Sei sulla pagina 1di 14

Engineering Structures 30 (2008) 24782491

www.elsevier.com/locate/engstruct

Response of a reinforced concrete infilled-frame structure to removal of two


adjacent columns
Mehrdad Sasani
Northeastern University, 400 Snell Engineering Center, Boston, MA 02115, United States
Received 27 June 2007; received in revised form 26 December 2007; accepted 24 January 2008
Available online 19 March 2008

Abstract
The response of Hotel San Diego, a six-story reinforced concrete infilled-frame structure, is evaluated following the simultaneous removal
of two adjacent exterior columns. Analytical models of the structure using the Finite Element Method as well as the Applied Element Method
are used to calculate global and local deformations. The analytical results show good agreement with experimental data. The structure resisted
progressive collapse with a measured maximum vertical displacement of only one quarter of an inch (6.4 mm). Deformation propagation over
the height of the structure and the dynamic load redistribution following the column removal are experimentally and analytically evaluated
and described. The difference between axial and flexural wave propagations is discussed. Three-dimensional Vierendeel (frame) action of the
transverse and longitudinal frames with the participation of infill walls is identified as the major mechanism for redistribution of loads in the
structure. The effects of two potential brittle modes of failure (fracture of beam sections without tensile reinforcement and reinforcing bar pull
out) are described. The response of the structure due to additional gravity loads and in the absence of infill walls is analytically evaluated.
c 2008 Elsevier Ltd. All rights reserved.

Keywords: Progressive collapse; Load redistribution; Load resistance; Dynamic response; Nonlinear analysis; Brittle failure

1. Introduction
As part of mitigation programs to reduce the likelihood
of mass casualties following local damage in structures, the
General Services Administration [1] and the Department of
Defense [2] developed regulations to evaluate progressive
collapse resistance of structures. ASCE/SEI 7 [3] defines
progressive collapse as the spread of an initial local failure from
element to element eventually resulting in collapse of an entire
structure or a disproportionately large part of it.
Following the approaches proposed by Ellingwood and
Leyendecker [4], ASCE/SEI 7 [3] defines two general methods
for structural design of buildings to mitigate damage due
to progressive collapse: indirect and direct design methods.
General building codes and standards [3,5] use indirect design
by increasing overall integrity of structures. Indirect design is
also used in DOD [2]. Although the indirect design method
can reduce the risk of progressive collapse [6,7] estimation of
Tel.: +1 (617) 373 5222; fax: +1 (617) 373 4419.

E-mail address: sasani@neu.edu.


c 2008 Elsevier Ltd. All rights reserved.
0141-0296/$ - see front matter
doi:10.1016/j.engstruct.2008.01.019

post-failure performance of structures designed based on such


a method is not readily possible.
One approach based on direct design methods to evaluate
progressive collapse of structures is to study the effects
of instantaneous removal of load-bearing elements, such as
columns. GSA [1] and DOD [2] regulations require removal
of one load bearing element. These regulations are meant to
evaluate general integrity of structures and their capacity of
redistributing the loads following severe damage to only one
element. While such an approach provides insight as to the
extent to which the structures are susceptible to progressive
collapse, in reality, the initial damage can affect more than
just one column. In this study, using analytical results that
are verified against experimental data, the progressive collapse
resistance of the Hotel San Diego is evaluated, following
the simultaneous explosion (sudden removal) of two adjacent
columns, one of which was a corner column. In order to explode
the columns, explosives were inserted into predrilled holes in
the columns. The columns were then well wrapped with a few
layers of protective materials. Therefore, neither air blast nor
flying fragments affected the structure.

M. Sasani / Engineering Structures 30 (2008) 24782491

2479

Fig. 1. A south view of hotel San Diego. Center structure is studied in this
paper.

Fig. 2. Second floor of building (Looking south).

