Sei sulla pagina 1di 18

Engineering Structures 113 (2016) 4158

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Investigation of the shear behavior of RC beams on the basis of measured


crack kinematics
Patrick Huber , Tobias Huber, Johann Kollegger
Institute for Structural Engineering, Vienna University of Technology, Karlsplatz 13/212-2, 1040 Vienna, Austria

a r t i c l e

i n f o

Article history:
Received 13 August 2015
Revised 7 December 2015
Accepted 20 January 2016
Available online 5 February 2016
Keywords:
Shear strength
Experiments
Crack opening and sliding relationship
Digital image correlation
Shear reinforcement
Flanges
Size effect

a b s t r a c t
This paper deals with an experimental and theoretical assessment of the shear strength of reinforced concrete beams with and without a minimum amount of transverse reinforcement. Various shear transfer
mechanisms contribute to the shear strength. The contributions of aggregate interlock, residual tensile
stresses at the crack tip, dowel action, and shear strength of the shear reinforcement depend mainly
on the opening and sliding behavior of the critical shear crack, which is directly linked to its shape. In this
paper, the kinematics of the diagonal crack is investigated in 20 test specimens by using threedimensional digital image correlation. In this test series the influences of the member depth, the cross
section, the type of concrete, and a minimum amount of transverse shear reinforcement on the shear
behavior are investigated. Based on the full-field optical measurements and the use of different constitutive laws from literature an estimation of the contributions of the various shear transfer mechanisms is
performed. Further, an evaluation of their impact on the shear strength at failure including the influences
of the investigated parameters is carried out. An explanation for the pronounced size effect for beams
without any transverse reinforcement is given. Moreover, the importance of aggregate interlock as well
as the effect of the presence of flanges for beams with and without stirrups is shown. A comparison of the
experimental results with the sum of the contributions to the shear strength by each mechanism (based
on the measured crack kinematics) yields reasonable agreement.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Even though research on the shear strength of reinforced and
prestressed concrete members with and without shear reinforcement has a long history, investigations on this complex topic are
still ongoing. Thanks to the extensive experimental and theoretical
work available, the research community agrees on the different
shear transfer mechanisms that can contribute to the shear
strength of reinforced concrete members [1,2]. However, there
are different opinions about the governing shear transfer mechanism and the main influencing parameters. Furthermore, no information is currently available on what shear transfer mechanisms
act just before failure of the member. For this investigation,
concrete members with and without shear reinforcement are considered separately, as a minimum amount of shear reinforcement
already results in a shear strength above the shear cracking load.
After diagonal cracking the shear behavior of members with shear
reinforcement changes completely. The contributions of the
Corresponding author. Tel.: +43 15880121255.
E-mail address: patrick.huber@tuwien.ac.at (P. Huber).
URL: http://www.betonbau.tuwien.ac.at (P. Huber).
http://dx.doi.org/10.1016/j.engstruct.2016.01.025
0141-0296/ 2016 Elsevier Ltd. All rights reserved.

individual shear transfer mechanisms in members with a minimum amount of shear reinforcement are very different compared
to those in members without shear reinforcement.
While the EulerBernoulli hypothesis represent a simple and
widely accepted theory for the prediction of the flexural strength
of reinforced concrete members, no such model exists for the shear
strength. Many different analytical approaches have been developed, each of which considers one individual or a combination of
different shear transfer mechanisms. Many of these theories are
based on empirical data only. Due to the large number of different
theories only some representative approaches are presented in this
section.
In many approaches the loss of the loadcarrying capacity of the
uncracked compression zone is the main reason for shear failure of
beams and slabs without transverse reinforcement [35]. The
aggregate interlock effect, on the other hand, plays a major role
in the critical shear crack theory [6], the modified compression
field theory [7] and the simplified approach [8]. These so-called
strain-based models are based on a relationship between the longitudinal strain ex (and crack spacing in the approaches presented
in [7,8]) and the crack width w. Thus, shear failure is defined by a
critical crack width and therefore predominantly by the loss of the

42

P. Huber et al. / Engineering Structures 113 (2016) 4158

Nomenclature
A, N
Acz
Agt
An
At
Cf
Esw
F
FExp
FS
GF
Iv
V
Vag
Vc
Vcr
Vcz
Vda
Vda,cr
Vda,max
VExp
Vi
Vlp
Vs,da
Vsw
a
an
at
a3, a4
b
bn
bf
bw
c
d
dg
f
fc
fc,cube
fct
fct,sp
ftw
fy
fyw
h
hf
l
lb

empirical constants for different bond characteristics


area of the uncracked compression zone
uniform elongation of the reinforcing bar
sum of all projected contact areas an normal to the crack
surfaces
sum of all projected contact areas at tangential to the
crack surfaces
aggregate effectivity factor
Youngs modulus of the transverse reinforcement
applied force
load applied in the test
force in the longitudinal reinforcing bar
fracture energy
moment of inertia of the longitudinal reinforcing bars
applied shear force
shear strength contribution of aggregate interlock
shear strength contribution of the concrete
shear strength contribution of the residual tensile
stresses in the fracture process zone
shear strength contribution of the uncracked
compression zone
shear strength contribution of dowel action
dowel cracking load
maximum dowel load
maximum shear strength in the tests
contributions of the different shear transfer
mechanisms
shear strength immediately before failure
shear strength contribution of dowel action, including
the contribution of a stirrup near the crack
shear strength contribution of the shear reinforcement
shear span (distance between the mid-point of the
applied load and the mid-point of the support plate)
projected contact area of an aggregate particle normal
to the crack surface
projected contact area of an aggregate particle
tangential to the crack surface
empirical constants of the rough crack model
width
effective width of the concrete between longitudinal
reinforcing bars
flange width
web width
cement content
effective depth
maximum aggregate size
distance of the stirrups to a diagonal crack
cylinder
compressive
strength
of
concrete
( = 150 mm, h = 300 mm)
cube compressive strength (b = h = 200 mm)
uniaxial tensile strength of concrete
splitting tensile strength of concrete
ultimate tensile strength of the transverse reinforcing
bars
yield strength of the longitudinal reinforcing bars
yield strength of the transverse reinforcing bars
specimen height
flange height
span length of the specimen
anchorage length

aggregate interlock. In the opinion of some researchers, the contribution of dowel action to the shear strength and the tensile stresses in the fracture zone cannot be neglected [911]. Both

lcr
lda
lFPZ
m
n
pk
r
s
sw
t
w
wcr
wsw
x
z

RVi

acr
d
db
dd

esw
ex
h

ql,w
qw
qw,min
rag
rct
rpu
rs
rsw
sag
sb
sb1
sb2
sExp
si

su
sxz
s0
l
s
w
COV
DIC
FPZ
LVDT
NC
RC
SCC

crack length
length of the tension stress block below the longitudinal
reinforcing bars
length of the fracture process zone
mean value
vertical axis normal to the crack surface
ratio of the volume of aggregate and volume of concrete
ratio of sliding s and opening w of the crack surfaces
sliding of the crack tangential to the crack surface
stirrup spacing
horizontal axis tangential to the crack surface
opening of the crack normal to the crack surface/water
content
maximum crack width for stress transfer
crack opening at the location of the transverse
reinforcement
horizontal axis
vertical axis
sum of the contributions of the different shear transfer
mechanisms
angle of the shear crack
vertical displacement
bond slip
vertical displacement of the longitudinal reinforcing
bars (dowel action)
stirrup strain
longitudinal strain in concrete
strut inclination
longitudinal reinforcement ratio in the effective web
area bwd
shear reinforcement ratio
minimum shear reinforcement ratio
normal stress due to aggregate interlock
tensile stress
cement matrix yield strength
reinforcing bar stress
transverse reinforcement stress
shear stress due to aggregate interlock
bond stress
bond stress before yielding of the stirrups
bond stress after yielding of the stirrups
shear stress in the effective web area bwd
shear stress in the effective web area bwd for each shear
transfer mechanism
shear stress in the effective web area bwd immediately
before failure
shear stress in the xz direction
basic shear stress of the rough crack model due to
aggregate interlock
friction coefficient between aggregate particles and
cement matrix
diameter of the longitudinal reinforcing bars
diameter of the transverse reinforcing bars
coefficient of variation
digital image correlation
fracture process zone
linear variable differential transformer
normal-strength concrete
reinforced concrete
self-compacting concrete

mechanisms are mostly used in theories which consider more than


one governing shear transfer mechanism. Finally, there are a number of approaches which do not subscribe to the theory that the

P. Huber et al. / Engineering Structures 113 (2016) 4158

loss of one or more shear transfer mechanisms is responsible for


shear failure; for example approaches based on fracture mechanics
[12,13], principal stress calculations on a potential band of the critical shear crack [14], or purely empirical formulations [15].
Truss models or stress fields are mainly used to describe the
shear behavior of reinforced or prestressed beams with shear reinforcement. Nowadays, two different approaches can be found in
the literature. The first approach is based on the lower-bound theorem of the theory of plasticity, which assumes that yielding of the
stirrups (tension ties) and crushing of the web concrete (compression strut) occurs simultaneously at failure [1618]. Thus, it is
assumed that the strut inclination h does not coincide with the
crack angle acr, which means that stress transfer across the cracks
is taking place. Since an explicit contribution of the concrete to the
shear strength (Vc) is not taken into account in these models, it is
not possible to quantify the contributions of the individual shear
transfer mechanisms. The second method assumes that the inclination of the compression strut h coincides with crack angle acr.
As opposed to the above mentioned model, the contribution of
the concrete to the shear resistance (Vc) can be derived directly
from one or a combination of the shear transfer mechanism relating to concrete and is added to the shear strength of the transverse
reinforcement (Vs). Whereas Bentz et al. [8] and Reineck [19] state
that the concrete can only contribute to the shear resistance if it is
capable of transmitting stresses across the cracks, Hegger et al. [20]
propose that besides the shear resistance of the transverse reinforcement, shear transfer in the uncracked compression zone must
be taken into account as an additional shear transfer mechanism.
In recent research the role of the various shear transfer mechanisms has been investigated more frequently. In this context, the
kinematics of the critical shear crack (opening and sliding of the
crack surfaces) is of particular interest [2124] and plays an essential role for the shear strength contribution of the aggregate interlock effect in particular [25,26].
In this paper an estimation of the contribution of the different
shear transfer mechanisms, based on the measured crack kinematics of tested reinforced concrete beams using a photogrammetric
measurement system (DIC), is presented. Different constitutive
models are used in order to evaluate each mechanism. More details
about the different laws and their theoretical backgrounds are presented in chapter 3 of this paper. The focus of this experiments is
on the influences of parameters (presence of flanges, member
depth, minimum number of stirrups, and type of concrete), for
which their role on the shear behavior is not yet fully clarified.

