Sunteți pe pagina 1din 26

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2009; 38:609–634


Published online 13 February 2009 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/eqe.898

Modeling progressive collapse in reinforced concrete buildings


using direct element removal

Mohamed Talaat1, ‡ and Khalid M. Mosalam2, ∗, †, §


1 Simpson Gumpertz and Heger Inc., San Francisco, U.S.A.
2 University of California, 721 Davis Hall, Berkeley, CA 94720-1710, U.S.A.

SUMMARY
This paper presents a novel analytical formulation of an element removal algorithm based on dynamic
equilibrium and the resulting transient change in system kinematics, by applying imposed accelerations
instead of external forces at a node where an element was once connected. The algorithm is implemented
into an open-source finite element code, numerically tested using a benchmark structural system with
simplified element removal criteria, and able to capture the effect of uncertainty in member capacity.
Realistic element removal criteria are introduced for mode-dependent gravity load collapse of seismically
deficient and retrofitted reinforced concrete (RC) columns and unreinforced masonry (URM) infill walls.
Two applications are conducted using structural systems of RC frames with URM infill walls. The first is
a probabilistic study of a one-story model subjected to an ensemble of 14 ground motion recordings from
similar neighboring sites during an earthquake event. The study produces empirical probability curves for
partial and complete collapse conditioned on different hazard levels, and concludes that the intra-event
variability is a major source of uncertainty affecting the outcome of progressive collapse simulations.
The second application is a deterministic sensitivity study of progressive collapse response in a five-story
structural model to uncertainty in live load, stiffness, damping, and seismic hazard level, subjected to
one ground motion record. The analysis identifies the time at incipient collapse as an adequate sensitivity
measure, and the uncertainty in ground motion intensity as the most important, followed by the stiffness
of the URM infill wall. Copyright q 2009 John Wiley & Sons, Ltd.

Received 11 July 2008; Revised 1 November 2008; Accepted 5 January 2009

KEY WORDS: dynamic analysis; element removal; finite element; progressive collapse; reinforced
concrete frame; unreinforced masonry infill

∗ Correspondence to: Khalid M. Mosalam, University of California, 721 Davis Hall, Berkeley, CA 94720-1710, U.S.A.

E-mail: mosalam@ce.berkeley.edu

Senior Engineer.
§ Professor and Vice Chair.

Contract/grant sponsor: Earthquake Engineering Research Centers Program; contract/grant number: EEC-9701568

Copyright q 2009 John Wiley & Sons, Ltd.


