You are on page 1of 11

ACI STRUCTURAL JOURNAL TECHNICAL PAPER

Title no. 103-S34

Compatibility Torsion in Spandrel Beams Using Modified


Compression Field Theory
by Khaldoun N. Rahal and Michael P. Collins

This paper describes a procedure that can be used to calculate the


compatibility torsion in spandrel beams and other indeterminate
structures. The procedure relies on the ability of the Modified
Compression Field Theory (MCFT) to calculate cracked torsional
and flexural stiffness for sections subjected to various combinations of
stress resultants. The calculated ultimate loads are compared with the
experimentally observed ultimate loads of 27 statically indeterminate
floor beam-spandrel beam frames and very good agreement is
observed. The comparison with the detailed experimental results
shows that the MCFT is capable of accurately calculating the
response of reinforced concrete sections subjected to combined
shear, torsion, bending moment, and axial compression. The
calculations of the proposed procedure also show that the strength
of the frames is significantly larger than that calculated using the
equations of the ACI and CSA codes due to the longitudinal
restraint of the beams.

Keywords: beams; cracking; reinforced concrete; shear; spandrel beams;


torsion.

INTRODUCTION
Many structural elements such as beams curved in plan,
eccentrically loaded bridge girders, and spandrel beams in
buildings are subjected to significant torsional moments. To
design such members, it is essential to recognize whether the
torsional moments are required to maintain equilibrium or
compatibility. These two cases are generally referred to1-4 as
equilibrium torsion and compatibility torsion, respectively.
Torsion in determinate structures is always equilibrium
torsion, while that in indeterminate structures can be of
either type.1
Figure 1 shows examples of the two cases of torsion. The
traffic load on the statically determinate multi-cell bridge in
Fig. 1(a) is eccentrically applied, and torsion on the transverse
cross-section is required to maintain equilibrium. Redistribution
of this torsional moment is not possible, and the total Fig. 1—Examples of torsion in reinforced concrete: (a) equi-
moment calculated using statics must be designed for. A librium torsion; and (b) and (c) compatibility torsion.
similar situation arises if the bridge is curved in plan.
Figure 1(b) shows torsion in a spandrel beam of a building
uncracked stiffness in the structural analysis greatly over-
caused by loading from a cast-in-place slab. This torsional
estimates the torsion for the specific level of twist.1,2,5-8 The
moment is due to imposed rotations from the slab, and is
redistribution of the torsional moment to positive flexural
equal to the negative bending moment in the slab. The
compatibility torsion due to the required twist in this statically moment near midspan of the floor beam leads to a more
indeterminate structure depends on the torsional stiffness of economic design because the flexural capacity of a typically
the spandrel beam relative to the flexural stiffness of the slab. reinforced concrete section is significantly larger than its
A similar phenomenon takes place if a floor beam frames torsional capacity. In addition, the cost of manufacturing and
into the spandrel beam, as shown in Fig. 1(c). placing the longitudinal steel in the floor beam or slab is less
In most of the cases, torsional moments encountered in than that for the transverse steel in the spandrel beam.
reinforced concrete buildings are of the compatibility type,
and their calculation poses a challenge. Once the spandrel ACI Structural Journal, V. 103, No. 3, May-June 2006.
MS No. 03-486 received February 22, 2005, and reviewed under Institute publication
beam cracks in torsion, its torsional stiffness reduces policies. Copyright © 2006, American Concrete Institute. All rights reserved, including
substantially. The reduction causes a significant redistribution the making of copies unless permission is obtained from the copyright proprietors. Pertinent
discussion including author’s closure, if any, will be published in the March-April
of the torque to the framing elements. Assuming a gross 2007 ACI Structural Journal if the discussion is received by November 1, 2006.

328 ACI Structural Journal/May-June 2006


ACI member Khaldoun N. Rahal is an associate professor in the Department of Civil
Engineering at Kuwait University, Kuwait. He is a member of the torsion subcommittee of
Joint ACI-ASCE Committee 445, Shear and Torsion, and is Past President of the ACI
Kuwait Chapter.

Michael P. Collins, FACI, is University Professor and Bahen-Tanenbaum Professor of


Civil Engineering at the University of Toronto, Toronto, Ontario, Canada. He is a member
of ACI Committees 318, Structural Concrete Building Code; 318-E, Shear and Torsion;
318-G, Precast and Prestressed Concrete; and Joint ACI-ASCE Committee 445, Shear and
Torsion.