43

For example, investigations on the effect of different types of concrete (different cylindrical compressive strength) provide contradictory results. While increasing the compressive strength fc
(21.396.1 MPa) in a test series of Mphonde et al. [27] increases
the shear strength significantly (55%), Angelakos et al. [28] showed
that increasing the compressive strength fc (21.080.0 MPa)
brought no additional shear carrying capacity. In this context, the
composition of the used concrete mixes would be of interest, but
regarding this, there are no data available in [28].
The influence of the presence of flanges is also uncertain due to
missing experimental data. On the one hand Placas and Regan [29]
could demonstrate, that a flange on the compression side, which is
4 times wider than the web, was only able to increase the shear
strength by 7%. On the other hand, in a test series of Rupf et al.
[30] the presence of a top and bottom flange turned out to be very
beneficial regarding the shear strength as well as the post-peak
behavior of members containing a minimum amount of transverse
reinforcement.
Many test series focused already on the issue regarding the
influence of member depth on the shear strength of reinforced concrete elements without transverse reinforcement (see e.g. databases in [31]). A world-wide accepted explanation on this topic
does not exist yet. Especially, for beams with a minimum quantity
of stirrups the influence of varying member depths has not been
investigated yet. As mentioned in [32], the only available experimental program on this topic was performed by Bahl [33]. However, only one of his tests on reinforced concrete beams failed in
shear. Therefore, a clear tendency is not discernible.
The effects of the investigated influencing parameters on the
different shear transfer mechanisms and the effect of the shape
of the critical shear crack on the kinematics in this paper will be
bring more light on these issues.
2. Experimental program
2.1. Specimens and test setup
In this section the main results of a test series on 20 reinforced
concrete beams are presented. The details of the beams (dimensions and reinforcement layout) are summarized in Fig. 1 and
Table 1. The specimens are labeled with an identifying code. The
first letter corresponds to the type of cross section (R = rectangular,
T = T-beam), the following number refers to the height h of the
beams in millimeters, and the subsequent letter indicates whether

(a)

(b)

Fig. 1. Test setup and specimen details for shear tests on: (a) beams without transverse reinforcement; and (b) beams with transverse reinforcement.

P. Huber et al. / Engineering Structures 113 (2016) 4158

the beams contain stirrups or not (o = no stirrups, m = a minimum


number of stirrups). The last two numbers stand for the target
value of the cylinder compressive strength of concrete fc. The
beams with a depth of h = 500 mm, a minimum number of stirrups,
and normal-strength concrete (NC) were given an additional consecutive number. The dimensions of the specimens and the reinforcement layout were scaled at a ratio of 1:2:4. Thus, the
specimens had member depths of 250 mm, 500 mm, and
1000 mm. Six beams had a compression flange, whereas the
remaining beams had a rectangular cross section. To investigate
the influence of the presence of a flange on the general shear
behavior, the compression flange bf was chosen twice as wide as
the web thickness bw, and the flange depth was hf = bf/3. A linear
transition from the web to the flange was chosen to ensure a better
force flow. The longitudinal reinforcement ratio ql,w was approximately 1.2% for all beams, which is a value often used in practice.
Steel plates were welded to both ends of the reinforcing bars so as
to avoid premature failure of the anchorage. Four longitudinal reinforcement bars with the same diameters as the stirrups were
placed in the compression zone of the beams. To study the effect
of a minimum amount of transverse reinforcement on the shear
behavior, 11 beams contained transverse reinforcement (see
Fig. 1(b)). The amount of shear reinforcement was chosen
according to the minimum reinforcement ratio specified in current standards, specifically according to the fib Model Code 2010
[34]. This document stipulates a minimum reinforcement ratio
p
qw;min 0:08  f c =f yw 0:070:11% for the material properties
used in this test series. The transverse reinforcement consisted of
either hooks (single leg) or stirrups (see Fig. 1(b) or Table 1). The
diameter w and the spacing sw of the transverse reinforcement
were scaled according to the member size, resulting in a shear
reinforcement ratio qw of 0.094%. The number of stirrups varied
slightly for beams R250m35, R500m352, and R500m353 to investigate the influence of a modified shear reinforcement ratio (see
Table 1).
The test setup is shown in Fig. 1. The beams were loaded (displacement controlled) by a point load applied at the mid-point of
a simply supported span of length 5.6h, resulting in a ratio of shear
span to effective beam depth a/d of 3.04. To force shear failure to
occur on one side of the beam, the other side was reinforced with
a higher number of stirrups, which allowed a more efficient use of
the measurement equipment. The loading plates under the
hydraulic jack and the support plates had the same dimensions
bf  0.2h  0.02h. All support plates were placed on rollers, thus
permitting rotation as well as longitudinal displacements.

0.211
0.286
0.168
0.249
0.264
0.351
0.187
0.256
0.189
0.189
0.181
1.51
1.72
1.20
1.53
1.58
2.10a
1.46
1.39
1.35
1.36
1.41
26.0
29.6
83.0
105.9
109.2
145.1
402.1
383.8
23.3
93.5
390.4
16
22
16
22
22
22
16
22
16
16
16
3.20
3.10
3.20
3.06
3.10
3.10
3.92
2.77
3.20
3.20
3.92
51.2
35.9
51.2
37.9
35.9
35.9
60.9
29.6
51.2
51.2
60.9
13/100
14/150
16/200
16/200
24/200
24/150
112/400
26/200
13/100
16/200
112/400
48
48
416
416
416
416
230 + 236
230 + 236
48
416
230 + 236

0.094
0.112
0.094
0.094
0.084
0.112
0.094
0.094
0.094
0.094
0.094

0.149
0.151
0.128
0.175
0.100
0.108
0.179
0.132
0.104
1.13
0.93
0.97
1.05
0.78
0.64
1.36
1.03
0.79
19.6
16.0
67.2
72.4
214.5
177.1
23.5
71.0
216.7
16
22
16
22
16
22
16
16
16
3.12
3.06
3.12
3.02
3.23
2.85
3.12
3.23
3.33
58.1
37.9
58.1
36.0
60.3
35.6
58.1
60.3
56.6

48
48
416
416
230 + 236
230 + 236
48
416
230 + 236

fc (MPa)

qw (%)
Shear reinf.
Longitudinal reinforcement

230
230
460
460
460
460
920
920
230
460
920
a

Flexural failure.

75
75
150
150
150
150
300
300
75
150
300
Beams with shear reinforcement
R250m60
SCC
R250m35
NC
R500m60
SCC
R500m351
NC
R500m352
NC
R500m353
NC
R1000m60
SCC
R1000m35
NC
T250m60
SCC
T500m60
SCC
T1000m60
SCC

shear reinforcement
SCC
75
NC
75
SCC
150
NC
150
SCC
300
NC
300
SCC
75
SCC
150
SCC
300
Beams without
R250o60
R250o35
R500o60
R500o35
R1000o60
R1000o35
T250o60
T500o60
T1000o60

230
230
460
460
920
920
230
460
920

Concrete type

bw (mm)

d (mm)

2.2. Material properties

Specimen

Table 1
Specimen dimensions, details of reinforcement, mechanical properties of concrete, and experimentally determined shear strength.

fct,sp (MPa)

dg (mm)

VExp (kN)

VExp/(bd) (MPa)

VExp/bw  d 

p p
f c MPa

44

To obtain a different roughness of the crack surfaces in this test


series two different types of concrete were used (see also Section 3.1): self-compacting concrete (SCC) with a maximum aggregate size dg of 16 mm and normal-strength concrete (NC) with
dg = 22 mm. The compressive strength fc, determined on cylinders
(diameter 150 mm, length 300 mm) on the day of testing, ranged
from 51.2 MPa to 60.9 MPa for the SCC beams and from 29.6 MPa
to 37.9 MPa for the NC beams. The uniaxial tensile strength fct of
the concrete was determined from the splitting tensile strength
fct,sp (determined on cylinders: diameter 150 mm, length
300 mm) obtained from a standard Brazilian test. The detailed
composition of the two different types of concrete is listed in
Table 2. In contrast to normal concrete (NC), the self-compacting
concrete (SCC) had a lower watercement ratio w/c (0.51 vs.
0.62), and the amount of fine aggregate particles as well as superplasticizer was much higher.

P. Huber et al. / Engineering Structures 113 (2016) 4158


Table 2
Composition of concrete (amounts in kg/m3).

Rock dust (01 mm)


Sand (04 mm)
Fine gravel (48 mm)
Coarse gravel (816 mm)
Coarse gravel (1622 mm)
Recycled aggregate (016 mm)
Cement
Superplasticizer
Water

Normal concrete
(NC)

Self-compacting
concrete (SCC)

811
296
580
131
105
339
1.12
208

130
822
254
603

380
4.95
197

kinematics of the crack surfaces it was possible to estimate the


contributions of the different shear transfer mechanisms and thus
get a better insight into the real shear behavior of concrete elements with and without shear reinforcement.

