Sunteți pe pagina 1din 15

THE STRUCTURAL DESIGN OF TALL AND SPECIAL BUILDINGS

Struct. Design Tall Spec. Build. 17, 757771 (2008)


Published online 27 July 2007 in Wiley Interscience (www.interscience.wiley.com). DOI: 10.1002/tal.375

PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE

MEHRDAD SASANI* AND JESSE KROPELNICKI


Northeastern University, Boston, Massachusetts, USA

SUMMARY
Progressive collapse denotes a failure of a major portion of a structure that has been initiated by failure in a rel-
atively small portion of the structure. One approach to evaluate progressive collapse of structures is to study the
effects of instantaneous removal of a load-bearing element such as a column. An experimental program is carried
out to study the behavior of a 3/8 scaled model of a continuous perimeter beam in a reinforced concrete frame
structure following the removal of a supporting column. A detailed finite element model (FEM) is developed and
verified to capture the behavior of the beam subjected to large deformation. In order to avoid a detailed FEM of
the whole building and to efficiently capture the system response, a three-dimensional nonlinear model of the
structure using beamcolumn and shell elements is also developed. The two models are integrated through hybrid
(substructuring) simulations. The potential progressive collapse of the structure and the dynamic load redistrib-
utions following column removals are studied. Copyright 2007 John Wiley & Sons, Ltd.

1. INTRODUCTION
The Ronan Point apartment building collapse in England in 1968 generated substantial interest in
general structural integrity for buildings and the prevention of progressive collapse. A second wave
of interest followed the attack on the Murrah Federal building in 1995 and the terrorist attacks on the
World Trade Center towers and the Pentagon on September 11, 2001. Progressive collapse is by nature
a system-level problem where the spread of an initial local failure from element to element eventu-
ally results in the collapse of an entire structure or a disproportionately large part of it (ASCE-7, 2002).
From an analytical point of view, progressive collapse occurs when a structure has its load pattern or
boundary conditions changed such that other structural elements are loaded beyond their capacity and
fail (Krauthammer et al., 2003).
Allen and Schriever (1972) defined progressive collapse as a situation where local failure of a
primary structural component(s) leads to the collapse of adjoining members, which in turn leads to
additional collapse. Thus, the extent of collapse is disproportionate to the original cause. In other
words, progressive collapse is a chain reaction of failures following damage to a relatively small
portion of a structure. Following the approaches proposed by Ellingwood and Leyendecker (1978),
ASCE-7 (2002) defines two general methods for structural design of buildings to mitigate damage due
to progressive collapse: indirect and direct design methods:

Indirect design: Incorporates implicit consideration of resistance to progressive collapse through the
provision of minimum levels of strength, continuity, and ductility.
Direct design: Incorporates explicit consideration of resistance to progressive collapse through
two methods. One is the alternative path method, in which local failure is allowed to occur,

* Correspondence to: Mehrdad Sasani, Department of Civil and Environmental Engineering, 400 Snell Engineering Center,
Northeastern University, Boston, MA 02115-5000, USA. E-mail: sasani@neu.edu

Copyright 2007 John Wiley & Sons, Ltd.


758 M. SASANI AND J. KROPELNICKI

but seeks to provide alternative load paths so that the damage is absorbed and major collapse is
averted. The other method is the specific local resistance method, which seeks to provide strength
to resist failure.
Breen (1975) has suggested that improved structural integrity is obtained by provision of integral ties
throughout the structure (indirect design) and that the number of ties can be determined from consid-
erations on debris loading and the amount of damage to be tolerated without determination of the mag-
nitude of the explosive or other abnormal load. Although the indirect design method can reduce the
risk of progressive collapse (FEMA 277, 1996; Corley et al., 1998; Sozen et al., 1998; Corley, 2004),
estimation of post-failure performance of structures designed on the basis of such a method is not
readily possible.
GSA (2003) provides guidelines for progressive collapse analysis of structures based on the alter-
native path method and mandates instantaneous removal of one load-bearing element with different
scenarios as the initiation of damage. Also the maximum allowable extents of collapse are described.
In a linear static analysis the following load combination is considered:

