You are on page 1of 15

Engineering Structures 29 (2007) 29873001

www.elsevier.com/locate/engstruct

Direct strut-and-tie model for single span and continuous deep beams
Ning Zhang, Kang-Hai Tan
School of Civil and Environmental Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
Received 8 October 2006; received in revised form 4 January 2007; accepted 2 February 2007
Available online 27 March 2007

Abstract
A modified strut-and-tie model (STM) for determining the shear strength of reinforced deep beams is proposed in this paper. The modified
model is based on Mohr Coulombs failure criterion. Several improvements have been made to the original model [Tan KH, Tong K, Tang CY.
Direct strut-and-tie model for prestressed deep beams. J Struct Eng 2001;127(9):107684; Tan KH, Tang CY, Tong K. A direct method for deep
beams with web reinforcement. Mag Concr Res 2003;55(1):5363]. The influence of the stress distribution factor k on the model prediction is
studied quantitatively, and a parametric study shows that the k value derived from a linear stress assumption is sufficiently accurate for predicting
deep beam shear strength. The use of an interactive failure criterion (Mohrs failure equation) to model the tensioncompression nodal zones in
STM means that no empirical stress limit is needed for calculating the ultimate strength, and that the softening effect of concrete compressive
strength due to transverse tensile strain is taken into consideration. Unlike the original model, the softening effect in the modified model is
explicitly derived and is comparable to MCFT and Belarbi and Hsus equations. The modified model for simply supported deep beams (SSDBs) is
evaluated using 233 test results and it gives better agreement than the original model. The modified STM is further applied to concrete continuous
deep beams (CDBs) and is in good agreement with a total of 54 experimental results.
c 2007 Elsevier Ltd. All rights reserved.

Keywords: Continuous deep beams; Interactive; Reinforced concrete; Single span; Strut-and-tie; Ultimate strength

1. Introduction
While the strut-and-tie model (STM) is a conceptually
simple design tool, common practices of STM involve
assuming different levels of stress limits under different stress
conditions on the STM components, i.e. nodal zones and struts.
However, although stress limits are set in such a way as to
include all the stress conditions, they are empirically defined
and may be over- or under-conservative for some special
conditions. There has been considerable debate about the value
of concrete efficiency factors. Researchers in the past have
tried to postulate various stress criteria for the concrete strut
and nodal zones in the STM (Marti [1], Schlaich et al. [2],
MacGregor [3], Bergmeister et al. [4], Alshegeir [5], Foster and
Malik [6]). It should be noted that ACI 318-02 [7], CSA [8]
and CEB-FIP model code 1990 [9] give different values on
the stress limits according to their respective investigations.
Another school of researchers (Mau and Hsu [10], Hwang
Corresponding author. Tel.: +65 679 052 97; fax: +65 679 166 97.

E-mail address: ckhtan@ntu.edu.sg (K.-H. Tan).


c 2007 Elsevier Ltd. All rights reserved.
0141-0296/$ - see front matter
doi:10.1016/j.engstruct.2007.02.004

et al. [11]) relate the stress limits with the transverse strain
perpendicular to the stress direction so as to consider the
softening effect of concrete compressive strength. However,
it may not be appropriate to apply the smeared crack model
concept to the D-Region, that is, the disturbed region where
the conventional plane-section-remains-plane principle is not
valid [2].
Tan et al. [12,13] have developed a direct STM for
simply supported deep beams, which can consider the
effects of different web reinforcement configurations (vertical,
horizontal, or inclined), and prestressing tendons. Other than
defining the stress limits for the STM components, the model
utilizes a failure criterion from the MohrColumb theory for
nodal zones (tensioncompression stress state) as below:
f2
f1
+ 0 =1
ft
fc

(1)

where f 1 and f 2 are principal tensile and compressive stresses


at the nodal zone respectively; f c0 is the cylinder compressive
strength representing the maximum compressive strength in
the f 2 direction; and f t represents the maximum tensile

2988

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Nomenclature
Shear span measured between concentrated load
and the support centre
Ac
Beam effective cross-sectional area, equal to
bw dc
As1 , As2 Area of top and bottom longitudinal reinforcement, respectively
Astr1 , Astr2 , Astr3 Cross-sectional areas of the concrete
diagonal struts at node A, B, C, respectively
(Fig. 9)
Astr3 , Astr4 Average cross-sectional areas of the outer and
inner concrete struts, respectively (Fig. 8)
Asw
Total area of web reinforcement that criss-crosses
the diagonal strut
Asw1
Area of single web steel bar that criss-crosses the
diagonal strut
Ash , Asv Total areas of vertical and horizontal steel reinforcement that criss-cross the strut, respectively
bw
Width of deep beam
c1 , c2 Distances from the centroid of the top and bottom
longitudinal steel to the beam top and beam soffit,
respectively
dc
Effective beam depth (Figs. 1 and 7)
dw
Distance from the beam top to the intersection of
web reinforcement with the line connecting the
support center and the load center.
E c , E s Elastic modulus of concrete and steel, respectively
f 1 , f 2 Principal tensile and compressive stresses, respectively
f c0
Concrete compressive strength
f ct
Tensile contribution of concrete to the composite
tensile capacity
f st
Tensile contribution of reinforcement to the
composite tensile capacity
ft
Combined tensile strength of reinforcement and
concrete
f y1 , f y2 Yield strength of top and bottom longitudinal
steel reinforcement, respectively
f yh , f yv Yield strength of horizontal and vertical web
reinforcement, respectively
f yw
Yield strength of web reinforcement
Fc
Compressive force in the diagonal strut of a
SSDB
Fc1 , Fc2 Compressive force in the outer and inner
diagonal strut of a CDB,
h
Overall height of deep beam
k
Stress distribution factor at the bottom nodal zone
k0
Stress distribution factor at top nodal zone
la , lb , lf Widths of the load bearing, outer support, and
inner support plates, respectively
lc , ld
Effective depths of the bottom and top nodal
zones, respectively
le
Effective span measured between center-tocenter of supports
a

m, n, p Ratios of axial stiffness of truss members in STM


ns
Nos. of web steel bars evenly distributed along
the strut
pt
Average equivalent tensile stress across the
diagonal strut
P
Load applied on the CDBs
Pexp
Experimental ultimate load of CDBs
Pn
Predicted ultimate load of CDBs
T
Tensile force in the reinforcement
T1 , T2 Tension force in the top and bottom steel,
respectively
T1 max , T2 max Yield strength of top and bottom steel,
respectively
T1a , T2a Corresponding tension force in the top and
bottom steel at the yielding of bottom and
top steel, respectively. They are limited by the
respective top and bottom steel yield strength.
Uc
Total complementary energy in the truss
Vexp
Experimental shear strength of SSDBs
Vn
Predicted shear strength of SSDBs
v
Concrete softening coefficient
X
Reaction force of the inner support of CDBs
1 , 2 Principal tensile and compressive stress at the
tensioncompression nodal zone, respectively
s
The strain in the reinforcement
w
Angle between the web reinforcement and the
horizontal axis of beams at the intersection of the
reinforcement and the diagonal strut (Fig. 1)
s
Angle between the longitudinal tension reinforcement and the diagonal strut
a
Total reinforcement ratio of beams, equal to
(As1 + As2 )/Ac
s
Effective longitudinal reinforcement ratio of the
beam, equal to As /Ac
capacity in the f 1 direction. It is worth noting that the term
f t is the combined tensile strength comprising contributions
from concrete and reinforcement (web and main). This failure
criterion is adopted because it offers a simple linear interactive
relationship between two limit failure modes, namely, brittle
concrete crushing and ductile reinforcement yielding. By
adopting the interactive stress-based failure criterion, i.e.
Mohrs failure criterion, no other empirically set stress limit is
needed in the model for calculating the ultimate strength. Case
studies have been carried out and the model gives consistent and
conservative predictions of the ultimate strengths of normal and
high strength concrete deep beams (Tan et al. [12,13]).
However, there are several limitations in Tans model, as
follows. Firstly, when deriving the stress distribution factor
k, only the force equilibrium was satisfied. Secondly, the
model was highly sensitive to the value of concrete tensile
strength, which is difficult to be determined accurately in
practice. Thirdly, the component force of bottom reinforcement
in the direction of the concrete diagonal strut is neglected
for simplicity. This term can be significant if the strut angle

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Fig. 1. Determination of tensile stress factors at nodal zones.