2.4. Test results

Table 3 shows a summary of the average mechanical properties


of the stirrups. The yield strength fyw of the cold-worked bars is
defined at the 0.2% residual strain. Since ribbed reinforced bars
were only available with a minimum diameter of 4 mm, plain
wires with a diameter w of 3 mm were used for the small specimens (R250m60, T250m60). The 3 mm and 4 mm stirrups had a
higher yield strength, resulting in a higher value of qwfyw. The
mechanical properties of the flexural reinforcement, which normally yields at a stress of 550 MPa, were not determined in this
test series.
2.3. Photogrammetric measurements (DIC)
Besides the use of conventional measuring techniques such as
linear variable differential transformers (LVDTs), strain gauges
and load cells the main focus of this research is on threedimensional digital image correlation (DIC) using photogrammetric techniques to record crack kinematics after the formation of
the critical shear crack. In contrast to conventional measurement
solutions, DIC can perform full-field measurement of the crack
opening normal to the crack surface and the crack sliding parallel
to the crack surface during the entire test procedure. The DIC software used was the ARAMIS v6.3.0-9 [35] non-contact optical strain
measurement system (GOM mbH, Braunschweig, Germany) with
two 4 M 60 Hz cameras (2352 by 1728 pixels resolution). The system uses a random speckle pattern (see Fig. 1) applied to the surface of the target object to create the unique features identifiable to
the system. To identify these features each image is divided into an
overlapping grid of facets (15  15 pixels with 2 overlapping pixels
for this test series). To determine surface displacements of the
investigated shear field, each facet is tracked from one image to
the next, creating a series of data points which are mapped to measure full-field displacement. Due to the different dimensions of the
specimens the position and size of the measuring field varies
within the test series. The image acquisition rate of the cameras
was changed at specific load steps to save computing time. In the
linear-elastic state and at the beginning of cracking images were
taken every five seconds. The frequency was increased up to 5 Hz
prior to failure. Hence, it was possible to investigate the opening
and sliding behavior of the critical shear crack from its initiation
up to failure of the specimen. From the detailed analyses of the

Table 3
Mechanical properties of cold-worked stirrups.

45

w (mm)

fywa (MPa)

ftw (MPa)

Agt (%)

Surface ()

3
4
6
12

748
653
569
552

776
710
638
654

3.9
4.9
3.1
3.4

Plain
Ribbed
Ribbed
Ribbed

Offset yield point at 0.2% strain for cold-worked reinforcing bars.

In Fig. 2 the experimentally determined shear stress sExp = VExp/


(bwd) is plotted against the dimensionless deflection (mid-span
deflection d divided by the span length l) in order to compare the
responses of the beams independent of their different dimensions.
Table 1 shows a summary of the maximum shear strengths VExp
determined from the tests. The results of the tested specimens
without transverse reinforcement showed a pronounced influence
of the beam depth, regardless of type of cross-section (see Fig. 2
(a)(c)). The maximum shear stress sExp, for example, decreased
from 1.13 MPa (R250o60) to 0.76 MPa (R1000o60) for rectangular
beams with self-compacting concrete without stirrups (see Fig. 2
(a) and Table 1). For the series with normal-strength concrete
(NC), beam R500o35 (see Fig. 2(c)) exhibited a higher shear stress
sExp compared to the smaller beam R250o35. In contrast, beam
R1000o35 reached only 70% of the critical shear stress sExp compared to the smallest specimen R250o35. As far as the shear
strength VExp of beams with a compression flange is concerned, it
can be concluded, that the presence of a flange brought no additional shear carrying capacity compared to a rectangular crosssection with the same type of concrete (compare Fig. 2(a) with
Fig. 2(b)). For the small beam T250o60 a higher shear strength
was determined compared to its rectangular counterpart
(R250o60). This discrepancy will be discussed more in detail later
(see chapter 4 or Fig. 14). As far as the concrete type is concerned,
decreasing the compressive strength fc resulted in a reduction of
the experimentally determined maximum shear stress sExp (compare Fig. 2(a) and (c)) for beams R250o35 and R1000o60 without
stirrups.
The presence of a minimum number of stirrups increased the
shear strength VExp of the beams by factors ranging from 1.00 to
2.17 and the deflection at mid-span d by a factor of between 1.00
and 3.00 compared to specimen without transverse reinforcement
with identical dimensions and material properties. No shear was
resisted by beam T250m60 after diagonal cracking, a fact which
can be attributed to the poor bond conditions of the plain stirrups.
As opposed to the beams without transverse reinforcement, the
beam depth h did not seem to influence the maximum shear stress
sExp of beams with stirrups (see Fig. 2(d)(f)). It can be assumed
that the size effect is suppressed due to the presence of a minimum
amount of shear reinforcement. Furthermore, neither the width of
the compression chord nor the type of concrete influenced the
experimentally determined shear stress sExp for beams with a minimum number of stirrups (see Fig. 2(d)(f)).
Three different failure modes were detected. Beams without
shear reinforcement exhibited typical brittle diagonal tension failure. At the maximum shear force VExp the critical shear crack
extended into the compression chord. Simultaneously, an incipient
dowel crack along the longitudinal reinforcement could be
observed. After reaching VExp in nearly all shear tests (with the
exception of R250o60) a small decrease in load (012%) occurred
before reaching the ultimate failure (Vlp). This final load stage is
characterized with significant kinematics of the crack surfaces
resulting in a pronounced shear rotation. This in turn led to a propagation of the critical shear crack deeply into the compression zone
as well as along the longitudinal reinforcement (see Fig. 2(g). The
formation as well as the propagation of the critical shear crack
could be tracked continuously based on the photogrammetric
measurements.

46

P. Huber et al. / Engineering Structures 113 (2016) 4158

(a)

(b)

1.5

(c)

Exp

[MPa]
R250o60
R500o60
R1000o60

0.5
0.0
0.000

0.001

0.002

Exp

[MPa]

failure
1.0

0.003

R250o35
R500o35
R1000o35

T250o60
T500o60
T1000o60
0.000

0.001

/l [-]

0.002

0.003

0.000

/l [-]

(d)
[MPa]

0.002

0.003

/l [-]

(e)

yielding

(f)

2.0

Exp

0.001

w=0.115%

1.5
1.0

R250m60
R500m60
R1000m60

0.5
0.0
0.000

0.002

0.004

0.006

R250m35
R500m351
R500m352
R500m353
R1000m35

T250m60
T500m60
T1000m60
0.000

0.002

/l [-]

(g)

0.004

/l [-]

0.006

0.000

0.002

0.004

0.006

/l [-]

(h)

Stirrup spacing sw different

Fig. 2. Shear stress sExp vs. dimensionless deflection d/l for: (a) rectangular beams (SCC); (b) T-beams (SCC); (c) rectangular beams (NC) without shear reinforcement; (d)
rectangular beams (SCC); (e) T-beams (SCC); (f) rectangular beams (NC) with stirrups; (g) crack pattern of the critical shear crack for specimens without shear reinforcement;
and (h) crack pattern of the critical shear crack for specimens with stirrups.

For beams with a minimum number of stirrups, however, diagonal cracking did not generally cause sudden failure, since crack
propagation was restrained by the transverse reinforcement. Here,
increasing the load led to a widening of the critical shear crack
combined with a crack propagation in the direction of the load
plate, thereby rupturing the stirrups. A third failure mode was
observed in beam R500m353. Due to a slightly larger amount of
transverse reinforcement (qw = 0.112%), yielding of the longitudinal reinforcement occurred (see Fig. 2(f)), followed by rupturing
of the stirrups.
Fig. 2(g) and (h) shows the shape of the critical shear crack at
failure. As far as the investigated influencing parameters
(member size, cross section, concrete type, and amount of shear

reinforcement) are concerned, no correlation between the location


and shape of the critical shear crack could be found. The crack pattern and location are important for the resulting crack kinematics
and hence for the different shear transfer mechanisms. Details on
the crack kinematics and its consequences on the behavior of the
member are described in detail in the following sections.

3. Shear transfer mechanisms


A very effective method for understanding the force and motion
characteristics of a concrete element is to draw a free body
diagram. Along a certain section, in this case at the location of

P. Huber et al. / Engineering Structures 113 (2016) 4158

the critical shear crack, the external and internal forces must be in
equilibrium. As far as the shear strength of a RC member is concerned, the applied shear load is resisted by a combination of all
acting shear transfer mechanisms. Fig. 3(a) shows the possible
shear transfer mechanisms for a RC beam without transverse reinforcement. The same mechanisms can also be found in members
with a low amount of stirrups (see Fig. 3(b)). Due to the pronounced differences in crack kinematics, however, the contribution
of each mechanism to the shear transfer is completely different for
beams with and without shear reinforcement.
For claritys sake the horizontally acting forces necessary for
equilibrium are omitted in Fig. 3.
It is widely accepted [1,2] that the following five mechanisms
can contribute to the shear transfer in a cracked RC element (see
Fig. 3):

Aggregate interlock Vag.


Dowel action Vda.
Residual tensile stresses in the FPZ Vcr.
Shear transfer in the uncracked compression zone Vcz.
Shear transfer by the transverse reinforcement Vsw.

The activation of the various shear transfer mechanisms


depends primarily on the shape and on the opening and sliding
behavior of the critical diagonal crack. The different shear transfer
mechanisms have been investigated by a number of authors [36
58] by using special test procedures and theoretical approaches.
Therefore, only a short summary of the theoretical backgrounds
and theories of these mechanisms is presented.
3.1. Aggregate interlock effect
Cracks in beams subjected to pure flexure open perpendicularly
to the direction of the crack. The opening of a shear crack is usually
accompanied by a sliding of the crack surfaces (also known as
shear displacement) due to the presence of shear stresses. The relation between opening w and sliding s is mainly influenced by the

(a)

(b)

Fig. 3. Vertical forces acting on the diagonal crack: (a) RC members without
stirrups; and (b) RC members with transverse reinforcement.