Beams subjected to shear, bending, and/or torsion, experience


significant elongation after cracking. Spandrel and floor
beams in reinforced concrete buildings are subjected to this
complex combination of stress resultants. In addition, they
are restrained in the longitudinal direction. Consequently,
they are subjected to a significant axial compressive force,
which affects their strength and stiffness and makes the study
of the redistribution of torsional moment more challenging.
Fig. 2—Test setup by Ozerdinc et al.5
Design for compatibility torsion: background
Reinforced concrete buildings were designed by neglecting
the torsion that takes place in the spandrel beams. In 1963 for Design Method C. The floor beam load was distributed along
example, the ACI code9 contained only a few lines on the the length, and an additional load was applied at the joint as
detailing of the torsional reinforcement in spandrel beams. shown in Table 1. The joint load increased the shear force in
The use of computers in the 1960s made the calculation of the spandrel, representing the high torsion and shear and
the torsional moments in space frame structures practical. bending near the support of the spandrel. Hsu and Hwang
This left designers with a puzzling question: How to design concluded that both design methods were satisfactory, and
for torsion? Between 1963 and 1973, a significant amount of recommended an expression for the minimum amount of
research10 focused on the strength of individual members transverse reinforcement required in the spandrel beam.
subjected to torsion, and led to the introduction of approxi- In 1978, Abul Mansur and Rangan8 tested seven floor
mately 16 sections of torsion design provisions in the 1971 beam-spandrel beam assemblies. Three of the specimens had
ACI code.11 The code, however, did not include special spandrel beam load as shown in Table 1. They compared the
provisions for compatibility torsion, implying the use of behavior of frames designed assuming GK = 0 (Collins and
gross uncracked torsional stiffness of the spandrel beams. Lampert’s approach was Method B) and Hsu and Burton’s
In 1971, Collins and Lampert1 defined the two different design approach described above with frames designed
types of torsion (compatibility torsion and equilibrium using GKgross. They concluded that both design approaches
torsion). Ozerdinc et al.5 reported the test results of six floor are satisfactory, and that Hsu and Burton’s approach
beam-spandrel beam assemblies shown in Fig. 2. The study (Method C) was slightly more economical.
investigated the difference in behavior between frames In 1991, Mo and Hsu13 developed a theoretical model for
designed using uncracked gross stiffness (GKgross) (Design analysis of compatibility torsion based on the use of trilinear
Method A) and zero stiffness (GK = 0) (Design Method B). torque-twist and bending-curvature relationships. Their
See also Table 1 for brief details on these tests. The study model captured the effects of the two redistributions that
also investigated the effects of variable relative stiffness and take place after torsional cracking in the spandrel and
length of the spandrel and floor beams. The results proved
yielding of the longitudinal reinforcement in the floor beam,
the feasibility of designing the frame based on redistribution
and good correlation with the experimental results was
of the torsional moment if the spandrel beam is provided
observed. However, the model did not take into consideration
with a minimum amount of properly detailed reinforcement.
the effects of longitudinal restraint, the effect of shear and
In 1972, Onsongo2 tested three frames similar to those
bending on the torsional stiffness of the spandrel, and the
shown in Fig. 2, but the spandrel beams were completely
restrained in the longitudinal direction. He concluded that effect of shear on the flexural stiffness of the floor beams.
restrained beams designed assuming (GKgross) or assuming The model was applicable to 11 of the specimens discussed
zero torsional stiffness (GK = 0) performed in a satisfactory above, and its results were in good agreement with the
manner, and that the restraining force has secondary effects. experimental results.
In 1974, Hsu and Burton6 tested 10 floor beam-spandrel
beam assemblies. They investigated an alternative design Design for compatibility torsion in current
approach where the spandrel is designed for a torsional shear ACI and CSA codes
stress of 4√fc′ psi (0.33√fc′ MPa), while the remaining The current ACI3 and CSA4 codes allow partial redistribution
torsional moment is redistributed to the framing elements of the torsional moment in the spandrel beam to flexural
(Design Method C). This approach was found satisfactory, moment in the framing slab or beam. The ACI3 design
and was later adopted by the 1983 ACI12 code. torque in the spandrel beam is the cracking torque given by
In 1977, Hsu and Hwang7 tested four floor beam-spandrel Eq. (1) while the remaining moment is redistributed to the
beam assemblies and compared the behavior of frames framing elements. A similar procedure is allowed in the
designed assuming (GK = 0) (Collins and Lampert’s Canadian code4 with the design torque equal to 67% of the
approach or Design Method B) with frames designed using cracking torque. The CSA design torque is given in Eq. (2).

ACI Structural Journal/May-June 2006 329


Table 1—Comparison of observed and calculated ultimate loads
Specimen size* and loading ρv fy spandrel,
arrangement Design method† MPa Pf-des, kN Ps /Pf Pf-exp, kN Pf-calc, kN Pf-exp /Pf -calc
S1 A 1.490 164.6 0 254.0 240.2 1.057
S2 B 0.507 191.3 0 l242.4 226.8 1.069
S3 A 0.705 231.3 0 294.9 267.0 1.104
S4 B 0.507 191.3 0 234.6 204.6 1.147
Ozerdinc et al.5 S5 B 0.507 133.4 0 174.8 178.0 0.982

S6 B 0.507 209.1 0 304.7 258 1.181

A1 A 3.30 135.2 0 163.7 151.2 1.082


A2 C 1.22 127.7 0 126.8 142.3 0.891
A3 C 1.22 127.7 0 149.5 149.0 1.003
A4 C 0.00 124.5 0 101‡ — —‡
A5 C 0.00 124.5 0 86.7‡ — —‡
Hsu and Burton6 B1 A 3.28 113.9 0 116.1 111.2 1.044
B2 C 1.14 113.9 0 125.4 129.0 0.972
B3 C 0.00 113.9 0 77.4‡ — —‡
B4 B-C 0.41 113.9 0 117.9 135§ 0.88§
B5 B-C 0.41 113.9 0 109.0 122§ 0.89§
SA-1 A 1.71 120 0 157.1 136 1.155
SA-2 C 1.28 116 0 142.1 133 1.068
SA-3 C 1.01 116 0 138.1 133 1.038
Abul Mansur and SA-4 B 1.08 116 0 142.1 140 1.015
Rangan8 SB-1 C 1.83 116 0.5 158.0 141 1.120
SB-2 C 1.51 116 0.5 166.0 150 1.107
B 1.51 116 0.5 154.0 145 1.062
SB-3