Fig. 2. Assumed tensile stress distribution arising from bottom steel.

is small. Fourthly, the softening factor which is implicit


in the model was not derived explicitly. The authors reexamine Tans model [12,13] and its shortcomings and propose
a modified STM. The stress distribution factor k is newly
derived from the consideration of both force and moment
equilibrium, and its influence on model prediction of beam
shear strength is discussed through a parametric study. Concrete
tensionstiffening properties are used instead of concrete
tensile strength to improve model prediction consistency. The
component force of tension tie in the direction of the concrete
diagonal strut is also included in the model for completeness.
The authors also illustrate that, by the use of interactive failure
criteria (such as Mohrs failure criterion), the softening effect
of concrete strength due to the presence of transverse tensile
strain is implicitly taken into consideration. The softening
effect in the authors model is comparable to MCFT (Vecchio
and Collins [14]) and Belarbi and Hsu [15]s equations. The
modified model is verified against 233 test results of normal and
high strength concrete simply supported deep beams (SSDBs).
Comparison results show that the modified model yields more
accurate and consistent predictions of beam ultimate shear
strength than the original one [12,13]. The modified STM is
further extended to concrete continuous deep beams (CDBs)
for the first time and is shown to be in good agreement with
54 published test results. It is also interesting to see that
the distribution of reaction forces among 3 supports can be
reasonably well predicted by the proposed strut-and-tie model.
2. Modified strut-and-tie model
From the equilibrium of forces at the bottom nodal zone of
the inclined strut (Fig. 1)
Vn
sin s
Vn
Ts =
tan s
Fc =

2989

(2)
(3)

where Fc and Ts are the forces in the concrete strut and bottom
reinforcement (tension tie), respectively. Vn is the shear strength
of the beam. The inclined angle of the diagonal strut s can be

computed from
tan s =

lc
2

ld
2

(4)
a
where h is the beam depth, lc and ld are the respective depths of
bottom and top nodal zones, and a is the shear span measured
from centre lines between the load and support bearing plates.
In the original model, the principal tensile stress f 1
at the tensioncompression nodal zone arises from the
component force of longitudinal reinforcement in the direction
perpendicular to the diagonal strut, namely, Ts sin s . Thus, f 1
can be expressed by
f1 =

kTs sin s
Ac /sin s

(5)

where Ts sin s /(Ac /sin s ) is the average equivalent tensile


stress across the diagonal strut and Ac is the effective beam
cross-sectional area, k in the numerator is a factor taking
account of the non-uniformity of the stress distribution. The
original model assumed a triangular stress distribution along
the diagonal strut due to the presence of the bottom steel.
According to force equilibrium, namely, by equating Ts sin s
to the force represented by the triangular stress block, Tan
et al. [12] proposed k = 2. Although the original model
satisfies force equilibrium, it is worth noting that the moment
equilibrium is violated. To satisfy both moment and force
equilibrium, the stress distribution is shown in Fig. 2 and the
factor k can be determined accordingly.
First, consider one reinforcing bar that criss-crosses the
diagonal strut and inclines at an angle w from horizontal
(Fig. 1). The effect of single reinforcement T is smeared
across the entire strut length. For simplicity, the tensile stress
due to tension force T is assumed to distribute linearly along
the diagonal strut. To achieve equivalence between the discrete
force T and the assumed tensile stress distribution, the latter
must satisfy both force and moment equilibrium. Applying
force equilibrium in the f 1 direction to tensile force T of the
steel bar and the idealized stress distribution along the diagonal
strut (Fig. 1), the following equation can be established:


k + k0
pt bw dc /sin s = T sin (s + w )
(6)
2

2990

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

where k and k 0 are the stress distribution factors at the respective


bottom and top nodal zones, s is the diagonal strut angle, bw is
the beam width, and dc is the beam effective depth taken as the
vertical distance between the centroids of top and bottom nodal
zones. It should be noted that pt = T sin(s + w )/(Ac /sin s )
is the average equivalent tensile stress across the diagonal strut
due to the component of T in the principal tensile direction of
the bottom nodal zone.
From moment equilibrium about the top node between T
and the idealized tensile stress distribution:
 0

k
k k0
+
(dc /sin s )2 bw pt = T sin (s + w )
2
3
(dw /sin s ) .

(7)

Comparing Eqs. (6) and (7), the following factors for


determining the principal tensile stress at the respective top and
bottom nodal zones can be obtained

dw

k 0 = 4 6
dc
(8)
d

k = 6 w 2.
dc
For the case of bottom reinforcement (Fig. 2), dw = dc and
w = 0, the stress distribution factor is:
 0
k = 2 (compression)
(9)
k=4
(tension).
Thus, the principal tensile stress f 1 in the model can be
expressed as below:
f1 =

4Ts sin s
.
Ac /sin s

(10)

The maximum tensile capacity in the f 1 direction is a composite


term and can be expressed by
f t = f st + f ct .

(11)

concrete strut and can be calculated from


1 = s + (s + 2 ) cot2 s

(13)

where s and 2 are the tensile strain of longitudinal steel


and peak compressive strain of the concrete strut at crushing,
respectively. Usually 2 is taken as 0.002 for normal strength
concrete. In the modified model, the term f ct is relatively
small compared to the tensile contribution from reinforcement
f st and can usually be neglected for conservatism. However,
for improved accuracy in comparison, it is included in this
proposed model. Thus, through incorporating Eq. (12), the
modified model is not so dependent on the magnitude of
concrete tensile strength and is more robust in prediction
consistency.
The term f st represents the contribution from steel
reinforcement. It consists of two parts: f sw from the web
reinforcement and f ss from the longitudinal reinforcement.
f st = f sw + f ss .

(14)

The presence of web reinforcement in the strut restricts the


diagonal crack from quickly propagating to either end of the
strut. Although this stitching action only occurs at discrete
locations of web reinforcement, the congregate sum effects are
considered in the STM at the interface between the top-andbottom nodal zones and the diagonal strut itself.
For web reinforcement, a composite factor can be obtained
from the summation of k factors of all individual web steel bars.
Assume that there are n s web steel bars evenly distributed along
the strut, the stress distribution factors for web steels can be
written as below:


ns 
X
dwi

= 4n s 3n s = n s
4

6
k
=

dc
i=1
(15)

ns 
X

dwi

k =
6

2
=
3n

2n
=
n
.

s
s
s

dc
i=1

The term f ct represents the contribution from concrete


tensile strength. The original model (Tan
p et al. [12,13]) uses
the concrete tensile strength f ct = 0.5 f c0 . However, at the
ultimate state, as concrete cracks extensively, concrete tensile
strength is significantly reduced. It would be un-conservative to
use full concrete tensile strength as f ct . Besides, the predictions
of the original model are highly dependent on the magnitude
of concrete tensile strength, which is to some extent difficult to
be determined accurately. This is one factor that has significant
influence on the model accuracy and consistency. To improve
on it, a concrete tension stiffening effect is used instead. The
following equation for the tensile strength of cracked reinforced
concrete proposed by Belarbi and Hsu [16] is adopted in the
modified model:
 0.4
p
cr
0
f ct = 0.31 f c
(12)
1