47

inclination acr of the diagonal crack (see Figs. 12 and 13). At the
crack tip as well as at the incipient dowel crack the opening w of
the crack is much more dominant compared to the sliding s. Thus,
the most parallel displacement of the crack surfaces could be
observed at the steep part of the critical shear crack. If the crack
surfaces slide against each other, the aggregate can interlock with
the cement matrix of the opposite crack surface, resulting in stress
transfer across the crack and hence in an increase of the shear
strength (see Fig. 4(a) and (b)).
Due to the low bond strength of the interface between the
aggregate particles and the cement matrix cracks in normalstrength concrete tend to go round the aggregates and so most of
the aggregate particles remain embedded in one crack flank (see
Fig. 4(c)).
In concrete with weak aggregate particles (for example lightweight concrete) or high-strength concrete, however, the aggregate particles primarily can split into two, resulting in a
smoother crack surface (see Fig. 4(d)). Thus, the crack roughness
depends on the one hand mainly on the bond strength between
the aggregate particles and the cement paste and on the other hand
on the tensile strength of the aggregates. Nevertheless, it can happen, that also in high strength concrete with a strong bond
strength between the two phases the crack is not able to intersect
the aggregates. Furthermore, the composition of concrete is an
important influencing parameter for the coarseness of a cracked
surface. In general, self-compacting concrete is characterized by a
high amount of fines and a reduced content of coarse aggregates
(see Table 2) resulting in smoother crack surfaces compared to regular concrete mixes. Even so, significant shear stresses can be
transferred across these cracks, presumably because of the undulated crack path (typical S-shape) and the resulting roughness on
a macro level [23,26].
To estimate the shear transfer across cracks, several researchers
performed push-off tests [26,3640], varying the type of aggregate
or reinforcement ratios. In Fig. 4(e) and (f) a common setup of such
a test is shown. Two precracked planes are moved against each
other, while the crack width w is controlled either with reinforcing
bars (Fig. 4(e)) or external restraint bars (Fig. 4(f)). Based on the
measured relationships between stresses (shear stress sag and normal stress rag), crack opening w, and sliding s, many different theories have been proposed.
The most comprehensive theoretical work in this field was
undertaken by Walraven [38,39]. The proposed model has a clear
physical background, regarding concrete as a two-phase system
in which the aggregate particles are idealized as rigid spheres
(see Fig. 4(b)). The cement matrix exhibits rigid plastic stress
strain behavior. The size of the contact areas between the particles
and the cement matrix depends mainly on the crack kinematics (w
and s), the aggregate size dg, and the ratio between the total volume of the aggregate particles and the volume of the concrete pk.
The basic equations for estimating the normal stress rag and the
shear stress sag in the interface are given as

rag rpu  At  l  An

sag rpu  An l  At

where rpu = 6.39f0.56


c,cube refers to the yield strength of the cement
matrix as a function of the cube compressive strength of concrete
fc,cube, l is the friction coefficient between the aggregate and the
matrix (the best agreement between experiments and theory is
achieved with l = 0.4), and An, At are the average contact areas
between aggregate particles and cement matrix perpendicular and
parallel to the crack surface, respectively. The contact areas depend
on the grading curve of the concrete and are stochastic variables.
The areas must be calculated numerically.

48

P. Huber et al. / Engineering Structures 113 (2016) 4158

Fig. 4. Contribution of aggregate interlock: (a) shear strength contribution due to aggregate interlock Vag in a typical shear test; (b) basic assumptions of the two-phase model
by Walraven [38,39] crack opening w and sliding s, contact areas an and at, and resulting normal stress rag and shear stress sag due to aggregate interlock; (c) stresses
transferred across a rough crack; (d) stresses transferred across a smooth crack (aggregate fracture); (e) push-off test with embedded bars; and (f) push-off test with external
restraint force.

Walraven [39] also published an empirical crack dilatancy


model with linear equations based on push-off tests with external
restraint bars. This approach is implemented in the fib Model Code
2010 [34] with an additional factor Cf to account for aggregate fracture. For concrete types in which the aggregate particles never split
because of concrete cracking, Cf equals 1.0. If most of the aggregate
particles fracture upon cracking, the factor Cf can be lowered to a
minimum value of 0.35. Using regression analysis, the stresses
due to aggregate interlock can be expressed as follows:

rag



C f 0:06  f c 1:35  w0:63 0:242  w0:55  0:19f c s

(a)

fc = 60 MPa, dg = 16mm
ag [MPa]
fc = 35 MPa, dg = 22mm
10
w = 0.1mm
8
6
0.5
4
2
0
1.0
-2
-4
0.5
-6
0.1
-8
0.5
1.0
1.5
2.0
0.0

3


sag C f 0:04  f c 1:8  w0:8 0:292  w0:7  0:25f c s

Another well-known approach was proposed by Bazant et al.


[41]. The so-called rough crack model was later refined by Gambarova et al. [42] by modifying the expression of the relationship
between normal stress rag and the crack kinematics (w and s). Further, in his refinement the aggregate size dg plays a greater role for
transferring shear. The stress transfer across the cracks can be calculated according to the following equations:

rag

p
0:62  w 

sag s0

r
1 r 2

0:25

 sag

s!
2w
a3 a4 jrj3
r
1
dg
1 a4 r 4

where r = s/w, s0 = 0.25fc, a3 = 2.45/s0, a4 = 2.44(14/s0).


Fig. 5 shows a comparison of the aggregate interlock effect
between the different models, using the material properties of
the two types of concrete used in the previously described test series. It can be seen that there is a pronounced difference between
the two-phase model by Walraven [38,39] (Fig. 5(a)) and the
refined rough crack model by Gambarova et al. [42] (Fig. 5(b)),

(b)
0.1
0.5
1.0
1.0

0.5
0.1

0.0

0.5

1.0

1.5

s [mm]

(c)

2.0

s [mm]

(d)

ag [MPa]
10
8
6
4
2
0
-2
-4
-6
-8
0.0

0.1

ag [MPa]

0.1

0.5

Cf = 0.5
Cf = 1.0

0.5

1.0
1.0

1.0
1.0
0.1

0.5

0.1

0.5

1.0

1.5

2.0

s [mm]

0.0

0.5

0.5

1.0

1.5

2.0

s [mm]

Fig. 5. Comparison of different crack dilatancy models for the two different types of
concrete used in the tests: (a) two-phase model by Walraven [38,39]; (b) rough
crack model by Gambarova et al. [42]; (c) linear aggregate interlock model by
Walraven [39]; and (d) linear aggregate interlock approach proposed in MC2010
[34].

especially for the SCC with a compressive strength fc of around


60 MPa. Whereas a shear stress sag of 2.39 MPa is predicted by
Walravens model for a crack opening w of 0.5 mm and sliding s
of 0.5 mm, the approach of Gambarova et al. arrives at a shear
stress sag of 7.88 MPa. Since partial fracturing of the aggregate

49

P. Huber et al. / Engineering Structures 113 (2016) 4158

particles could be observed based on visual inspection in both the


splitting tensile tests as well as in the crack surfaces of the SCC
beams for the shear tests, the factor Cf in Eqs. (3) and (4) was
reduced to 0.5 to account for smoother crack surfaces. For the
normal-strength concrete (NC) Cf was set to 1.0.
Based on the measured crack kinematics, the inclination of the
critical crack acr(t) at every single segment, and the constitutive
law for the aggregate interlock effect, the resulting shear strength
is defined as

Z
V ag b 

lcr
t0

sag  sin acr t  dt 

lcr

t0

rag  cos acr t  dt

where lcr refers to the crack length, where shear can be transferred
by aggregate interlock. To avoid an overlap with the contribution of
residual tensile stresses in the fracture process zone (see Section 3.3), the contribution of aggregate interlock at the crack tip is
disregarded.

Fig. 6(b)). As can be seen in Fig. 6(c), large plastic deformations


occurred in specimens with and without stirrups. It can further
be seen from Fig. 6(f) that the stirrups are activated as a crack
opens horizontally, resulting in an increase of the shear strength
(Vs,da) (see Fig. 6(c)). To model the formation of a dowel crack it
is common practice to model the dowel as a beam on an elastic
foundation [44], as illustrated in Fig. 6(d). Based on this model
assumption Baumann and Rsch [48] propose the following simple
equation to estimate the dowel cracking load Vda,cr:

V da;cr f ct  bn  lda

where bn = bw  Rs is the effective width of the concrete between


the longitudinal reinforcement and lda is the effective length of the
concrete in tension (assuming a rectangular stress block) below the
reinforcing bars. The effective length lda was determined empirically, resulting in the following formula for the dowel crack load
Vda,cr:

V da;cr 1:72  bn  s 

3.2. Dowel action


At the beginning of the 20th century Mrsch [43] suggested that
the longitudinal reinforcement is capable of transferring forces
perpendicular to their lengths and can thus contribute to the shear
strength of RC members. To investigate this phenomenon in more
detail, researchers have carried out many experiments with various types of test setups and specimens (block-type specimens
[44] and beams [4549]). The most extensive experimental work
in this field was done by Baumann and Rsch [48], using a test configuration for beams proposed by Krefeld and Thurston [47]. They
investigated the dowel action by loading a wedge-shaped part of a
beam, which was separated from the rest of the beam by a preformed crack, until a crack alongside the longitudinal reinforcement (dowel crack) developed (see Fig. 6(a) and (b)). Besides
investigating various influencing parameters (such as bar diameter, number of reinforcement layers, and concrete cover), the
researchers focused on the loadcarrying behavior after dowel
cracking in beams with stirrups close to the dowel crack (see

(a)

(c)

q
3
fc

Since the differences to other approaches [11,50] are small


(they are all fitted to the same test data), detailed descriptions of
these models are not given in this paper. If the beams contain no
stirrups, the dowel crack propagates horizontally while the dowel
load remains constant (see Fig. 6(c)). As the dowel crack opens a
shift of the tensile stresses rct toward the support occurs. Finally,
the increasing shear rotation leads to a total delamination of the
concrete cover.
The vertical deflection of the longitudinal reinforcement is
restrained by the transverse reinforcement. The stirrups are activated (Vs,da) by resisting the shear force and they restrict the propagation of the dowel crack. The strain and hence the force in the
transverse reinforcement can be estimated from the measured
crack width wsw at the location of the stirrup and the selected bond
stressslip relationship (further details are given below). Alternatively, the empirical relation proposed by Baumann and Rsch
[48] can be used to describe the loaddeflection behavior for the
relevant curve in Fig. 6(c):

(b)

(d)

(e)

(f)

Fig. 6. Dowel action of the longitudinal reinforcement according to Baumann and Rsch [48]: (a) test setup for beams without stirrups; (b) test setup for beams with stirrups;
(c) loaddisplacement curve for dowel action; (d) assumption of a beam on an elastic foundation for determining the dowel crack load Vda,cr; (e) dowel action after cracking;
and (f) dowel action with stirrups near the crack.

