You are on page 1of 9

Engineering Structures 29 (2007) 2233–2241

www.elsevier.com/locate/engstruct

Investigations on the punching behaviour of reinforced concrete footings


J. Hegger a , M. Ricker a,∗ , B. Ulke b , M. Ziegler b
a RWTH Aachen University, Institute of Structural Concrete, Mies-van-der-Rohe-Str. 1, D-52074 Aachen, Germany
b RWTH Aachen University, Geotechnical Engineering, Mies-van-der-Rohe-Str. 1, D-52074 Aachen, Germany

Received 24 July 2006; received in revised form 29 October 2006; accepted 12 November 2006
Available online 26 December 2006

Abstract

The punching shear capacity of footings varies significantly for different codes. One reason is that the amount of the soil reaction to be
deducted from the punching load differs from one code to another. The aim of the present investigation is to develop a design model for the
punching strength of footings taking into account the soil-structure-interaction. The results of five punching tests on reinforced concrete footings
supported on soil are presented. The experimental results indicate that the angle of the shear failure plane is steeper than observed in punching
tests on flat slabs and the shear slenderness seems to affect the punching shear capacity significantly. Based upon the findings and a test data bank
an advanced design model is derived, which is based on the BS 8110-1:1997 provisions.
c 2006 Elsevier Ltd. All rights reserved.

Keywords: Footings; Punching shear capacity; Reinforced concrete; Soil-structure-interaction; Soil pressure redistribution

1. Introduction

Since the beginning of the 20th century only few punching


tests on footings have been performed. Since the experimental
study of a footing under realistic boundary conditions is
associated with considerable expenditure, most researchers
avoid using real soil in their experiments. The aim of the present
investigation is to derive a design model for footings taking into
account the soil stress distribution underneath the footing.

2. Soil stress distribution under footings

The bearing capacity of foundations is influenced by the


Fig. 1. Soil pressure distribution under a rigid footing according to (a)
stress distribution underneath the footing. For the calculation Boussinesq and (b) Prandtl–Buisman.
of the punching shear capacity, the design codes suggest a
reduction of the shear force by the soil pressure within the correct as long as the soil behaviour is assumed to be elastic
control perimeter. The assumption of a uniform, or linear isotropic.
stress distribution is commonly accepted. However, the stress If the load increases, the soil will plasticize under the
distribution depends on the magnitude of the effective load. In foundation edges and the stresses will be redistributed towards
1885, Boussinesq [1] showed that for rigid, vertically loaded the centre of the footing. This redistribution is completed when
and rotation-symmetric foundations a concave stress distribu- the bearing capacity is achieved. At limit state, the stress
tion underneath the footing develops (Fig. 1(a)). This theory is distribution is convex (Fig. 1(b)) according to the theory of
Prandtl–Buisman. For a convex stress distribution, the soil
∗ Corresponding author. Tel.: +49 241 80 26330; fax: +49 241 80 22335. reaction underneath the punching cone is greater than for
E-mail address: mricker@imb.rwth-aachen.de (M. Ricker). a uniform one. This results in a more economic design.

0141-0296/$ - see front matter c 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2006.11.012
2234 J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241