Thus, the tensile contribution of web reinforcement at the


interface of the nodal zone can be calculated as

where cr is the strain of concrete at cracking, taken as 0.00008


(Belarbi and Hsu [16]). 1 is the principal tensile strain of

The principal compressive stress f 2 in the direction of the


left diagonal strut at the bottom nodal zone is computed from:

f sw =

Asw f yw sin(s + w )
Ac /sin s

(16)

where Asw = n s Asw1 represents the total area of web


reinforcement criss-crossing the concrete strut, where Asw1 is
the area of an individual web steel. For the common cases of
vertical or horizontal web reinforcement, Eq. (16) reduces to
Asv f yv sin 2s
2Ac

sin2

or sh yhAc s , respectively, where, Asv and Ash


are the respective total areas of vertical and horizontal web
reinforcement within the shear span.
The contribution of bottom longitudinal steel f ss can be
obtained in a similar fashion as Eq. (10):
f ss =

4As f y sin s
.
Ac /sin s

(17)

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

f2 =

Fc T cos s
Astr

2991

(18)

where Astr is the cross-sectional area at the bottom end of


the diagonal strut. In the original STM, the component force
T cos s of the main longitudinal reinforcement is omitted for
simplicity and conservatism. It was argued that for a small a/d
ratio of deep beams, this component would be insignificant,
as cos s approaches zero. However, this may not always
be the case. In many instances the a/d ratio is relatively
high so that T cos s becomes significant and its influence
on ultimate strength is no longer negligible. By omitting it,
the original model underestimates the shear strength of deep
beams.
Substituting Eqs. (2) and (3) into Eqs. (10) and (18), and
combining with Eq. (1), the following expression for the
nominal shear strength Vn is proposed:
Vn =

1
4 sin s cos s
Ac f t

sin s
Astr f c0

(19)

It is noteworthy that the top nodal zone is subjected to a biaxial


compressioncompression stress state. The authors suggest that
the compressive stress along the diagonal strut at the interface
of the top nodal zone should not exceed f c0 . Therefore, due to
the interactive nature between tension and compression stresses
in the failure criterion, Eq. (19) safeguards the compression
failure of the top nodal zone, given that the width of the top
nodal zone is comparable with that of the bottom support
region. Thus, no further consideration is given to the top node.
The accurate value of angle s needs to be determined
through an iterative procedure as documented in the original
model (Tan et al. [12,13]), because ld is initially unknown in
Eq. (4). In practice, the iterative procedure can be simplified
by assuming the depth of the top nodal zone to be equal to
that of the bottom nodal zone, i.e. ld = lc in Fig. 1. By
doing so, it is found that the error introduced in Vn is generally
less than 2%, which is in accordance with the findings of
previous researchers (Tan et al. [12], Tang and Tan [17]). This is
because ld is typically ten times smaller than the beam height h.
Therefore, even a rough estimate of ld will have little effect on
the accuracy of the inclined strut angle s . However, for better
agreement with test results, the iterative procedure is adopted
in this paper for model verifications. Usually, the convergence
is very rapid. An iteration of at most 3 or 4 rounds will be
sufficient to obtain a convergent result. The entire procedure
can be easily implemented using a spreadsheet. The iterative
procedure for calculating the ultimate strength of deep beams
by the modified model is shown in Fig. 3 for the purpose of
implementation.
In summary, the major modifications made to the original
STM [12,13] are listed below. First, the stress distribution factor
k in the model is obtained considering both force and moment
equilibrium. Second, the component of tension force of the
reinforcement in the direction parallel to the concrete diagonal
strut is included in the calculation of principal compressive
stress f 2 . Third, the modified STM incorporates the more
realistic tension stiffening effect rather than the full concrete

Fig. 3. Iteration procedure for computing the ultimate strength of SSDBs.

tensile strength when evaluating the composite concrete tensile


strength.
3. Verification of the modified STM
The shear strength predictions of 233 simply supported
deep beams, shown in Table 1, have been evaluated by the
original STM and the proposed modified STM (Eq. (19)). The
concrete strength of the deep beams varied from 16 to 120 MPa,
including normal and high strength concrete. The beams had
an overall depth h ranging from 200 to 1750 mm, and an a/d
ratio from 0.28 to 2.0. The longitudinal main reinforcement
ratio ranged from 0.90% to 4.07%, and vertical and horizontal
web reinforcement ratios ranged from zero to 2.86%, and
zero to 3.17%, respectively. The predicted shear strengths Vn
versus experimental shear strengths Vexp are plotted in Fig. 4b,
and c for both original and modified models. The average
(AVG) and coefficient of variation (C.O.V.) Vn /Vexp ratios have
also been reported in the figure. The modified STM predicts
the ultimate shear strength more accurately and consistently
than the original STM, because the C.O.V. obtained with the
modified STM is 38% lower [(0.130 0.212)/0.212] while the
AVG is 7% higher [(0.91 0.85)/0.85].
The results obtained in terms of shear strength from the
modified model on the 212 specimens have also been compared
with results obtained from ACI 318-99. Ten specimens from

2992

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Table 1
SSDBs used for experimental comparison
Kong et al. [18]: S-30, S-25, S-20, S-15, S-10, D-30, D-25, D-20, D-15, D-10
Smith and Vantsiotis [21]: 0A0-44, 0A0-48, 1A1-10, 1A3-11, 1A4-12, 1A4-51, 1A6-37, 2A1-38, 2A3-39, 1A4-40, 2A6-41, 3A1-42, 3A3-43, 3A4-45, 3A6-46,
0B0-49, 1B1-01, 1B3-29, 1B4-30, 1B6-31, 2B1-05, 2B3-06, 2B4-07, 2B4-52, 2B6-32, 3B1-08, 3B1-36, 3B3-33, 3B4-34, 3B6-35, 4B1-09, 0C0-50, 1C1-14,
1C3-02, 1C4-15, 1C6-16, 2C1-17, 2C3-03, 2C3-27, 2C4-18, 2C6-19, 3C1-20, 3C3-21, 3C4-22, 3C6-23, 4C1-24, 4C3-04, 4C3-28, 4C4-25, 4C6-26, 0D0-47,
4D1-13
Subedi et al. [24], Subedi [25]: 1A2, 1B1, 1B2, 1C2, 1D1, 1D2, 2A2, 2B2, 2D1, 2D2, 4G1, 4G2, 4G3, 4G4
Walraven and Lehwalter [26]: V711/4, V022/4, V511/4, V411/4, V711/3, V022/3, V511/3, V411/3, V711, V022, V511
Tan et al. [20]: A-0.27-2.15, A-027-3.23, A-0.27-4.30, A-0.27-5.38, B-0.54-2.15, B-0.54-3.23, B-0.54-4.30, B-0.54-5.38, C-0.81-2.15, C-0.81-3.23, D-1.08-2.15,
D-1.08-3.23, D-1.08-4.30, D-1.08-5.35, E-1.62-3.23, E-1.62-4.30, E-1.62-5.38, E-2.16-4.30, G-2.70-5.38
Tan et al. [27]: I-1/0.75, I-3/0.75, I-4/0.75, I-6S/0.75, II-2N/1.00, II-3/1.00, II-5/1.00, II-6N/1.00, III-2N/1.50, III-2S/1.50, III-6N/1.50
Tan et al. [23]: 1-2.00/0.75, 1-2.00/1.00, 2-2.58/0.25, 2-2.58/0.50, 2-2.58/0.75, 3-4.08/0.25, 3-4.08/0.50, 3-4.08/0.75, 3-4.08/1.00
Foster and Gilbert [28]: B1.2-1, B1.2-2, B1.2-3, B1.2-4, B2.0-1, B2.0-2, B2.0-3, B2.0A-4, B2.0B-5, B2.0C-6, B2.0D-7, B3.0-1, B3.0-2, B3.0-3, B3.0A-4, B3.0B-5
Tan and Lu [29]: 1-500/0.50, 1-500/0.75, 1-500/1.00, 2-1000/0.50, 3-1400/0.50, 4-1750/0.50
Shin et al. [30]: MHB1.5-50, MHB1.5-75, MHB1.5-100, MHB2.0-50, MHB2.0-75, MHB2.0-100, MHB2.5-50, MHB2.5-75, MHB2.5-100, HB1.5-50, HB1.5-75,
HB1.5-100, HB2.0-50, HB2.0-75, HB2.0-100, HB2.5-50, HB2.5-75, HB2.5-100
Oh and Shin [31]: N4200, N42A2, N42B2, N42C2, H4100, H41A2(1), H41B2, H41C2, H4200, H42A2(1), H42B2(1), H42C2(1), H4300, H43A2(1), H43B2,
H43C2, H45A2, H45B2, H45C2, H41A0, H41A1, H41A2(2), H41A3, H42A2(2), H42B2(2), H42C2(2), H43A0, H43A1, H43A2(2), H43A3, H45A2(2), U41A0,
U41A1, U41A2, U41A3, U42A2, U42B2, U42C2, U43A0, U43A1, U43A2, U43A3, U45A2, N33A2, N43A2, N53A2, H31A2, H32A2, H33A2, H51A2, H52A2,
H53A2
Yang et al. [32]: L5-40, L5-60, L5-60R, L5-75, L5-100, L10-40R, L10-60, L10-100, UH5-40, UH5-60, UH5-75, UH5-100, UH10-40, UH10-40R, UH10-60,
UH10-100