Load = 2 0(DL + 0 25LL ) (1)

where DL and LL are dead load and live load, respectively. For an elastic dynamic or nonlinear analy-
sis, the coefficient 20 in Equation (1) is removed. The acceptance criteria for a linear analysis are
similar to the criteria in FEMA 356 (2000), which are based on internal force demandcapacity ratios
(DCR). GSA (2003) also recommends application of a nonlinear analysis, particularly for buildings
having more than 10 stories above the grade. For such analysis, acceptance criteria based on rotation
or rotation ductility as given in DOD (2005) are provided.
DOD (2005) provides two design methods: one employs the tie force method (indirect design) and
the other employs the alternative path method (direct design). Distinguishing between ductile and
brittle modes of failure, acceptance criteria consist of strength requirements and deformation limits.
If an element failed to satisfy deformation limits or its behavior is brittle and fails to satisfy strength
requirements, the element is removed and its internal forces are (dynamically) redistributed. Detailed
guidelines for analysis procedures are presented under the alternative path method. The guidelines
allow an (iterative) linear static analysis, as well as nonlinear static and dynamic analyses. Note that
based on DOD (2005) linear analyses are not allowed.
Potential progressive collapse of structures due to terrorist attack (or accidental loads) and the
dynamic redistribution of loads that follows need to be evaluated at the system level. In this paper,
using the alternative path method, the response of a seven-story reinforced concrete (RC) structure
following loss of load-bearing elements is studied. In order to evaluate the progressive collapse resist-
ance of a structure, its critical beams need to be reliably modeled. An experimentally verified detailed
finite element model of beams bridging over a removed column is developed and hybrid analyses of
the whole structure are carried out. The dynamic redistribution of the gravity loads and the response
of the building following the removal of a single column as well as two adjacent columns are obtained
and discussed. The importance of proper floor modeling is examined. The application of DCR is also
evaluated.

2. CHARACTERISTICS OF BUILDING TO BE STUDIED


A seven-story building with ordinary RC frames is designed. The building is assumed to be located
on a site class C, very dense soil and soft rock (IBC, 2003) in Atlanta, GA. The plan of the building
is shown in Figure 1. The floor is a one-way joist system in the transverse direction. The span in the
transverse direction is set equal to 30 ft in order to have an economical joist floor system (Alsamsam

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 759

A B C D E F G

30'

30'
Joist
Floor
3
19' 19' 19' 19' 19' 19'
Region to
be tested

Figure 1. Plan of building

Figure 2. Removal of column due to explosion/impact

and Kamara, 2004). The total depth of the floor system is 20 in. with a solid slab of 45 in. The depth
of all beams is equal to 20 in. (equal to the depth of joist floor system) to minimize the formwork cost.
Reinforcement of grade 60 ksi is used along with concrete compressive strength of 4 ksi. In the design
of the building the integrity requirements are satisfied (ACI 318, 2002). In order to examine effects
of splices on the development of catenary action in beams and progressive collapse of structures, the
longitudinal reinforcements are spliced.

3. REMOVAL OF A FIRST-FLOOR COLUMN


As discussed in the Introduction, one of the scenarios to simulate damage to a support in a structure
is the instantaneous removal of an exterior column. Figure 2 schematically shows the removal of such
a column, which results in a dynamic redistribution of loads through the beams, floor systems and
columns. Following the redistribution of loads, if the structure comes to an equilibrium position pro-
gressive collapse does not occur. However, if some elements are strained beyond their capacities, pro-
gressive collapse can be initiated.
The unsupported spans of beams in Figure 2 that were previously supported by the removed column
are increased by a factor of two. Furthermore, the gravity loads dynamically affect such beams, impos-
ing even more demand than just static equilibrium would suggest. Given such high demand, the beam
integrity requirements are meant to provide these elements with catenary action to resist the applied

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
760 M. SASANI AND J. KROPELNICKI

forces (Item R7.13.2 in ACI 31802) under imposed large deformations. However, to the best knowl-
edge of the authors, there are no experiments that have been conducted on RC beams to evaluate their
potential for development of catenary action as well as their rotational capacity under such extreme
loading conditions. Therefore, an experimental program was conducted to evaluate the behavior of the
perimeter beams following the loss of the supporting column.

4. EXPERIMENTAL PROGRAM
In conducting an experiment to study the behavior of beams bridging over the removed column, there
is a need for proper consideration of the boundary conditions of such beams. Following the removal
of column D-3 (see Figure 1), the ends of the beams running between columns C-3 and E-3 can move
in six directions: three translational and three rotational movements. Among these, the horizontal dis-
placement and the rotation in the plane of frame 3 are the two more important movements, which are
related to the axial force and the major bending moment of the beams.
The horizontal constraints of joints C-3 and E-3 (Figure 1) along frame 3 are mainly provided by
the beams of this frame (i.e., beams A-3 to C-3 and E-3 to G-3). The movements of these beams in
the planes of floor diaphragms are in turn constrained primarily by the floor slab as well as by the
columns of not only frame 3 but also those of frames 2 and 1. Such constraints are required for the
development of the catenary action in the beams bridging over the removed columns with larger ver-
tical displacements. Furthermore, these constraints can significantly affect the axial compressive forces
developed in the beams under smaller vertical displacement that in turn affect beam flexural strength.
The rotational constraints of joints C-3 and E-3 (Figure 1) in the plane of frame 3 are mainly pro-
vided by the columns and beams of the frame, connected to these joints. The inclusion of columns C-3
and E-3 in the physical model of the beam C-3 to E-3 would improve the modeling of the rotational
stiffness of beam end boundary conditions. In this case, however, in order to provide the beam with the
axial force required for the development of catenary action, the beams beyond points C-3 and E-3 also
need to be modeled. To avoid such a costly experiment, it is decided to model the beam bridging over
the lost column with fixed boundary conditions. The effects of the movement of the boundaries on the
beam and in turn on the structure will be modeled analytically as will be described later in this paper.
A 3/8 scale model of the second floor beam, bridging the removed column, was constructed with
fixed boundary conditions. Figure 3 shows the reinforcement detailing of the beam. The yield and ulti-