610 M. TALAAT AND K. M. MOSALAM

INTRODUCTION

Progressive collapse assessment using nonlinear time-history finite element (FE) simulation is
gaining popularity over traditional methods such as alternate path analysis. Applications of tradi-
tional methods include decommissioned [1] and existing [2, 3] reinforced concrete (RC) buildings,
retrofitted steel frames [4], retrofitted buildings [5], multi-span bridges [6], and probabilistic eval-
uation methods [7]. There is limited literature on this field of research as few experimental inves-
tigations have been conducted on RC frames that are redundant enough to experience progressive
collapse, e.g. [8–10]. Experimental data from these studies are intended to establish the behavior
of RC systems after the loss of one or more deficient columns designed to lose axial load capacity.
However, observed collapse modes are difficult to generalize to cases involving disconnection and
subsequent collision by collapsed elements.
A recent study reported in [11] describes an analytical approach to use post-yield strength and
stiffness degradation to conduct quasi-static progressive failure analysis. Another study reported
in [12] defines a macro-level damage index to predict collapse of yielding beam–column elements
based on maximum element deformations and accumulated plastic energy. Collapsed elements are
removed from the structural system, and external nodal forces are applied at the removed element’s
end-nodes using a step function for the duration of the subsequent load step. This approach is
valid for quasi-static behavior but is sensitive to the time step size during dynamic simulations and
may not be representative of the energy imparted into the damaged structure due to the release of
internal forces. In the approach presented in [13], the downward motion of the collapsed element
is tracked using a condensation method that avoids redefining degrees of freedom (DOFs) and
splitting one node into two. This approach is extended in [12] to include a simplified account for
impact on the structure by a collapsed element.
The effect of locally released internal energy upon element collapse on exciting transient vibra-
tions in the damaged structure was presented in [14] using an energy balance approach. The study
concludes that a quasi-static equilibrium-based progressive failure analysis is not conservative,
and confirms this with both experiments and computations on a simple spoked wheel structure.
Another study on a multiple degree-of-freedom (MDOF) truss system [15] demonstrated similar
conclusions. An analytical formulation was developed in [16] which uses an energy balance method
to amplify the displacement demands on a structural system following the removal of a gravity
load-carrying column. It conservatively ignores the energy dissipated by viscous damping in the
system due to the short duration involved, and characterizes the kinetics of the damaged system
at the onset of column collapse using nodal displacements only.
Two studies in [17] and [18] include the effect of uncertainty on simulated progressive collapse
assessment of MDOF building systems. They assume ductile details and adopt for the definition
of collapse a side sway mechanism of columns within an entire floor, not explicitly conducting
any removal of elements during the simulation.
In order to capture the progression of collapse, one must model the behavior of individual
components under extreme loading and identify the criteria for their removal from the structural
model. In this paper, two modes of RC column collapse, namely shear-axial and flexure-axial,
are investigated. Shear-axial collapse was recently investigated in a number of experimental and
analytical studies reviewed in [10]. The shear friction-based model developed in [19] is utilized in
this paper to establish limit-states for shear-axial collapse. The damage indices developed in [20]
for fiber-discretized cross-sections are used for flexure-axial collapse. These indices are material
model dependent and reflect the damage accumulation at individual fibers. For non-ductile and

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 611

retrofitted—using fiber-reinforced polymers (FRP)—column behavior, new models are adopted


from [21] for confined fiber-discretized cross-section and confinement-sensitive concrete material
behavior, and from [20] for confinement-sensitive bar buckling and lap splice behavior. The
computational implementation in this study was carried out in OpenSees [22].
Two test-bed structural systems are presented as applications. Both systems consist of RC
moment frames with seismically deficient columns and infilled with unreinforced masonry (URM)
wall(s) in one or more bays. Application ‘A’ conducts a probabilistic assessment on a one-story
building and application ‘B’ conducts a sensitivity analysis on a five-story building.

DIRECT ELEMENT REMOVAL

Mechanics of element removal


Consider the dynamic equilibrium of the node illustrated in Figure 1(a). At time t, beams B1
and B2 and columns C1 and C2 are attached to the node. For clarity, the free-body diagrams are
presented without externally applied or damping-induced nodal forces. In Figure 1, F and M refer,
respectively, to resisting forces (both shear and axial) and bending moments from the attached
elements. The node has a lumped translational mass m and a rotational mass moment of inertia I .
Dynamic equilibrium is satisfied by the nodal inertia forces max , ma y , and I , where ax and a y
are translational accelerations in the respective x and y directions and  is rotational acceleration
in the x–y plane.
At time t, column element C1 reaches its collapse limit-state, and is removed from the compu-
tational model at time t  = t +t, where t is the prescribed time step. The resulting free-body
diagram is illustrated in Figure 1(b), where the nodal masses and accelerations are updated to m  ,
I  , ax , a y , and  , respectively. In practical applications, the majority of the lumped nodal masses
are derived from the floor system, and the updated masses may not be significantly different from

MC2,FC2 M'C2 F'C2

Mass = m Mass = m'


Time = t Time = t' = t+ t

MB1,FB1 MB2,FB2 M'B1,F'B1 M'B2,F'B2


I I' '

max m'a'x X
X
may m'a'y

a'x >> ax Removed element


a'y >> ay releases internal forces
' >>
(a) MC1,FC1 (b)

Figure 1. Dynamic equilibrium of a node connected to a collapsed element: (a) before element
removal and (b) after element removal.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
612 M. TALAAT AND K. M. MOSALAM

the original ones. The dynamic equilibrium of the node under resisting and inertia forces can be
expressed by
⎧ ⎫ ⎧ ⎫ ⎧ ⎫ ⎧  ⎫
⎪ Fx ⎪ ⎪ max ⎪ ⎪
⎪ F ⎪ ⎪ m ax ⎪
⎨ ⎬ ⎨ ⎬  ⎨ x⎪ ⎬ ⎪
⎨ ⎪

