reinforcing bars fyk ≤ 600 MPa and the maximum aggregate size dg > 10 mm, the
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

parameter kv (LoAI) can be determined by :


180
kv (LoA I) =
(1000 + 1.25d)
4.2.2.  Levels of Approximation (LoA II)
The second level of approximation (LoA II) allows to calculate the shear strength on the basis of a
more accurate determination of the parameter kv, which is now, in addition to the effective depth d,
a function of the strain value εx taken at the mid-depth of the effective section (this value has a clear
physical meaning, it represents the average longitudinal strain state in the web) and also a function of
the maximum aggregate size. The parameter kv (LoA II) can be determined as below :
0.40 1300
kv (LoA II) = ( )×( )
1 + 1500𝜀x 1000 + kdg d

32
kdg = ≥ 0.75
(16 + dg)

( MEd )
d
+ VEd
𝜀x =
2Es × As

The factor kv (LoA II) includes the “strain effect” (εx) and the “size effect” (member size d), with the
aggregate size being taken into account within kdg, where dg is the specified maximum size of the
aggregate. In fact, Level I (LoA I) makes the simplification that kdg = 1.25 (assuming a maximum aggre-
gate size > 9.6 mm) and εx = 0.00125, i.e. half the yield strain of the longitudinal reinforcing bars with
fyk = 500 MPa. Note that MEd and VEd are, respectively, the bending moment and the shear force at the
control section. The control section is generally defined at a distance d (the effective flexural depth of
the specimen) from the face of the support. In our study, the modulus of elasticity of the reinforcing
steel is assumed to be 210 GPa.
To compare test data to the code predictions, two calculations were conducted (Table 3). Firstly, the
calculation using the characteristic concrete compressive strength fck was conducted. It is assumed that
fck is equal to the nominal value of class concrete used. So, fck = 25 MPa for C25/30 concrete class. In
this calculation with fck, all partial safety factors are equal to 1.5 (γc = 1.5). The second calculation with
the mean value (“fcm measured” in table 4) of the measured material properties are used√ and all partial
safety factors are equal to 1 (γc = 1). The VuFNA is calculated
√ with the value of υ min
of 0.34
𝛾c
fck . The VuEC2
is calculated with the value of 𝜈min = 0.035 ⋅ k 3∕2 ⋅ fck .
The horizontal load spreading from Figure 3 with the 45° horizontal load spreading from the far
side of the loading plate towards the support was used for these approaches. The results obtained in
Table 4 show that the shear predictions from EN 1992-1-1:2005 with the recommended value lead to
14    T.-T. BUI ET AL.

Table 3. Comparison between experimental and calculated shear failure loads with fck and γc = 1.5.

Recommended value by Recommended value by Model Code 2010


French National Annexe EC2 (LoA I and LoA II)
Vfailure/VuFNA Vfailure/VuEC2 Vfailure/Vu(LoA I) Vfailure/Vu (LoA II)
fck = 25 MPa fck = 25 MPa fck = 25 MPa fck = 25 MPa
Specimens 𝛾c = 1.5 𝛾c = 1.5 𝛾c = 1.5  𝛾c = 1.5
1 1.26 2.31 2.64 1.48
2 1.64 2.64 3.42 1.81
3bis 1.59 2.53 3.33 1.80
4 (beam) 1.83 2.82 3.82 1.80
5 1.82 2.94 3.80 2.18
5bis 1.87 3.02 3.90 2.26
6 1.50 2.42 3.14 1.69
6bis 1.35 2.17 2.82 1.46
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

AVG 1.61 2.61 3.36 1.81


STD 0.23 0.30 0.47 0.29
5% percentile 1.29 2.22 2.70 1.46

Table 4. Comparison between experimental, calculated and numerical shear failure loads with fcm and γc = 1.

French National Recommended value Recommended value by Model Code


Annexe by EC2 2010 (LoA I and LoA II) FEM
Vfailure/VuFNA Vfailure/VuEC2 Vfailure/Vu(LoA I) Vfailure/Vu(LoA II) Vfailure/VuFEM
fcm measured fcm measured fcm measured fcm measured fcm measured
Specimens 𝛾c = 1 𝛾c = 1 𝛾c = 1 𝛾c = 1 𝛾c = 1
1 0.83 1.52 1.73 0.97 0.95
2 0.99 1.65 2.07 1.09 1.01
3bis 0.96 1.58 2.01 1.09 –
4 (beam) 1.20 1.86 2.51 1.18 –
5 1.10 1.84 2.31 1.32 1.04
5bis 1.13 1.89 2.37 1.37 1.07
6 1.14 1.76 2.39 1.29 1.03
6bis 1.03 1.58 2.14 1.11 0.93
AVG 1.05 1.71 2.19 1.18 1.01
STD 0.12 0.14 0.25 0.14 0.06
5% percentile 0.88 1.54 1.83 1.01 0.93

underestimate predictions (AVG = 1.71; STD = 0.14). The EN 1992-1-1:2005 with values from the French
National Annexe (safety factors are equal to 1) give closer estimates for the shear resistance of a slab
under concentrated loads (AVG = 1.05; STD = 0.12). The results also show that the Model Code 2010 with
the second level of approximation (LoA II) predicts well the shear capacity of specimens. A reasonable
safety margin was also obtained (AVG = 1.18; STD = 0.14). However, the Model Code 2010 with the first
level of approximation (LoA I) seems to be very conservative (AVG = 2.19; STD = 0.25). Note that the
calculation for slab S3 with the punching failure mode was not conducted in this study.