2. Building characteristics
Hotel San Diego was constructed in 1914 with a south annex
added in 1924. The annex included two separate buildings.
Fig. 1 shows a south view of the hotel. Note that in the picture,
the first and third stories of the hotel are covered with black
fabric. The six story hotel had a non-ductile reinforced concrete
(RC) frame structure with hollow clay tile exterior infill walls.
The infills in the annex consisted of two wythes (layers) of
clay tiles with a total thickness of about 8 in (203 mm). The
height of the first floor was about 190 800 (6.00 m). The height
of other floors and that of the top floor were 100 600 (3.20 m) and
160 1000 (5.13 m), respectively. Fig. 2 shows the second floor of
one of the annex buildings. Fig. 3 shows a typical plan of this
building, whose response following the simultaneous removal
(explosion) of columns A2 and A3 in the first (ground) floor is
evaluated in this paper.
The floor system consisted of one-way joists running in the
longitudinal direction (NorthSouth), as shown in Fig. 3. Based
on compression tests of two concrete samples, the average
concrete compressive strength was estimated at about 4500
psi (31 MPa) for a standard concrete cylinder. The modulus
of elasticity of concrete was estimated at 3820 ksi (26 300
MPa) [5]. Also, based on tension tests of two steel samples
having 1/2 in (12.7 mm) square sections, the yield and ultimate
tensile strengths were found to be 62 ksi (427 MPa) and 87 ksi
(600 MPa), respectively. The steel ultimate tensile strain was
measured at 0.17. The modulus of elasticity of steel was set
equal to 29 000 ksi (200 000 MPa).
The building was scheduled to be demolished by implosion.
As part of the demolition process, the infill walls were removed
from the first and third floors. There was no live load in
the building. All nonstructural elements including partitions,

Fig. 3. Typical plan of Hotel San Diego (South Annex). First floor removed
columns are crossed.

plumbing, and furniture were removed prior to implosion. Only


beams, columns, joist floor and infill walls on the peripheral
beams were present.
3. Sensors
Concrete and steel strain gages were used to measure
changes in strains of beams and columns. Linear potentiometers
were used to measure global and local deformations. The
concrete strain gages were 3.5 in (90 mm) long having a
maximum strain limit of 0.02. The steel strain gages could
measure up to a strain of 0.20. The strain gages could operate
up to a several hundred kHz sampling rate. The sampling rate
used in the experiment was 1000 Hz. Potentiometers were used
to capture rotation (integral of curvature over a length) of the
beam end regions and global displacement in the building, as
described later. The potentiometers had a resolution of about
0.0004 in (0.01 mm) and a maximum operational speed of about
40 in/s (1.0 m/s), while the maximum recorded speed in the
experiment was about 14 in/s (0.35 m/s).

2480

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 4. Reinforcement detail of columns and (a) Beam A3B3 in second floor; and (b) Beam A1A2.

4. Finite element model


Using the finite element method (FEM), a model of the
building was developed in the SAP2000 [8] computer program.
The beams and columns are modeled with Bernoulli beam
elements. Beams have T or L sections with effective flange
width on each side of the web equal to four times the
slab thickness [5]. Plastic hinges are assigned to all possible
locations where steel bar yielding can occur, including the
ends of elements as well as the reinforcing bar cut-off and
bend locations. The characteristics of the plastic hinges are
obtained using section analyses of the beams and columns and
assuming a plastic hinge length equal to half of the section
depth. The current version of SAP2000 [8] is not able to track
formation of cracks in the elements. In order to find the proper
flexural stiffness of sections, an iterative procedure is used as
follows. First, the building is analyzed assuming all elements
are uncracked. Then, moment demands in the elements are
compared with their cracking bending moments, Mcr . The

moment of inertia of beam and slab segments are reduced by


a coefficient of 0.35 [5], where the demand exceeds the Mcr .
The exterior beam cracking bending moments under negative
and positive moments, are 516 k in (58.2 kN m) and 336 k
in (37.9 kN m), respectively. Note that no cracks were formed
in the columns. Then the building is reanalyzed and moment
diagrams are re-evaluated. This procedure is repeated until all
of the cracked regions are properly identified and modeled.
The beams in the building did not have top reinforcing bars
except at the end regions (see Fig. 4). For instance, no top
reinforcement was provided beyond the bend in beam A1A2,
12 inches away from the face of column A1 (see Figs. 4 and 5).
To model the potential loss of flexural strength in those sections,
localized crack hinges were assigned at the critical locations
where no top rebar was present. Flexural strengths of the hinges
were set equal to Mcr . Such sections were assumed to lose their
flexural strength when the imposed bending moments reached
Mcr .

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 5. Location of bends in beam top reinforcement (in an adjacent annex


building at a location similar to beam A1A2, close to column A1).