MTV1 A 2.18 171.2 0 240.2 211 1.14

MTV2 B 0.74 160 0 231.7 218 1.062


Onsongo2

MTV3 B 0.74 171.2 0 250.4 231.3 1.083

C1 C 1.80 113.9 0 144.6 139 1.040


C2 C 1.80 113.9 0.55 142.3 135 1.054
Hsu and Hwang7
C3 B 1.86 113.9 1.50 111.2 108 1.030
C4 B 0.98 113.9 0.81 108.5 100 1.085
Average 1.05
Coefficient of variation, % 7.40
*
Beam lengths drawn to scale, cross section dimensions of spandrel and floor beam in mm.

Design Method A: gross stiffness; Design Method B: zero stiffness (Collins and Lampert)1; and Design Method C: Hsu and Burton.6
‡No torsional reinforcement in spandrel beam.
§Inadequate torsional reinforcement.13

2 of the transverse reinforcement ensures ductility of the spandrel


A
T ACI = φ0.33 f c′ -----c (1) and its ability to provide the twist required for redistribution.
pc
While the code approach is simple to use, it does not take
into consideration various significant factors such as the
A
2 amounts of transverse and longitudinal reinforcement in the
T CSA = 0.268φ c f c′ -----c (2) beams, the restraining effects in the longitudinal direction,
pc and the spans of the beams. Consequently, the code approach
does not take into account the actual torsional and flexural
If U.S. customary units are used, the factors 0.33 and 0.268 stiffness of the spandrel and the floor beams. The satisfactory
are replaced with 4 and 3.23, respectively. Proper detailing performance of spandrel beams designed according to the

330 ACI Structural Journal/May-June 2006


code provisions proves that these provisions are conservative. In
the case of analysis of existing structures, however, more
rational models are required to obtain accurate results.

Calculation of stiffness in reinforced concrete beams


Following the publication of the experimental results on
compatibility torsion in the 1970s,1,2,6-8 there has been a
significant advancement in the analysis methods of reinforced
concrete beams subjected to complex loading. One such
behavioral model, the Modified Compression Field Theory
(MCFT) is capable of calculating the full response of reinforced
and prestressed concrete beams subjected to shear, bending,
and axial load.14 A further development15,16 included the
effects of torsion and enabled the calculation of the response
of rectangular sections subjected to all six possible stress
resultants on a beam cross section. A computer program
named COMBINED was developed based on this model,
and was shown to give accurate results on steel and concrete
stresses and strains, and overall deformations such as twist,
elongation, and curvatures for combined shear and torsion15
and combined torsion and bending.16 Such a program can
hence be used to calculate the torsional and flexural stiffness
in rectangular beam cross sections subjected to complex loading
in general, and in spandrel beams in particular. It is to be noted
that the MCFT has not been thoroughly tested for the
combination of torsion, shear, bending, and axial compression.
This paper describes the use of the MCFT and a standard
linear-elastic structural analysis program to calculate the
compatibility torsion in spandrel beams and other indeterminate
structures. The comparison with the test results will also
enable checking the ability of the MCFT to calculate the
response of sections subjected to the complex combination
of shear, torsion, bending, and axial compression.

RESEARCH SIGNIFICANCE
The ACI and CSA provisions dealing with compatibility
torsion such as that encountered in spandrel beams are simple Fig. 3—Flow chart of proposed procedure.
and conservative for design. There is a lack of analytical
models, however, that can account for the influence of the
variable factors that affect the performance of spandrel in each of the elements obtained from the structural analysis
beams in buildings, and of members subjected to compat- output. These values, along with the geometrical and
ibility torsion in general. The MCFT is used to calculate reinforcement details of the element’s cross sections, are then
more realistic cracked torsional and flexural stiffness that used in COMBINED, the sectional analysis computer program,
can be used in the structural analysis procedure for more to calculate the state of strains and stresses in the materials and
accurate results. The procedure applies to the practical the overall sectional deformations (twist ψ and flexural
case where the beams contain properly detailed transverse and curvature ζ). Refined values of the assumed effective torsional
longitudinal reinforcement. and flexural stiffness (GK)eff and (EI)eff are calculated from the
twist and flexural curvature using the following equations
PROPOSED ANALYSIS PROCEDURE
Figure 3 shows a flow chart of the proposed procedure. A (GK)eff = T/ψ (3)
typical structural analysis of the indeterminate frame is
performed based on assumed flexural and torsional stiffness
of the structural members. These assumed values can be the (EI)eff = M/ζ (4)
uncracked gross values, or smaller values if cracking is
expected at the level of the load applied. Generally, the where E is the modulus of elasticity of concrete and G is the
stiffness varies considerably along the length of the structural shear modulus of concrete, which may be taken as 0.5E.
members such as beams. Hence, modeling a beam or a Refined values of effective K and I are then calculated for
column using only one element does not lead to an accurate each of the elements and used as input in the structural analysis.
representation of the variation in stiffness. Hence, similar to Shear deformations can also be included, but are neglected
nonprismatic sections, structural members need to be for the type of frames considered in this study.
modeled using numerous elements, with lengths recommended Members subjected to shear, bending, and torsion, acting
not to exceed twice the depth. individually or in combination, experience significant elongation
The average values of the shearing force V, axial load N, after cracking. In addition, members subjected to combined
and bending and torsional moments (M and T, respectively) shear and torsion experience lateral curvature.15,17 The floor

ACI Structural Journal/May-June 2006 331


Fig. 5—Variation in torsional stiffness of spandrel beams
and flexural stiffness of floor beams at P = 178 kN (40 kips).