5.  Finite element model (FEM)


The numerical investigations were conducted with the Abaqus software. All calculations were made dis-
placement controlled with an explicit solution technique. The failure of RC slabs is sudden, in particular
for failure in shear. This could cause numerical instability and convergence problems in the traditional
static analysis (implicit analysis) (ABAQUS Version 6.12 Documentation, 2013). Therefore, the explicit
method was used with very small load increments considering a so-called smooth-step function to
develop quasi-static analysis. In addition, the computing time corresponds to the critical increment
size and for FEM with a great number of degrees of freedom; it is considerably shorter in comparison
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   15

to the computing time under application of implicit solvers. The concrete damaged plasticity model
(CDP) in Abaqus was used for concrete and the elastic–plastic model was applied to reinforcement steel.

5.1.  Concrete damage plasticity model


The CDP model in Abaqus is based on the models proposed by Lubliner, Oliver, Oller, and Oñate (1989)
and Lee and Fenves (1998). The model proposed by Lubliner et al. (1989) for monotonic loading and
has been developed later by Lee and Fenves (1998) to consider the dynamic and cyclic loadings. The
constitutive theory, in this section, aims to capture the effects of irreversible damage associated with
the failure mechanisms that occur in concrete and other quasi-brittle materials under fairly low con-
fining pressures (less than four or five times the ultimate compressive stress in uniaxial compression
loading). The non-linear behaviour of concrete is attributed to the damage and plasticity processes.
The plasticity behaviour can be characterised by several phenomena such as strain softening, pro-
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

gressive deterioration. The damage process can be attributed to micro-cracking. Damage is associated
with the concrete’s failure mechanisms and therefore, results in a reduction in the elastic stiffness.
The model uses concepts of isotropic damaged elasticity in combination with isotropic tensile and
compressive plasticity to represent the inelastic behaviour of concrete i.e. tensile cracking and com-
pressive crushing.

5.2.  Concrete damage and plasticity parameters


5.2.1.  Concrete damage parameters
The stiffness degradation of material has been considered, in this model, for both tension and com-
pression behaviours by scalar-damage variables, respectively, dt and dc which can take value from
zero, representing the undamaged material, to one, which represents total loss of strength. The typical
stress–strain (or displacement) relationship under uniaxial tension and compression loading which are
characterised by CDP material model in ABAQUS are given below and illustrated in Figure 14. The equa-
tions for estimation of uniaxial stress–strain (displacement) curves based on tested material properties
will be described in the next sections.
𝜎t = (1 − dt )E0 (𝜀t − 𝜀̃pt ) (5)

𝜎c = (1 − dc )E0 (𝜀c − 𝜀̃pc ) (6)


where, E0 is the initial undamaged elastic modulus, 𝜀̃pt is the equivalent plastic strain in tension, 𝜀̃pc is the
equivalent plastic strain in compression.
In this study, the evolution of the concrete damage dt, dc was assumed in a linear relationship with
the inelastic strain εt and εc, respectively (simplified in linear form). In other words, for the definition
of the damage parameters simplified linear relationship was adopted by given the minimum damage
parameter equal to zero at the strain level ε0 and the maximum value at the strain level εu.
The choice of the damage properties (in the range 0  <  d  ≤  0.99) is important since, generally,
excessive damage may have a critical effect on the rate of convergence. It is recommended to avoid
using values of the damage variables above 0.99, which corresponds to a 99% reduction of the stiff-
ness (ABAQUS Version 6.12 Documentation, 2013). The maximum value for the damage parameters
in both tension and compression was chosen to be 0.9. This maximum value is recommended by
Genikomsou and Polak (2015). In the study of Genikomsou and Polak (2015), the results obtained from
the analysis considering the damage parameters displayed that the failure of the control specimen
happened earlier compared to the analysis results without considering damage parameters. Based on
the observations of the effect of the damage parameters, it can be said that the damage parameters
in the concrete damaged plasticity model in ABAQUS are similar to the hardening parameters used
in the classic plasticity theory.
16    T.-T. BUI ET AL.
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 14. Typical behaviour of concrete under axial compressive (a) and tension(b) strength in the CDP model (ABAQUS Version
6.12 Documentation, 2013).

5.2.2.  Concrete plasticity parameters


5.2.2.1.   Yield surface definition.  A Drucker–Prager criterion-based yield function is implemented in
the CDP model. The function is able to determine failure both by normal and shear stress. It is a pressure-
dependent criterion based on the two stress invariants of the effective stress tensor; the hydrostatic
pressure p̄ and the Mises equivalent stress q̄ . The equation below represents the implemented yield
function in terms of effective stresses,
1 ( ( ) ) ( )
F= q̄ − 3𝛼 p̄ + 𝛽 𝜀̌ pl 𝜎̂̄ max − 𝛾 − 𝜎̂̄ max − 𝜎̄ c 𝜀̌ pl = 0
1−𝛼
where
( )
𝜎b0
−1 𝜎c0
𝛼= ( ) 0 ≤ 𝛼 ≤ 0.5
𝜎
2 𝜎b0 − 1
c0