The floor system consisted of joists in the longitudinal


direction (NorthSouth). Fig. 6 shows the cross section of
a typical floor. In order to account for potential nonlinear
response of slabs and joists, floors are molded by beam
elements. Joists are modeled with T-sections, having effective
flange width on each side of the web equal to four times the slab
thickness [5]. Given the large joist spacing between axes 2 and
3, two rectangular beam elements with 20-inch wide sections
are used between the joist and the longitudinal beams of axes 2
and 3 to model the slab in the longitudinal direction. To model
the behavior of the slab in the transverse direction, equally
spaced parallel beams with 20-inch wide rectangular sections
are used. There is a difference between the shear flow in the
slab and that in the beam elements with rectangular sections
modeling the slab. Because of this, the torsional stiffness is set
equal to one-half of that of the gross sections [9].
The building had infill walls on 2nd, 4th, 5th and 6th floors
on the spandrel beams with some openings (i.e. windows and
doors). As mentioned before and as part of the demolition
procedure, the infill walls in the 1st and 3rd floors were removed
before the test. The infill walls were made of hollow clay tiles,
which were in good condition. The net area of the clay tiles
was about 1/2 of the gross area. The in-plane action of the
infill walls contributes to the building stiffness and strength
and affects the building response. Ignoring the effects of the
infill walls and excluding them in the model would result in
underestimating the building stiffness and strength.
Using the SAP2000 computer program [8], two types of
modeling for the infills are considered in this study: one uses
two dimensional shell elements (Model A) and the other uses
compressive struts (Model B) as suggested in FEMA356 [10]
guidelines.
4.1. Model A (infills modeled by shell elements)
Infill walls are modeled with shell elements. However, the
current version of the SAP2000 computer program includes
only linear shell elements and cannot account for cracking. The
tensile strength of the infill walls is set equal to 26 psi, with a
modulus of elasticity of 644 ksi [10]. Because the formation of

2481

cracks has a significant effect on the stiffness of the infill walls,


the following iterative procedure is used to account for crack
formation:
(1) Assuming the infill walls are linear and uncracked, a
nonlinear time history analysis is run. Note that plastic hinges
exist in the beam elements and the segments of the beam
elements where moment demand exceeds the cracking moment
have a reduced moment of inertia.
(2) The cracking pattern in the infill wall is determined by
comparing stresses in the shells developed during the analysis
with the tensile strength of infills.
(3) Nodes are separated at the locations where tensile stress
exceeds tensile strength.
These steps are continued until the crack regions are
properly modeled.
4.2. Model B (infills modeled by struts)
Infill walls are replaced with compressive struts as described
in FEMA 356 [10] guidelines. Orientations of the struts are
determined from the deformed shape of the structure after
column removal and the location of openings.
4.3. Column removal
Removal of the columns is simulated with the following
procedure.
(1) The structure is analyzed under the permanent loads and
the internal forces are determined at the ends of the columns,
which will be removed.
(2) The model is modified by removing columns A2 and A3
on the first floor. Again the structure is statically analyzed under
permanent loads. In this case, the internal forces at the ends of
removed columns found in the first step are applied externally to
the structure along with permanent loads. Note that the results
of this analysis are identical to those of step 1.
(3) The equal and opposite column end forces that were
applied in the second step are dynamically imposed on the ends
of the removed column within one millisecond [11] to simulate
the removal of the columns, and dynamic analysis is conducted.
4.4. Comparison of analytical and experimental results
The maximum calculated vertical displacement of the
building occurs at joint A3 in the second floor. Fig. 7 shows the
experimental and analytical (Model A) vertical displacements
of this joint (the AEM results will be discussed in the next
section). Experimental data is obtained using the recordings of
three potentiometers attached to joint A3 on one of their ends,
and to the ground on the other ends. The peak displacements
obtained experimentally and analytically (Model A) are 0.242
in (6.1 mm) and 0.252 in (6.4 mm), respectively, which
differ only by about 4%. The experimental and analytical
times corresponding to peak displacement are 0.069 s and
0.066 s, respectively. The analytical results show a permanent
displacement of about 0.208 in (5.3 mm), which is about 14%
smaller than the corresponding experimental value of 0.242 in
(6.1 mm).

2482

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 6. Typical joist floor system.

Fig. 7. Experimental and analytical vertical displacements of joint A3 in second floor.

Fig. 8. Vertical displacement histories of joint A3 in second floor estimated analytically based on Models A and B (FEM).