Figure 3. Reinforcement detailing of longitudinal beams of second floor

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 761

mate stresses of the longitudinal bars were 75 ksi and 103 ksi, respectively. The concrete compressive
strength was about 6 ksi. The test was conducted utilizing displacement control at the center span.
In order to evaluate the test set-up, the vertical displacement was applied slowly up to a deflection of
04 in. Beyond this initial displacement of 04 in., the displacement was applied at a rate of 2 in./s
(90% of the actuator capacity).
Figure 4 shows the vertical force versus the displacement of the beam center point. At vertical dis-
placements of about 60 in. and 75 in., the two bottom bars fractured. Figure 5 shows a picture of the
beam after second bar fracture. This bar fracture was observed on one side of the center stub (loca-
tion of removed column). Up to the initiation of the yielding of the bottom bars at the center of the
beam, the tensile strain in these bars on both sides of the center stub was almost equal. The yielding,
however, started on one side of the beam and caused the damage to be concentrated at that location.
The rotation over a length of 75 in. on the damaged side when the second bar ruptured was measured
at about 8 degrees (slope of 14%).

18
Experimental
Analytical
Force (kips)

12

0
0 4 8 12 16
Displacement (in)

Figure 4. Forcedisplacement relationships

Figure 5. Damage at center following bar fractures

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
762 M. SASANI AND J. KROPELNICKI

No indication of splice failure was observed. The following three factors resulted in higher splice
strength at the center of the beam than one would expect: (1) the concrete design strength was 4 ksi,
while the concrete strength on the day of testing was measured at about 6 ksi; (2) although a splice
type A (equal to the development length) satisfies the integrity requirements, splice type B, which is
30% longer than splice type A, was used because the bottom reinforcing bars were under tension due
to seismic loads; and (3) because the reinforcing bars used in the test were #3 (versus #8 bars used in
the prototype structure), an over-strength of 25% is expected in the test (based on ACI-318). If these
factors had not affected the splice, the splice length could have been reduced to about 725 in. (one
half of the 145 in. used in the test) and as result increased the likelihood of splice failure.
As can be seen in Figure 4, following the bar fractures, catenary action provided by the top rein-
forcement results in the increasing resistance of the beam. At a vertical displacement of about 85 in.,
the top continuous bars at the center of the beam which were previously in compression yielded in
tension. The tensile strain in these bars increased to about 0032 at a maximum vertical displacement
of 16 in. Because of geometric constraints, the test was stopped at 16 in. of vertical deformation. At
this point the tensile strain in the top reinforcement at the face of the supports was about 012, which
was close to the fracture strain of 013015. Note that the beam end rotation at the conclusion of the
test was about 113 degrees (slope of 20%).

5. FINITE ELEMENT ANALYSIS OF BEAM


In order to reliably model the beams bridging over the removed column, a detailed finite element
model is developed using the computer program ANSYS. This model utilizes eight-node solid ele-
ments for the concrete and two-node elements for the reinforcing bars. A two-parameter
DruckerPrager yield criterion (Chen, 1982)

f ( I1 , J 2 ) = aI1 + J 2 k = 0 (2)

is used to model the behavior of concrete under three-dimensional states of stress. Note that a smeared
concrete cracking model and a strain-based concrete crushing model, as described later, are used along
with DruckerPrager yield criterion. In Equation (2) I1 and J2 are the first invariant of the stress tensor
and second invariant of the stress deviator tensor, respectively. a and k are the positive model param-
eters that are functions of cohesion and the friction angle. A friction angle of 35 degrees with a cohe-
sion value of 15 ksi ( fc/4) is used (Chen, 1982). Note that in this study an associative flow rule is
utilized. In ANSYS, the DruckerPrager yield criterion supports an elasticperfectly plastic
stressstrain relationship. A modulus of elasticity of 07 fc/e0 is used, which is a reasonable lin-
earization of the ascending branch of the concrete stressstrain relationship up to the peak stress. (e0
is the concrete strain associated with maximum compressive strength in a uniaxial compressive test.)
The concrete crushing capability available in ANSYS is stress-based; however, a strain-based crush-
ing model is more appropriate in modeling confined as well as unconfined concrete. Therefore, a strain-
based crushing routine is implemented in ANSYS utilizing the APDL programming language.
The minimum principal strain in each element is evaluated and the element birth and death capa-
bilities in ANSYS: kill all elements exceeding a predetermined input strain. These strain values are
determined to be 0006 and 004 for the cover and core concrete, respectively. The available smeared
concrete cracking model in ANSYS is used. Based on experimental results on the concrete, a modulus
of rupture of 8 3 fc is found, where fc is the concrete compressive strength. The tensile strength was
determined to be equal to the modulus of rupture based on test results.
The longitudinal reinforcement for the beam is modeled discretely using truss elements, while
the transverse reinforcing is modeled utilizing the smeared reinforcement. The stressstrain