50

V da

P. Huber et al. / Engineering Structures 113 (2016) 4158

r
1
wsw  Iv

f
0:45

10

In Eq. (10) f refers to the distance from the stirrup to the nearest
vertical crack, wsw refers to the opening of the dowel crack, and Iv is
the moment of inertia of the reinforcing bar and the concrete cover
directly below the bar.
Fig. 7 shows the shear stress sda (dowel action Vda, calculated
with Eq. (10), divided by the effective web area bwd) for different
distances of the stirrups to the crack f in dependence on the vertical deflection of the longitudinal reinforcement. The loadcarrying
capacity of the stirrup Vda,max (= Aswfyw) can be regarded as the
upper limit of the dowel force. Activation of the stirrup, however,
occurs only when the dowel cracking load Vda,cr is reached. It can
also be shown that the location of the stirrup f significantly influences the loaddeflection behavior of the longitudinal
reinforcement.
3.3. Vertical component of the residual tensile stresses in the fracture
process zone
Shear can be transferred in the beams by residual tensile stresses rct at the crack tip of the critical shear crack (see Fig. 8(a)). As
opposed to cracks in metals, cracks in concrete do not exhibit a
well-defined crack tip. Therefore, in concrete a so-called fracture
process zone (FPZ) forms, in which stress is transferred across a
crack until the crack reaches a certain limit width (see Fig. 8(b)).
A widely used approach for characterizing the relationship
between residual tensile stresses rct and the crack w (also known
as tensionsoftening behavior) was proposed by Hordijk [52] (see
Fig. 8(c)):

("

rct f ct

)

3 #
w
w
6:93w=wcr
e
1 3
 0:0274 
wcr
wcr

11

In this equation, wcr = 5.14GF/fct refers to the crack limit width,


beyond which no stress is transferred across the crack. As an alternative to the exponential law in Eq. (11) the simple bilinear relationship proposed in the fib Model Code 2010 [34] can be used
(see Fig. 8(c)). Using the measured crack opening w the residual

da =Vda/bw d [MPa]

0.5
Vda,max

0.4

tensile stresses at the crack tip rct can be estimated. Furthermore,


the shear strength transmitted at the crack tip can be determined
by integration of the residual tensile stresses rct along the length of
the fracture process zone lFPZ

Z
V cr b 

t0

f = 50mm

rct  cos acr t  dt

12

where b is the width of the beam and acr is the crack angle.
The area under the tension vs. crack opening curve corresponds
to the fracture energy GF, which reflects the energy dissipated by
the opening of the crack. GF mainly depends on the compressive
strength of the concrete fc, the maximum aggregate size dg, the
water/cement ratio w/c, as well as the type of aggregate (round
or crushed aggregate) [53]. Fig. 8(d) shows a comparison of the calculated values of fracture energy GF according to different
approaches versus compressive strength fc. It can clearly be seen
that the predictions differ widely: for the SCC (fc  60 MPa,
dg = 16 mm, w/c = 0.51) used in the previously described test series, for example, the calculated fracture energy GF ranges from 96
to 153 N/m.
3.4. Contribution of the stirrups
If the beam contains a sufficient amount of transverse reinforcement, it can sustain further load after shear cracking, as the tensile
stresses acting on the diagonal crack are redistributed to the stirrups (see Fig. 9(a)). On the one hand, the stress in the shear reinforcement, is influenced by the bond conditions of the
reinforcement. On the other hand the inclination of the diagonal
crack acr is important, as it determines the number of stirrups
crossing the crack. Whereas small crack angles acr lead to the activation of more stirrups, the contribution of the aggregate interlock
effect is minimized due to the reduced sliding s of the crack surfaces. More details will be given in chapter 4.
Fig. 9(b) shows typical bond stressslip relationships for a
ribbed and a plain reinforcing bar determined from standard
pull-out tests. In ribbed bars secondary cracks develop at the ribs
as the bar reaches a peak value sb,max, after which the bond stress
decreases while the bond slip increases. For plain bars, however,
the behavior is significantly softer due to the missing mechanical
interaction of the ribs with the surrounding concrete.
To model the bond behavior of reinforcement a remarkable
number of different bond stressslip relationships have been proposed [55]. To simplify the calculation process simple rigidplastic
behavior with a constant mean bond stress sb1 according to [56] is
assumed in this paper. After yielding of a ribbed bar the bond stress
sb is reduced to sb2 = sb1/2 in order to consider the decreasing bond
stress. For plain bars the bond stress is constant and therefore
independent of slip db (see Fig. 9(b) and (d)). The mean bond stress
sb1 for a given stress of rsw = fyw can be determined from the
expression given in the work by Noakowski [57]:

100mm

0.3

lFPZ

sb1

N
!1N
1

1N
2
f yw
1N
8A
2=3
f


 w
8
1N c
Esw

13

150mm

0.2
Vda,cr

0.1
0.0
0.0

1.0

2.0

3.0

d [mm]
Fig. 7. Shear stress sda = Vda/bwd vs. displacement for different distances f between
the crack and the nearest stirrup.

where the mean bond stress sb,1 depends mainly on the type of concrete (in terms of the concrete compressive strength fc), the transverse reinforcement diameter w, and the mechanical properties
fyw and Esw of the reinforcing bar, as well as the empirical constants
for the different bond characteristics A and N (A = 0.42, N = 0.10 for
plain bars and A = 0.95, N = 0.12 for ribbed bars).
From the measured crack width wsw at the location of the stirrups it is possible to obtain the stress state and the forces in each
reinforcing bar. Such a calculation requires the use of a bondslip
relationship (see Fig. 9(c and d)). In 1973 already Leonhardt et al.

51

P. Huber et al. / Engineering Structures 113 (2016) 4158

(a)

(b)

detail

(d) 200
GF [N/m]

(c)

SCC

MC2010 [34]

150
100
50

GF

NC

Mari et al. [9]


Baant et al. [53]
MC1990 [54]

0
20

40

dg = 16 mm
60

80

100

f c [MPa]
Fig. 8. Residual tensile stresses at the crack tip: (a) resulting vertical component Vcr at the FPZ; (b) tensile stress state at the FPZ; (c) stresscrack opening relationship for a
crack (tension-softening behavior); and (d) calculated fracture energy GF as a function of the compressive strength of concrete fc according to different approaches.

(a)

(c)

(b)

(d)

(e)
[34] [56]

Fig. 9. Contribution of stirrups: (a) activated stirrups at the critical shear crack; (b) pull-out test and measured bond stressslip relationships for ribbed and plain reinforcing
bars; (c) stress state and bond behavior of an embedded stirrup; (d) assumed rigidplastic bond behavior; and (e) stress vs. crack opening behavior of a stirrup.

[58] determined the stress state in stirrups from the measured


crack opening w and a bond model. The crack width of a single
crack is given by



rsw
wsw 2  db 1
Esw

14

where wsw is the crack width at the location of the stirrups, rsw is
the stress in the stirrups, and Esw is the Youngs modulus of the
transverse reinforcement. The bond slip db can be determined by
integrating the steel strains esw using a bilinear stressstrain relationship (see Fig. 9(e)) for the transverse reinforcement:

db

w

4

Z rsw
0

1
 e r dr
sb;i rsw sw sw sw

15

As pointed out by Campana et al. [21], even a crack with a small


width can cause a reinforcing bar to yield. For example, a ribbed
stirrup with a diameter of 6 mm theoretically yields at a crack

opening of 0.3 mm in a beam with the given material properties


(see Fig. 9(e)). In contrast, smooth bars with a diameter of 3 mm
need a larger crack width wsw (in the order of 0.5 mm) to yield.
With the measured crack width wsw and the previously
described bond model the stress in each stirrup crossing the critical shear crack can be determined. If all resulting forces in the stirrups are added, the shear strength contribution Vsw of the
transverse reinforcement is

V sw

V sw;i

rsw;i 

2w  p
4

16

where Vsw,i and rsw are the shear resistance and the normal stress of
a single stirrup i. In the shear tests presented in this paper the crack
width wsw at failure is so large that all stirrups can be assumed to
have yielded. Stress levels below fyw in the transverse reinforcement only occur at the fracture process zone (FPZ) at the tip of
the critical crack.

52

P. Huber et al. / Engineering Structures 113 (2016) 4158

For beams with a very low amount of transverse reinforcement


the shape of the critical shear crack plays a crucial role for the contribution of the stirrups to the shear strength. Due to the propagation of the crack toward the load application point additional
stirrups can be activated, provided that they cross the diagonal
crack and are anchored sufficiently well in the concrete. The influence of the stirrups at the delamination crack along the longitudinal reinforcement is considered in the contribution of dowelling
(see Section 3.2).
3.5. Shear strength contribution of the uncracked compression zone
Shear can also be transferred by the uncracked compression
zone. This mechanism must not be confused with the arching
action that can be observed in deep beams. In slender beams the
compression strut is intersected by a diagonal crack, and therefore
it cannot contribute to the shear strength (see Fig. 10(a)). Nevertheless, it can be assumed that the uncracked compression zone
can carry shear as assumed in the classical beam theory (see
Fig. 10(b)):

Z
V cz
Acz

17

sxz dAcz

where Acz refers to the area and sxz to the acting shear stress of the
compression zone.
The contribution of the uncracked compression zone is strongly
influenced by the height of the compression zone, which is defined
by the shape of the critical shear crack.
For the test specimens in this research, this shear transfer
mechanism is not considered in more detail, since due to the chosen image size the performed DIC measurements are not accurate
enough to allow an interpretation of the strain state and hence the
stress state in the uncracked compression zone.
4. Discussion of the test results
The different shear transfer mechanisms in concrete beams are
investigated by using the images of an optical measurement system and the observed crack patterns. Since crack localization is
not yet implemented in the post-processing of the employed DIC
software, the critical shear crack was idealized as a polyline in
which the length of each segment corresponds to approximately
three facets (3  15 pixels) In each of these segments the crack
angle acr was determined. Thus, about 50 data points are recorded
for the critical shear crack. For each data point the movements of
two opposite facets on the concrete surface are tracked during
the entire test. Using crack angle acr and the measured displacements of the crack surfaces in the x and z directions, the crack
opening w and crack sliding s can be determined. Since the measurement accuracy of the DIC system depends on the size of the
image, the horizontal dimension of the measuring field was limited