Fig. 8 compares vertical displacement histories of joint A3


in the second floor estimated analytically based on Models A
and B. As can be seen, modeling infills with struts (Model
B) results in a maximum vertical displacement of joint A3
equal to about 0.45 in (11.4 mm), which is approximately 80%
larger than the value obtained from Model A. Note that the
results obtained from Model A are in close agreement with
experimental results (see Fig. 7), while Model B significantly
overestimates the deformation of the structure. If the maximum
vertical displacement were larger, the infill walls were more
severely cracked and the struts were more completely formed,

the difference between the results of the two models (Models A


and B) would be smaller.
Fig. 9 compares the experimental and analytical (Model A)
displacement of joint A2 in the second floor. Again, while the
first peak vertical displacement obtained experimentally and
analytically are in good agreement, the analytical permanent
displacement under estimates the experimental value.
Analytically estimated deformed shapes of the structure
at the maximum vertical displacement based on Model A
are shown in Fig. 10 with a magnification factor of 200.
The experimentally measured deformed shape over the end
regions of beams A1A2 and A3B3 in the second floor

M. Sasani / Engineering Structures 30 (2008) 24782491

2483

Fig. 9. Experimental and analytical vertical displacement of joint A2 in second floor.

Fig. 10. Analytical (FEM, Model A) deformed shapes of structure (Second floor experimentally estimated deformed shapes are also shown).

are represented in the figure by solid lines. A total of 14


potentiometers were located at the top and bottom of the end
regions of the second floor beams A1A2 and A3B3, which
were the most critical elements in load redistribution. The
beam top and corresponding bottom potentiometer recordings
were used to calculate rotation between the sections where
the potentiometer ends were connected. This was done by first
finding the difference between the recorded deformations at the
top and bottom of the beam, and then dividing the value by
the distance (along the height of the beam section) between

the two potentiometers. The expected deformed shapes between


the measured end regions of the second floor beams are shown
by dashed lines. As can be seen in the figures, analytically
estimated deformed shapes of the beams are in good agreement
with experimentally obtained deformed shapes.
Analytical results of Model A show that only two plastic
hinges are formed indicating rebar yielding. Also, four sections
that did not have negative (top) reinforcement, reached cracking
moment capacities and therefore cracked. Fig. 10 shows the
locations of all the formed plastic hinges and cracks.

2484

M. Sasani / Engineering Structures 30 (2008) 24782491

5. Applied element model


The Applied Element Method (AEM) can track the
structural collapse behavior during early stages of loading
and can account for nonlinear behavior of structures including
element separation [12]. Structural elements are modeled with
3-dimensional cubical sub-elements (cuboids). Cuboids are
rigid and connection between them is achieved by a series of
spring triples (one axial and two shear element springs), which
represent the structural characteristics of connected cuboids. In
other words, all structural characteristics are concentrated in the
springs.
A three dimensional model of the building using the AEM
is developed in the computer program ELS [13]. Beams are
modeled with 200 spring triples at each cross section as
an array of 10 (over the width) by 20 (over the height).
The same number of springs is used for the cross sections
of the columns. Reinforcement of the beams and columns
are explicitly modeled. Characteristics of the springs are
determined from the geometry and the material properties of
the connected cuboids. For instance, if there is a reinforcing
bar running through the interface of two cuboids, a spring
representing the rebar is assigned to the interface.
Floor joists are modeled in a similar manner to the beams
(i.e. with 200 spring triples). Slabs between joists are also
modeled using cuboid elements. Since beams and joists are
connected to slabs, the contribution of slabs on the behavior
of beams and joists (i.e. T or L beams) is accounted for.
Reinforcement of joists, as well as the reinforcement of slabs,
is explicitly modeled.
As described before, the building had infill walls in 2nd,
4th, 5th and 6th floors. The two layer running bond infills were
made of 12 in 12 in 4 in (305 mm 305 mm 102 mm)
hollow clay tiles. In the analytical model, tiles are modeled with
cuboids having half of the infill thickness (to account for the
tile hollowness) and with the same height and width of the tiles.
Connections between tile elements are modeled with interface
mortar elements.
5.1. Comparison of results
Fig. 7 shows experimental and analytical displacements
of joint A3. The first analytical peak displacement is 0.235
in (6.0 mm), which occurs at 0.072 s. This analytical
peak displacement is only about 3% smaller than the
experimental value. Analytical results show a second peak
of 0.250 in (6.4 mm) at 0.16 s, which overestimates the
corresponding experimental peak displacement. Analytical
results also show a permanent vertical displacement of about
0.206 in (5.2 mm), which is about 15% smaller than the
corresponding experimental value of 0.242 in (6.1 mm).
Fig. 9 compares experimental and analytical vertical
displacements of joint A2 in the second floor. While the
first peak vertical displacements obtained experimentally and
analytically are in good agreement, the analytical permanent
displacement again underestimates the experimental value.
The analytically estimated deformed shapes of the structure
at the maximum vertical displacement are shown in Fig. 11