Reference 1 and 5 for Specimens S1 and S2, and from


Reference 2 for Specimen MTV2.
The frames had 2895 mm (9 ft, 6 in.) long spandrel beams
and 4572 mm (15 ft) long floor beams. Specimen S1 was
designed for a floor beam load of 164.6 kN (37 kips) based
on the uncracked gross torsional stiffness GKgross of the
spandrel beam (Method A). Specimen S2 was designed for a
load of 191.3 kN (43 kips) based on a zero torsional stiffness
assumption (Method B), and hence allowing redistribution
of the compatibility torsion to flexure in the floor beam. It
contained minimal transverse reinforcement as shown in
Table 1. Specimen MTV2 was designed for 160 kN (36 kips)
floor beam load based on a zero torsional stiffness assumption.
Hence, it was similar to Specimen S2 (that is, GK = 0), except
that the spandrel beam was restrained in the longitudinal
direction to simulate the effects of the columns and the
spandrels in adjacent bays. The actual case in a spandrel
beam lies within these three extreme cases, but with additional
longitudinal restraint in the floor beam. Additional analyses
of Specimen MTV2 with longitudinal restraint in the floor
beam are described later.
Fig. 4—Cross section details of specimens and idealization The frames were modeled as a series of 20 elements, 10 in
of structure (Specimens S1, S2, and MTV2).2,5 each beam, as shown in Fig. 4. To maintain clarity in
presenting the results, the structural modeling did not take
advantage of the symmetry in the spandrel along the axis of the
beam in Fig. 2 elongates, while the spandrel beam experiences floor beam. The analysis was performed for different values of
both elongation and lateral curvature due to the combination the applied load, and the analytical results are checked against
of shear and torsion. The elongation and lateral curvature the experimental results in the following sections.
calculated by the computer program COMBINED can be The transverse curvature in the spandrel beam due to the
included in the structural analysis as initial deformations combination of shear and torsion calculated by COMBINED
in the elements. does not bear an effect on the results of the structural analysis
The structural analysis is repeated for the structure using and hence was not included. Moreover, a compressive force
the refined values of the stiffness, elongation, and curvature is applied to the spandrel to maintain a negligible average
until convergence is achieved. longitudinal strain in the 10 elements. For the simple frame,
this was more suitable than including the longitudinal strain
of each element as an initial strain in structural analysis
CORRELATION WITH EXPERIMENTAL RESULTS
input. In addition, the frame shown in Table 1 and Fig. 2 is
The experimental results from the tests listed in Table 1 indeterminate to the first degree, and hence it was sufficient
are used for the verification of both the proposed procedure to check for convergence in the value of the compatibility
and the MFCT in the case of complex loading. The ultimate torsion in the spandrel beams.
load capacities of the frames are compared with the calculated
capacities. In addition, three of the frames are selected for Variation in torsional and flexural stiffness
detailed comparison of the response. Figure 5 shows the calculated torsional stiffness in the
Indeterminate frames similar to that shown in Fig. 2 have spandrel beams and the flexural stiffness in the floor beams
been tested at the University of Toronto.1,2,5 Three of these at a load of 178 kN (40 kips). The calculated gross uncracked
frames, S1, S2, and MTV2, were selected for the verification. stiffness is also shown. The variation in the stiffness along
Figure 4 gives some of the details of the reinforcement in the the length of the beams is considerable and shows the significant
three frames. More detailed results can be obtained from effect of the variable bending moment on the value of GK in

332 ACI Structural Journal/May-June 2006


the spandrel. This shows the need of the numerous elements
to model the beams, especially in the spandrel beams after
torsional cracking. The same was found to be true for the
floor beams after yielding of the longitudinal steel. The
longitudinal restraint limited the spread of torsional
cracking, and increased the torsional stiffness in the
uncracked regions of the spandrel beam significantly beyond
the gross uncracked values.