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 763

relationships of both the transverse and the longitudinal steel bars are input based on tensile tests of
the steel bars.
The finite element mesh is refined in areas with high stress gradients and verified through sensitiv-
ity analysis of multiple mesh schemes. Geometric nonlinearity is accounted for. A modified
NewtonRaphson solution technique with line search was found to be the most robust method of those
offered in ANSYS to analyze large displacement response of the beam.
Figure 4 compares the analytical forcedeformation relationship of the test beam, which is in close
agreement with the experimental results. At a vertical displacement of about 5 in., the analytical results
show the bar fracture at the center point, beyond which the analysis was not continued. Note that bond
slip is not modeled, which would have resulted in a larger estimated vertical displacement. Figure 6
shows the longitudinal concrete stress in the beam (the vertical plane is at the center of the beam) at
a vertical displacement of 23 in. This displacement corresponds to a sudden drop in strength (see
Figure 4), which is due to the crushing of concrete cover close to the center stub. The boundary of
concrete core (confined by transverse reinforcement) is shown by solid lines. Figure 6 shows that the
axial compressive stress in the cover concrete at the top and close to the center stub is reduced to
zero following concrete crushing. Furthermore, the concrete stress in the top of the core close to the
center stub remains high. Due to three-dimensional confining effects, the stress is increased above
the uniaxial strength of concrete. Figure 7 compares the experimental and analytical end rotation

Figure 6. Longitudinal concrete stress in tested beam

0.06

Experimental
Analytical
0.04
Rotation

0.02

0.00
0 1 2 3 4 5
Displacement (in)

Figure 7. Beam rotation over a length of 75 in. at end support

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
764 M. SASANI AND J. KROPELNICKI

18

Force (kips)
12

0
0 4 8 12 16
Displacement (in)

Figure 8. Analytical forcedisplacement relationship for beam with symmetric deformation

of the beam. The experimental rotation at the center of the beam is also closely predicted by the
analytical results.
As discussed in the experimental program, the deformation of the beam was not symmetric and
damage in the vicinity of the center stub was concentrated on one side. In fact, the rotation and dis-
placement of the center stub were measured and accounted for in the evaluation and verification of
the finite element model. Figure 8 shows the analytical force deformation relationship if the defor-
mation of the beam was symmetric. As can be seen, catenary action starts to overcome the loss of
beam strength at about 8 in. vertical displacement. The analysis shows that at a vertical displacement
of about 13 in. the bottom reinforcing bars fracture. This is associated with a rotation of about 9 degrees
(slope of 16%). Note that following the bottom bar fracture, and as was observed in the experimen-
tal program, the top reinforcing bars can continue providing beam strength through catenary action.
Analytical modeling beyond bar fracture requires significant and sudden redistribution of stresses and
strains which is not carried out in this study. Studies on conducting such modeling are underway.

6. THREE-DIMENSIONAL MODEL OF STRUCTURE


In order to study the response of the building following the removal of the column, an analytical model
of the structure needs to be developed. To avoid a detailed FEM of the whole building and to effi-
ciently capture the system response, EulerBernoulli beamcolumn elements are used for modeling
beams and columns. The program ANSYS does not provide a library of such elements to model non-
linear behavior of RC beamcolumns. Therefore, using the computer program Open Source for Earth-
quake Engineering Simulation (OpenSEES, 2002) a three-dimensional model of the building is
developed.
Except for the beams between columns C-3 and E-3 in different floors (see Figure 1),
EulerBernoulli beamcolumns are used to model beams and columns and their cross sections are dis-
cretized to fibers. Both material and geometric nonlinearities are included in the model. Note that the
bar pull-out (from joints) are explicitly accounted for in the critical beams in the detailed finite element
models. The co-rotational coordinate transformation (Felippa, 2000) is used to account for geometric
nonlinearity. The co-rotational formulation of geometrically nonlinear analysis has been developed in
recent years to simplify the treatment of arbitrarily large rotations for structural finite elements such
as beams and shells that contain rotational degrees of freedom. Given the effectiveness of such devel-
opment, it has essentially superseded the Updated Lagrangian formulation for such elements.