( )
𝜎̄ c 𝜀̌ plc
𝛽 = ( ) (1 − 𝛼) − (1 + 𝛼)
𝜎̄ t 𝜀̌ plt

( )
3 1 − Kc
𝛾=
2Kc − 1

�̄max is the maximum principal effective stress and σb0/σc0 is the user-specified ratio of the equibiaxial
𝜎
compressive yield stress and the initial uniaxial compressive yield stress, which by default is set to 1.16
according to (ABAQUS Version 6.12 Documentation, 2013). The equation, shows that the evolution of
the yield surface is controlled by the hardening variables 𝜀̌ plc and 𝜀̌ plt.
Kc is a user-defined parameter that depends on the stress invariants. It must be fulfilled that
0.5 ≤ Kc ≤ 1.0 and the factor is per default Kc = 2∕3 making the yield criterion approach Rankine’s
formulation. The difference of the yield surfaces in the deviatoric plane for Kc = 2∕3 and Kc = 1.0 is
shown in Figure 15. The biaxial yield surface in plane stress is also illustrated in Figure 15. In the CDP
model, the variation of the parameters describing the yield function had no significant influence on
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   17
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 15. Drucker–Prager yield criteria in the deviatoric plane for different Kc (Kc = 2/3 corresponds to the Rankine formulation and
Kc = 1 corresponds to the Drucker–Prager criterion.

the shear behaviour. The σb0/σc0 and Kc parameters were taken according to Lubliner et al. (1989),
described in Table 5.

5.2.2.2.  Flow potential function.  In CDP model, the Drucker–Prager hyperbolic plastic potential
function is used as illustrated in Figure 16(a). The plastic flow in the function, G, is a non-associated
flow rule that means it is not matching with the yield surface (Figure 16(b)). In this case, the plastic
flow develops along the normal to the plastic flow potential and not to the yield surface. The plastic
potential function is defined as:


( )2
G= ∈ ft0 tan𝜓 + q̄ 2 − ptan𝜓
̄ (13)

The dilation angle ψ is used as a material parameter in Abaqus. This parameter controls the amount
of plastic volumetric strain developed during plastic shearing and is assumed constant during plastic
yielding. It is measured in the p−q plane as the inclination angle of the plastic potential function, for
high confining pressure. Low values of the dilation angle will produce brittle behaviour, while higher
values will produce more ductile behaviour. Contrary to the parameters describing the yield function
which had no significant influence on the shear behaviour, the shear capacity was clearly influenced
by the variation of the parameters defining the flow rule, particularly by the variation of the dilation
angle. A parameter study of 20°–40° variation shows that the shear capacity increased with increasing
dilation angle due to increasing ductility (Malm, 2006). A 37° dilatation angle value was advised by
Reissen and Hegger (2013). In our study in the next sections the dilation angle will be calibrated. The
flow potential eccentricity parameter ∈ varied between 0.01 and 1.0, resulting in an increase in shear
capacity as the eccentricity parameter increased. This parameter defines the rate at which the function
approaches the asymptote (the flow potential tends to a straight line as the eccentricity tends to zero).
The default value is 0.1 (ABAQUS Version 6.12 Documentation, 2013), meaning that the material has
almost the same dilatation angle over a wide range of configuring pressure stress values.
As the uniaxial behaviour of material models defines the evolution of the yield criterion in a
FE-analysis, the definition of material parameters and uniaxial material behaviour curves becomes
more important. In our study, the concrete strength (compressive and tensile strength) and the concrete
elastic stiffness measured from the experiments was used in the numerical model, while the uniaxial
behaviour curves of concrete in both tension and compression are obtained using models proposed
in the literature which are given in the following sections.
18    T.-T. BUI ET AL.

Table 5. Concrete damaged plasticity parameters.

Parameter Description Default value


Ψ Dilation angle User-defined
∈ Flow potential eccentricity 0.1
σb0/σc0 Ratio of initial equibiaxial compressive yield stress to initial uniaxial compressive yield stress 1.16
Kc Ratio of the second stress invariant on the tensile meridian to that on the compressive 0.6667
meridian at initial yield for any given value of the pressure invariant such that the maximum
principal stress is negative
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 16. (a) The Drucker–Prager hyperbolic plastic potential function in p–q plane, illustration of dilation angle and eccentricity;
(b) Illustration of the plastic potential in relation to a yield surface.

5.2.2.3. Synthesis.  The synthesis of the recommended values of the concrete damaged plasticity
parameters are presented in the Table 5, except the dilation angle value which will be calibrated in
the next section.