Load-compatibility torque diagrams


Figure 6(a) shows the compatibility torsional moment in
the spandrel of Specimen S1 at different levels of the load P.
It also shows the ACI and the CSA design torsional moments
calculated using Eq. (1) and (2) based on unit resistance
factors and the response based on the gross uncracked
stiffness, and on the cracked stiffness proposed by Collins
and Lampert.1 Before torsional cracking, the behavior was
similar to what can be obtained from a linear elastic analysis
based on gross and uncracked sectional properties. The
proposed procedure accurately calculated this precracking
response, but slightly underestimated the cracking loads.
Two distinct changes affected the behavior of the specimen;
at torsional cracking of the concrete in the spandrel and at
flexural yielding of steel in the floor beam. Upon torsional
cracking, the torsional stiffness of the spandrel decreased
dramatically, allowing redistribution of the torque into the
floor beam. Additional loading caused a relatively smaller
increase in the torsional moment and more significant
increase in the flexural moment in the floor beam. This was
accurately captured by the proposed method. When the
longitudinal reinforcement in the floor beam yielded, the
demand on the rotation in the joint increased dramatically.
The spandrel was designed to resist larger torque, and hence
the torque increased at a larger rate. In other words, the direction
of the redistribution reversed. The figure shows that the
model accurately captured both changes in behavior,
including the compatibility torsion at design load, and the
yielding in the bottom steel in both beams and in the stirrups
of the spandrel. The calculated ultimate load was attained
when the section at and near midspan (under the load)
reached its flexural and shear capacity. The ultimate load
was underestimated by less than 6%.
Figure 6(b) shows a similar behavior in Specimen S2 until
the occurrence of yielding in the bottom longitudinal
reinforcement of the floor beam. The spandrel contained
relatively small amounts of transverse steel, and hence it was
capable of twisting considerably without an increase in the
torsional moment. The demand on the twist increased when
the bottom reinforcement of the floor beam yielded, but the
transverse steel of the spandrel yielded at the same load
level, which did not allow for reverse redistribution and Fig. 6—Torque-applied floor load diagrams for spandrel
increase in load. A good correlation is found in the comparison beams.
between the calculated and observed response. The model
captured the precracking response, the change in behavior as was shown in Fig. 5, the longitudinal restraint caused an
upon torsional cracking in the spandrel, and the occurrence increase in the torsional stiffness and strength of the spandrel.
of yielding in the longitudinal and transverse steel of the This is reflected on the relatively stiff post-cracking behavior
spandrel and the bottom longitudinal steel in the floor beam. of the specimen, and then on the increase in torsion after
The model accurately predicted the compatibility torsion yielding in the bottom steel of the floor beam. The ultimate load
at design load, and underestimated the ultimate load by was not significantly affected because it was dominated by
approximately 7%. the flexural strength of the floor beam.
Figure 6(c) shows a similar diagram for Specimen MTV2, Figure 6(c) shows that the model is capable of calculating
where the longitudinal spandrel was restrained in the the full response of the specimen, including the change in
longitudinal direction. The response was relatively softer behavior caused by the longitudinal restraint and yielding in
than expected before cracking. After cracking, however, and the bottom reinforcement of the floor beam. The model

ACI Structural Journal/May-June 2006 333


of the spandrel of Specimens S1 and S2. The twist in the
different elements was variable due to the presence of the
variable bending moment. To be consistent with the
experimental results used in Fig. 7, the calculated values
reported are the average of the twist from Elements 12 to 15
and the part of Element 11 that falls outside the joint. Figure 7(a)
shows a very good agreement between the observed and
calculated response, including the occurrence of torsional
cracking and the precracking and post-cracking response.
The values of the torsional stiffness at torques corresponding
to design and ultimate loads were 37 and 14%, respectively,
of the measured uncracked stiffness.
Figure 7(a) also shows the torque-twist response of the
various elements in the spandrel beams. The elements subjected
to higher bending moment near the joint showed softer response
and hence smaller torsional stiffness. At design and calculated
ultimate loads, respectively, the effective torsional stiffness
values in the element with highest bending moment (M/V =
0.45Lf in Element 11) were 50 and 84% of those in the elements
with lowest bending moment (M/V = 0.05Lf in Element 15).
This shows the significant effect of the bending moment on the
torsional stiffness, especially at design load.
The response of Specimen S2 shown in Fig. 7(b) was
slightly different because of the relatively low amount of
transverse reinforcement. The torque remained virtually
unchanged after cracking, allowing a relatively ductile
behavior and redistribution of the torque as the load increased.
The proposed model accurately captured the response of
Specimen S2, including the precracking stiffness, the cracking
torque, and the post-cracking stiffness. The values of the
torsional stiffness at torques corresponding to design and
ultimate loads were respectively 15 and 10% of the measured
uncracked stiffness. These values are significantly smaller
than Specimen S1 designed based on gross stiffness.
Figure 7(c) shows a similar diagram for Specimen MTV2,
except that the twist was measured over a 1.22 m (4 ft) length
of the specimen and the calculated values reported in the
figure corresponds to the twist averaged over this length.
Similar to the trend observed in the load-torque diagrams, the
longitudinal restraint caused a significant stiffening in the
response. The torsional stiffness at torques corresponding to
design and ultimate loads were 43 and 33% of the measured
uncracked stiffness, respectively, which are higher those
observed in Specimens S1 and S2. Figure 7(c) shows that the
proposed method accurately calculated the precracking and
the post-cracking stiffness, and the change in behavior due to
the longitudinal restraint. Hence, the MCFT is capable of
predicting the response of members subjected to combined
shear, torsion, bending moment, and axial compression.

Fig. 7—Average torque-twist diagrams for spandrel beams. Restraining axial force in spandrel beam
The spandrel beam in Specimen MTV2 was restrained from
underestimated the ultimate load by approximately 6%. expanding longitudinally using a 100 ton jack bearing on end
Comparing the results of the three specimens, it is found that plates and 1-1/4 in. Dywidag bar positioned in the center of a
the behavior and strength are significantly affected by the 100 mm (4 in.) hole running at the center of the beam. This bar
longitudinal restraint and the amount of reinforcement in the was prestressed at the different levels of the loading to ensure
spandrel beam. that the longitudinal strains in the spandrel beam were negli-
gible. The restraining force was measured during the test. As
Torque-twist diagrams observed from Fig. 6 and 7, the restraining force had a signif-
The twist along the length of the spandrel beams in icant effect on the load-torque and torque twist response.
Specimens S1 and S2 was measured using a Metrisite Figure 8 shows the calculated and the experimentally
extending from the center the joint to the clamping truss at measured restraint axial force in the spandrel of Specimen
the edge of the beams as shown in Fig. 2. Figure 7(a) and (b) MTV2 at different levels of the applied floor load P. The
compares the observed and calculated torque-twist response restraint force was negligible until the occurrence of torsional

334 ACI Structural Journal/May-June 2006


Fig. 8—Restraining axial compressive force in Specimen
MTV2.

cracking, and increased almost linearly after that due to the


elongation in the spandrel beam. A good correlation is observed
between the measured and the calculated response at different
levels of loading, including the design load.