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 765

The modified Kent and Park model (Park et al., 1982) is employed to describe the stressstrain rela-
tions for unconfined and confined concrete. Similarly, each floor joist is explicitly modeled. An implicit
integration technique (Newmark b method) is used in the analysis. A classical (Rayleigh) mass pro-
portional damping matrix associated with 5% damping in the first mode of vibration (following the
column removal) is used. A lumped mass matrix is assumed. The floor joists are modeled as nonlin-
ear beams with T-sections. The 45 in. thick concrete slabs are modeled by shell elements.
After analyzing the structure under gravity loads, initial local failure as a result of the removal of
a column is modeled by a sudden release of the resultant end forces of the column (Powel1, 2004).
First the structure is analyzed under the gravity loads and the internal forces in the column to be
removed are determined. Next, the column is removed and, instead, the column top reactions are
applied to the structure, along with the gravity loads. Note that the results of such analysis are iden-
tical to those of the previous analysis, where the column was not removed. Finally, forces in the oppo-
site direction to the forces applied to the structure at the top of removed column are suddenly applied
to the structure to model the removal of the column and a dynamic analysis is conducted. Potential
failure of elements, following dynamic redistribution of loads, is modeled through appropriate force
deformation (stressstrain) relationships.

7. HYBRID ANALYSIS
Hybrid (pseudo-dynamic) testing is a procedure that combines experimental techniques with computer
simulation of structural response. The concept of hybrid testing was first proposed by Hakuno et al.
(1969). The test procedure in the present form was first introduced by Takanashi et al. (1977). Con-
siderable development of the method was due to Mahin and Shing (1985) and Shing et al. (1990). In
hybrid testing, the objective is to determine the response of a simulation composed of experimental
and numerical elements. The complete structural model is idealized as a discrete system with a finite
number of degrees of freedom. The governing equation of motion is

m + c u + F = Peff (3)

where m is the mass matrix, c is the damping matrix, u is the displacement vector at the degrees of
freedom, F is the vector of element restoring forces, and Peff is the effective loading vector, which for
progressive collapse studied here is associated with the removal of a column. The superscript dot refers
to derivative with respect to time. That is, and u are relative acceleration and velocity vectors, respec-
tively. Under severe loading, the nonlinear restoring force can become increasingly difficult to esti-
mate for some structural elements. For these cases, hybrid testing can be used to eliminate some of
the uncertainty in modeling by replacing the element restoring forces with measured data from an
experiment.
In a hybrid test method, similar to a conventional numerical analysis in the time domain, a proto-
type structure is modeled in the computer to conduct a dynamic structural analysis. The governing
equation of motion (3) is solved using time-stepping integration algorithms. The computer analysis
program is linked to one or more physical sub-assemblages of the structure. The experimental element
is incorporated into the analytical model by means of sub-structuring techniques (Dermitzakis and
Mahin, 1985). A dynamic analysis of the complete structure is carried out whereby the experimental
subassembly is considered as part of the analytical model.
In order to experimentally obtain the element restoring forces, analytically estimated displacements
are imposed on the physical model. For beam elements, each boundary point is associated with six
degrees of freedom. Imposing such boundary conditions in the laboratory requires elaborate testing
facilities. To model the behavior of critical beams following the loss of a load-bearing element, in this

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
766 M. SASANI AND J. KROPELNICKI

study the boundary conditions in the laboratory were imposed as fixed. Having developed an analyt-
ical model that has been calibrated and verified against the experimental data as described above, the
actual boundary conditions are applied to the analytical model.
A communication protocol is developed to automatically manage the interaction between the two
computer programs OpenSEES and ANSYS. At each step of analysis, OpenSEES provides ANSYS
with the displacement boundary conditions of beams bridging over the removed column. ANSYS ana-
lyzes the continuous beam subsystems under the given displacements and obtains force vectors at the
boundaries and reports them back to OpenSEES. OpenSEES utilizes the forces and checks conver-
gence requirements. If convergence was not achieved, the communication will continue and ANSYS
would use new displacements obtained from OpenSEES. After achieving convergence, the next step
of analysis is conducted similarly.
The vertical displacement of joint D-3 at the second floor is shown in Figure 9. The thick and thin
solid curves are for the cases where the first-story column is removed in 001 s and 0001 s, respec-
tively. In both cases a maximum vertical displacement of about 22 in. is observed. Figure 9 also shows
the vertical displacement of joint D-3 at the roof, which apart from some initial time lag is almost the
same as that at the second floor. In order to evaluate the effects of a rigid floor assumption, the verti-
cal displacement of joint D-3 using rigid floors is also shown. As can be seen, the maximum dis-
placement of about 18 in. is estimated. The reduction in the maximum displacement compared to the
model with non-rigid floor is mainly due to the amount of axial compressive force developed in the
beams bridging over the lost column. When the floor is rigid, such beams that tend to elongate due to
cracking cannot elongate and they develop larger axial compressive forces. The maximum axial force
in the beam with rigid floor assumption is about 297 kips, which is more than twice that with non-
rigid floor. The higher axial force in the beam results in larger bending moment capacity by about
20%, which in turn leads to a smaller vertical displacement.