The common approximation methods for the calculation of 2d 2d


v Rd,c = C Rd,c · k · (100 · ρl · f ck )1/3 · ≥ vmin · (2)
the soil pressure distribution in soil mechanics, for example aEC2 aEC2
the classical Winkler solution or the coefficient of subgrade
where
reaction method, simulate the ground as an elastic, isotropic
aEC2 = distance from the periphery of the loaded area to the
half-space with a system of linear springs. These methods
control perimeter considered;
do not sufficiently describe the redistribution of the soil
C Rd,c = 0.18/γc ; with γc being the material resistance factor
pressure towards the centre of a footing at an increased load
for concrete
q (1.5);
level. The problem of calculating the punching shear capacity 200
of footings is in assessing the soil stress distribution and k =1+ d ≤ 2.0 (d in mm) size factor;
1/2
the soil reaction force. Thus, only nonlinear finite element vmin = 0.035 · k 2/3 · f ck .
analyses (not discussed in this paper) as well as experimental
examinations represent the only possibility to determine these 4. Experimental program
effects realistically.
Five reinforced concrete footings were tested under realistic
3. Design codes boundary conditions. The notations DF1 to DF5 will be used
for the test specimens. Principal variables between the test
The design codes allow a part of the soil reaction to be specimens were the compactness and stiffness of the sand, the
subtracted from the punching load. However, the amount to shear span ratio, and the shear reinforcement.
be deducted differs from one code to the other. Most design
codes with the exception of Eurocode 2 and Model Code 90 do 4.1. Test specimens
not differentiate between the treatment of the punching shear
strength of flat slabs and footings. The same equations are The dimensions of the test specimens were chosen to model
used. In the following, the provisions of BS 8110-1:1997 and an ordinary footing at a scale of 1/2–1/3 and to fit into the
Eurocode 2 are briefly described. experimental sandbox. All tested footings had a footprint of
900 mm × 900 mm and were designed to fail in punching. The
dimensions and reinforcement details of a typical test specimen
3.1. BS 8110-1:1997 [2]
are shown in Fig. 2. Full details of the tested footings are given
in Table 1. The square column stubs were cast monolithically at
The factored design shear stress v is calculated along the
the centre of the slabs.
critical shear section at a distance of 1.5d from the boundaries
The footing DF3 included heavy shear reinforcement
of the loaded area. The control perimeter is of a rectangular
consisting of vertical bars with a diameter of 12 mm and a yield
shape. A reduction of the shear force by the effective soil
strength of 548 MPa. At the top and the bottom of the slab the
pressure within the control perimeter is allowed. The soil
bar ends were anchored by welded plates because the height of
pressure may be assumed to be uniformly distributed. Provided the slab was only 200 mm. The footings DF4 and DF5 were
that the shear force along the control perimeter is less than vc more compact and were designed to investigate the effect of a
(Eq. (1)) no shear reinforcement is required. small shear slenderness λ = a/d on the punching behaviour.
As 1/3 400 1/4
   
0.79
vc = · 100 · · (1) 4.2. Material properties
γm bv · d d

where 100 · bAv ·d


s
≤ 3; 400 Ready-mixed concrete with a maximum coarse aggregate
d ≥ 1. size of 16 mm and ordinary CEM III A 32.5 N Portland cement
For a characteristic concrete strength greater than 25 MPa,
were used in all footings. Table 1 summarizes the properties
the value vc may be multiplied by ( f cu /25)1/3 . The value
of the concrete used. Steel bars BSt 500 (A) with a measured
of f cu should not be taken as greater than 40 MPa. The
maximum possible shear stress that can be supported is given by
1/2
vmax ≤ 0.8 · f cu or 5 MPa along the column face.

3.2. Eurocode 2 [3]

The punching resistance of footings should be verified at


control perimeters within 2.0d from the periphery of the col-
umn. The lowest value of resistance at the different sections
controls the design. Our own calculations showed that the low-
est values are found at a distance within 0.55d and 1.75d from
the column face. The soil pressure within the control perimeter
is allowed to be deducted from the punching load. The punch-
ing shear capacity of the concrete v Rd,c is calculated as: Fig. 2. Dimensions and reinforcement of footings DF1 and DF2.
J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241 2235

Table 1
Details of the test specimens

ρl
Footing d (mm) c (mm) λ (−) f c (MPa) f ct,sp (MPa) f ct,flex (MPa) E c (GPa) Bar size (mm) f y (MPa)
(ρl0 ) (%)
1.03
DF1 150 150 2.5 20.2 1.63 2.81 24.0 14 552
(−)
1.03
DF2 150 150 2.5 22.0 1.76 3.02 22.6 14 552
(−)
1.03
DF3a 150 150 2.5 30.7 2.33 3.59 25.8 14 552
(0.33)
0.62
DF4 250 150 1.5 24.5 1.97 3.89 24.0 14 552
(−)
0.73
DF5 250 175 1.45 17.6 1.51 3.30 23.3 12/16 515
(−)
f c : Cylinder compressive strength; f ct,sp : Splitting tensile strength; f ct, f l : Flexural tensile strength; E c : Young’s modulus of concrete; ρl (ρl0 ): Flexural tensile
(compressive) reinforcement ratio; f y : Yield stress of steel.
a DF3 included shear reinforcement (A 2
sw = 4070 mm , s0 = 75 mm distance between the column face and the first row of shear reinforcement, sr = 112.5 mm
radial spacing between successive rows of shear reinforcement).