Load-deflection diagrams
Figure 9 compares the observed and calculated load-
deflection curve in the floor beam for the three specimens.
The response showed considerable softening when the
longitudinal steel in the floor beam yielded. The longitudinal
restraint in the spandrel beam of Specimen MVT2 increased
the torque and hence reduced the flexure and consequently
the vertical deflection in the floor beam. The proposed
method accurately predicted the general trends in the
response, but overestimated the stiffness.

Crack width calculation


One of the major concerns with allowing redistribution
was the suspicion that using less transverse steel in the spandrel
beams would lead to less satisfactory crack width
behavior.1,2,5,6,8 Specimen S2 designed assuming GK = 0
contained approximately 1/3 the transverse steel provided in
Specimen S1 designed assuming GKgross. The observed
load-twist diagrams in Specimens S1 and S2 were, however,
very similar,1,5 and the recorded maximum crack widths
were similar up to design load. Estimates of the crack width
are necessary to satisfy the design requirements, though it
should be realized that crack width calculations do not Fig. 9—Applied floor load-midspan deflection in floor beam.
usually lead to accurate results.
Average and maximum diagonal crack widths were
recorded in the three specimens. Reference 2 reports that width in accordance with the CEB-FIP code.18 The ACI
these widths were measured near mid-height. The computer code does not set a limit on the width of diagonal cracks, but
program COMBINED calculates average crack widths at it bases the serviceability performance of flexural members
three different levels along the height of the cross section, on a 0.3 mm crack width limit.
and those widths vary because of the longitudinal strain Figure 10 shows the observed and calculated load versus
variation across the height16 from Element 11. The maximum crack width diagrams near midheight of the
maximum crack width occurs near the bottom of the section section for the three specimens and the 3 mm limit. The
where the effect of the bending is significant, but no reported calculated crack width was that from Element 11,
measurements of crack widths are reported at that level. The shown in Fig. 4. The proposed procedure seems to give
comparison is hence checked for the maximum width at adequate results, especially at design loads. The measured
midheight. Due to the wide variation in crack spacing, there crack widths are highest in members designed using Method B,
will be a similar variation in the crack width. The maximum but were smaller than 0.3 mm at service load (assumed to be
crack width is taken as 1.7 times the calculated average 70% of the design load). Longitudinal restraint and design

ACI Structural Journal/May-June 2006 335


Fig. 10—Maximum width of diagonal cracks near mid-
depth of spandrel beams. Fig. 11—Average strains in stirrups of spandrel beams.

using Method A (providing larger amounts of transverse reported are the average of the three values calculated along
reinforcement in the spandrel) significantly reduce the width the height of the section16 from Element 11. Figure 11(a) shows
of the diagonal cracks. that the stirrup strains in Specimen S1 were negligible before
torsional cracking, and increased upon cracking and reached
Strains in transverse reinforcement yielding levels. The proposed procedure accurately calculated
The strains in the stirrups on the inside face of the spandrel this behavior including the occurrence of yielding before the
beams were measured using 200 mm (8 in) targets as shown ultimate conditions were reached.
in Fig. 2. These strains are critical because the stirrups in the Figure 11(b) shows a flat post-cracking behavior in
spandrel are designed to resist the combined shearing Specimen S2, which twisted as the load was increased,
stresses from torsion and shear. Figure 11 compares the allowing redistribution of the torsional moment. The model
observed and the calculated strains. The theoretical values captured the precracking and the post-cracking behavior, and