5.3.  Mechanical behaviour


5.3.1.  Concrete tension behaviour
The tensile response of RC is modelled using a non-linear tension stiffening model. Tension stiffening
is influenced by the reinforcement ratio, cracking spacing and the quality of the bond between the
concrete and reinforcement. Before the occurrence of cracks, the tension behaviour of concrete was
assumed to be linear elastic and can be defined only by initial elastic modulus and peak tensile stress.
The post-peak behaviour of concrete in tension can be characterised by a stress-crack displacement
response or a stress–strain relationship. There are three main stress–displacement relationship options
for modelling the concrete tensile behaviour: exponential, bilinear and linear. The exponential rela-
tionship used in our study and judged by Barr (2003) is the most accurate model for post-peak tension
softening behaviour is that experimentally derived by Cornelissen, Hordijk, and Reinhardt (1986), see
Figure (17) and is expressed as below. The models are based on a fracture energy criterion to model
the post-peak branch. The stress–displacement relationship is expressed as:
[ ( )3 ]
𝜎 w −c w w( )
= 1 + c1 e 2 wc − 1 + c13 e−c2 (7)
ft w c w c

G
where, c1 = 3; c2 = 6.93, wc = 5.14 f f .
w is crack opening displacement; wc is crack opening displacement at the complete release of stress
t

or fracture energy; ft is concrete uniaxial tensile strength; Gf is the fracture energy required to create a
stress-free crack over a unit area; and c1 = 3.0 and c2 = 6.93 are constants determined from tensile tests
of concrete. In this study, Gf could be calculated according to CEB-FIP model code 1990 (FIB, 1990) in
the following manner:
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   19
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 17.  Tension softening behaviour suggested by (Cornelissen et al., 1986); (b) tensile damage parameter–crack opening
relationship for concrete.

( )0.7
fc
Gf = Gf 0 if fc ≤ 80 MPa (8)
10

Gf = 4.3Gf 0 if fc > 80 MPa (9)


with fc, ft in MPa; Gf in Nmm/mm2; w, wc in mm.
Gf 0 (Nmm/mm2) is the base value of the fracture energy. It is a coefficient related to maximum
aggregate size, as shown in Table 6. By polynomial interpolation based on the data, the value of Gf 0
can be calculated as below:
( 2
)
Gf 0 = 0.00005dmax − 0.0005dmax + 0.026
with dmax in millimetres and Gf 0 as a unit of N/mm. Consequently, the value of Gf 0 is 0.036 N/mm for all
specimens as the maximum aggregate size dmax is equal to 20 mm.

5.3.2.  Concrete compression behaviour


In this study, the concrete stress–strain law in compression suggested by Feenstra et al. (1993) has been
implemented. The post-peak behaviour is defined by a parabolic equivalent stress-equivalent strain
diagram and based on the fracture energy. The end point of the softening curve is defined by means
of the ultimate equivalent strain εu. Before peak stress, the stress–strain relation can be divided into
two domains. The model was first represented by a linear-elastic branch ending at a stress level of 1/3
fcm. Then, this linear-elastic branch is followed by a second ascending branch up to the peak stress.
The formulation of the equivalent stress reads:
� �
⎧ fc 1 + 4 𝜀̄ − 2 𝜀̄2 if 𝜀̄ < 𝜀
⎪ 3
𝜎̄ = ⎨ � � e �2 �e
𝜀 𝜀 2 e

𝜀−𝜀
̄ e (10)
⎪ fc 1 − 𝜀u −𝜀e if 𝜀e < 𝜀̄ < 𝜀u

Table 6. Gf0 depending on dmax (Rots, 1988).

dmax (mm) Gf0 (Nmm/mm2)


8 0.025
16 0.030
32 0.058
20    T.-T. BUI ET AL.

The maximum compressive strength will be reached at an equivalent strain εe, which is determined
irrespective of element size or compressive fracture energy and reads:
4 fc
𝜀e = (11)
3 Ec
The maximum equivalent strain εu is related to the compressive fracture energy and the element size
and reads:
Gc 11
𝜀u = 1.5 − 𝜀 (12)
hfc 48 e
The pre-peak energy has been taken into account with the correction factor 11 𝜀 in the above equation.
48 e
A possible snap-back on the constitutive level if the equivalent length becomes too large has been
avoided by the assumption that the ultimate equivalent strain εu is limited by: εu ≥ 1.75εe. The total
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

compressive fracture energy found in the experiments ranges from 10 to 25 [Nmm/mm2], which is
about 50–100 times the tensile fracture energy: Gc = 50–100 Gf (Feenstra et al., 1993) (see Figure 18).

6.  Validation of FEM by experimental results


6.1.  Identify of parameters
The slabs were modelled using eight-node volume elements (C3D8I) and only half of the slab was mod-
elled using symmetry in the transverse direction. The mesh was discretised with a fine mesh measuring
approximately 15 mm, which reached mesh convergence. Figure 19 shows the mesh of one of the slabs
modelled. The slab’s boundary conditions are simply supported edges, which ensured a shear span ratio
av/d = 2 (av is the horizontal distance of the section from the face of support; d is the effective depth
of the slab). Modelling the boundary conditions was simplified by supporting the slab directly on the
rigid steel tube. The RC beam was not modelled. The pertinence of this simple model was confirmed
by the numerical results obtained.
To model the reinforcement, two-node truss elements (T3D2) were used. The reinforcement bars
were embedded in the concrete, simulating a perfect bond by linking the nodal degrees of freedom.
Assuming a perfect bond between reinforcements and concrete is in fact an idealisation of the inter-
action between these two materials, which in some cases could influence the predictions of load–dis-
placement curves and cracking patterns (Xenos & Grassl, 2014). This is, for example, the case of beams
reinforced with stirrups (Chen, Chen, & Teng, 2012). It would then be necessary to use a concrete-steel
bond model in such a case. This aspect is more complex and requires further investigation. However,

Figure 18. Compression and softening model (Feenstra et al., 1993).


EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   21

(a) (b)

Figure 19. (a) Mesh of symmetric RC slab S2; (b) Transparent perspective view of RC slab S2.
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

400
Slab S5 - Dilation angle
350

300

250
Load (kN)

200
Slab S5 - EXP
150
20 degrees
100 30 degrees
50
37 degrees
45 degrees
0
0 1 2 3 4
Displacement (mm)

Figure 20. The load–deflection response of slab S5 under influence of the dilation angle.

5.00
EC2
4.50 France
Vexperimental/Vcalculated

4.00 Model Code 2010 (LoA I)


Model Code 2010 (LoA II)
3.50
FEM
3.00
2.50
2.00
1.50
1.00
0.50
0.00
S1 S2 S3bis S4 S5 S5bis S6 S6bis

Figure 21. Comparison of the experiment, numerical simulation (FEM) and design codes (EC2, FNA, Model Code) for shear resistance
(with fcm and γc = 1).

in (Jendele & Cervenka, 2006), it has been shown that the adoption of an elaborate bond model could
lead to numerical stability problems and would not significantly influence the results if a fine mesh
is used. The elastic–plastic material behaviour was assumed for the reinforcement. The yield strength
fy = 567 MPa and the ultimate strength fu = 672 MPa were used for the reinforcement behaviour.
22    T.-T. BUI ET AL.

(a) (b)
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

(c) (d)

Figure 22. Experimental and numerical load/deflection curves for S1, S2, S5, S6.

(a) (b)

Figure 23. Crack pattern at peak load observed by maximum principal strain of slab S1 (L = 100 cm): (a) in symmetry axis; (b) bottom
surface.

6.2.  Calibration of the dilation angle ψ


As discussed before, the dilation angle measured in the p−q plane as the inclination angle of the plastic
potential function may have a significant influence on the prediction of the shear capacity of the FEM.
Consequently, the effect of dilation angle on the behaviour of model was studied. Figure shows the
influence of different values of dilation angle on the load–displacement response of slab S5 chosen as
a control slab. The results show that in this model, to achieve a good agreement between the experi-
mental results and FE analysis, the dilation angle can be taken as 37° (Figure 20).
The other parameters used in the models were presented in the Table 5 above (Section 5.2).
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   23

(a) (b)
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Figure 24. Crack pattern at peak load observed by maximum principal strain of slab S2 (L = 100 cm): (a) in symmetry axis; (b) bottom
surface.

(a) (b)

Figure 25. Crack pattern at peak load observed by maximum principal strain of slab S6 (L = 40 cm): (a) in symmetry axis; (b) bottom
surface.

6.3.  Numerical results


Based on the proposed numerical configurations presented above, non-linear finite element analysis
of the test specimens was conducted. The comparison between the ultimate loads from experimental
tests and numerical analysis is shown in Table 4 and Figure 21 above. It was found that the predictions
of collapse loads from FEM showed good correlations with the experimental results.
In the comparison between the load/deflection curves in experimental tests and FEM as shown
in Figure 22, it can be seen that the numerical model adopted was able to predict the experimental
responses. The trends of the load–displacement responses in experimental tests and numerical analysis
are similar. However, Figure 22(a) and (b) shows that the prediction in FEM is slightly stiffer than that
in experimental tests (S1, S2). This could be due to the small sliding of the RC beams supported, which
were based directly on the laboratory floor in the S1 and S2 tests. In modelling, this effect was not taken
into account because the boundary conditions were simplified by modelling the single-support edge. In
the S5 and S6 tests, there was a mortar layer between the RC beams supported and the laboratory floor,
which helped to reduce this sliding. As a result, the most accurate prediction of the structural response
between the experimental and numerical results was found in the S5 and S6 tests (Figure 22(c) and
(d)). Figures 23–25 shows the cracking patterns obtained from the FE analysis. In the numerical models,
the crack patterns were visualised through the maximum principal plastic strains which give a good
representation of the cracks. The X–Z plane view in these figures shows a shear mechanism in which
the diagonal shear crack inducing failure is originating from the loading point towards the support
24    T.-T. BUI ET AL.

and corresponds to a shear failure. Compared to the crack patterns observed in the experimental tests,
these figures show the ability of the proposed non-linear FE model to accurate predict the locations
and the directions of the crack propagations.
The numerical results show that the proposed model is able to predict the shear behaviour of con-
crete slabs without shear reinforcement in terms of shear strength and failure mode. Consequently,
the proposed model can be used in future parametric studies on various aspects influencing the shear
capacity of concrete slabs.