(a)

(b)

Fig. 10. Contribution of the compression zone: (a) shear strength in the uncracked
compression zone and direct compression strut intersected by the critical shear
crack; and (b) shear stress in the compression zone.

to approximately 1.2 m so as to obtain reasonable results. Because


of this restriction it was not possible to record the kinematics of
the entire critical diagonal crack of the large specimens. From the
test results it can be seen that shear failure of the beams without
transverse reinforcement is quite brittle. Hence, from the last
recorded image before failure it was not possible to define a free
body for beams R500o60, T1000o60, and R1000o35. Since the critical shear cracks of beams R500o60, T1000o60, and R1000o35
developed from several cracks, a higher acquisition rate would
have been necessary to record the instant in which these cracks
combined to form the critical shear crack. Further, specimen
T250m60 was excluded from the analysis, since the beam could
not sustain any additional load after diagonal cracking had
occurred.
From the detailed crack kinematics and the shape of the critical
shear crack an estimation of the contributions of the different
shear transfer mechanisms can be carried out. In Fig. 11 the measured crack kinematics for 12 representative beams is shown. The
displacements of the crack surfaces were tracked before failure, at
maximum shear load VExp (failure load) and at the last recorded
picture prior to the total loss of equilibrium Vlp (see e.g. Fig. 2(c)
loaddeflection curve of beam R250o35). The kinematics based
on the photogrammetric measurements at failure load (VExp) as
well as at the last recorded image prior to the unstable crack propagation (Vlp) are used to determine the contributions of the different shear transfer mechanisms based on the previously described
theories. Fig. 11(a) shows the resulting shear forces Vi (indizes i
refers to the several shear-transfer actions) acting on a free body
for beams without shear reinforcement for the load stage at the
last recorded image prior to the total loss of equilibrium Vlp.
Fig. 11(b) shows the shear transfer mechanisms for six representative beams with a minimum number of stirrups. In Table 4 the contributions of the shear transfer mechanisms for 16 tested beams at
VExp and Vlp are summarized. Further, a comparison is shown
between the experimentally measured and estimated shear
strength VExp/RVi and Vlp/RVi the latter being determined from
the optical measurements. Since there are many different crack
dilatancy models for analyzing the contribution of the aggregate
interlock effect, it was decided to evaluate four different
approaches with regard to their suitability for predicting the stresses transferred across a crack. The results for the different
approaches for beams without transverse reinforcement are illustrated in Fig. 11(a); Table 4 shows the different results for beams
with stirrups. To compare the experimental observed shear forces
(VExp and Vlp) and the sum of the contributions of the different
shear transfer mechanisms RVi from Table 4, the contribution of
the aggregate interlock effect to the shear strength is determined
with the two-phase model proposed by Walraven [38,39]. The fracture energy GF was determined according to the approach suggested by Mari [9] to determine the residual tensile stress in the
fracture process zone Vcr.
When the relationship between the measured crack opening w
and sliding s is linked with the chosen constitutive laws, reasonable agreement between the shear force at the last recorded picture before unstable crack propagation Vlp and the sum of the
shear transferred by each mechanism RVi can generally be
achieved. In contrast, the accuracy of the prediction between the
maximum shear force in the experiments VExp and the shear
strength based on the shear-transfer actions RVi is significantly
worse. Based on this fact the following conclusions can be drawn:
 The contribution to the shear strength by aggregate interlock
plays a decisive role for beams without transverse reinforcement (096% of VExp and 44112% Vlp). This contribution
depends mainly on the crack kinematics, which is directly
dependent on the shape of the critical shear crack (see Figs. 12

P. Huber et al. / Engineering Structures 113 (2016) 4158

(a)

53

(b)

Fig. 11. Shear transfer mechanisms observed in the free body defined by the critical diagonal crack, and measured crack kinematics at the maximum load (determined from
the last image taken before beam failure) for: (a) beams without transverse reinforcement; and (b) beams with a minimum number of stirrups.

(a) and (b) and 13(a) and (b)). Steeply inclined cracks experience pronounced sliding of the crack edges s, whereas in cracks
with small crack angles the dominant action is the opening of
the crack. Thus, the aggregate interlock effect was most pronounced for beam R1000o60, with the aggregate interlock effect

constituting 96% of the maximum shear force VExp and 112% of


the shear force Vlp, respectively. This can be attributed to the
very steep inclination of the crack at web mid-height (see
Fig. 12(b)), where the sliding of the crack surfaces s was larger
than the crack opening (w). According to Walravens crack

54

Table 4
Comparison of the experimentally obtained shear strength and the calculated contributions to the shear strength of the different shear transfer mechanisms.
Vag (kN)

Vcr (kN)

Vda (kN)

Vsw (kN)

VExp(Vlp) (kN) VExp/RVi (Vlp/RVi)


(kN)

21.2(20.9)
7.0(16.5)
45.9(49.8)
243.7
(267.1)
19.5(24.8)
46.9(59.5)

19.6(19.6)
16.0(15.2)
72.4(66.9)
214.5(204.3)

0.92(0.94)
2.29(0.92)
1.58(1.34)
0.88(0.76)

23.5(23.0)
71.0(62.5)

1.20(0.93)
1.51(1.05)

m
COV

1.40(0.99)
0.34(0.18)

26.0(25.3)
29.6(27.7)
83.0(73.0)
105.9(102.2)
109.2(105.6)

0.88(0.86)
0.97(0.91)
0.67(0.93)
0.96(0.93)
0.96(0.94)

145.1(143.6)
402.1
383.8
93.5(87.4)
390.4(382.5)

1.10(1.10)
0.99
0.97
0.77(0.84)
0.92(0.89)

m
COV

0.92(0.94)
0.13(0.08)

Two-phase model
Modified rough
(see Eqs.
Linear relation
Dowel action acc.
Linear relation acc.
Tension softening acc.
crack model [42] [38,39] (see Eqs. (1), to Hordijk [52]
acc. to Walraven [38]
to Baumann et al. [48] (13)(16))
to MC 2010 [34]
a
(see Eqs. (3), (4), and (7)) (see Eqs. (3), (4), and (7)) (see Eqs. (5)(7)) (2), and (7))
(see Eqs. (11) and (12)) (see Eqs. (9) and (10))
Beams without shear reinforcement
R250o60
4.1(5.4)
R250o35
0.0(13.7)
R500o35
4.6(27.8)
R1000o60 153.6(253.6)

5.3(9.2)
0.0(12.3)
0.0(25.2)
248.3(441.7)

26.7(27.0)
5.1(21.2)
23.8(54.3)
464.0(656.0)

12.7(13.0)
0.0(12.0)
24.8(29.1)
205.6(229.0)

6.1(5.5)
5.0(2.5)
13.1(12.7)
0.0(0.0)b

2.4
2.0
8.0
38.1

0.0(0.0)
0.0(0.0)
0.0(0.0)
0.0(0.0)

T250o60
T500o60

7.6(10.2)
8.1(19.5)

19.9(36.1)
60.3(85.9)

8.1(12.9)
17.7(33.4)

9.0(9.5)
19.7(16.6)

2.4
9.5

0.0(0.0)
0.0(0.0)

Beams with
R250m60
R250m35
R500m60
R500m351
R500m352
R500m353
R1000m60
R1000m35
T500m60
T1000m60

4.4(7.0)
10.0(12.3)

shear reinforcement
1.8(1.7)
0.0(0.0)
29.3(0.0)
0.0(0.0)
0.0(0.0)
0.0(0.0)
0.0
0.0
0.0(0.0)
91.1(77.2)

3.7(3.5)
0.0(0.0)
55.2(0.0)
0.0(0.0)
0.0(0.0)

24.3(23.8)
7.2(7.2)
154.0(86.1)
14.1(11.3)
19.0(14.3)

7.6(7.4)
1.9(1.8)
62.0(18.5)
2.0(2.0)
4.1(2.9)

0.0(0.0)
0.0(0.0)
8.8(5.9)
0.0(0.0)
5.1(3.3)

5.5
2.0
18.0
17.8
17.8

0.0(0.0)
0.0
0.0
0.0(0.0)
156.2(134.8)

16.3(15.9)
46.6
2.1
95.0(73.7)
171.2(138.0)

3.7(3.4)
4.9
1.6
25.8(14.7)
43.4(30.3)

4.5(2.8)
0.0b
3.5
8.8(0.0)
7.5(0.0)

17.8
38.5
30.3
18.0
38.5

The values refer to shear transfer actions based on VExp, the values in brackets refer to Vlp.
a
Chosen for comparison between experiments and analytical predictions.
b
No measurements available.

16.5(16.5)
26.5(26.7)
35.2(36.0)
90.0(90.1)
86.5(88.6)

29.6(29.4)
30.4(30.5)
124.0(78.4)
109.8(109.9)
113.5
(112.6)
105.9(106.2) 131.9(130.2)
362.6
406.0
359.1
394.5
68.5(71.4)
121.1(104.1)
333.0
422.4
(359.3)
(428.1)

P. Huber et al. / Engineering Structures 113 (2016) 4158

RVi (kN)

55

P. Huber et al. / Engineering Structures 113 (2016) 4158

(b)
h [mm]

R250o60

(c)

250
200
150
100
50
0

1.2

w (VExp)
s (VExp)
w (Vlp)
s (Vlp)

Vlp

VExp

1.0

Vi / Vi [-]

(a)

0.8
0.6
0.4
0.2

h [mm]

R1000o60

1000
800
600
400
200
0

0.0

Vag

Vda

Vc r

w/d, s/d [mm/m]


Fig. 12. Comparison between beams R250o60 and R1000o60 without shear reinforcement: (a) shape of the critical shear cracks; (b) normalized crack opening w and sliding s
relationship for the diagonal crack; and (c) contribution of each shear transfer mechanism to the shear resistance.