Kong et al. [18] have been excluded from the case study because
ACI 318-99 does not include the design of deep beams with
inclined web reinforcement. The predicted shear strength Vn
versus experimental shear strength Vexp are plotted in Fig. 4(a)
for the ACI code. It can be observed that while ACI 318-99 [19]
gives AVG = 0.72 and C.O.V. = 0.526, the modified model
provides AVG = 0.91 and C.O.V. = 0.130, highlighting its
accuracy and consistency.
4. Influence of factor k on the prediction of ultimate
strength Vn
The actual tensile stress distribution along the diagonal strut
is highly nonlinear and difficult to be determined mechanically.
For simplicity, it is assumed that the tensile stress transverse
to the diagonal strut is kpt at the bottom nodal zone. Thus the
maximum tensile capacity in the f 1 direction is:
ft =

f yw Asw sin(s + w )
k f y As sin s
+
+ f ct .
Ac /sin s
Ac /sin s

(20)

The beam ultimate shear strength can be expressed from


Eq. (19)
Vn =
=

1
k sin s cos s
f t Ac

sin s
f c0 Astr

1
cos s

A f
sin( + )
ct Ac
f y As sin s + sw yw k s w + kfsin
s

sin s
f c0 Astr

(21)

From the mathematical form of Eq. (21), it is noted that if the


deep beam is not provided with any web reinforcement Asw and

the concrete tensile contribution f ct is ignored, the prediction


of ultimate shear strength Vn is completely independent of
the k factor. That is to say, whatever the form of stress
distribution is along the diagonal strut, the model predicts the
same value of Vn for the beam. However, as mentioned in
the previous section, for optimal predictions, f ct is considered.
Thus, from Eq. (21), it can be deduced that with an increase
in k factor, Vn will slightly decrease. This is because as k
increases, the contributions from web steel and concrete (which
are independent of k) to the composite tensile strength f t
become relatively insignificant. Thus the f 1 / f t ratio increases
as k increases, resulting in a relatively smaller failure load. To
demonstrate how the value of k influences the ultimate shear
strength quantitatively, a parametric study is carried out. Twohundred-and-sixteen fictitious beams with the same a/d ratio
of 0.75 and a main reinforcement ratio of 1.5% are analyzed
using the proposed STM. The variables are web reinforcement
ratio (w ) ranging from 0% to 1.0% and concrete strength ( f c0 )
from 20 to 80 MPa. The influence of the factor k on the model
prediction of beam shear strength Vn can be represented by the
percentage of Vn , expressed by (Vn Vk=4 )/Vk=4 , where Vk=4
represents the beam shear strength obtained by the proposed
STM with k = 4. The plotted curves are shown in Fig. 5.
For higher web reinforcement ratios, the model predictions
decrease more quickly as the factor k increases. Similarly,
for higher concrete strengths, the model predictions decrease
more quickly as k increases. Generally, there is about 3%
reduction in shear strength when w goes from 0% to 1.0%
and about 2% reduction when f c0 goes from 20 to 80 MPa. For
2 k 4, the predictions are greater than Vk=4 , but when
4 k 10, the predictions are smaller than Vk=4 . However,

2993

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

5. Softening effect in the STM


Experimental studies have provided strong evidence that the
ability of diagonally cracked concrete to resist compression
decreases as the amount of tensile straining increases. A
number of analytical models (Vecchio and Collins [14,22],
Belarbi and Hsu [15]) have been proposed to represent the
compression softening effect observed in cracked reinforced
concrete in tensioncompression stress states. According to the
proposed STM, the equation that yields the concrete softening
coefficient v can be expressed as:
Fc
= v f c0 .
Astr

(22)

Substituting Eqs. (10), (18) and (22) into Eq. (1), the
following equation can be obtained:
v =1+

4Ts sin2 s
Ts cos s

Astr f c0
Ac f t

(23)

where Ts is the tension force in the bottom reinforcement


and can be expressed as s E s As . It is noteworthy that the
second term in Eq. (23) represents the contribution of the
tension force of the bottom steel resolved in the direction
of the diagonal strut, which negates a certain amount of
compression force in the strut itself. The third term in Eq. (23)
represents the contribution from transverse tensile stress due
to the bottom steel, which is the main factor for concrete
softening. Substituting Eq. (20) into Eq. (23) and utilizing the
relationship between Ts and s , the following expression for the
concrete softening coefficient is established,
v = 1+

E s As s cos s
f c0 Astr

4 fy +

Fig. 4. Ultimate strength predictions by means of: (a) ACI 318-99 (212 beams);
(b) Original STM; (c) Modified STM for 233 reinforced concrete tested deep
beams.

as shown in Fig. 5 the reduction is generally less than 5% of


the predicted beam shear strength when k is greater than 4.
Thus, it can be concluded that k = 4 derived from the linear
stress distribution assumption is sufficiently accurate for the
model predictions. Besides, the modified model now satisfies
both force and moment equilibrium.

4E s s
Asw sin(s +w )
f yw
As sin s

f ct
s sin2 s

(24)

where E s and As are respectively the elastic modulus and crosssectional area of longitudinal reinforcement; s is the strain in
the longitudinal reinforcement; Asw is the cross-section area of
web reinforcement crossing the diagonal strut; s is the inclined
angle of the strut; f y and f yw are the yield strengths of the
longitudinal and web reinforcement, respectively; Astr is the
cross-sectional area of the diagonal strut; s = As /Ac is the
effective longitudinal reinforcement ratio.
From Eq. (24), it is worth noting that the softening factor
v in the STM is not only influenced by the strut angle s but
also by other variables such as concrete compressive strength,
yield strength and reinforcing ratios of steel reinforcement
(main and web). However, the strut angle s is the main
variable that affects the softening coefficient v significantly.
For a larger reinforcement ratio or a greater yield strength of
steel, the slope becomes gentler, i.e. the decreased portion of
concrete compressive strength is less than lightly reinforced
beams. Web reinforcement seems to reduce the softening slope,
but the effect is minimal. A typical deep beam specimen,
say, 34.08/0.75 is chosen and the softening relationship by
the STM is plotted in Fig. 6 together with Vecchio and

2994

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Fig. 5. Influence of factor k on ultimate shear strength predictions.

Fig. 6. Typical maximum concrete compressive stress as function of tensile


strain in reinforcement.