   
Pex + Fy − ma y = Pex + Fy − m a y (1)

⎩ ⎪ ⎭ ⎪⎩ ⎪
⎭ ⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪

I ⎩  ⎭ ⎩   ⎭
M M I

where Pex is the vector of externally applied nodal loads, if any, assumed independent of the
element connectivity and not affected by the removal of a collapsed element. The summation is
conducted for the elements connected to the node. Upon abrupt element collapse and removal, this
dynamic equilibrium is disturbed by the sudden release of internal forces from element C1 and must
be restored before the solution continues. The resisting forces in the connected elements, namely
F  and M  , can only change as a result of updated element deformations, which require changes in
relative displacements (and possibly velocities) between the elements’ end-nodes. These changes
result from updated nodal velocities and accelerations and do not take place instantaneously. The
inertia forces are directly related to nodal accelerations ax , a y , and  , which need to change abruptly
to satisfy dynamic equilibrium at time t  . Since the connected elements’ end-nodes other than their
shared node are under dynamic equilibrium themselves, this abrupt change is first localized at the
shared node at time t  . In the following time steps, element removal superposes a case of transient
loading on the damaged system, whereby the resulting changes in nodal kinetics lead to an updated
set of resisting (and inertia) forces that propagate outward through the connected elements into
neighboring nodes. These updated nodal forces must satisfy dynamic equilibrium of their respective
DOFs during every time step, which results in updated inertia forces and nodal accelerations. This
process continues to propagate in the damaged system through element connectivity until updated
nodal configurations corresponding to a new equilibrium state are reached.
Upon reaching the new equilibrium state, the nodal velocities of the damaged system result in
overshooting the corresponding displacements, amplifying deformation demand in the structural
elements. In the hypothetical absence of external excitation, free vibration of the system ensues.
An initial transient phase is followed by a steady-state phase oscillating about the new equilibrium
state. The new deformation demands may lead to the collapse of additional elements, which would
be ‘safe’ in the new equilibrium state using quasi-static analysis. If so, a second transient phase is
excited, and this process continues until the system either (a) reaches an equilibrium state about
which it can safely oscillate and the progression of collapse is arrested; (b) experiences overall
gravity load collapse; or (c) experiences partial collapse that is nevertheless compartmentalized
while the rest of the structural system survives. In case (c), the structure will probably need to
be ultimately demolished, yet the prospect of maintaining life safety is enhanced compared with
case (b).
In the presence of an earthquake excitation, the input ground motion is treated as equivalent
external loads. The input time step may be larger than that required for capturing transient effects,
particularly if reaching a new equilibrium state requires local axial vibrations in stiff structural
elements with significantly short natural periods. Hence, shorter time steps are needed immediately
after element removal, during which the input ground motion is interpolated. An adaptive time
stepping scheme can be used for computational efficiency.
The advantages of explicitly removing elements instead of assigning them low stiffness are
threefold. First, it avoids numerical problems associated with ill-conditioned stiffness matrices.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 613

Second, enforcing dynamic equilibrium enables the computation of the resulting increase in nodal
accelerations and the inclusion of the system’s complete kinematic state at the time of element
collapse in determining whether it can successfully survive to a new equilibrium state. Third, the
motion of the collapsed element can be tracked relative to the damaged system to estimate the
time and kinetics at subsequent collision, e.g. using rigid body analysis of the collapsed element
motion or a condensation approach [13] where DOFs redefinition in the damaged system is not
required. If collision is detected, a ‘soft-impact’ analysis can characterize the duration, forces,
and redistribution of masses [23]. This approach assumes negligible relative displacement during
impact between the collapsed element and the damaged system at the impact point and low-velocity
impact that precludes punching and penetration [24].