7.  Conclusions and prospects


This paper contributes data for RC slab behaviour subjected to a concentrated load. The study addressed
the case of loading close to supports with a ratio of av/d = 2. Experiments were conducted to study
the influence of the boundary conditions, the loading plate length, and the transverse reinforcement
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

ratio. The failure mechanisms and the load–deflection curves including the failure loads were discussed
and compared. Numerical modelling using the finite element method was investigated to analyse and
compare with the experiment results. The following conclusions can be drawn from this study:

• The sensitivity of the boundary conditions was observed. Slab stiffness of the load/displacement
response was improved in two phases (elastic and then post-elastic) when the slab was supported
on four sides in comparison with the slab supported on two sides. However, no significant change
in ultimate shear strength was found. The failure modes obtained were developed in two different
ways. For the slab supported on four sides, the cracks on the bottom face seemed to be limited in
the transverse direction and developed mainly in the longitudinal direction. Nevertheless, for the
slab supported on two sides, the cracks tended towards two lateral sides and to being distributed
uniformly in two directions: longitudinal and transverse.
• The shear capacity increases as the loading plate lengthens. The behaviour in shear of these slabs
loaded with a concentrated load was found different from the behaviour of the two-dimensional
beam. For the beam, cracks develop only perpendicular to the span direction, while in slabs the
cracks develop perpendicular and parallel to the span direction, indicating the ability of the slab
to distribute the load over the longitudinal and transverse directions. The influence of the loading
plate length for slabs supported on four sides requires further study. With a greater loading length,
the effective width could be greatly influenced by the presence of the two lateral sides.
• The results calculated using the EC2 approach with the value proposed by FNA combined with
the assumption of the effective width give the best correlation with the experimental results.
The effective width was assumed to spread to 45° disseminated from the far corners of the load.
• The FEM numerical model proposed is capable of predicting the shear behaviour of RC slabs under
a concentrated load such as the load/displacement response and the failure modes.
• The differences between experimental and NLFEA results are related to sliding effects but maybe
also related to the rotation of two RC beams on two supports. In a future work, this issue shall
be fixed and presence of the RC beams should be simulated on the basis of realistic boundary
conditions of the slab.

Nomenclature

av  clear shear span: face-to-face distance between the load and the support
a  horizontal distance from the axis of the load to the axis of the support
As  tensile reinforcement (mm2)
d  effective depth of the cross-sections
Es  modulus of elasticity of the reinforcing steel
ρl  ratio of flexural reinforcement in the longitudinal direction
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   25

ρt  ratio of flexural reinforcement in the transverse direction


bw  smallest width of the cross-section in the tensile area
beff  effective width in shear
fck  nominal characteristic cylinder compressive strength (MPa)
fcm  measured cylinder compressive strength of the concrete at the age of testing
fctm  measured cylinder tensile strength of the concrete at the age of testing
Pfailure  measured peak load in an experiment
εx  strain at mid-depth of effective shear depth
MEd  design bending moment at the control section
VEd  design shear force at the control section
Vfailure  shear force at failure in the experiment
VRd,c  design shear resistance attributed to the only concrete contribution
Vu,EC2  shear capacity calculated according to NEN-EN 1992-1-1:2005
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Vu,FNA  shear capacity calculated according to French National Annexe


Vu,MC  shear capacity calculated according to Model Code 2010
VFEM  shear force at failure obtained in the finite element analysis
Kc  parameter defining the shape of the yield surface in the deviatoric plane
Ψ  concrete dilation angle
Gf  tensile concrete fracture energy
Gc  compressive concrete fracture energy
wc  maximum tensile concrete crack opening width
εe  compressive concrete equivalent strain at peak load
εu  compressive concrete ultimate equivalent strain
dmax  concrete maximum aggregate diameter
L  loading plate length

Acknowledgements
This work, carried out by INSA Lyon, was funded by contract from FFB (Fédération Française du Bâtiment) and EDF Septen.
The authors gratefully acknowledge Mr André Coin and Mr Henry Thonier for their contribution to this project.

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
This work was supported by the INSA Lyon, was funded by contract from FFB (Fédération Française du Bâtiment) and
EDF Septen.

ORCID
Tan-Trung Bui   http://orcid.org/0000-0002-8279-7230

References
318, A. C. (2008). Building code requirements for structural concrete (ACI 318-08) and Commentary. Farmington, MI: American
Concrete Institute.
ABAQUS Version 6.12 Documentation. (2013). ABAQUS Analysis user’s manual 6.12-EF, Dassault Systems Simulia Corp.,
Providence, RI, USA.
ASCE-ACI Committee 445 on Shear and Torsion (1998). Recent approaches to shear design of structural concrete. Journal
of Structural Engineering, 124, 1375–1417. doi:10.1061/(ASCE)0733-9445(1998)124:12(1375)
26    T.-T. BUI ET AL.

Barr, B. I. G. (2003). Failure of concrete structures. In I. Milne, R. O. Ritchie, & B. Karihaloo (Eds.), Comprehensive structural
integrity (pp. 241–269). Oxford: Pergamon. doi:10.1016/B0-08-043749-4/01118-6
Belletti, B., Damoni, C., Hendriks, M. A. N., & de Boer, A. (2014). Analytical and numerical evaluation of the design shear
resistance of reinforced concrete slabs. Structural Concrete, 15, 317–330. doi:10.1002/suco.201300069
Belletti, B., Pimentel, M., Scolari, M., & Walraven, J. C. (2015). Safety assessment of punching shear failure according to the
level of approximation approach. Structural Concrete, 16, 366–380. doi:10.1002/suco.201500015
Belletti, B., Walraven, J. C., & Trapani, F. (2015). Evaluation of compressive membrane action effects on punching shear
resistance of reinforced concrete slabs. Engineering Structures, 95, 25–39. doi:10.1016/j.engstruct.2015.03.043
Belletti, B., Cantone, R., Fernández Ruiz, M., & Muttoni, A. (2016). Approaches for suitable modelling and strength prediction
of reinforced concrete slabs. 2016 fib Symposium. Cape Town. Retrieved from https://infoscience.epfl.ch/record/226740
Bui, T. T., Abouri, S., Limam, A., NaNa, W. S. A., Tedoldi, B., & Roure, T. (2016). Experimental investigation of shear strength
of full-scale concrete slabs subjected to concentrated loads in nuclear buildings. Engineering Structures, 131, 405–420.
doi:10.1016/j.engstruct.2016.10.045
Cantone, R., Belletti, B., Manelli, L., & Muttoni, A. (2016). Compressive membrane action effects on punching strength of
flat RC slabs. Key Engineering Materials, 711, 698–705. doi:10.4028/www.scientific.net/KEM.711.698
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