Collins [14] and Belarbi and Hsu [15] softening curves. Details
of the specimen are reported in Tan et al. [23]s experimental
programme. In the figure, s is the strain in the bottom main
steel and the strut angle s is 45 . It is shown that the concrete
compressive strength has a nearly linear descending trend with
an increase of strain in the bottom steel. This is due to the
adoption of a linear interactive failure criterion in the STM.
Therefore, although the failure criterion adopted in the STM
is based on stress values (Eq. (19)), the softening effect of
concrete compressive strength due to transverse tensile strain is

implicitly taken into account through Eq. (24). It is comparable


to the softening curves proposed by Vecchio and Collins [14]
and Belarbi and Hsu [15]. Furthermore, the descending trend
with an increase of steel strain is determined not only by the
strut angle s but also by the web and main reinforcement
details. This seems more reasonable for deep beam specimens.
As can be seen from Fig. 6, the softening coefficient of STM
is slightly larger than the other two curves when steel strain
s is relatively small. This is the situation when the beam is
over-reinforced so that the strain in longitudinal reinforcement
is small when the beam approaches failure. When s approaches
the yield strain y of the longitudinal reinforcement (for f y =
499 MPa, E = 20 000 MPa, y = 0.0025), the STM
softening coefficient v is comparable to the other softening
curves (Fig. 6), and when s exceeds the yield strain, the STM
coefficient is less than those determined by the other three
equations, indicating a greater softening effect.
6. Modelling continuous deep beams
A STM for two-span continuous deep beams (CDBs) with
a top point load at each mid-span is given in Fig. 7. It can
be idealized as a statically indeterminate truss as shown in
Fig. 8. Assuming perfect elasticplastic material properties for
concrete and steel bars, the internal forces of the truss are solved
(detailed derivation can be found in Appendix A). They are
denoted as Fc1 = A P, T1 = B P, T2 = C P, Fc2 = D P,
where Fc1 and Fc2 represent the respective forces in the outer

2995

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Fig. 7. Two-span continuous deep beam under two-point loading.

Fig. 8. Truss model for two-span CDB under two-point loading.

and interior concrete strut, T1 and T2 represent the respective


tension forces in the top and bottom longitudinal reinforcement,
and P represents the point load acting on the beam (Fig. 8).
Again, Mohrs linear interaction failure criterion is adopted at
the interface between the strut and tensioncompression nodal
zones:
f2
f1
+ 0 = 1.
ft
fc

(25)

Three tensioncompression nodal zones are identified in the


STM for CDBs as shown in Fig. 7. Similar to the modelling
of SSDBs, Morhs failure criterion is applied to the three nodal
zones, as follows:

The principal tensile stress f 1A across the diagonal strut and


the principal compressive stress f 2A in the diagonal strut can be
obtained in a similar manner as the SSDB:

f 2A

P
4T2 sin s
= 4C sin2 s
=
Ac /sin s
Ac
Fc1 T2 cos s
P
= (A C cos s )
=
Astr1
Astr1

(26)
(27)

where s is the inclined angle of diagonal strut with respect to


the horizon, Astr1 is the cross-sectional area at the bottom of the
outer concrete strut (Fig. 9).
From (25), (26) and (27), the following expression can be
derived for the ultimate force PnA :
PnA =

1
4C sin2 s
f tA Ac

f tA =

(AC cos s )
f c0 Astr1

(28)

4As2 f y sin s
Asw f yw sin(s + w )
+
+ f ct .
Ac /sin s
Ac /sin s

6.2. Nodal zone B


The principal tensile stress f 1B across the diagonal strut at
nodal zone B consists of two components: the contributions
from the top and bottom reinforcement. As the k factor is 2
(compression) for top steel and 4 (tension) for bottom steel, the
combined f 1B can be expressed as below
f 1B =

6.1. Nodal zone A

f 1A

where f tA is the maximum tensile capacity of nodal zone A in


the f 1 direction and can be similarly expressed by:

4T2 sin s 2T1 sin s


P
= (4C 2B) sin2 s
. (29)
Ac /sin s
Ac

The principal compressive stress at nodal zone B is obtained


similarly as nodal zone A
f 2B =

Fc2 T2 cos s
P
= (D C cos s )
Astr2
Astr2

(30)

where Astr2 is the cross-sectional area at the bottom of the


interior concrete strut (Fig. 9).
From Eqs. (25), (29) and (30), the following expression can
be derived for the ultimate load PnB :
PnB =

1
(4C2B) sin2 s
f tB Ac

(DC cos s )
f c0 Astr2

(31)

It is noteworthy that the maximum tensile capacity of nodal


zone B f tB is expressed by
f tB =

4T2 max 2T1a


Ac /sin s
2

Asw f yw sin(s + w )
+ f ct
Ac /sin s

2996

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Fig. 9. Details of nodal zones in CDB.

where T2 max is the yield strength of bottom steel and T1a is the
corresponding tension force in the top steel at the yielding of
bottom steel (= CB T2 max ). The term T1a should not exceed the
yield strength of top steel.
6.3. Nodal zone C
P
4T1 sin s 2T2 sin s
= (4B 2C) sin2 s
Ac /sin s
Ac
P
Fc2 T1 cos s
=
= (D B cos s )
.
Astr3
Astr3

f 1C =

(32)

f 2C

(33)

From Eqs. (25), (32) and (33), the following expression can be
derived for the ultimate load PnC :
PnC =

1
(4B2C) sin2 s
f tC Ac

(DB cos s )
f c0 Astr3

(34)

The term f tC in Eq. (34) is the maximum tensile capacity of


nodal zone C:
Asw f yw sin(s + w )
4T1 max 2T2a
f tC =
+
+ f ct .
2
Ac /sin s
Ac /sin s
Similarly, T1 max is the yield strength of top steel and T2a
is the corresponding tension force in the bottom steel at the
yielding of top steel (= CB T1 max ). T2a should not exceed the
yield strength of bottom steel.
Thus the predicted ultimate load P will be the minimum
among Eqs. (28), (31) and (34), denoted as Pn .
Pn = Min(PnA , PnB , PnC ).

(35)

7. Reliability of STM for CDBs


The computing of STM for CDBs can be easily implemented
by hand calculations or a spreadsheet. An example of hand
calculations is shown in the appendix for illustrative purposes.

Fig. 10. Ultimate strength predictions for 54 continuous deep beams.

Fifty-four concrete continuous deep beams reported by other


researchers have been evaluated by the authors model.
The details of the specimens and the predicted-versus-actual
ultimate strength ratios are summarized in Table 2. The beams
had an overall depth h ranging from 400 to 1000 mm, and an
le /d ratio from 0.95 to 4.49. The top-and-bottom longitudinal
main reinforcement ratios ranged from 0.07% to 1.88% and
0.32% to 1.88%, respectively. The vertical and horizontal web
reinforcement ratios ranged from zero to 0.90% and zero to
1.71%, respectively. The STM for CDBs generally performs
well in predicting the ultimate strengths with an overall average
prediction to test result of 0.95 and a C.O.V of 0.130. The
ultimate strength Pn versus the experimental strength Pexp are
plotted in Fig. 10. It shows the comparison of model predictions
with 54 test results. Generally speaking, the proposed model is
on the safe side and gives consistent predictions. For simplicity,
the model assumes linear material properties for steel and
concrete when solving the statically-indeterminate truss. It

2997

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Table 2
Predictions of ultimate load for CDBs
No.