(b)
h [mm]

R500m60

h [mm]

R500m351

(c)

500
400
300
200
100
0
500
400
300
200
100
0

Vsw
1.2

VExp

Vag

Vda

Vcr

Vlp

1.0

Vi / Vi [-]

(a)

0.8
0.6
0.4
0.2
0.0

R500m60

R500m351

w (VExp
s (VExp)
w (Vlp)
s (Vlp)

w, s [mm]
Fig. 13. Comparison between beams R500m60 and R500m351 with stirrups: (a) shape of the critical shear cracks; (b) crack opening w and sliding s relationship for the
diagonal crack; and (c) contribution of each shear transfer mechanism to the shear resistance.

(a)

(b)

2,0

Exp

i [MPa]

1,6
Exp

lp

lp

1,2
0,8
0,4
0,0

R250o60 T250o60 T500o60


sw
ag

R250m60 T500m60 T1000m60


da
cr

Fig. 14. Investigation of the influence of the depth and flanges at different load stages (VExp and Vlp): (a) shear transfer actions for beams without stirrups and (b) shear
transfer actions for beams with transverse reinforcement.

dilatancy model [38,39] shear stresses sag of up to 6 MPa can be


transferred by aggregate interlock in this part of the crack. As
pointed out by Campana et al. [21], the two-phase model may
overestimate the contribution of the aggregate interlock in the

case of very steep cracks. For beam R1000o60 the predicted


shear resistance due to aggregate interlock Vag is even higher
than the observed shear force Vlp (see Fig. 12(c)). The analyses
also revealed that the crack kinematics depends strongly on

56

P. Huber et al. / Engineering Structures 113 (2016) 4158

the depth of the member (see Fig. 11(a)). Thus, a size effect can
be observed in larger specimens regarding the aggregate interlock effect, due to the presence of wider cracks in these larger
beams.
The role of aggregate interlock is less important for specimens
with a low number of stirrups (the contribution to the shear
strength by the aggregate interlock effect ranges from 075% of
VExp and 029% Vlp), since the large opening of the critical shear
crack reduces the probability that aggregate particles interlock.
Fig. 13(a) shows the crack shapes of the critical shear cracks, and
Fig. 13(b) illustrates the crack openings w and sliding s for beams
R500m60 and R500m351 prior to failure. Both figures clearly show
the previously mentioned interaction between crack shape and
kinematics. Additionally, Fig. 13(c) emphasizes their influence on
the contribution of the aggregate interlock effect.
 Using other crack dilatancy models, the ratio of the experimentally determined and analytically predicted shear strength VExp/
RVi and Vlp/RVi deviates significantly from unity, and the COV
increases clearly. As shown in Table 4 and explained in detail
in the previous section, the modified rough crack model [42]
considerably overestimates the stress that can be transferred
across cracks.
 For beams without transverse reinforcement the contribution to
the shear strength by dowel action of the longitudinal reinforcement is rather small (1018% of VExp and 1019% Vlp). For the
test series presented in this paper this is due to the fact that
the effective width between reinforcing bars bn was quite small.
Since the relationship between the specimen dimensions and
reinforcement ratios etc. are the same for all the specimens,
the contribution of the dowel action is independent of the
member height. If there are stirrups near the crack, the contribution of the dowel action to the shear strength increases (7
22% of VExp and 725% Vlp). Depending on the distance of the
nearest stirrup to the dowel crack, this additional stirrup could
also contribute to the shear transfer.
 The contribution of the residual tensile stresses in the fracture
process zone (FPZ) is of particular relevance for beams without
transverse reinforcement (1838% of VExp and 1641% Vlp). The
stresses transmitted at the crack tip contribute to the size effect
that can be observed in beams without stirrups. This is illustrated in Fig. 14(a), in which the shear transfer mechanisms
(expressed as shear stresses si = Vi/(bwd)) for beams with different member depths are compared. Besides the aggregate interlock effect, the contribution of the residual tensile stresses at
the crack tip exhibit a considerable size effect. Furthermore,
Fig. 14(a) reveals that the influence of flanges is correlated to
the contribution of the residual tensile stresses at the fracture
process zone. The higher shear strength VExp of specimen
T250o60 compared to T500o60 can be attributed to the fact that
in general the crack width is not so pronounced for small members and, thus, the length lcr, where residual tensile stresses can
potentially arise, is longer in relation to the depth of the member. The difference concerning the shear strength one picture
prior failure (Vlp) between the specimen T250o60 and
R250o60 can be explained by the higher contribution of the
residual tensile stresses at the crack tip due to presence of the
flange. In contrast, the residual tensile stresses in the fracture
process zone play a minor role (011% of VExp and 08% Vlp)
for beams with a minimum amount of shear reinforcement
(see Fig. 14(b)). This can be attributed to the fact, that large
crack openings can be observed for beams with a low amount
of stirrups.

 Even a minimum number of stirrups is sufficient to provide the


main contribution to the shear strength (4294% of VExp and
4996% Vlp). How much shear is carried by the stirrups depends
mainly on the shape of the critical shear crack. Even though
cracks with small crack angles lead to an activation of a larger
number of stirrups, the shear strength contribution of the
aggregate interlock effect is minimized, since the sliding of
the crack surfaces, which is an essential requirement for interlocking, is reduced (see Figs. 11(b) and 13). Due to the propagation of the critical shear crack in the horizontal direction
additional stirrups can be activated. For beam R500m60 the
diagonal crack propagated so deeply into the compression zone
that no stirrups crossed the critical shear crack (see Fig. 13(a)),
resulting in a lower shear strength compared to beams
R500m351 and R500m352.
 As far as the influence of the different types of concrete on the
shape and kinematics of the critical shear crack are concerned,
no tendency could be observed. As shown in Table 4, the summation of the several shear transfer action Vi leads in some
cases to an underestimation of the experimental shear strength
VExp, which may be attributed to the neglect of the contribution
of the uncracked compression zone.
5. Conclusions
This investigation deals with an experimental and theoretical
assessment of the shear strength contributed to by various shear
transfer mechanisms for beams with and without a minimum
amount of transverse reinforcement. By linking the crack opening
and sliding behavior, which was determined by using digital image
correlation, with several constitutive laws from literature it was
possible to perform an estimation of the contributions of different
shear transfer mechanisms. Based on the findings of this analysis
the following conclusions are drawn:
 Digital image correlation is an adequate tool for performing
full-field measurement of the crack kinematics. In contrast to
conventional measurement techniques, DIC is capable of continuously tracking the critical shear crack. Thus it is possible
to get a better insight into the shear behavior of RC beams, even
immediately before failure.
 The crack opening and sliding behavior is directly correlated
with the shape of the critical shear crack. Steeply inclined cracks
result in pronounced sliding of the crack edges, whereas for
small crack angles crack opening is more dominant. While the
aggregate interlock effect is more pronounced in specimens
with large crack angles, the number of stirrups crossing the critical shear crack is minimized in beams with stirrups.
 For beams without shear reinforcement the contribution of the
aggregate interlock effect governs. The contributions to the
shear strength by the residual tensile stresses in the fracture
process zone (FPZ) and the aggregate interlock effect exhibit a
pronounced size effect, which is confirmed by the measured
crack kinematics. In contrast, the effect of dowel action of the
longitudinal reinforcement was found to be independent of
the specimen size in the presented test series. Further, the
amount of shear transferred by dowel action was found to be
relatively small.
 The influence of the presence of a flange on the compression
side is not very pronounced. This can be explained by the fact,
that the shear strength due to the residual tensile stresses at
the fracture process zone, is the main shear-transfer action,
which can be attributed to the flanges. Whereas for small
specimen with small crack widths this effect is significant, the

P. Huber et al. / Engineering Structures 113 (2016) 4158

influence of flanges for large beams and especially for members


with a minimum amount of transverse reinforcement (due to
larger crack openings) is negligible.
 For beams with a low number of stirrups the aggregate interlock effect plays a minor role due to the larger opening of the
critical shear crack. Nevertheless, for steeply inclined cracks
the aggregate interlock effect helps to transfer the shear which
would otherwise be transferred by the larger number of stirrups
in cracks with low crack angles. Even with a minimum number
of stirrups they represent the dominant shear transfer mechanism. From the experimental investigation and theoretical analysis it could be seen that the presence of stirrups lessens the
size effect.
 Whereas the crack kinematics was clearly dependent on member depth and the presence of stirrups, especially with respect
to the crack opening behavior, it showed no dependence on
the concrete type for either crack opening or sliding behavior.