Chen, G. M., Chen, J. F., & Teng, J. G. (2012). On the finite element modelling of RC beams shear-strengthened with FRP.
Construction and Building Materials, 32, 13–26. doi:10.1016/j.conbuildmat.2010.11.101
Collins, M. P., Bentz, E. C., & Sherwood, E. G. (2008). Where is shear reinforcement required? Review of research results and
design procedures. ACI Structural Journal, 105, 590–600. doi:10.14359/19942
Cornelissen, H. A. W., Hordijk, D. A., & Reinhardt, H. W. (1986). Experimental determination of crack softening characteristics
of normalweight and lightweight concrete. HERON, 31. Retrieved from https://resolver.tudelft.nl/uuid:08c29b39-5c60-
4ab6-b9d5-643d11007f7c
Elstner, R. C., & Hognestad, E. (1956). Shearing strength of reinforced concrete slabs. ACI Journal Proceedings, 53, 29–58.
doi:10.14359/11501
European Committee for Standardisation. (2005). EN 1992-1-1:2005 Eurocode 2: Design of concrete structures – Part 1-1:
General rules and rules for buildings. Retrieved from https://archive.org/details/en.1992.1.1.2004
FD P 18-717. (2013). Eurocode 2 – Calcul des structures en béton – Guide d’application des normes NF EN 1992.
Feenstra, P. H., de Borst, R., & TU Delft: Civil Engineering and Geosciences, & TU Delft, Delft University of Technology. (1993,
November 29). Computational aspects of biaxial stress in plain and reinforced concrete. Delft University Press. Retrieved
from https://resolver.tudelft.nl/uuid:faf2fd16-1c43-4711-b783-9e8e00d10c21
FIB. (1990). FIB model code for concrete structures 1990. Wiley-VCH Verlag GmbH & Co. KGaA, (pp. 1–400). Retrieved from
http://www.fib-international.org/ceb-fip-model-code-90-pdf.
Fib. (2013). FIB model code for concrete structures 2010 (pp. 190–350). Wiley-VCH Verlag GmbH & Co. KGaA. Retrieved from
https://onlinelibrary.wiley.com/doi/10.1002/9783433604090.ch7/summary
Genikomsou, A. S., & Polak, M. A. (2015). Finite element analysis of punching shear of concrete slabs using damaged plasticity
model in ABAQUS. Engineering Structures, 98, 38–48. doi:10.1016/j.engstruct.2015.04.016
Grassl, P., & Jirásek, M. (2006). Damage-plastic model for concrete failure. International Journal of Solids and Structures, 43,
7166–7196. doi:10.1016/j.ijsolstr.2006.06.032
Gurutzeaga, M., Oller, E., Ribas, C., Cladera, A., & Marí, A. (2014). Influence of the longitudinal reinforcement on the shear
strength of one-way concrete slabs. Materials and Structures, 48,  1–16, doi:10.1617/s11527-014-0340-5
Hansen, N. R., & Schreyer, H. L. (1994). A thermodynamically consistent framework for theories of elastoplasticity coupled
with damage. International Journal of Solids and Structures, 31, 359–389. doi:10.1016/0020-7683(94)90112-0
Jendele, L., & Cervenka, J. (2006). Finite element modelling of reinforcement with bond. Computers & Structures, 84, 1780–
1791. doi:10.1016/j.compstruc.2006.04.010
Kachanov, L. M. (1986). Introduction to continuum damage mechanics (Vol. 10). Dordrecht: Springer. doi:10.1007/978-94-
017-1957-5
Kani, G. N. J. (1964). The riddle of shear failure and its solution. ACI Journal Proceedings, 61, 441–468. doi:10.14359/7791
Lantsoght, E. O. L. (2013, June 14). Shear in reinforced concrete slabs under concentrated loads close to supports (dissertation).
Delft University of Technology.
Lantsoght, E. O. L., van der Veen, C., & Walraven, J. C. (2012). Shear capacity of slabs and slab strips loaded close to the
support. Special Publication, 287, 1–18.
Lantsoght, E. O. L., van der Veen, C., & Walraven, J. C. (2013). Shear in one-way slabs under concentrated load close to
support. ACI Structural Journal, 110. Retrieved from https://trid.trb.org/view.aspx?id=1245253
Lantsoght, E. O. L., Van der Veen, C., Walraven, J. C., & De Boer, A. (2013, Winter). Recommendations for the shear assessment
of reinforced concrete slab bridges from experiments. Experiments Structural Engineering International. Retrieved from
https://resolver.tudelft.nl/uuid:dde7c3e6-e500-446e-b22d-45480d63da09
Lantsoght, E. O. L., de Boer, A., van der Veen, C., & Walraven, J. C. (2015). Effective shear width of concrete slab bridges.
Proceedings of the Institution of Civil Engineers – Bridge Engineering, 168, 287–298. doi:10.1680/bren.13.00027
Lee, J., & Fenves, G. (1998). Plastic-damage model for cyclic loading of concrete structures. Journal of Engineering Mechanics,
124, 892–900. doi:10.1061/(ASCE)0733-9399(1998)124:8(892)
EUROPEAN JOURNAL OF ENVIRONMENTAL AND CIVIL ENGINEERING   27