Reference

f c0 (MPa)

h (mm)

bw (mm)

le (mm)

f y As1 / f y As2
(kN/kN)

f yv Av / f yh Ah
(kN/kN)

Pexp (kN)

Pn (kN) STM

Pn /Pexp

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54

CD-0.5/V
CD-0.5/H
CD-0.5/I
CD-1.0/H
CD-1.0/I
CD-0.0
MC2-200
MC2-150
MC2-100
MC2-75
MC2-50
MC2-75-2
CDB1
CDB2
CDB3
CDB4
CDB5
CDB6
CDB7
CDB8
BM 7/1.0(T1)
BM 7/1.0(T2)
BM 6/1.0(T1)
BM 6/1.0(T2)
BM 5/1.0(T1)
BM 5/1.0(T2)
BM 4/1.0(T1)
BM 4/1.0(T2)
BM 3/1.0(T1)
BM 3/1.0(T2)
HW
VW
N2-1.0-WO
N2-1.0-WV
N2-1.0-WVH
N2-1.5-WO
N2-1.5-WV
N2-1.5-WVH
2CB4
2CB3
1CB2
1CB1
BM 1.0/1/1(r)
BM 1.0/1/2
BM 1.0/1/3
BM 1.0/2/1
BM 1.0/2/2
BM 1.0/2/3
BM 1.5/1/1
BM 1.5/1/2
BM 1.5/1/3
BM 1.5/2/1
BM 1.5/2/2
BM 1.5/2/3

40.4
40.6
45.1
42.2
41.0
50.4
50.9
48.4
51.7
53.1
56.0
48.8
30.0
33.1
22.0
28.0
28.7
22.5
26.7
23.6
34.5
34.5
35.8
35.8
36.9
36.9
28.5
28.5
28.9
28.9
51.6
49.5
36.9
37.2
37.3
33.9
35.1
35.4
38.0
38.0
48.0
48.0
29.6
27.5
25.8
25.2
28.2
30.4
29.6
28.9
25.9
27.8
26.0
28.5

960
960
960
960
960
960
1000
1000
1000
1000
1000
1000
625
625
625
625
625
425
425
425
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
700
700
600
600
600
600
600
600
600
600
400
400
1000
1000
1000
1000
1000
1000
600
600
600
600
600
600

47
47
47
47
47
47
40
40
40
40
40
40
120
120
120
120
120
120
120
120
200
200
200
200
200
200
200
200
200
200
100
100
150
150
150
150
150
150
75
75
50
50
150
150
150
150
150
150
150
150
150
150
150
150

860
860
860
860
860
860
860
860
860
860
860
860
1340
1340
1340
1340
1340
1340
1340
1340
2100
2100
2100
2100
2100
2100
2100
2100
2100
2100
1400
1400
1200
1200
1200
1800
1800
1800
1680
1680
1000
500
2300
2300
2300
2300
2300
2300
2300
2300
2300
2300
2300
2300

163/163
163/163
163/163
163/163
163/163
163/163
15/314
15/314
15/314
15/314
15/314
29/314
305/226
305/226
305/226
305/226
113/113
192/192
192/192
113/113
484/363
484/363
484/363
484/363
484/363
484/363
456/342
456/342
456/342
456/342
338/225
338/225
735/735
735/735
735/735
735/735
735/735
735/735
116/323
323/323
99/99
99/99
356/265
356/265
356/265
265/356
265/356
265/356
445/356
445/356
445/356
356/445
356/445
356/445

43/0
0/93
(93)
0/185
(162)
0/0
23/42
23/53
23/87
23/118
23/182
23/237
262/149
135 /74
0/74
135/0
135/74
139/39
72/20
72/20
0/0
0/0
0/194
0/194
518/0
518/0
0/65
0/65
130/0
130/0
0/247
330/0
0/0
162/0
162/216
0/0
206/0
206/165
182/114
182/114
48/29
29/29
354/0
257/0
161/0
354/0
257/0
161/0
354/0
257/0
161/0
354/0
257/0
161/0

400
425
500
425
510
400
598
590
599
596
590
570
550
475
285
443
410
248
223
193
705
1085
1095
1059
1280
1147
1083
988
1084
1163
646
580
650
750
923
500
725
650
420
425
180
330
800
754
595
914
743
686
595
519
398
580
535
384

427
454
485
385
485
466
610
593
598
594
624
652
492
457
314
394
329
249
239
186
840
840
995
995
1188
1188
798
798
856
856
522
563
663
678
694
597
667
673
361
333
175
259
671
621
573
728
722
697
514
479
425
550
500
480

1.07
1.07
0.97
0.91
0.95
1.16
1.02
1.00
1.00
1.00
1.06
1.14
0.90
0.96
1.10
0.89
0.80
1.00
1.08
0.97
1.19
0.77
0.91
0.94
0.93
1.04
0.74
0.81
0.79
0.74
0.81
0.97
1.02
0.90
0.75
1.19
0.92
1.04
0.86
0.78
0.97
0.79
0.84
0.82
0.96
0.80
0.97
1.02
0.86
0.92
1.07
0.95
0.93
1.25

Mean
S.D.
C.O.V.

0.95
0.123
0.130

Note: SourceSpecimens no. 16, Chemrouk [33]; 712, Maimba [34]; 1320, Ashour [35]; 2130, Rogowsky et al. [36]; 3132, Choo and Lim [37]; 3338,
Poh [38]; 3942, Subedi [39]; 4354, Asin [40].

2998

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Table 3
Comparison bewteen calculated and measured interior support reaction
Reference

Predicted interior
support reaction
X n (kN)

Tested interior
support reaction
X exp (kN)

X n / X exp

CDB1
CDB2
CDB3
CDB4
CDB5
CDB6
CDB7
CDB8
BM 7/1.0(T1)
BM 7/1.0(T2)
BM 6/1.0(T1)
BM 6/1.0(T2)
BM 5/1.0(T1)
BM 5/1.0(T2)
BM 4/1.0(T1)
BM 4/1.0(T2)
BM 3/1.0(T1)
BM 3/1.0(T2)
HW
VW
N2-1.0-WO
N2-1.0-WV
N2-1.0-WVH
N2-1.5-WO
N2-1.5-WV
N2-1.5-WVH
BM 1.0/1/1(r)
BM 1.0/1/2
BM 1.0/1/3
BM 1.0/2/1
BM 1.0/2/2
BM 1.0/2/3
BM 1.5/1/1
BM 1.5/1/2
BM 1.5/1/3
BM 1.5/2/1
BM 1.5/2/2
BM 1.5/2/3

672.2
624.2
427.1
537.4
433.0
325.6
313.7
244.7
1145.0
1145.0
1357.4
1357.4
1621.5
1621.5
1085.1
1085.1
1164.3
1164.3
719.6
775.4
1026.2
1081.3
1133.5
738.6
826.6
858.0
935.3
862.0
794.1
913.4
905.8
875.4
699.9
652.3
578.0
699.9
635.7
610.5

702.0
612.0
360.4
567.8
516.0
312.2
281.0
247.4
842.0
1376.0
1281.0
1186.0
1741.0
1491.0
1330.0
1228.0
1374.0
1540.0
828.0
726.0
844.0
936.0
1156.0
548.0
887.0
856.0
1053.0
989.0
776.0
1171.0
937.0
845.0
798.0
694.0
691.0
745.0
675.0
499.0

0.96
1.02
1.19
0.95
0.84
1.04
1.12
0.99
1.36
0.83
1.06
1.14
0.93
1.09
0.82
0.88
0.85
0.76
0.87
1.07
1.22
1.16
0.98
1.35
0.93
1.00
0.89
0.87
1.02
0.78
0.97
1.04
0.88
0.94
0.84
0.94
0.94
1.22

Mean
S.D.
C.O.V.

0.99
0.147
0.149

would be interesting to see how the model-predicted support


reactions compared with the test results. Table 3 shows the
comparison between model predictions and test results of
reaction forces at the interior support. Clearly, it shows that
the STM for continuous deep beams can predict the reaction
forces reasonably well. The fifty-four Pn /Pexp ratios obtained
from the proposed model are plotted versus le /d ratio, concrete
strength f c0 , and total reinforcement ratio a in Fig. 11. It is
shown that the scatter is very low and uniform for the entire set
of three variables. It can be concluded that the application of
the modified STM to CDBs is feasible and successful. It also
shows the validity and versatility of the proposed STM from
simply-supported to continuous deep beams.
It is also noteworthy that the overall average prediction-totest result ratios for the 233 SSDBs and 54 CDBs are very close:
0.91 and 0.95, respectively, and the C.O.V. of both studies are

Fig. 11. Effect of: (a) concrete strength ( f c0 ); (b) effective span to overall depth
ratio (le /d); and (c) total reinforcement ratio on ultimate strength predictions
(a ).

same at 0.130. This shows that the predictions are consistent


and accurate for both SSDBs and CDBs with different
geometrical properties, concrete strengths and reinforcement
configurations.