Fig. 3. Grain-size distribution of the sand.

yield stress of f sy = 552 MPa and a tensile strength of


f su = 634 MPa were used as reinforcement.
The sand used in this research is medium sand, washed,
dried, and sorted by particle size. It is composed of rounded
to subrounded particles. The grain-size distribution was
determined using the dry sieving method. The results are
revealed in Fig. 3. The friction angles of the sand determined
by consolidated undrained triaxial shear tests varied between Fig. 4. Section of the test setup.
33.5◦ , and 39.4◦ . The friction angles were determined at the
same relative densities as used in the different tests. The 0.707 for test DF5. Thus, all footings can be regarded as rigid
maximum and the minimum dry unit weights were found to compared to the soil. Table 2 summarizes the sand properties
be 17.9 and 14.4 kN/m3 and the corresponding values of in detail including the system rigidity ks and the oedometric
the minimum and the maximum void ratios were 0.482 and modulus E s .
0.842. The experimental tests were conducted on sand with
average unit weights between 15.6 kN/m3 and 17.6 kN/m3 4.3. Test setup and testing procedure
representing medium-dense and dense conditions, respectively.
Fig. 4 shows the experimental setup. In order to achieve
The system rigidity ks is calculated as reasonably homogeneous sand beds of reproducible packing,
controlled pouring and tamping techniques were used to deposit
E c · Ifooting sand in 50 mm layers into the experimental sandbox. The
ks = (3)
Es · l 3 · b load was applied in increments of approximately 25–50 kN
where E c and E s are the modulus of elasticity of the concrete by a hydraulic jack (maximum capacity 2000 kN). During
and the oedometric modulus of the soil, respectively; Ifooting is testing, the vertical displacements at the slab centre and the slab
the moment of inertia of the cross section of the footing; l is corners as well as the steel and concrete strains were measured.
the side length; and b is the width of the footing. In the present Furthermore, pressure gauges were used to monitor the soil
tests the system rigidity varies between 0.188 for test DF3 and pressure distribution.
2236 J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241

Table 2
Characteristics and properties of the sand

Footing ρd (kN/m3 ) e (−) Dr (−) E s (kN/m2 ) ϕ 0 (◦ ) D (m) ks (−)


DF1 15.6 0.704 0.383 44,163 33.5 0.60 0.497
DF2 17.3 0.532 0.862 80,089 39.4 0.60 0.258
DF3 17.4 0.519 0.896 125,766 39.4 0.60 0.188
DF4 17.6 0.502 0.944 127,500 39.4 0.70 0.581
DF5 17.0 0.559 0.787 96,070 38.1 0.85 0.707

ρd : Dry unit weights; e: In situ void ratio; Dr : Relative density; E s : Oedometric modulus of the soil; ϕ 0 : Friction angle; D: Founding depth; ks : System rigidity.

Table 3
Failure loads achieved in tests and maximum flexural capacities

Footing (−) Vtest (kN) Vtest /Vcode Vflex /Vtest (−)


EC 2 (−) BS 8110-1 (−)
DF1 551 0.87 1.19 2.27
DF2 530 0.81 1.17 2.39
DF3 1197 1.51 2.07a 1.11
DF4 1251 0.89 –b 1.81
DF5 1130 0.77 –b 1.98
av
max = min(γm · f cu · u 0 d; γm · 5 · u 0 d; 2vc ) according to BS 8110-1,
0.5
3.7.7.2, 3.7.7.5.
b Critical section outside the footing.
Fig. 5. Typical failure patterns of the test footings.
4.4. Cracking and failure characteristic