Time (sec)
0 0.1 0.2 0.3 0.4 0.5 0.6
0
2nd Floor Displacement (Column removed in 0.01 sec)
2nd Floor Displacement (Column removed in 0.001 sec)
Roof Displacement (Column removed in 0.001 sec)
-0.5
Vertical Displacement (in)

2nd Floor Displacement (Column removed in 0.01 sec) Rigid Floor

-1

-1.5

-2

-2.5
Figure 9. Vertical displacement of second floor and roof right above removed column

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 767

If the concrete strength and the steel yield stress are reduced to 4 ksi and 60 ksi (values used in
design), respectively (from 6 ksi and 75 ksi obtained from the experiment), a maximum vertical dis-
placement of about 31 in. is found. This displacement is associated with a beam rotation of only about
086 degrees (slope of 15%).

8. DYNAMIC REDISTRIBUTION OF LOADS


Following the removal of column D-3 (Figure 1), the gravity loads transferred to the support through
this column need to be dynamically redistributed to the adjacent columns. In this section, the mech-
anism of such redistribution is examined. Figure 10 shows the variation of the axial loads in the
columns right above the removed column in the first 0015 s (15 ms). Compressive forces are shown
as positive values. The first story column is removed in 0001 s. As can be seen, the axial force in the
second-story column drops to zero after only about 0005 s. This is due to the fact that as the bottom
of the second-story column starts to move downwards, the column axial compressive stress and strain
are relaxed. As the axial force in the second-story column drops, the upward force provided by this
column to joint D-3 on the third floor drops. As a result, the downward force of the third-story column
moves this joint on the third floor downwards. A similar phenomenon occurs in the stories above, with
some time lags (see Figure 10). It takes about 0008 s until the movement starts to affect the axial force
of the seventh-story column (see Figure 10).
Figure 11 shows the velocity of joint D-3 in the second and third floors. When the axial force in
the second story column is about zero (at about 0005 s), the velocity of joint D-3 is at its peak value
of about 255 in/s. From 001 to 009 s, the velocity varies between 135 in./s to 223 in./s, with a mean
of about 181 in./s. Such rather limited variation in the velocity results in an approximate linearly
increasing downward displacement of joint D-3 in the second floor (see Figure 9). As can be seen in
Figure 11, the velocity of the third floor after some time lag is close to that of the second floor. A
similar pattern occurs in other floors as well. As a result, and as can be seen in Figure 9, the dis-
placements of the second and the roof, other than some time lag at the beginning, are almost identi-
cal. The same behavior is also observed for the other floors.
As discussed above, and apart from some time lags at the beginning, joints D-3 in all floors move
similarly in the vertical direction. While the beam cross-sections are constant over the height of the
building, different amounts of reinforcement are used in the beams in floors 13, 45, and 67, making
the beams in the top floors weaker than in the lower floors. Therefore, some of the loads of the top
floors are transferred to the beams below, which are stronger. The mean resultant vertical force applied
to joint D-3 at different floors by the columns from 005 to 060 s are shown in Table 1. The positive
values are upwards. The values show that while the top two beams are affected by upward loads of

Table 1. Mean values of resultant column forces applied to


joint D-3

Floor Force (kips)


Roof 140
7 151
6 20
5 11
4 101
3 108
2 113

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
768 M. SASANI AND J. KROPELNICKI

about 140 and 151 kips, the bottom three beams are affected by downward loads from 101 to
113 kips. The mean axial forces in the columns during this time period are all compressive, with a
maximum of 322 kips in the fourth-floor column and a minimum of 113 kips in the second-floor
column. Note that the axial forces in the columns before the removal of the first story column are
shown in Figure 10 at time zero. The analysis shows that about 95% of the axial force of the removed
column was transferred to columns C-3 and E-3 and only about 5% was transferred to column D-2.