with a magnification factor of 200. The beam cracked section


is also shown. The experimentally measured deformed shape
over the end regions of beams A1A2 and A3B3 in the second
floor are also shown (as described before). As can be seen in
Fig. 11, analytically estimated deformed shapes of the beams
are in close proximity with experimentally obtained deformed
shapes. Unlike the results obtained from FEM based Model A,
analytical results show no yielding of reinforcing bars.
Fig. 12 shows the analytical vertical displacement histories
of joint A3 in different floors above the removed column. As
displayed in the figure, nodes above the removed columns
experience approximately the same relative end displacements,
with nodes in the floors above having slightly smaller
displacement.
Following the removal of columns A2 and A3 in the first
floor (Fig. 3), gravity loads transferred through these columns
needed to be dynamically redistributed to the adjacent columns.
Table 1 shows axial compressive forces in the first floor
columns before and after the column removal, as well as the
change in the axial forces. As can be seen, column B3 had
the largest increase in its axial force followed by columns A1
and B2, respectively. Note that the sum of the increase in axial
forces of these columns is more than the sum of axial forces
of the removed columns A2 and A3. This is due to the fact
that axial forces of columns in Axes C and D (see Fig. 3) were
reduced.
Fig. 13 shows the variation in axial forces of column A3 in
different floors above the removed corner column. The figure
shows that at the end of vibration, the permanent axial forces in
the columns (other than the third floor) are approximately the
same, with tensile forces varying from 7.7 to 10.7 kips (34.2 to
47.6 kN). The permanent axial tensile force in the third floor
column A3 is about 2.2 kips (9.8 kN). Note that among the
floors above the first floor, only the third floor does not have
infill walls. Variation of the axial forces in column A3 in the
floors above the removed column within the first 0.01 s (10
milliseconds) is included in Fig. 13. As can be seen, the axial
force in the second floor column drops to zero after only about
0.002 s. It takes about 0.003 s until the movement starts to affect
the axial force of the sixth floor column.
Following column removal in the first floor, sudden
unbalanced forces equal to the axial compressive force in the
removed columns formed at joint A2 and A3 in the second
floor. As a result, the joints started to move downwards, and the
second floor columns above the removed columns elongated.
This in turn led to a reduction of compressive forces in second
floor columns. This phenomenon is pointed out by Sasani and
Kropelnicki [14], Sasani [15] and Sasani et al. [11].
Fig. 14 shows axial force variation in columns A2 and A3 in
the second floor and in adjacent columns A1 and B3 in the first
floor. As shown, while the drop in axial force in columns A2
and A3 in the second floor takes only about a few milliseconds,
it takes more than 0.05 s (50 milliseconds) for the load to
be transferred to neighboring columns A1 and A3. Analytical
data for these columns are compared with experimental results
below.

2485

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 11. Analytical (AEM) deformed shapes of structure (Second floor experimentally estimated deformed shapes are also shown).

Fig. 12. Vertical displacement histories of joints A3 in different floors.


Table 1
Analytical estimation of axial forces in first floor columns before and after (permanent) column removal
Column compressive force, kips (kN)
A1
A2

A3

B1

B2

B3

Before

73.0
(324.4)

84.4
(375.1)

66.1
(293.7)

94.0
(417.7)

112.7
(500.8)

95.6
(424.8)

After

128.2
(569.7)

n/a

n/a

94.3
(419.1)

148.9
(661.7)

187.9
(835.0)

Change

55.2
(245.3)

n/a

n/a

0.4
(1.8)

36.2
(160.9)

92.3
(410.2)

2486

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 13. Variation of axial forces in column A3 in different floors.