336 ACI Structural Journal/May-June 2006


the occurrence of yielding in the stirrups. Figure 11(c) shows calculated ultimate load for FSR was 50% larger than that of
a similar correlation between the calculated and observed Specimen MTV2, and 106% larger than the design load.
torque-stirrups strain diagrams. The figure also shows that Figure 7(c) shows that releasing the force (NR analysis)
the longitudinal restraint reduced the transverse strains. softened the response and significantly reduced the torsional
ultimate capacity. Restraining the floor beam (FSR) allowed the
Correlation with other test results increase in the ultimate capacity of the frame and caused an
The proposed procedure was also used to calculate the increase in the post-cracking stiffness and torsional capacity.
ultimate floor loads in 27 other test specimens. Table 1 Figure 8 compares the axial compressive forces in the
reports some of the main characteristics of the test specimens restrained beams of analysis FSR with that from Specimen
and the results of the comparison. The main variables studied MTV2. To maintain clarity of the results of Specimen
in these tests are: 1) the design method; 2) the cross section MTV2, the FSR results are shown only for P of approximately
size and reinforcement; 3) the relative length of the spandrel 230 kN and higher. The response was nearly linear past the
and floor beams; 4) the longitudinal restraint in the spandrel cracking in the beams. At ultimate load, the restraint in the
beam; 5) the concentrated load at the joint; and 6) the floor and the spandrel beams caused compressive forces of
uniform load on the floor beam. 560 and 590 kN, respectively, which are equivalent to a
It is to be noted that frames without joint loads simulate stress of approximately P/Agross = 0.3fc′ .
floor beams framing near midspan of the spandrels, where Figure 9(c) shows the vertical deflections from the NR and
the spandrel shearing stresses are mainly due to the torsional FSR analyses. Releasing the restraint in the spandrel beam
moment, with smaller contribution from the shearing force. NR resulted in a softer response. Adding the restraint in the
To simulate the more critical conditions near the supports of flexurally critical floor beam (FSR) allowed to maintain the
the spandrels, Hsu and Hwang7 and Abul Mansour and stiff response of the frame up to failure load, and did not
Rangan8 tested specimens with floor loading in addition to a display the softening typically associated with yielding in the
concentrated load at the joint. This joint load simulated the bottom longitudinal steel in the floor beam.
difference in column load near the support, and increased the Figure 10(c) shows that releasing the restraint in the spandrel
shearing force in the spandrel.7 This is similar to the conditions (NR) caused an increase in the diagonal cracks, even though
near the support regions of the spandrel. at a specific level of load resulted in a smaller torque. Adding
Three out of the 30 frames listed in Table 1 did not contain the longitudinal restraint in the floor beam (FSR) allowed to
transverse reinforcement in the spandrels, and hence were maintain the stiff response of the frame up to failure load.
omitted from the comparison. For the remaining 27 specimens, Figure 10 shows that, while frames designed using Method
the average of the ratio Pf-exp /Pf-calc was 1.05 and the coefficient B can develop relatively larger cracks, these cracks remain
of variation was 7.4%. Table 1 shows that the proposed below 0.3 mm at estimated service load, especially when
procedure gave accurate results even when the transverse steel in longitudinal restraint is taken into consideration.
the spandrels was close to minimum shear reinforcement values.
Mo and Hsu13 studied only 20 tests, six by Ozerdinc et al.,5 CONCLUSIONS
10 by Hsu and Burton,6 and four out of the eight by Abul Based on the research, the following conclusions can
Mansur and Rangan.8 Their method was applicable only to be made:
11 out of the 20 tests. For these specimens, the average and 1. The MCFT is capable of accurately predicting the full
coefficient of variation of Pf-exp/Pf-calc were reported13 to be response of beams subjected to combined shear, torsion,
1.27 and the 10.2%, respectively. For the same 11 specimens, flexure, and axial compression;
the procedure proposed in this study gives an average of 1.04 2. The proposed procedure is capable of predicting the
and a coefficient of variation of 4.9%. response of indeterminate frames where the effects of
compatibility torsion are dominant;
Effect of longitudinal restraint 3. The effective torsional stiffness in spandrel beams
To check the impact of the longitudinal restraint on the depends mostly on the axial compressive force caused by the
behavior, Specimen MTV2 was reanalyzed twice, the first longitudinal restraint provided by the columns and by the
without the longitudinal restraint in the spandrel (NR), and adjacent bays. It is also affected by the amount of transverse
the second with longitudinal restraint in both the spandrel reinforcement and by the magnitude of the bending
and the floor beams (FSR). The results of these analyses are moment and shearing force acting in combination with the
compared with the results of Specimen MTV2 in Fig. 5 to 11. compatibility torsion;
Figure 5 shows that the restraint in FSR increased the flex- 4. The torsional (and flexural) stiffness varies considerably
ural stiffness of the floor beams. The average increase along between torsionally (and flexurally) cracked and uncracked
the length was 55%. This increase was smaller than the regions. To capture the variations in stiffness, relatively
increase in the torsional stiffness in the spandrel, which averaged smaller elements should be used in the structural analysis
219%. This indicates that the increase in stiffness due to modeling in areas where cracked elements are adjacent to
longitudinal restraint was proportional to the decrease in uncracked elements. A maximum length equal to twice
stiffness due to cracking. depth of the beam is recommended;
Figure 6(c) shows that releasing the longitudinal restraint 5. The axial compressive force caused by the longitudinal
from the spandrel of Specimen MTV2 (analysis NR) resulted restraint significantly increases the effective torsional stiffness in
in a softer response. The ultimate floor load was not significantly the spandrel, and consequently increases the magnitude of the
affected because the flexural capacity of the floor beam compatibility torsion. Hence, the current ACI and CSA
was critical in determining the failure load. Adding equations might underestimate the torque in the spandrel beams.
restraint in the floor beam (FSR analysis) increased its The code procedures remain conservative for design purposes,
flexural stiffness and its flexural (and shearing) capacities. The however, because the compressive force also increases the