Implementation of element removal algorithm


An algorithm is designed and implemented for automated removal of collapsed elements during
an ongoing simulation, Figure 2. The implementation is carried out as a new OpenSees module,
designed so that it is called by the main analysis module after each converged load step, to check
each element for possible violation of its respective removal criteria, presented later in this paper.
A violation of any criterion triggers the activation of the element removal algorithm on the violating
element before returning to the main analysis module. The algorithm is split into two main sets
of procedures. The first set has been fully implemented within the source-code of OpenSees.
It includes updating nodal masses, checking if the removal of the collapsed element results in
leaving behind dangling nodes or floating elements, Figure 3, which must be removed as well,
and removing all associated element and nodal forces, imposed displacements, and constraints.
The second set involves the tracking of the removed element motion for possible collision. The
kinetics of the separated end-node(s) at the time of separation are identified from the last converged
state of the system, representing the initial conditions for the free falling or rotating motion of
the collapsed element. This motion is tracked, assuming rigid body motion of the element under
effects of its own weight, and compared with the position of the damaged system to determine if
and when collision takes place. At the time of collision, the relative impact velocity is computed
and used to determine the impact duration and the magnitude and temporal variation of impact
force to be superposed on the damaged system during the subsequent time steps. This second set
of procedures is performed in the present study by conducting the necessary computations using
MATLAB [25], and finally inputting the results to OpenSees.
After completing both sets of procedures, the algorithm updates the state of the system to the new
masses, geometry, and impact forces. The assumption of Rayleigh viscous damping (mass and/or
stiffness proportional) is implicitly affected. Since numerical convergence may face difficulties
following such extreme events, especially in softening structural systems, the solution parameters
may need to be updated. This is conducted by iteratively switching solver type, convergence
criteria, and other analysis options, e.g. time step, so that an ultimate failure to converge would
likely take place only if the damaged structural system experiences global instability. An OpenSees
input script for adaptive time stepping and iterative solver modification is developed and used in
the applications presented herein.

Benchmark demonstration problem setup and results


The above algorithm implementation is numerically tested using a benchmark problem of an
idealized structural system consisting of two trussed cantilever beam canopies, Figure 4, with

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
614 M. TALAAT AND K. M. MOSALAM

Check for dangling nodes, • Remove dangling nodes


floating elements, and • Remove floating elements
element loads and masses • Delete element/node loads
• Remove element
• Update nodal masses
Start
from
Identify nodal kinetics Update structural model, time
main
at separation step, and solution parameters
code
End
E nd
Italicized text back to
executed outside Track collapsed element Identify location, compute force, main
of OpenSees motion until impact duration, and mass redistribution code

Figure 2. Automatic element removal algorithm.

Intact structure
Nodal loads

Dangling
node
Floating
element

Dashes represent
collapsed elements

Figure 3. Elements, nodes, and loads requiring removal due to element removal.

listed element properties (elastic modulus, E, cross-sectional area, A, and moment of inertia, I )
in Table I. The (inclined) truss members exhibit a linear elastic behavior up to an axial load Fy ,
then exhibit either brittle or perfectly plastic behavior up to a ductility limit, Table II, before
removal. Four case studies are used to explore the sensitivity of the algorithm to uncertainty in
element capacities. The (horizontal) beam elements are linear elastic. Nonlinear geometry and large
deformation effects are included in the FE model by using co-rotational geometric transformation
from element quantities, e.g. orientation, to global coordinates [26]. The mass is lumped at the
nodes, Table III. The gravity loads are applied followed by support excitation at nodes 2–5 and 12

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 615

Figure 4. Geometry (in meters) and FE model of benchmark problem.

Table I. Element properties of benchmark problem.

Element EA (kN) E I (kN m2 ) Fy (kN)


1 2000 3000 100
2 100 000 5000 ∞ (Elastic)
3 1000 1000 50
4–9 100 000 10 000 ∞ (Elastic)
10,11 3000 5000 150

Table II. Ultimate truss element ductility limits for benchmark problem.
Truss element Case 1 Case 2 Case 3 Case 4
1, 11 5 5 5 5
3 5 5 1 1
10 5 1 5 1

Table III. Lumped masses at unrestrained nodes of benchmark problem.