Fig. 14. Variation of axial forces in column A3 and A2 in second floor and in adjacent columns A1 and B3 in first floor.

Fig. 15. Experimental and analytical strain changes on East side of first floor column A1.

Figs. 15 and 16 compare analytical and experimental strain


changes in first floor columns A1 and B3. Strain gages were
located 6 in (152 mm) below second floor beams on the South
and East faces of columns B3 and A1, respectively. Overall,
Figs. 15 and 16 show a good agreement between experimental
and analytical results.
Fig. 17 compares experimental and analytical strain changes
in the bottom reinforcement of second floor beam A3B3,
close to the face of column A3. While the strain at the end is
predicted rather reasonably, analytical results do not capture
the recorded data in the first 0.3-0.4 s following column

removal. Such rapid variations in recorded steel strain were


not observed in the potentiometer recording that measured
the longitudinal deformation of the bottom of the beam over
21 in (533 mm) from the face of column A3. Therefore, it
is concluded that variations of recorded strains were a local
effect (local vibration), in part due to the removal of concrete
from the bottom side of the rebar. Note that the permanent
experimental and analytical tensile strains indicate a change in
the direction of bending moment in the beam, in the region
close to the removed column, which will be described in the
next section.

M. Sasani / Engineering Structures 30 (2008) 24782491

2487

Fig. 16. Experimental and analytical strain changes on South side of first floor column B3.

Fig. 17. Experimental and analytical strain changes on bottom reinforcement of second floor beam A3B3 at face of joint A3.

Fig. 18. Experimental and analytical strains at top of third floor beam A1A2 at face of joint A2.

Fig. 18 compares experimental and analytical concrete


strain changes at the top of third floor beam A1A2 close to
the face of column A2. The figure displays good agreement
between experimental and analytical results. The figure further
establishes the change in the direction of bending moments in
beams, which will be described in the next section.
6. Load redistribution and resisting system
Fig. 19 shows maximum bending moment diagrams of
spandrel frames in transverse and longitudinal directions
following column removal. The figure demonstrates a

development of bi-directional Vierendeel (frame) action in this


structure. Vierendeel action can be characterized by relative
vertical displacement between beam ends and corresponding
double curvature deformations of beams. In this structure, such
a deformation pattern is developed because of the existence of
moment connections and the interaction between beams with
columns and infill walls. The double curvature deformed shape
of beams provides the shear force needed to redistribute the
loads, following the column removal.
As can be seen in Fig. 19(a), bending moments in beams
A1A2, at the face of joints A2 (above the removed column),

2488

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 19. Maximum beam bending moment diagrams of spandrel frames following column removals in (a) Transverse direction; and (b) Longitudinal direction.

are positive (tension at the bottom), following column removal.


Similarly, bending moments in beams A3B3, at the face
of joints A3 in different floors are positive (Fig. 19(b)). At
these locations, prior to the removal of the columns, bending
moments were negative (tension at the top). For instance,
Fig. 20 shows the bending moment in beam A1A2 in the
second floor before and after column removal. The direction of
bending moments of these beams in the vicinity of the removed
columns changed after column removal.
The change of bending moment direction may cause
unexpected failure. For instance, if the beam bottom
reinforcement is not well anchored to the joint above a removed
column, the reinforcement will be susceptible to pull out,
and brittle local failure can develop. Because no structural
drawings were available, it is not clear if the beam bottom bars

were properly anchored or not. The small values of recorded


local deformations and strains at these critical locations in the
second and third floors suggest that no bottom bar pull out
had occurred. In a previous study where the beam bottom
rebars were not properly anchored into the joints, no bar pull
out was observed [11], which was in part due to the inherent
redundancy of the RC structure, its system level response, and
three dimensional load redistribution.
The shapes of bending moment diagrams in beams depend
on the amount of gravity load, the relative displacement
between the two ends of the beams, and the effects of the
infill walls. Although modeling the infill walls with struts was
shown not to be appropriate for this structure, strut action can
be schematically and qualitatively used to explain the variation
of bending moments in the beams. Fig. 21 schematically shows