ACI Structural Journal/May-June 2006 337


torsional strength, which enables the beams to resist the ψ = twist of cross section due to torsional moment
larger torque; ζ = bending curvature in cross section
6. The longitudinal restraint in the spandrel and floor
beams increases the resistance of the frames significantly REFERENCES
1. Collins, M. P., and Lampert, P., “Redistribution of Moments at
beyond the design transverse loads. The calculated increase Cracking—The Key to Simpler Torsion Design,” Analysis of Structural
in one of the frames exceeded 100% of the design load; Systems for Torsion, SP-35, American Concrete Institute, Farmington Hills,
7. Design based on limit state and redistribution of the Mich., 1973, pp. 343-383.
torque leads to relatively larger diagonal crack widths in the 2. Onsongo, W. M., “Longitudinally Restrained Beams in Torsion,”
MASc thesis, Department of Civil Engineering, University of Toronto,
spandrels. The longitudinal restraints present in the beams of Toronto, Ontario, Canada, 1972, 43 pp.
actual structures, however, reduce the widths to acceptable 3. ACI Committee 318, “Building Code Requirements for Structural
limits; and Concrete (ACI 318-02) and Commentary (318R-02),” American Concrete
8. There is a lack of a simple procedure to estimate the Institute, Farmington Hills, Mich., 2002, 443 pp.
effective torsional stiffness due to the numerous factors that 4. CSA Standard, “Design of Concrete Structures (A23.3-94),” Canadian
Standards Association, Rexdale, Ontario, Canada, 1994, 199 pp.
affect it. Experience with the use of the current ACI and CSA 5. Ozerdinc, M. E.; Lampert, P.; and Collins, M. P., “Torsion-Bending-
code provisions show that they are conservative for design. Redistribution at Cracking—An Experimental Investigation,” Civil
For the calculation of the ultimate strength of existing structures, Engineering Publication No. 72-73, University of Toronto, Toronto,
however, the code provisions are unduly conservative, and Ontario, Canada, Dec. 1972, 23 pp.
6. Hsu, T. T. C., and Burton, K. T., “Design of Reinforced Concrete
more accurate calculations can be obtained using rational Spandrel Beams,” Journal of the Structural Division, V. 100, No. ST1, Jan.
procedures such as that presented in this paper. 1974, pp. 209-229.
7. Hsu, T. T. C., and Hwang, C., “Torsional Limit Design of Spandrel
ACKNOWLEDGMENTS Beams,” ACI JOURNAL, Proceedings V. 74, No. 2, Feb. 1977, pp. 71-79.
The research reported in this paper was carried out at the Department of 8. Abul Mansur, M., and Rangan, V., “Torsion in Spandrel Beams,” Journal
Civil Engineering at Kuwait University. Support, and making the computer of the Structural Division, V. 104, No. ST7, July 1978, pp. 1061-1075.
facilities available, is gratefully acknowledged. The development of the 9. ACI Committee 318, “Building Code Requirements for Reinforced
MCFT has been made possible by a series of grants from the National Concrete (ACI 318-63),” American Concrete Institute, Farmington Hills,
Science and Engineering Research Council of Canada. This support is Mich., June 1963, 144 pp.
gratefully acknowledged. 10. Torsion of Structural Concrete, SP-18, American Concrete Institute,
Farmington Hills, Mich., 1968, 505 pp.
11. ACI Committee 318, “Building Code Requirements for Reinforced
NOTATION Concrete (ACI 318-71),” American Concrete Institute, Farmington Hills,
Ac = area enclosed by outside perimeter of concrete cross section Mich., 1971, 78 pp.
Agross = gross area of concrete cross section 12. ACI Committee 318, “Building Code Requirements for Reinforced
E = modulus of elasticity of concrete Concrete (ACI 318-83),” American Concrete Institute, Farmington Hills,
fc′ = specified concrete strength of concrete Mich., 1983, 111 pp.
fy = yield stress of reinforcing steel 13. Mo, Y. L., and Hsu, T. T. C., “Redistribution of Moments in Spandrel
G = shear modulus of concrete Beams,” ACI Structural Journal, V. 88, No. 1, Jan.-Feb. 1991, pp. 22-30.
I = moment of inertia of floor beam cross section 14. Vecchio, F. J., and Collins, M. P., “Predicting the Response of
K = torsional stiffness of spandrel beam cross section Reinforced Concrete Beams Subjected to Shear Using the Modified
Pf = floor beam total load Compression Field Theory,” ACI Structural Journal, V. 85, No. 3, May-
Pf-calc= floor beam calculated total load June 1988, pp. 258-268.
Pf-des = floor beam design total load 15. Rahal, K. N., and Collins, M. P., “Analysis of Sections Subjected to
Pf-exp = floor beam observed total load Combined Shear and Torsion—A Theoretical Model,” ACI Structural Journal,
Ps = spandrel beam midspan load V. 92, No. 4, July-Aug. 1995, pp. 459-469.
pc = outside perimeter of concrete cross section 16. Rahal, K. N., and Collins, M. P., “Combined Torsion and Bending in
M = bending moment Reinforced Concrete Beams,” ACI Structural Journal, V. 100, No. 2, Mar.-
N = axial force Apr. 2003, pp. 157-165.
T = torsional moment 17. Rahal, K. N., and Collins, M. P., “Effect of Thickness of Concrete
TACI = design compatibility torsion (ACI code) Cover on Shear and Torsion Interaction—An Experimental Investigation,”
TCSA = design compatibility torsion (CSA code) ACI Structural Journal, V. 92, No. 3, May-June 1995, pp. 334-342.
φ = strength reduction factor (ACI code) 18. CEB-FIP, “Model Code for Concrete Structures: CEB-FIP International
φc = material reduction factor (CSA code) Recommendations,” 3rd Edition, Comité Euro-International du Béton,
ρv = ratio of transverse reinforcement Paris, 1978, 348 pp.

338 ACI Structural Journal/May-June 2006

You might also like