Node 1 8 6, 7, 9–11
Mass (ton) 10 6 4


in the vertical direction using the earthquake record from Northridge 1994, Tarzana station, 90
component, with scaled peak ground acceleration (PGA) of 1.04 g. For the analysis objective, a
time step of 0.005 s is chosen with no viscous damping.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
616 M. TALAAT AND K. M. MOSALAM

It is assumed that collapsed truss elements separate from the structural system at the location
of their connection with the beams. Hence, the collapse of element 3 causes it to fall and to
collide with the lower canopy, slightly to the right of node 9. For simplicity of the analysis,
collision is assumed to take place exactly at node 9. After collision, element 3 is assumed to
deform in flexure, simply supported at nodes 4 and 9. The energy of the impact is assumed
to dissipate during one half-cycle of such flexural deformation before rapidly decaying, e.g.
due to friction, and element 3 remains passively supported by nodes 4 and 9 until collapse of
the lower canopy. In addition to impact forces, it is assumed that collision with (or loss of) a
truss element at a node results in updating its assigned mass and gravity loads by adding (or
subtracting) 2 ton.
All simulations were successfully conducted without exhibiting numerical convergence prob-
lems. Cases 1 and 2 result in similar responses where no element collapse takes place. In both
cases, the upper canopy elements 1 and 3 exhibit brief yielding, inducing a ductility demand of
1.17 and 1.11, respectively. Cases 3 and 4 result in collapse of element 3 onto the lower canopy
at 8.875 s. As a result, the ductility demand on element 1 increases to only 2.1; and the upper
canopy exhibits no further collapse. Figure 5 illustrates the time history of the normalized axial
forces using their respective values at yield in truss elements 1 and 3 for all four cases. The
response of element 3 is identical for all cases prior to reaching its ultimate ductility limit and
collapsing in Cases 3 and 4. The response of element 1 is identical for all cases prior to 8.875 s.
Subsequent response exhibits higher average force values in Cases 3 and 4 and a longer period of
yielding before the damaged upper canopy starts to oscillate about a new equilibrium state. The
collapse of element 3 results in larger permanent displacement in the vertical direction at node 1,
Figure 6(a), where the nodal displacements are normalized using their respective values caused by
the application of the gravity load.
In the lower canopy, element 3 collides with node 9 at 11.26 s, and the resulting demand causes
yielding in elements 10 and 11. In Case 3, the ductility capacity in both elements is sufficient
to prevent the progression of collapse in the system, which reaches a new equilibrium state and
starts oscillating about it, Figure 6(b). In Case 4, brittle element 10 collapses at 12.37 s. Shortly
afterwards, element 11 reaches its ultimate ductility limit and collapses at 14.57 s, leading to
complete collapse of the lower canopy, Figure 6(b). The deformed shapes at 14.5 s and collapse

1.5
Element 1 Time = 8.875
1
Normalized Force

0.5
Cases 1&2
0 Element 3 Collapse time for brittle
(4 cases overlap element 3 (Cases 3&4) Cases 3&4
-0.5 up to 8.875 sec)

-1

-1.5
0 5 10 15 20 25
Time [sec]

Figure 5. Simulated response of elements 1 and 3 of benchmark problem.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 617

-0.5 -0.5
Time = 8.875 Time = 11.260

Normalized Displacement

Normalized Displacement
-1 -1

-1.5 -1.5
Cases1&2
Cases 1&2
Cases3&4
Cases 3&4 Cases1&2
Cases 1&2
-2 -2
Case33
Case

Case44
Case
-2.5 -2.5
0 5 10 15 20 25 0 5 10 15 20 25
(a) Time [sec] (b) Time [sec]

Figure 6. Simulated nodal displacement time-histories for benchmark


problem: (a) node 1 and (b) node 11.

(a) (b) (c)

Figure 7. Snapshots of benchmark structural model deformation at Time= 14.5 s: (a) cases 1
and 2; (b) case 3 and (c) case 4.

modes, if any, are illustrated in Figure 7 (no displacement magnification) with deformed beam
elements simplified by straight lines connecting the end-nodes. Figure 8 compares the maximum
ductility demands imposed on each truss element during simulations and lends insight to the
sensitivity of the outcome and extent of system collapse to its element properties. There is a
significant contrast between the small difference in ductility demands on members 3 and 10 and
the large difference in simulated progressive collapse outcome for each case. This highlights the
sensitivity of the simulation procedure, because of its inherent path dependency, to uncertainty
in the structural model parameters. Hence, there is an obvious need to conduct simulation-based
progressive collapse assessment within a probabilistic framework.