M. Sasani / Engineering Structures 30 (2008) 24782491

2489

original applied element model. The weight of infill walls is


also excluded in the model. Nonlinear dynamic analysis is
performed under removal of columns A2 and A3 in the first
floor. The maximum vertical displacement of the structure is
increased by about 2.4 times (compared to the response of the
model with infills), yet the structure resists progressive collapse.
A maximum displacement of 0.6 in (15.2 mm) occurs at joint
A3 in the second floor at about 0.1 s.
Because of the lack of top reinforcement in the transverse
beam A1A2 in the vicinity of column A1 (see Fig. 4), all
beams except for roof level beams, crack and lose almost all of
their flexural strength at such sections. In spite of the formation
of cracks in beams, beam bending moments in floors 26 are
developed such that the beam is deformed in double curvature,
providing resistance against progressive collapse.
9. Conclusions
Fig. 20. Bending moment in beam A1A2 in second floor before and after
column removal.

the components of bending moments in the longitudinal second


floor beam. After column removal, the beam end at the top
of the removed column experienced vertical displacement.
Fig. 21(b) shows the bending moment diagram of the beam
due to distributed gravity load. Fig. 21(c) shows the moment
diagram of the beam, subjected to a relative vertical end
displacement. In the absence of infill walls, the superposition
of bending moment diagrams in Fig. 21(b) and (c) determines
the beam bending moment diagram. The effect of the infill
wall is schematically shown in Fig. 21(d). In reality, the infill
wall has a distributed effect on the beam, particularly under
small displacements. A simplified description of the infill wall
effect can be presented through the use of compressive struts
with a concentrated load. The bending moment diagram of the
beam is the sum of the three bending moments. Note that this
description is valid if the beam bending moment does not reach
the yielding moment, which in fact is the case for the beam
described above based on the results obtained from AEM.
7. Response of structure in presence of aditional dead and
live loads
As described before, all partitions, ceilings and mechanical
systems were removed before column removal. Also no live
loads existed. To investigate the effects of these additional
gravity loads on the building response, a unit floor load of 30
psf (1.44 kN/m2 ) is added to the applied element model to
account for partitions, ceiling, and mechanical systems. Also
an additional live load of 12.5 psf (0.60 kN/m2 ) equal to 25%
design live load is considered [1]. Analytical results show that
the maximum vertical displacement joint A3 in the second floor
is increased by about 32% to 0.330 in (8.4 mm) with permanent
displacement of 0.285 in (7.2 mm), which shows an increase of
about 38%.
8. Response of structure without infill walls
To investigate the effect of infill walls on building response,
another model is developed by removing infill walls from the

Bi-directional Vierendeel (frame) action of transverse and


longitudinal frames with the participation of infill walls is
identified as the major mechanism for redistribution of loads
in this structure. That is, because of moment connections,
columns caused beams to deform in double curvature, which
in turn developed sufficient beam shear forces to redistribute
gravity loads. As a result, the direction of bending moment
in these beams changed in the vicinity of the joint above the
removed column. Furthermore, infill walls provided the beams
with constraints and supports to help carry additional loads.
The maximum measured vertical displacement of the structure
after removal of two adjacent columns was only about 0.25 in
(6.4 mm); directly above the removed columns.
When Vierendeel (frame) action develops, in order for a
beam bottom reinforcement to be fully effective and to develop
stresses up to the ultimate stress at the face of a removed
column, the reinforcement should be well anchored into the
joint. In this structure, it was not clear if the beam bottom
reinforcement was properly anchored. The recorded maximum
tensile strain of the bottom reinforcement of the second floor
longitudinal beam at the face of the joint above the removed
column was small. Such a small strain suggests that the bar
would not have pulled out, even if it did not have proper
anchorage.
The mechanism for propagation of deformation over the
height of the structure is analytically described. It is shown that
soon after (about 1/20th of the time required for the structure
to reach the peak displacement) column removal, axial forces
in the columns above the removed columns reduce to almost
zero and then different floors practically move together. This is
due to the difference between the speed of the propagation of
the axial waves and that of flexural waves; the latter requires
the system to deform significantly more than the deformation
associated with axial deformation of columns.
Analytical results show that the joints above a removed
column in two different floors move almost identically, with the
floor above having slightly smaller displacement. The smaller
displacement is due to the loss of axial force in the column
connecting the two floors and its corresponding elongation.

2490

M. Sasani / Engineering Structures 30 (2008) 24782491

Fig. 21. Bending moment of a beam under different types of loading.