All footings failed in a punching shear mode. The failure


loads are listed in Table 3. The comparison with the flexural
carrying capacities of the footings Vflex , calculated according
to the yield-line theory, reveals the assumption that the flexural
carrying capacities were not reached and hence confirms the
fact that failure occurred by punching. The failure loads were
far from the experimental soil bearing capacity, too. (a) Test DF1. (b) Test DF3.
The BS√8110-1:1997 limit on the column face shear of
γm · 0.8 · f cu and γm · 5 MPa, respectively, was introduced Fig. 6. Punching cone of test DF1 (a) and saw-cut of footing DF3 (b).
to limit the punching strength in case of small loaded areas.
This limit was extensively exceeded by the shear-reinforced
specimen DF3. This is due to the strong dependency of the code
rule on the column dimensions. It is the authors’ view that this
restriction is not a reasonable limit for the design of footings
including shear reinforcement.
Typical crack patterns of the footings are shown in Fig. 5.
After the test, the specimens were sawn into two halves. In all
slabs, the failure surface consisted of a wide shear crack which
formed the surface of a truncated cone. A view of the resulting
punching cone for specimen DF1 is presented in Fig. 6(a). The
crack pattern in the saw-cut of test DF3 is shown in Fig. 6(b).
The crack formation in the saw-cut of the shear-reinforced
footing DF3 indicates that the failure can be explained by
a simple strut and tie model. Furthermore, the failure of the
compressive strut at the column stub section can be seen well.

4.5. Concrete and steel strains

For all footings, measurements were taken to determine


the distribution of the concrete strain along a radius of the
slab. Due to the rotation-symmetric loading the reinforcement
is subjected to radial tensile stresses (Fig. 7). Thus, in the Fig. 7. Typical strain distribution of the steel in the radial direction for
compression zone radial compression stresses can be expected specimen DF4.
J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241 2237

(a) Test DF1. (b) Test DF5.

Fig. 8. Typical strain distribution of the concrete in the radial direction.

Fig. 9. Strain gauges arranged inside the test specimen (left) and strain measurements of the inclined compression struts in the radial direction for specimen DF5
(right).

as well. The strain measurements of the slender footings DF1 tried to explain the absence of compression strains at the top of
and DF2 confirmed this assumption (Fig. 8(a)) whereas the the slab by the eccentric loading of the inclined concrete struts.
more compact specimens DF4 and DF5 revealed a different This is comparable with an eccentrically loaded prism leading
behaviour. In radial direction, small tension strains up to to high compression stresses under the loading point and only
+0.3h were measured (Fig. 8(b)). small compression strains or even small tension strains on the
The strain distribution at the top of the slab is not a reliable opposite side of the prism.
indicator for the real stress distribution in the compression zone.
4.6. Measured soil pressure distribution
Therefore, strain gauges were arranged inside the specimens
as shown in Fig. 9, left. The strain measurements inside the Pressure gauges were used to measure the soil pressure dis-
slab confirm that there are indeed radial compression stresses tribution beneath the footing. For test DF1 17 pressure gauges
(Fig. 9, right). However, the radial compression zone consists were used, 20 for test DF2 and 21 for the remaining tests. The
of inclined compression struts. These inclined struts expand arrangement of the load cells is revealed in Fig. 5. The soil
inside the slab thickness from the corners and the edges of the pressure distribution beneath the footing DF2 is represented in
slab–column intersection towards the tension zone. Dieterle [4] Fig. 10 for the sections 3-3 and C-C. The equilibrium of the
2238 J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241

Fig. 10. Soil pressure distribution for section 3-3 (left) and section C-C (right) of specimen DF2.