Lubell, A., Bentz, E., & Collins, M. (2009). Influence of longitudinal reinforcement on one-way shear in slabs and wide beams.
Journal of Structural Engineering, 135, 78–87. doi:10.1061/(ASCE)0733-9445(2009)135:1(78)
Lubell, A. S., Bentz, E. C., & Collins, M. P. (2009). Shear reinforcement spacing in wide members. ACI Structural Journal, 106,
205–214. doi:10.14359/56359
Lubliner, J., Oliver, J., Oller, S., & Oñate, E. (1989). A plastic-damage model for concrete. International Journal of Solids and
Structures, 25, 299–326. doi:10.1016/0020-7683(89)90050-4
Malm, R. (2006). Shear cracks in concrete structures subjected to in-plane stresses. Retrieved from https://www.diva-portal.
org/smash/record.jsf?pid=diva2:11277
Massonnet, C., Olszak, W., & Phillips, A. (1979). Plasticity in structural engineering, fundamentals and applications. Vienna:
Springer. doi:10.1007/978-3-7091-2902-9
Nana, W. S. A., Bui, T. T., Limam, A., & Abouri, S. (2017). Experimental and numerical modelling of shear behaviour of full-
scale RC slabs under concentrated loads. Structures, 10, 96–116. doi:10.1016/j.istruc.2017.02.004
Natário, F., Fernández Ruiz, M., & Muttoni, A. (2014). Shear strength of RC slabs under concentrated loads near clamped
linear supports. Engineering Structures, 76, 10–23. doi:10.1016/j.engstruct.2014.06.036
Regan, P. E., & Rezai-Jorabi, H. (1988). Shear resistance of one-way slabs under concentrated loads. ACI Structural Journal,
Downloaded by [Australian Catholic University] at 20:00 28 September 2017

85, 150–157. doi:10.14359/2704


Reißen, K., & Hegger, J. (2013a). Experimental investigations on the effective width for shear of single span bridge deck
slabs. Beton- Und Stahlbetonbau, 108, 96–103. doi:10.1002/best.201200064
Reißen, K., & Hegger, J. (2013b). Experimental investigations on the shear-bearing behaviour of bridge deck cantilever
slabs under wheel loads. Beton- Und Stahlbetonbau, 108, 315–324. doi:10.1002/best.201200072
Reissen, K., & Hegger, J. (2013). Numerical investigations on the shear capacity of reinforced concrete slabs under
concentrated loads. In Research and applications in structural engineering, mechanics and computation (Vol. 1–0, pp.
1507–1512). CRC Press. Retrieved from https://www.crcnetbase.com/doi/abs/10.1201/b15963-271
Rombach, G., & Kohl, M. (2013). Shear design of RC bridge deck slabs according to Eurocode 2. Journal of Bridge Engineering,
18, 1261–1269. doi:10.1061/(ASCE)BE.1943-5592.0000460
Rots, J. G. (1988). Computational modeling of concrete fracture. Retrieved from https://repository.tudelft.nl/islandora/object/
uuid%3A06985d0d-1230-4a08-924a-2553a171f08f
Sherwood, E. G., Bentz, E. C., & Collins, M. P. (2007). Effect of aggregate size on beam-shear strength of thick slabs. ACI
Structural Journal, 104, 180–190. doi:10.14359/18530
Sherwood, E. G., Lubell, A. S., Bentz, E. C., & Collins, M. P. (2006). One-way shear strength of thick slabs and wide beams. ACI
Structural Journal, 103. 794–802. doi:10.14359/18229
Timoshenko, S., & Woinowsky-Krieger, S. (1959). Theory of platesand shells, (540 pages). McGraw-Hill.
Theodorakopoulos, D. D., & Swamy, R. N. (2002). Ultimate punching shear strength analysis of slab–column connections.
Cement and Concrete Composites, 24, 509–521. doi:10.1016/S0958-9465(01)00067-1
Vaz Rodrigues, R., Fernández Ruiz, M., & Muttoni, A. (2008). Shear strength of R/C bridge cantilever slabs. Engineering
Structures, 30, 3024–3033. doi:10.1016/j.engstruct.2008.04.017
Xenos, D., & Grassl, P. (2014). Comparison of nonlocal and crack-band damage-plasticity approaches for modelling the
failure of reinforced concrete structures. In Edited by Nenad Bícanić, Herbert Mang, Günther Meschke and René de Borst.
Computational modelling of concrete structures (Vol. 1–0, pp. 513–520). CRC Press. doi:10.1201/b16645-57

You might also like