2999

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

8. Conclusions
In this paper, the authors proposed a modified strut-andtie model (STM) based on previous work of Tan et al. [12,
13]. The modified STM for SSDBs is evaluated using 233 test
results collected from the literature. The modified model yields
better agreement than the original STM [12,13] in terms of
accuracy and consistency. The influence of stress distribution
factor k on model prediction is studied numerically. It is shown
that a k value of 4 derived from a linear stress distribution
assumption is sufficiently accurate for predicting the deep beam
shear strength. The modified model also implies a softening
effect in concrete compressive strength due to transverse tensile
strain, by adopting an interactive failure criterion, i.e. Mohrs
failure criterion. The softening effect given by the model is
shown comparable to that obtained from MCFT and Belarbi
and Hsus equations.
The modified STM is further extended to predict the ultimate
strengths of concrete continuous deep beams (CDBs) for the
first time. The predictions of the STM for CDBs including beam
strengths and support reactions are in good agreement with a
total of 54 test results of concrete CDBs.
Acknowledgements
The authors would like to express their gratitude to
the Nanyang Technological University for research account
RG9/99. The first author also acknowledges a scholarship from
the Nanyang Technological University, Singapore.
Appendix A. Derivation of internal forces of strut-and-tie
model

strut; E c and E s are the elastic modulus of concrete under


compression and steel bar under tension, respectively. The
terms Astr1 , Astr2 and Astr3 are the cross-sectional areas at the
ends of the tapered concrete struts (Fig. 9), while Astr4 and
Astr5 represent the average cross-sectional areas of the outer
and inner tapered concrete struts, respectively (Fig. 8). They
are expressed as follows:
Astr1 = bw (lc cos s + lb sin s )
Astr2 = bw (lc cos s + lf sin s )
Astr3 = bw (ld cos s + la sin s )
Astr1 + Astr3
Astr4 =
2
Astr2 + Astr3
Astr5 =
2
2(h c1 c2 )
s = arctan
le

(A2)
(A3)
(A4)
(A5)
(A6)
(A7)

where
la , lb and lf are the widths of the support and load bearing
plates (Fig. 7);
lc and ld are the respective effective depths of the bottom and
top nodal zones, lc = 2c2 , ld = 2c1 (Fig. 9);
c1 and c2 are the distances from the centroid of the top and
bottom longitudinal steel bars to the beam top and beam
soffit respectively (Fig. 7).
From the Theorem of Minimum Complementary Potential
(TMCP), at interior support B:
Uc
= 0.
(A8)
X
Thus, from Eqs. (A1) and (A8), an expression for the reaction
force X can be obtained as follows:

B =
The stressstrain relationship of concrete under compression
is taken as linear:
c = E c c
where E c is the modulus of elasticity of concrete in
compression
and can be taken from the ACI code (E c =
p
4730 f c0 , MPa) for normal strength concrete.
The stressstrain relationship of steel in tension is given by:

The truss members in Fig. 8 are internally determinate but


externally indeterminate to 1 degree. Releasing the interior
reaction force X and applying the CrottiEngesser theorem, the
total complementary energy in the truss is:

X Z Z
Uc =
d dV
V

2(m + 2n cos3 s + 2 p cos3 s )


P
1 + m + 4n cos3 s + 2 p cos3 s

()

P X2

2

le

2 sin2 s cos s

E c Astr4

X 2 le
8 sin2 s cos s

E c Astr5

(X P)2 le
2 tan2 s

E s As1

P X2

2

le

tan2 s

E s As2
(A1)

where le is the effective span measured between centre-tocentre of supports; s is the inclined angle of the diagonal

(A9)

where m, n, p are the ratios of axial stiffness of truss members:


Astr5
;
Astr4
E c Astr5
n=
;
E s As1
E c Astr5
.
p=
E s As2
m=

s = E s s .

X=

(A10)
(A11)
(A12)

After solving the reaction force X , the internal forces of truss


members (Fig. 8) can be obtained as follows:
1 + 2n cos3 s
P
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
= A P (compression)

Fc1 =

cos s (m + 2 p cos3 s 1)
P
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
= B P (tension)

(A13)

T1 =

(A14)

3000

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

cos s (1 + 2n cos3 s )
P
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
= C P (tension)
m + 2n cos3 s + 2 p cos3 s
P
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
= D P (compression)
.

(A15)

Fc2 =

(A16)

Appendix B. Worked example on continuous deep beam


The notation of the beam is BM 5/1.0 (Rogowsky et al. [36]),
details of specimens are as follows:
Geometry: le = 2100 mm, c1 = 50 mm, c2 = 25 mm,
lc = 50 mm, ld = 100 mm, h = 1000 mm, bw = 200 mm,
2
lb = 200 mm, la = 300 mm, lf = 400 mm, Ac = 185 000
p mm ;
2
0
Concrete strength: f c 0 = 36.9 N/mm , E c = 3694 f c =
p
22 439 MPa, f sp = 0.461 f c0 = 2.80 MPa;
Longitudinal reinforcement: As1 = 1195 mm2 , As2 =
896 mm2 , f y As1 = 484 kN, f y As2 = 363 kN;
Vertical web reinforcement: Asv f yv = 518 kN, w = 90 .
Step 1: Determine s from Eq. (A7);
s = arctan

cos s (1 + 2n cos3 s )
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
0.750 (1 + 2 5.401 0.422)
=
0.661 (1 + 1.298 + 4 5.401 0.422 + 2 7.204 0.422)
= 0.360

C =

T2 =

2 (1000 50 25)
2 (h c1 c2 )
= arctan
le
2100

= 41.4 .

m + 2n cos3 s + 2 p cos3 s
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
1.298 + 2 5.401 0.422 + 2 7.204 0.422
=
0.661 (1 + 1.298 + 4 5.401 0.422 + 2 7.204 0.422)
= 1.032.

D =

Step 4: Determine PnA , PnB , and PnC from Eqs. (28), (31) and
(34), respectively;
p

f ct = 0.31


T1 max


f c0

cr
1

0.4
= 0.31

36.9
0.4

0.00008
0.002 + 0.004/ tan2 41.38
= f y As1 = 484 kN

= 0.31 MPa

T2 max = f y As2 = 363 kN




B
T1a = min f y As1 , T2 max = 416 kN
C


C
T2a = min f y As2 , T1 max = 363 kN.
B

Step 2: Determine m, n, p, from Eqs. (A10) to (A12);


Astr1 = bw (lc cos s + lb sin s ) = 200 (50 0.750
+ 200 0.661) = 33 945 mm2
Astr2 = bw (lc cos s + lf sin s ) = 200 (50 0.750
+ 400 0.661) = 60 386 mm2
Astr3 = bw (ld cos s + la sin s ) = 200 (100 0.750
+ 300 0.661) = 54 669 mm2
33 945 + 54 669
Astr1 + Astr3
=
= 44 307 mm2
Astr4 =
2
2
Astr2 + Astr3
60 386 + 54 669
Astr5 =
=
= 57 528 mm2
2
2
57 528
Astr5
m=
=
= 1.298
Astr4
44 307
22 439 57 528
E c Astr5
=
= 5.401
n=
E s As1
200 000 1195
E c Astr5
22 439 57 528
p=
=
= 7.204.
E s As2
200 000 896
Step 3: Determine A, B, C, D from Eqs. (A13) to (A16);
1 + 2n cos3 s
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
1 + 2 5.401 0.422
=
0.661 (1 + 1.298 + 4 5.401 0.422 + 2 7.204 0.422)
= 0.481

A =

cos s (m + 2 p cos3 s 1)
sin s (1 + m + 4n cos3 s + 2 p cos3 s )
0.750 (1.298 + 2 7.204 0.422 1)
=
0.661 (1 + 1.298 + 4 5.401 0.422 + 2 7.204 0.422)
= 0.414

B =

Nodal zone A:
Asv f yv sin 2s
+ f ct
2Ac
Ac /sin s
4 363 0.437 103
518 0.992 103
=
+
+ 0.31
185 000
2 185 000
= 3.43 + 1.39 + 0.31 = 5.13 MPa
1
PnA =
(AC cos s )
4C sin2 s
f tA Ac +
f c0 Astr1
1/1000
= 40.3600.437 0.4810.3600.750 = 1202 kN.
5.13185 000 +
36.933 945
f tA =