in Karlsruhe are comparable with our own experiments. The


footings were designed in such a way that no punching failure
occurred. The tests were incrementally loaded until the soil
bearing capacity was reached.
At a load equal to the soil bearing capacity predicted
by BS EN 1997-1 [8] the measured stresses underneath
the punching cone were about 8% higher than the average
value of the soil pressure (V /A). These results confirmed
our own experimental investigations. However, the measured
concentration of the soil pressure beneath the column was not
as distinctive as the theory of Prandtl–Buisman would have
expected (Fig. 1). The experimental soil bearing capacity was
Fig. 11. Three-dimensional representation of soil pressure distribution for greater than predicted by BS EN 1997-1 and other international
footing DF2 near ultimate limit state. design rules. This observation was validated by two further
tests on rigid footings. The outline of these footings was
vertical forces was used to eliminate the measurement devia- chosen smaller (450 × 400 mm) compared to the outline
tion. For loads below 300 kN/m2 the soil stress distribution ac- of the sandbox (3250 × 3250 mm). Thus, the confining
cording to Boussinesq can vaguely be recognized in the section effect of the container was significantly reduced. Tests known
C-C. A local stress maximum beneath the column can already from literature also confirmed this assumption [5–7]. The
be seen at the serviceability state (approximately 300 kN/m2 ). redistribution of the soil pressure towards the centre of the
The concentration progresses with increasing load level. The footing was not completed when the calculated bearing capacity
ratio of the average measured stress beneath the punching cone was achieved. An evaluation of the soil stress distribution close
to the average soil pressure (σm = V /A) is about 1.06 close to ultimate load showed that the measured stresses beneath the
to limit state. A three-dimensional representation of the stress punching cone were 15.7% greater than the average value of the
distribution close to failure is revealed in Fig. 11. The results of soil pressure (V /A).
the other tests (specimen DF1, DF3–DF5) show similar tenden-
cies and verify the assumption that the ratio between the applied 5.2. Test data bank comprising punching tests on footings from
load and the soil bearing capacity as well as the compactness of literature
the soil clearly influence the soil stress distribution.
The punching databank for footings without shear reinforce-
5. Comparison of model predictions and experimental ment contains about 200 punching tests [4,10–14]. The com-
results parison between these tests and selected codes is based on the
ratio of the observed failure load Vtest and the punching load
5.1. Experimental investigations dealing with the soil pressure V Ru,code at ultimate limit state. This ratio was calculated for
distribution underneath footings each test. Finally, the ratios Vtest /V Ru,code were plotted versus
the main punching parameters (d, ρl , f c ) and the ratio of the
35 years ago an experimental investigation was carried out areas of the critical section A0 and the footing A to give some
to determine the soil pressure distribution at the University information about the trend of the performance of the different
of Karlsruhe [5,6]. The results of these tests were repeatedly codes. The punching shear capacities were calculated using the
evaluated in a database to verify the results of the present characteristic concrete strength according to BS EN 106-1 [9]
experimental investigations. The measuring technique, the ( f cu = ( f c,cyl,m − 4)/0.79) and a critical shear section at a dis-
properties of the soil and the foundation measurements used tance 1.5d from the boundaries of the loaded area. In this paper,
J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241 2239

Fig. 12. Comparison of punching tests and the punching shear capacity according to BS 8110-1:1997.

only the results of this evaluation according to BS 8110-1:1997 6. Proposed model


are discussed. Further information can be taken from [14].
6.1. Consideration of the soil-structure-interaction
5.2.1. BS 8110-1:1997 The concentration of the soil pressure below the punching
BS 8110-1:1997 underestimates the influence of the cone depends on the ratio of the applied load to the computed
effective depth as shown in Fig. 12. For an effective depth value of the soil bearing capacity (Fig. 13(a)). The scatter
greater than 0.2 m, the safety level falls below 1.0. The code of the experimental values can be reduced considering the
satisfactorily reflects the influences of the concrete strength. degree of compression (Fig. 13(b)). In Table 4 the equations
However, BS 8110-1:1997 does not reflect the influence of for the regression lines for sand with medium, high or very
the ratio of reinforcement trendlessly. Due to the large control high density are presented. This is sufficient because in situ
perimeter the reduction of the shear force by the soil pressure foundations are bedded on soil with at least medium density.
and also the punching shear resistance is extensive. Therefore, The soil stress concentration underneath the punching cone
the maximum shear stress on the column perimeter limits the depending on the applied load is shown in Fig. 13(b) for a
punching shear resistance. Yet, the equation for the maximum footing bedded on soil with high density. When the punching
shear capacity does not govern the influence of the flexural shear capacity is reached the soil reaction force can be
reinforcement and for this reason it is not surprising that calculated by the following equation:
British Standard cannot capture the influence of the ratio of V Rk
reinforcement correctly. The provisions of BS 8110-1:1997 σpunching cone = · RS where
A
tend to become less conservative for footings with a ratio A0 /A V Rk
greater than 0.4. With respect to the statistical evaluation BS R S = (b1 · R f + b2 ) and Rf = ≤1 (4)
Rk
8110-1:1997 shows a small coefficient of variation of v = 0.18.
However, Eurocode [15] demands a 5%-fractile value greater V Rk column force at limit state;
than one. British Standard cannot satisfy this requirement for Rk characteristic calculated soil bearing capacity accord-
footings without shear reinforcement (ζ5% = 0.65). ing to BS EN 1997-1;
2240 J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241