Modeling the infill walls with struts based on FEMA


356 [10] results in a maximum vertical displacement at
the top of the removed column equal to about 0.48 in.
This displacement is approximately 1.8 times the value
obtained when the infills are modeled by shell elements,
and cracking of the elements are accounted for. For small
deflections, struts do not realistically model beam and column
constraints, overestimating vertical displacement. If the vertical
displacement of the structure were large, so that significant
cracking had formed in the infill walls, modeling the infills with
struts might have predicted the structural deformation more
closely.
The response of the structure under additional dead loads
from partitions and mechanical systems as well as 25% of
design live load is analytically estimated. The additional gravity
loads result in about 32% and 38% increases in the maximum
and permanent displacement of the structure, respectively.
In order to further evaluate the effect of the infill walls on
the response of the system, the structure is analyzed without
the walls. The results show that while the maximum vertical
displacement of the structure is increased by almost 2.4 times,
the system still resists progressive collapse. This occurs in spite
of formation of brittle flexural failure (i.e. cracking in beam
sections without top reinforcement) in several beams of the
exterior transverse frame.
Acknowledgements
The author would like to thank Catherine K. Lee, Bela I.
Palfalvi, and Mario Ramirez (General Services Administration)
for funding of the study presented in this paper through the
GSA contract No. GS09P06KTM0019. The author greatly
appreciates the support provided by Mark Loizeaux (Controlled

Demolition Inc); without his help this study could not have
been completed. The help provided by Serkan Sagiroglu and
Marlon Bazan in the experimental and analytical studies is
acknowledged. The help provided by the contractors Clauss
Construction (Patrick M. Clauss, Malcolm Lee, and William
Musbach) and Jacobs (Bill Zondorak) is also appreciated.
References
[1] GSA. Progressive collapse analysis and design guidelines for new federal
office buildings and major modernization projects. Washington (DC): US
General Service Administration; 2003.
[2] DOD. Design of building to resist progressive collapse. Unified Facility
Criteria, UFC 4-023-03. Washington (DC): US Department of Defense;
2005.
[3] ASCE/SEI 7. Minimum design loads for buildings and other structures.
Reston (VA): Structural Engineering Institute, American Society of Civil
Engineers; 2005.
[4] Ellingwood B, Leyendecker EV. Approaches for design against
progressive collapse. Journal of the Structural Division ASCE 1978;
104(3):41323.
[5] ACI 318. Building code requirement for structural concrete. MI:
American Concrete Institute; 2005.
[6] Sozen MA, Thornton CH, Corley WG, Mlakar PF. The Oklahoma city
bombing: Structure and mechanisms of the murrah building. Journal of
Performance of Constructed Facilities, ASCE 1998;12(3):12036.
[7] Corley WG. Lesson learned on improving resistance of buildings to
terrorist attacks. Journal of Performance of Constructed Facilities ASCE
2004;18(2):6878.
[8] SAP2000, Three dimensional static and dynamic finite element analysis
and design of structures, Analysis Reference, version 9.2. Berkeley (CA):
Computer and Structures, Inc.; 2005.
[9] MacLeod IA. Analytical modeling of structural systems An entirely
new approach with emphasis on behavior of building structures. Ellis

M. Sasani / Engineering Structures 30 (2008) 24782491


horwood series in civil engineering. England. 1990.
[10] FEMA 356. Prestandard and commentary for the seismic rehabilitation of
buildings. Washington (DC): Federal Emergency Management Agency;
2000.
[11] Sasani M, Bazan M, Sagiroglu S. Experimental and analytical progressive
collapse evaluation of an actual RC structure. Structural Journal, ACI
2007;104(6):7319.
[12] Meguro K, Tagel-Din HS. Applied element method used for large
displacement structural analysis. Journal of Natural Disaster Science

2491

2002;24(1):2534.
[13] ELS, Extreme loading for structures technical manual. Raleigh (NC):
Applied Science International, LLC; 2006.
[14] Sasani M, Kropelnicki J. Progressive Collapse Analysis of an
RC Structure. The Structural Design of Tall and Special Buildings, John Wiley & Sons; 2007, in press [doi:10.1002/tal.375].
http://www3.interscience.wiley.com/cgi-bin/jhome/102522232.
[15] Sasani M, Sagiroglu S. Progressive collapse resistance of hotel San Diego.
Journal of Structural Engineering, ASCE 2008;134(3):47488.

Potrebbero piacerti anche