ELEMENT REMOVAL CRITERIA

Three element removal criteria are defined for two different modes of failure in seismically
deficient and retrofitted RC beam–columns and for truss members. These criteria are implemented
in OpenSees for force-based and displacement-based distributed plasticity fiber elements, lumped

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
618 M. TALAAT AND K. M. MOSALAM

5
Cases Case Case
4.5 1&2 3 4
4

Maximum Ductility Demand


Collapsed

3.5
3
2.5
2
Collapsed Collapsed
1.5
1
0.5
0
1 3 10 11
Element

Figure 8. Simulated maximum element ductility demands of benchmark problem.

plasticity beam–column elements with fiber-discretized plastic hinges, shear-axial coupled spring
elements [19] and truss elements.

Removal criteria for RC beam–columns due to flexure-axial interaction


An analytical confined RC cross-section model has been developed in [20]. This analytical model
iteratively computes the confining stress distribution within the fiber-discretized cross-section at
each solution step using deformation compatibility between the dilation strain and the confining
material, e.g. internal steel hoops or external FRP jacket for RC columns. The model accounts
for the effect of this compatibility on individual fibers, which are constituted using confinement-
sensitive uniaxial material models for reinforcing steel bars subject to buckling, lap-spliced steel
bars, and concrete. A set of material damage indices, 0Dfiber 1, is developed and calibrated
in [20] for fibers representing these materials. These damage indices are based, respectively, on
the accumulation of steel plastic strains following the Coffin–Manson fatigue model [27], the
maximum lap splice slip displacement, and the ratio of hysteretic energy dissipation to compressive
fracture energy in concrete. These material-level damage indices define two aggregated cross-
section damage indices, D A and D M , as expressed below, to reflect the loss of a cross-section’s
capacity to resist axial loads and bending moments, respectively,



D A = 1− Iconf 1− (Dfiber Afiber /Across−section ) (2)
fibers



D M = 1− Iconf 1− (Dfiber Afiber h 2fiber /Icross−section ) (3)
fibers

where A and I refer to transformed area and moment of inertia, and h to the distance between the
fiber’s center and the uncracked section centroid. Areas and inertias are transformed using the ratio
between the steel and concrete fibers’ initial stiffness moduli. Iconf is an indicator for the loss of
confinement whose value is 0.0 if the confining medium (e.g. transverse steel ties or FRP jacket)

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 619

fracture is detected and 1.0 otherwise. The proposed damage indices are hence suitable for use
with deficient and FRP-retrofitted RC columns. When D A reaches a threshold value, the associated
column cross-section is considered to have lost its axial load capacity and the element removal
algorithm is invoked on the associated element. The threshold value is defined as D A = 1 in this
paper. A future extension may explicitly reflect the axial load level in the column. An analogous
approach can be defined for D M , but is not of interest in this paper.

Removal criteria for RC beam–columns due to shear-axial interaction


These element removal criteria correspond to violating the drift capacity model developed in [19],
where shear-axial coupled springs are connected in series to a beam–column element. When the
lateral drift-axial load limit is reached, the corresponding spring is removed from the FE model.
Consequently, the connectivity between the associated beam–column element of an RC column
and the structural system is severed at one end-node. Similarly, when the shear limit curve is
reached in a beam–column element of an RC beam, the corresponding spring is removed. Hence,
a shear-damaged column is allowed to maintain its axial load while shedding shear force until it
reaches the axial load-lateral drift limit-state, while a shear-damaged beam is assumed to collapse
in a brittle manner in the absence of a large compressive force that may prevent shear failure from
rapid propagation.

Removal criterion for truss members


This element removal criterion corresponds to violating maximum (positive) or minimum (negative)
axial deformation values. When the axial deformation in the truss element reaches either threshold
value defined by the analyst, the element is removed from the structural model. The choice of
such values enables modeling both brittle and ductile behavior, as demonstrated in the benchmark
problem above.