9. TWO-COLUMN REMOVAL
In order to further evaluate the potential progressive collapse of the building with severe initial damage,
Figure 12 shows the vertical displacement of the mid-span point of beams D-3 to E-3, if columns D-

350
2nd Story Column
300 3rd Story Column
4th Story Column
250
Axial Force (kips)

5th Story Column


200 6th Story Column

150 7th Story Column

100

50

-50
0 0.003 0.006 0.009 0.012 0.015
Time (sec)

Figure 10. Axial force in columns D-3

Time (sec)
0 0.02 0.04 0.06 0.08 0.1
0

-4 2nd Floor Velocity

-8 3rd Floor Velocity


Velocity (in/sec)

-12

-16

-20

-24

-28

Figure 11. Velocity of joint D-3 in second and third floors

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 769

Time (sec)
0 0.1 0.2 0.3 0.4 0.5 0.6
0

Vertical Displacement (in)


-1
-2
-3
-4
-5
-6
-7 Nonlinear
-8
-9 Linear-Elastic
-10

Figure 12. Vertical displacement of middle point of beam D-3 to E-3 in second floor following simultaneous
removal of columns D-3 and E-3

3 and E-3 are simultaneously removed. The concrete strength and the steel yield stress are 4 ksi and
60 ksi (design values), respectively. As can be seen, the maximum vertical displacement is about 9 in.,
which is associated with a beam rotation of about 15 degrees (slope of 27%).
If a linear dynamic analysis based on GSA (2003) is carried out, the maximum vertical displace-
ment of the mid-span point of beam D-3 to E-3 of about 73 in. will be found; see Figure 12. Note
that based on ACI 318 (2002) the flexural stiffness of beam and column sections is considered as 35%
and 70% of those of gross sections, respectively. The maximum positive bending moment from this
linear analysis in the beam is found equal to 6540 k in. The nominal bending moment capacity of the
beam is about 1600 k in., which is almost 1/4 of the demand. Because the DCR is more than 2, based
on GSA (2003) the structure is not acceptable for such an event, while the inelastic analysis demon-
strate a rather moderate rotation demand on the beam.

10. CONCLUSIONS
An experimental test is conducted on the behavior of RC continuous beams up to collapse following
the loss of a supporting column. It is shown that in spite of tensile reinforcement fracture of beam
bottom reinforcement the beam has significant remaining strength and deformation capacity. By sat-
isfying the integrity requirements of ACI-318, catenary action develops in top reinforcement. The
beam end rotation at the conclusion of the test was measured at about 11 degrees (slope of about 20%).
Although the splices in bottom reinforcement did not fail, because of different sources that resulted
in stronger splices no significant conclusions can be made regarding the effectiveness of class B
splices, as is accepted by ACI-318.
A cost-effective hybrid analysis is conducted to study potential progressive collapse of an RC struc-
ture. A detailed finite element model of the beam is developed that closely predicts the behavior of
the beam up to the fracture of the bottom reinforcing bars. Such a calibrated and verified analytical
modeling along with hybrid analyses can be utilized to evaluate potential progressive collapse of RC
structures.
It is shown that a rigid floor assumption leads to significantly larger axial compressive forces in
the critical beams (more than twice) compared to that in a non-rigid floor system. Such larger
forces will result in overestimating the bending moment capacity of the beam, which in turn results
in underestimating the maximum vertical displacement of the beams bridging over the removed
column.

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
770 M. SASANI AND J. KROPELNICKI

The dynamic redistribution of loads is described. It is concluded that following the removal of load-
bearing columns the axial forces in the columns directly above the removed column reduce almost to
zero in a very short period of time (515 ms). Following these reductions in column axial forces, the
beams bridging over the lost column oscillate practically in phase and with almost the same ampli-
tude, with the weaker top beams somewhat supported by the bottom beams. As a result of the removal
of column D-3, only about 5% of the column load is transferred to column D-2 in the adjacent lon-
gitudinal frame and the remaining load is practically supported by columns C-3 and E-3.
For the structure studied, it is shown that even simultaneous removal of two adjacent columns from
an exterior frame will not result in progressive collapse of the structure. The response of the building
is also examined using the DCR method. Based on the DCR method, the removal of two adjacent
columns results in internal forces four times the capacity of the beam and, unlike the results of non-
linear dynamic analysis, the structure is considered unsafe if two columns are removed. Therefore it
is concluded that the DCR method can be overly conservative. By implementing the analytical inves-
tigation as presented in this paper, better structural analysis efficiency can be achieved, thereby reduc-
ing engineering costs during design against potential progressive collapse.

ACKNOWLEDGEMENT

Mr Dominic J. Kelly (SGH) and Dr S. K. Ghosh (S. K. Ghosh Associates Inc.) provided valuable
advice on the design of the building and its detailing. The authors are greatly thankful for their help.
This study is in part supported by NSF grant no. CMS-0601258. The authors greatly appreciate this
support. Any opinions, findings, and conclusions or recommendations expressed in this material are
those of the authors and do not necessarily reflect the views of the National Science Foundation.