6.2. Design model for the punching behaviour of footings

The development of empirical models is based on the


evaluation of experimental data employing the method of
regression analysis. The tests of Dieterle [4,12], Kordina [11]
and Richart [10] as well as the present experimental
investigations revealed a strong influence of the ratio A0 /A
and the shear slenderness λ = a/d, respectively, on the
punching load. With increasing ratio A0 /A, the portion of the
soil pressure which is subtracted from the applied shear force
increases as well as the calculated punching shear capacity.
For this reason, the empirical design rule in BS 8110-1:1997
was modified with a term taking into account the ratio A0 /A.
The partial function F(A0 /A) = (A0 /A)−0.42 = (A/A0 )0.42
considers the influence of the ratio A0 /A well. The new
(a) All tests.
empirical design model is based on a calculated shear strength
along a critical perimeter at a distance of 1.0d from the column
face. The advantage is that the critical section will be inside
the footing even for more compact footings. A reduction of
the shear force by the complete soil pressure within the control
perimeter is allowed. After the derivation and verification of the
individual partial functions, the final regression analysis leads
to Eq. (5). The calculation takes into account the required safety
level of Eurocode—Basis of structural design, Annex D [15].
As 1/3 400 1/4
     0.42
0.59 A
Vc = · 100 · · ·
γm bv · d d A0
· u 1.0d · d (5)
where
A0 /A ratio of the areas of the critical section A0 and the
(b) Tests supported on densely packed sand. footing A;
u 1.0d critical perimeter (rectangular shape) at a distance of
Fig. 13. Correlation between the stress concentration underneath the punching 1.0d from the column face.
cone against load level (a) for all specimens and (b) for all specimens supported
on densely bedded soil. The characteristic concrete strength, the reinforcement ratio,
and the size effect of the effective depth should be considered
according to BS 8110-1:1997 (Section 3.1).
Table 4 As shown in Fig. 14 the new empirical model is able to
Equations for the regression lines depending on the ratio of load and the approximate the influence of the ratio A0 /A and of the shear
calculated bearing capacity of the soil
slenderness λ = a/d well. The influence of the main punching
Degree of compression Equation of regression line Rs (for R f = 1) parameters (d, ρl , f c ) is also reflected well. With respect to the
(−) R s = b1 · R f + b2 (−) statistical evaluation, it can be concluded that our own model
will result in a better mean value for the ratios Vtest /Vcode , if
Low to high density (all Rs = 0.2253 R f + 0.8603 1.086
compared to BS 8110-1:1997. Our own empirical model shows
tests)
Medium density Rs = 0.3058 Rf + 0.7195 1.025 less scattering of the ratio ζ . This leads to a smaller coefficient
High density Rs = 0.2049 Rf + 0.8702 1.075 of variation (v = 0.12). The demanded 5%-design value of 1.0
Very high density Rs = 0.2325 Rf + 0.8341 1.067 is reached.
High and very high Rs = 0.2171 Rf + 0.854 1.071
density 7. Conclusions
Our own tests DF1–DF5 Rs = 0.1365 R f + 0.9245 1.061
Based on the results of the investigations the following
conclusions can be drawn:
A area of footing;
(1) The ratio A0 /A and the shear slenderness affect the
b1 gradient of the regression lines depending on the punching shear resistance of footings.
density of the soil (Table 4); (2) The observed angle of failure cone was about 45◦ in the
b2 axis section of the regression lines depending on the present tests. This failure angle seems to be steeper than for
density of the soil (Table 4). flat slabs.
J. Hegger et al. / Engineering Structures 29 (2007) 2233–2241 2241

Fig. 14. Comparison of punching tests and the punching shear capacity according to the proposed empirical model.