APPLICATION ‘A’: ONE-STORY STRUCTURAL SYSTEM

The structural system is selected to benefit from existing experimental data from a similar physical
model subjected to shake-table ground motion records of increasing intensity at the University of
California, Berkeley, Figure 9(a). The analysis of results is transparent by considering a one-story
system and avoiding dynamic effects of higher vibration modes. The physical model represents
a substructure of the first story of a five-story prototype building. It consists of three seismically
detailed moment-resisting RC frames identical in geometry and reinforcement, placed parallel to
each other and connected by an RC slab and transverse RC beams. The middle RC frame is
infilled with a URM wall. Details of the experiment can be found in [28]. A corresponding FE
model, Figure 9(b), is calibrated and then modified to introduce seismically deficient reinforcement
details in the RC columns. The objective of application ‘A’ is to investigate the effect of intra-event
variability, i.e. uncertainty in the characteristics of ground motion due to site location and building
orientation, on the results of progressive collapse simulations using 14 ground motion records from
a specific seismic event. An empirical probability measure is defined to construct fragility-like
curves for different collapse limit-states.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
620 M. TALAAT AND K. M. MOSALAM

16 B7 14 B5 12

B3 C6
B2 B1
C2
S1
C4
S2
B6 B4
15 11
13

N C5 C3 C1

(a) (b)

Figure 9. Shake-table test structure: (a) physical model in [28] and (b) computational model.

Table IV. Geometry and details of shake-table test structure [28].


Structural Dimensions Transverse reinforcement
member (mm) Main reinforcement (closed stirrups)
Slab 95 10-mm bars @305mm t & b∗ Not applicable
Eight 19-mm bars 10-mm@95mm in joint & 610 mm
Column 305×305 32-mm post-tensioned rod from joint face, then @152 mm
Longitudinal 10-mm@70 mm over 711 mm
beam 267×343 Three 19-mm bars t & b∗ from joint face, then @203 mm
Transverse beam 229×305 Two 19-mm bars t & b∗ 10-mm@305 mm
Footing 457×356 Four 22-mm bars t & b∗ 10-mm@102 mm
URM wall 102 Not applicable Not applicable
∗ Top and bottom.

Table V. Material properties of shake-table test structure [28].


Material Concrete Concrete Concrete Steel  y Steel u Steel E s Steel E sh
property f c (MPa) f t (MPa) cover (mm) (MPa) (MPa) (GPa) assumed
Value 31 3 15 459 600 210 1.0% E s

Description and modeling of test-bed structure


The dimensions used in the physical model represent a 75%-scale of typical dimensions. The
RC frames had 4115 mm span, 3219 mm height, and 1829 mm spacing. To account for the effect
of the upper stories, post-tensioned rods were used to apply axial forces of 290 and 145 kN on
each column of the infilled and bare RC frames, respectively. Lead weights totaling 320 kN were
uniformly distributed on the slab, which had 229 mm overhangs in each direction. The dimensions
and reinforcement are listed in Table IV. Concrete and steel properties are listed in Table V. The
102 mm thick masonry wall had average tested shear strength of 1.81 MPa. The shake-table test
used unidirectional ground motion records from Loma Prieta (1989), Northridge (1994) and Düzce
(1999) in the North (N) direction with PGA ranging from 0.27 to 1.93 g.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 621

σ 13 104 103 105


104 106 14

f mo 133 143
Cover concrete fiber
213 214
Parabola Core concrete fiber
Straight line S2
Steel tie
f mu Bending axis

ε mo ε mu ε Steel bar 142


Tension Compression Column cross-sections
Stress-strain model for truss elements 132
Node
Beam-column FE in plastic hinge zone
General beam-column FE 113 S1 114
131 141
Shear-axial coupled zero-length FE
30 40
Compression-only strut 130 140
3 Beam cross-sections 4
Bending moment-releasing connection

Figure 10. OpenSees FE mesh of the one-story test-bed structural system (infilled frame).