REFERENCES

ACI 318-02. 2002. Building Code Requirement for Structural Concrete. American Concrete Institute: Farming-
ton Hills, MI.
Allen DE, Schriever WR. 1972. Progressive Collapse, Abnormal Loads, and Building Codes. Division of Build-
ing Research, National Research Council: Washington DC.
Alsamsam IM, Kamara ME. 2004. Simplified Design: Reinforced Concrete Buildings of Moderate Size and Height
(3rd edn). Portland Cement Association: Skokie, IL.
ASCE-7. 2002. Minimum Design Loads for Buildings and Other Structures, ASCE Standard, ASCE: Reston, VA.
Breen JE. 1975. Research Workshop on Progressive Collapse of Building Structures, University of Texas at Austin.
National Bureau of Standards: Washington, DC.
Chen WF. 1982. Plasticity in Reinforced Concrete. McGraw-Hill: New York.
Corley WG. 2004. Lesson learned on improving resistance of buildings to terrorist attacks. Journal of Perfor-
mance of Constructed Facilities, ASCE 18(2): 6878.
Corley WG, Mlakar PF, Sozen MA, Thornton CH. 1998. The Oklahoma City bombing: summary and recom-
mendations for multihazard mitigation. Journal of Performance of Constructed Facilities, ASCE 12(3):
100112.
Dermitzakis SN, Mahin SA. 1985. Development of substructuring techniques for on-line computer controlled
seismic performance testing, UCB/EERC-85/04 Report, Earthquake Engineering Research Center, University
of California: Berkeley, CA.
DOD. 2005. Design of building to resist progressive collapse. Unified Facility Criteria, UFC 4-023-03. US Depart-
ment of Defense: Washington, DC.
Ellingwood B, Leyendecker EV. 1978. Approaches for design against progressive collapse. Journal of the Struc-
tural Division, ASCE 104(ST3): 413423.
Felippa CA. 2000. A Systematic Approach to the Element-Independent Corotational Dynamics of Finite Elements.
Center for Aerospace Structures, College of Engineering, University of Colorado: Boulder, CO.
FEMA 277. 1996. The Oklahoma City Bombing: Improving Building Performance Through Multi-
hazard Mitigation. Building Performance Assessment Team, Federal Emergency Management Agency:
Washington, DC.

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal
PROGRESSIVE COLLAPSE ANALYSIS OF AN RC STRUCTURE 771

FEMA 356. 2000. Prestandard and Commentary for the Seismic Rehabilitation of Buildings. Federal Emergency
Management Agency: Washington, DC.
GSA. 2003. Progressive Collapse Analysis and Design Guidelines for New Federal Office Buildings and Major
Modernization Projects. US General Service Administration: Washington, DC.
Hakuno M, Shidawara M, Hara T. 1969. Dynamic destructive test of a cantilever beam controlled by an analog
computer. Transactions of Japan Society of Civil Engineers 171: 19.
IBC. 2003. International Building Code. International Code Council: Country Club Hills, IL.
Krauthammer T, Hall RL, Woodson SC, Baylot JT, Hayes JR, Sohn Y. 2003. Development of progressive col-
lapse analysis procedure and condition assessment for structures. In National Workshop on Prevention of Pro-
gressive Collapse in Rosemont, IL. Multihazard Mitigation Council of the National Institute of Building
Sciences: Washington, DC.
Mahin SA, Shing P-SB. 1985. Pseudodynamic method for seismic performance testing. Journal of Structural
Engineering, ASCE 111(7): 14821503.
OpenSEES. 2002. Open system for earthquake engineering simulation, Pacific Earthquake Engineering Research
Center, University of California, Berkeley, CA. http://opensees.berkeley.edu. (accessed: June 15, 2006).
Park R, Priestley MJN, Gill WD. 1982. Ductility of square-confined concrete columns. Journal of the Structural
Division, ASCE 108(ST4): 929950.
Powell G. 2004. Progressive collapse: case studies using nonlinear analysis. In SEAOC Annual Convention, Mon-
terey, CA, August 2004.
Shing PB, Vannan MT, Cater E. 1990. Implicit time integration for pseudo-dynamic tests. Earthquake Engineering
and Structural Dynamics 20: 551576.
Sozen MA, Thornton CH, Corley WG, Mlakar, PF. 1998. The Oklahoma City bombing: structure and mecha-
nisms of the Murrah Building. Journal of Performance of Constructed Facilities, ASCE 12(3): 120136.
Takanashi K, Udagawa K, Tanaka H. 1977. A simulation of earthquake response of steel buildings. In Proceed-
ings of the 6th World Conference on Earthquake Engineering, Vol. 3, New Delhi; 31563162.

Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 17, 757771 (2008)
DOI: 10.1002/tal

S-ar putea să vă placă și