Acknowledgements
The authors would like to acknowledge the financial support of
the Austrian infrastructure operators BB-Infrastruktur AG and
ASFINAG Bau Management GmbH as well as the Austrian Federal
Ministry for Transport, Innovation and Technology.
References
[1] Fdration Internationale du Bton (fib). Shear and punching shear in RC and
FRC elements. fib, Bulletin 57. Lausanne, Switzerland; 2010. p. 268.
[2] ASCE-ACI Committee 445 on Shear and Torsion. Recent approaches to shear
design of structural concrete. ASCE, J Struct Eng 1998;124(12):1375417.
[3] Park HG, Kang S, Choi KK. Analytical model for shear strength of ordinary and
prestressed concrete beams. Eng Struct 2013;46:94103.
[4] Tureyen AK, Frosch RJ. Concrete shear strength: another perspective. ACI Struct
J 2003;100(5):60915.
[5] Zararis P. Shear strength and minimum shear reinforcement of reinforced
concrete slender beams. ACI Struct J 2003;100(2):20314.
[6] Muttoni A, Fernndez Ruiz M. Shear strength of members without transverse
reinforcement as function of critical shear crack width. ACI Struct J 2008;105
(2):16372.
[7] Vecchio FJ, Collins MP. The modified compression-field theory for reinforced
concrete elements subjected to shear. ACI J Proc 1986;83(2):21931.
[8] Bentz EC, Vecchio FJ, Collins MP. Simplified modified compression field theory
for calculating shear strength of reinforced concrete elements. ACI Struct J
2006;103(4):61424.
[9] Mar A, Bairn J, Cladera A, Oller E, Ribas C. Shear-flexural strength mechanical
model for the design and assessment of reinforced concrete beams. In:
Structure and infrastructure engineering: maintenance management lifecycle
design and performance. Taylor & Francis Online; 2014. p. 21.
[10] Knig G, Zink M. Diagonal shear cracking in slender concrete beams (in
German: Zum Biegeschubversagen schlanker Betonbauteile). Bautechnik
1999;76(11):95969.
[11] Reineck K-H. Ultimate shear force of structural concrete members without
transverse reinforcement derived from a mechanical model. ACI Struct J
1991;88(5):592602.
[12] Bazant ZP, Kim J-K. Size effect in shear failure of longitudinally reinforced
beams. ACI J Proc 1984;81(5):45668.
[13] Bazant ZP, Yu G. Designing against size effect on shear strength of reinforced
concrete beams without stirrups: I. Formulation. ASCE, J Struct Eng 2005;131
(12):187785.
[14] Tue NV, Theiler W, Tung ND. Shear response of reinforced concrete members
without stirrups under bending (in German: Schubverhalten von
Biegebauteilen ohne Querkraftbewehrung). Beton-Stahlbetonbau 2014;109
(10):66677.
[15] Zsutty T. Shear strength prediction for separate categories of simple beam
tests. ACI J Proc 1971;68(2):13843.
[16] Nielsen MP, Braestrup MW, Bach F. Rational analysis of shear in reinforced
concrete beams. In: IABSE colloquium proc. P-15, vol. 2. Bergamo, Italy; 1978.
p. 16.
[17] Thrlimann B. Shear strength of reinforced and prestressed concrete. CEB
Bulletin, vol. 126. Paris. France: CEB; 1978. p. 1638.
[18] Sigrist V. Generalized stress field approach for analysis of beams in shear. ACI
Struct J 2011;108(4):47987.
[19] Reineck K-H. Modelling of members with transverse reinforcement. In: IABSE
colloquium on structural concrete. IABSE Report. vol. 62. Zurich, Switzerland:
IABSE; 1991. p. 48188.

57

[20] Hegger J, Grtz S. Shear capacity of beams made of normal and high
performance concrete (in German: Querkraftmodell fr Bauteile aus
Normalbeton und Hochleistungsbeton). Beton- Stahlbetonbau 2006;101
(9):695705.
[21] Campana S, Fernndez Ruiz M, Anastasi A, Muttoni A. Analysis of shear
transfer actions on one-way RC members based on measured cracking pattern
and failure kinematics. Mag Concr Res 2013;56(6):386404.
[22] Cavagnis F, Fernndez Ruiz M, Muttoni A. Shear failures in reinforced concrete
members without transverse reinforcement: an analysis of the critical
shear crack development on the basis of test results. Eng Struct 2015;103:
15773.
[23] Fernndez Ruiz M, Muttoni A, Sagaseta J. Shear strength of concrete members
without transverse reinforcement: a mechanical approach to consistently
account for size and strain effects. Eng Struct 2015;99:36072.
[24] Yang Y. Shear behaviour of reinforced concrete members without shear
reinforcement, a new look at an old problem. PhD thesis, Delft University of
Technology, Delft, Netherlands; 2014. p. 344.
[25] Knig G, Dehn F, Hegger J, Grtz S. Influence of shear friction on the shear
bearing capacity (in German: Der Einfluss der Rissreibung auf die
Querkrafttragfhigkeit). Beton-Stahlbetonbau 2000;95(10):58491.
[26] Sagaseta J, Vollum RL. Influence of aggregate fracture on shear transfer
through cracks in reinforced concrete. Mag Concr Res 2011;63(2):11937.
[27] Mphonde AG, Frantz GC. Shear tests of high and low-strength concrete beams
without stirrups. ACI J Proc 1984;81(4):3507.
[28] Angelakos D, Bentz EC, Collins MP. Effect of concrete strength and minimum
stirrups on shear strength of large members. ACI Struct J 2001;98(3):290300.
[29] Placas A, Regan P. Shear failure of reinforced concrete beams. ACI J Proc
1971;68(10):76373.
[30] Rupf F, Fernndez Ruiz M, Muttoni A. Post-tensioned girders with low
amounts of shear reinforcement: shear strength and influence of flanges.
Eng Struct 2013;56:35771.
[31] Reineck K-H, Bentz EC, Fitik B, Kuchma DA, Bayrak O. ACI-DAfStb database of
shear tests on slender reinforced concrete beams without stirrups. ACI Struct J
2011;111(5):86776.
[32] Yu Q, Bazant ZP. Can stirrups suppress size effect on shear strength of RC
beams? ASCE J Struct Eng 2011;137(5):60717.
[33] Bahl NS. Concerning the influence of the member depth on the shear strength
of simply supported reinforced concrete beams with and without shear
reinforcement (in German: ber den Einfluss der Balkenhhe auf die
Schubtragfhigkeit von einfeldrigen Stahlbetonbalken mit und ohne
Schubbewehrung. PhD thesis, University of Stuttgart, Stuttgart, Germany;
1968. p. 124.
[34] Fdration Internationale du Bton (fib), fib Model Code for Concrete
Structures 2010, Ernst & Sohn; 2013. p. 434.
[35] GOM. ARAMIS 3.1 und hher, manual; 2009.
[36] Paulay T, Loeber PJ. Shear transfer by aggregate interlock. ACI Spec Publ
1974;42(1). p. 15.
[37] Mattock JA, Johal L, Chow HC. Shear transfer in reinforced concrete with
moment or tension acting across the shear plane. PCI J 1975;20(4):7693.
[38] Walraven JC. Fundamental analysis of aggregate interlock. ASCE J Struct Div
1981;107(11):224570.
[39] Walraven JC. Aggregate interlock: a theoretical and experimental analysis. PhD
thesis, Delft University of Technology, Delft, Netherlands; 1980. p. 208.
[40] Daschner F, Kupfer H. Tests concerning shear transfer in cracks of normal and
lightweight concrete (in German: Versuche zur Schubkraftbertragung in
Rissen von Normal- und Leichtbeton). Bauingenieur 1982;57(2):5760.
[41] Bazant ZP, Gambarova PG. Rough crack models in reinforced concrete. ASCE J
Struct Eng 1980;106(4):81942.
[42] Gambarova PG, Karako C. A new approach to the analysis of the confinement
role in regularly cracking concrete elements. In: Trans. 7th struct mech in
reactor tech, vol. H; 1983. p. 25161.
[43] Mrsch E. Concrete-steel construction theory and application (in German:
Der Eisenbetonbau seine Theorie und Anwendung). Verlag von Konrad
Wittwer, 4th ed. Stuttgart, Germany; 1912. p. 710.
[44] Dei Poli S, Di Prisco M, Gambarova P. Cover and stirrup effects on the shear
response of dowel bar embedded in concrete. ACI Struct J 1993;90(4):44150.
[45] Jones R. The ultimate strength of reinforced concrete beams in shear. Mag
Concr Res 1956;8(23):6984.
[46] Fenwick RC, Paulay T. Mechanisms of shear resistance of concrete beams. ASCE
J Struct Div 1968;94(10):232550.
[47] Krefeld W, Thurston CW. Contribution of longitudinal steel to shear resistance
of reinforced concrete beams. ACI J Proc 1966;63(3):32544.
[48] Baumann T, Rsch H. Tests for the analysis of dowel action of flexural
reinforcement of reinforced concrete beams (in German: Versuche zum
Studium der Verdbelungswirkung der Biegezugbewehrung eines
Stahlbetonbalkens). Deutscher Ausschuss fr Stahlbeton, vol. 210. Berlin,
Germany: Ernst & Sohn; 1970. p. 4383.
[49] Taylor HPJ. The fundamental behaviour of reinforced concrete beams in
bending and shear. ACI Spec Publ 1974;42(3). p. 35.
[50] Vintzeleou EN, Tassios TP. Behavior of dowels under cyclic deformations. ACI
Struct J 1987;84(1):1830.
[51] Hillerborg A, Moder M, Petersson PE. Analysis of crack formation and crack
growth in concrete by means of fracture mechanics and finite elements. Cem
Concr Res 1976;6(6):77381.
[52] Hordijk DA. Tensile and tensile fatigue behaviour of concrete, experiments,
modelling and analyses. Heron 1992;37(1):79.

58

P. Huber et al. / Engineering Structures 113 (2016) 4158

[53] Bazant ZP, Becq-Giraudon E. Statistical prediction of fracture parameters of


concrete and implications for choice of testing standard. Cem Concr Res
2002;32(4):52956.
[54] Comit Euro-Internationale du Bton (CEB/FIP). CEB/FIP Model Code 1990.
Thomas-Telford; 1993. p. 460.
[55] Fdration Internationale du Bton (fib). Bond and anchorage of embedded
reinforcement: background to the fib Model Code for Concrete Structures
2010. fib, Bulletin 72, Lausanne, Switzerland; 2014. p. 170.
[56] Sigrist V. Deformation capacity of reinforced concrete beams (in German: Zum
Verformungsvermgen von Stahlbetontrgern). PhD thesis, ETH Zurich, vol.
11169. Zurich, Switzerland; 1995. p. 159.

[57] Noakowski P. Reinforcing of reinforced concrete elements under constraint


forces due to temperature (in German: Die Bewehrung von
Stahlbetonbauteilen bei Zwangsbeanspruchung infolge Temperatur).
Deutscher Ausschuss fr Stahlbeton, vol. 296. Berlin, Germany: Ernst &
Sohn; 1978. p. 144.
[58] Leonhardt F, Koch R, Rostsy FS. Shear tests on prestressed concrete beams (in
German: Schubversuche an Spannbetontrgern). Deutscher Ausschuss fr
Stahlbeton, vol. 227. Berlin, Germany: Ernst & Sohn; 1973. p. 180.

Potrebbero piacerti anche