(3) In two small-scale tests as well as in tests from literature the [2] BS 8110-1. Structural use of concrete—Part 1: Code of practice for design
experimental soil bearing capacities exceed the calculated and construction. London: British Standard Institution; 1997.
values according to current code provisions. Thus, the soil [3] BS EN 1992-1-1. Eurocode 2: Design of concrete structures. London:
British Standard Institution; 2005.
pressure redistribution towards the centre seems not to
[4] Dieterle H, Rostásy F. Tragverhalten quadratischer Einzelfundamente aus
be completed when the calculated bearing capacity was Stahlbeton. In: Deutscher Ausschuss für Stahlbeton, Heft 387. Berlin:
reached. Beuth Verlag; 1987.
(4) The assumption of an uniformly distributed soil pressure is [5] Leussink H, Blinde A, Abel P. Versuche über die Sohldruckverteilung
safe for most practical cases. unter starren Gründungskörpern auf kohäsionslosem Sand. Karlsruhe:
(5) The punching loads predicted by different codes tend to be Veröffentlichung des Institutes für Bodenmechanik und Felsmechanik der
conservative for slender footings. Nevertheless, the codes Technischen Hochschule Fridericana in Karlsruhe. Heft 22. 1966.
[6] Blinde A, Wibel A. Sohldruckverteilung unter starren quadratischen
tend to overestimate the punching resistance for compact
Fundamenten. Karlsruhe: Veröffentlichungen des Institutes für Boden-
footings with small shear slenderness. mechanik und Felsmechanik der Technischen Hochschule Fridericana in
(6) Due to the strong dependency on the column dimensions Karlsruhe. Heft 48. 1971.
√ 8110-1:1997 limit on the column face shear of γm ·
the BS [7] von Wolffersdorff P-A. Probebelastung zur Baugrundtagung 1990: Ver-
0.8· f cu and γm ·5 MPa, respectively, is not able to predict suchsergebnisse und Auswertung des Prognosewettbewerbs. Geotechnik
the punching shear resistance of footings trendlessly. 1991;14:16–21.
(7) A new equation was derived which ensures the safety [8] BS EN 1997-1. Eurocode 7. Geotechnical design. General rules. London:
British Standard Institution; 2004.
level of Eurocode. The punching shear capacity of compact
[9] BS EN 206-1. Concrete—Part 1: Specification, performance and
footings can be calculated with this equation. conformity. London: British Standard Institution; 2001.
[10] Richart FE. Reinforced concrete wall and column footings. Journal
Acknowledgements of the American Concrete Institute 1948;45(2):97–127. Part 1; Part 2:
45(3):237–60.
This research program (AiF-No. 13620, DBV-No. 245) was [11] Kordina K, Nölting D. Tragverhalten von ausmittig beanspruchten
supported via the Deutsche Beton- und Bautechnik Verein E.V. Einzelfundamenten aus Stahlbeton. Abschlußbericht zum DFG-Vorhaben
(DBV) by the funds of the Federal Ministry of Economy of Ko 204/27+30. Braunschweig; 1981.
Germany with a contribution from the Arbeitsgemeinschaft [12] Dieterle H, Steinle A. Blockfundamente für Stahlbetonfertigteilstützen.
industrieller Forschung (AiF). The authors express thanks to the In: Deutscher Ausschuss für Stahlbeton, Heft 326. Berlin: Beuth Verlag;
1981.
AiF and the participating companies for their financial support,
[13] Hallgren M, Kinnunen S, Nylander B. Punching shear tests on column
as well as to the committee accompanying the project for their footings. Nordic Concrete Research 1998;21(1):1–23.
helpful suggestions. [14] Hegger J, Ziegler M, Ricker M, Ulke B. Entwicklung eines
Bemessungskonzeptes zum Durchstanzen von Fundamentplatten unter
References Berücksichtigung der Boden-Bauwerk-Interaktion. Abschlußbericht zum
AiF-Vorhaben 13620. Aachen; 2005.
[1] Daniel L. Application des potentiels à l’étude de l’équilibre et du [15] EN 1990. Eurocode—Basis of structural design. Brussels: European
mouvement des solides élastique. Imprimérie. Lille. 1885. Committee for Standardization; 2002.

You might also like