4T2 max

Nodal zone B:
Asv f yv sin 2s
+ f ct
2Ac
Ac /sin s
(4 363 2 416) 103
=
185 000/0.437
518 0.992 103
+
+ 0.31
2 185 000
= 1.46 + 1.39 + 0.31 = 3.16 MPa
1
PnB =
cos s )
4C2B
+ (DC
f c0 Astr2
f tB Ac /sin2 s
1/1000
= 40.36020.414
= 1251 kN.
1.0320.3600.750
3.16185 000/0.437 +
36.960 386
f tB =

4T2 max 2T1a

Nodal zone C:

N. Zhang, K.-H. Tan / Engineering Structures 29 (2007) 29873001

Asv f yv sin 2s
+
+ f ct
2Ac
Ac /sin2 s
(4 484 2 363) 103
=
185 000/0.437
518 0.992 103
+
+ 0.31
2 185 000
= 2.86 + 1.39 + 0.31 = 4.56 MPa
1
PnC =
cos s
4B2C
+ DB
f c0 Astr3
f tC Ac /sin2 s
1/1000
= 1187 kN.
= 40.41420.360
1.0320.4140.750
4.56185 000/0.437 +
36.954 669
f tC =

4T1 max 2T2a

Step 5: Determine the ultimate load Pn from Eq. (35).


Pn = Min(PnA , PnB , PnC ) = 1187 kN.
Thus the predicted ultimate load of Pn is 1187 kN. Total
ultimate load is 2Pn = 2374 kN. In this example, the ultimate
load 2Pexp from experiment was 2559 kN (T1). Thus, the ratio
of Pn /Pexp is 0.93.
References
[1] Marti P. Basic tools of reinforced concrete beam design. ACI J 1985;
82(1):4656.
[2] Schlaich J, Schaefer K, Jennewein M. Toward a consistent design of
structural concrete. PCI J 1987;32(3):74150.
[3] MacGregor JG. Reinforced concrete: Mechanics and design. Prentice
Hall; 1988. p. 848.
[4] Bergmeister K, Breen JE, Jirsa JO. Dimensioning of the nodes and
development of reinforcement. In: Rep IABSE colloquium on structural
concrete. p. 5516.
[5] Alshegeir A. Analysis and design of disturbed regions with strut-tie
models. Ph.D. thesis. West Lafayette (IN): Purdue University; 1992.
[6] Foster SJ, Malik AR. Evaluation of efficiency factor models used in
strut-and-tie modeling of nonflexural members. J Struct Eng 2002;128(5):
56977.
[7] ACI Committee 318. Building code requirements for structural concrete
(ACI 318-02) and commentary (ACI 318R-02). Farmington Hills (MI):
American Concrete Institute; 2002.
[8] Canadian Standards Association. Design of concrete structures: Structures
design. Rexdale (ON): Canadian Standards Association; 1994.
[9] CEB-FIP model code 1990. Design code. London: Thomas Telford Ltd;
1993.
[10] Mau ST, Hsu TTC. Formula for the shear strength of deep beams. ACI
Struct J 1989;86(5):51623.
[11] Hwang S-J, Lu W-Y, Lee H-J. Shear strength prediction for deep beams.
ACI Struct J 2000;97(3):36776.
[12] Tan KH, Tong K, Tang CY. Direct strut-and-tie model for prestressed deep
beams. J Struct Eng 2001;127(9):107684.
[13] Tan KH, Tang CY, Tong K. A direct method for deep beams with web
reinforcement. Mag Concr Res 2003;55(1):5363.
[14] Vecchio FJ, Collins MP. The modified compression-field theory for
reinforced concrete elements subjected to shear. ACI J 1986;83(2):
21931.

3001

[15] Belarbi A, Hsu TTC. Constitutive laws of softened concrete in biaxial


tensioncompression. ACI Struct J 1995;92(5):56273.
[16] Belarbi A, Hsu TTC. Constitutive laws of concrete in tension and
reinforcing bars stiffened by concrete. ACI Struct J 1994;91(4):46574.
[17] Tang CY, Tan KH. Interactive mechanical model for shear strength of
deep beams. J Struct Eng 2004;130(10):153444.
[18] Kong FK, Robins PJ, Kirby DP, Short DR. Deep beams with inclined web
reinforcement. ACI J 1972;69(3):1726.
[19] ACI Committee 318. Building code requirements for structural concrete
(ACI 318-99) and commentary (ACI 318R-99). Farmington Hills (MI):
American Concrete Institute; 1999.
[20] Tan KH, Kong FK, Teng S, Guan L. High-strength concrete deep beams
with effective span and shear span variations. ACI Struct J 1995;92(4):
395405.
[21] Smith KN, Vantsiotis AS. Shear strength of deep beams. ACI J 1982;
79(3):20113.
[22] Vecchio FJ, Collins MP. Compression response of cracked reinforced
concrete. J Struct Eng 1993;119(12):3590610.
[23] Tan KH, Teng S, Kong FK, Lu HY. Main tension steel in high strength
concrete deep and short beams. ACI Struct J 1997;94(6):75268.
[24] Subedi NK, Vardy AE, Kubota N. Reinforced concrete deep beams
some test results. Mag Concr Res 1986;38(137):20619.
[25] Subedi NK. Reinforced concrete deep beams: A method of analysis. Proc
ICE J 1988;85:130. [Part 2].
[26] Walraven J, Lehwalter N. Size effects in short beams loaded in shear. ACI
Struct J 1994;91(5):58593.
[27] Tan KH, Kong FK, Teng S, Weng LW. Effect of web reinforcement on
high-strength concrete deep beams. ACI Struct J 1997;94(5):57282.
[28] Foster SJ, Gilbert RI. Experimental studies on high-strength concrete deep
beams. ACI Struct J 1998;95(4):38290.
[29] Tan KH, Lu HY. Shear behavior of large reinforced concrete deep beams
and code comparisons. ACI Struct J 1999;96(5):83645.
[30] Shin S-W, Lee K-S, Moon J-I, Ghosh SK. Shear strength of reinforced
high-strength concrete beams with shear span-to-depth ratios between 1.5
and 2.5. ACI Struct J 1999;96(4):54956.
[31] Oh J-K, Shin S-W. Shear strength of reinforced high-strength concrete
deep beams. ACI Struct J 2001;98(2):16473.
[32] Yang K-H, Chung H-S, Lee E-T, Eun H-C. Shear characteristics of highstrength concrete deep beams without shear reinforcements. Eng Struct
2003;25(10):134352.
[33] Chemrouk M. Slender concrete deep beams: Behaviour, serviceability
and strength. Ph.D. thesis. Newcastle: The University of Newcastle upon
Tyne; 1988.
[34] Maimba PP. Effects of support continuity on the behaviour of
slender reinforced concrete deep beams. M.Eng. thesis. Newcastle: The
University of Newcastle upon Tyne; 1989.
[35] Ashour AF. Tests of reinforced concrete continuous deep beams. ACI
Struct J 1997;94(1):312.
[36] Rogowsky DM, MacGregor JG, Ong SY. Tests of reinforced concrete
deep beams. ACI J 1986;83(4):61423.
[37] Choo QL, Lim HL. Design and behaviour of continuous concrete
deep beams. BEng. Final year project report. Singapore: Nanyang
Technological University; 1993.
[38] Poh SP. Strength and serviceability of prestressed concrete deep beams.
M.Eng. thesis. Singapore: Nanyang Technological University; 1995.
[39] Subedi NK. Reinforced concrete two-span continuous deep beams. Proc
Instn Civ Engrs Struct & Bldgs 1998;128(1):1225.
[40] Asin M. The behaviour of reinforced concrete continuous deep beams.
Delft University Press; 1999. p. 168.

You might also like