The element mesh of the URM-infilled frame C3-B2-C4 is illustrated in Figure 10. Nodes 3
and 4 are restrained from all translations and rotations. Nodes 30 and 40 coincide with nodes 3
and 4, respectively. The beam–column elements representing the column plastic hinge zones have a
length of 366 mm, based on L p = 0.077 L c +8.16 db , where L c is the length between the column’s
contra-flexure points and db is the longitudinal steel bar diameter [20], and two Gauss numerical
integration points. The cross-section is discretized using the confined concrete section and material
models from [21] with a grid of 10×10 fibers for the core concrete and 2 layers of fibers in each
direction for the concrete cover. The longitudinal steel is modeled using buckling-enabled steel bar
fibers [20] with bar mechanical stress–strain relationship according to [29]. The transverse rein-
forcement is modeled using a bilinear material model. A confinement efficiency factor of 0.75 is
estimated according to [30] for the square shape of the cross-section with four tie branches relative
to the circular shape. The beam–column elements in the middle parts of the columns have nine
Gauss numerical integration points and cross-section discretization identical to that of the plastic
hinge zones. The beam–column elements used to represent the longitudinal beam plastic hinge
zones have a length of 303 mm (using the formula for L p above), and two Gauss numerical integra-
tion points. Their cross-sections are discretized using a grid of 5×10 fibers for the core concrete (10
layers parallel to the bending axis) and one layer in each direction for the concrete cover. According
to recommendations in [31], a 1030-mm effective width of the RC slab is considered in resisting
compressive stresses and meshed using one layer of fibers. Experimental observations showed no
damage taking place in the beams or the middle parts of the RC columns, and hence the steel
bars in these cross-sections are modeled with bar buckling disabled for computational efficiency,
and a constant confining stress computed assuming yielding of the transverse ties is preset and
applied to the core concrete fibers. The URM wall is modeled using the two truss elements S1 and
S2, whose axial force-deformation behavior resembles a compression-only concrete-like material,
Figure 10. The parameters for this stress–strain curve were calibrated in [28] from experimen-
tally measured material properties and are reported as εmo = 0.0028, εmu = 0.0042, f mo = 17 MPa,
and f mu = 2 MPa, corresponding to a strut area of 223 mm2 . Elements 130 and 140 representing

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
622 M. TALAAT AND K. M. MOSALAM

shear-axial interaction of the infilled frame columns are constructed according to [19]. The meshing
of the bare frame members is similar, with the exclusion of the wall struts and the shear-axial
springs.
The transverse beams B4–B7 and the one-way RC slabs were observed to exhibit only minor
flexural cracking during the shake-table tests. Thus, their flexural stiffness is modeled using elastic
beam–column elements with a 1030-mm effective width of the slab contributing to the beam
cross-section and 0.7 times the uncracked concrete Young’s modulus of elasticity [31]. The in-
plane shear stiffness of the RC slab is observed to remain approximately constant throughout the
shake-table tests, and estimated as 406.0 kN/mm [28]. It is modeled using equivalent elastic truss
elements (Figure 9(b)) of stiffness 243.15 kN/mm each, assuming linear geometric transformation
in the plane of the RC slab.
The post-tensioning loads on the columns are applied as external nodal forces at nodes 11–16,
which would result in P- effects more similar to realistic structures. Masses and distributed
gravity loads are lumped at the nodes. The infilled frame is assumed to carry twice as much gravity
load as the bare frames. Numerical time integration uses the Newmark -method with constant
acceleration. Based on snap back test results in [28], a stiffness-proportional viscous damping
ratio  = 0.045 is estimated at the fundamental mode of vibration. The stiffness used to determine
the damping matrix during each analysis step is computed as the mean of the initial stiffness
and the tangent stiffness of the last converged analysis step, in order to avoid rapid numerical
reduction in damping caused by decreased and possibly momentary non-positive tangent stiffness
matrix due to structural damage. The FE model’s estimation of the fundamental vibration period is
0.121 s, compared with the experimentally recorded value of 0.135 s, and it reasonably reproduces
the force–displacement time histories of the test structure at levels of ground motion up to and
including collapse of the URM wall [20]. As an example of such reasonable match, Figure 11
shows the comparison of experimental and simulated lateral roof displacement at the middle frame
of the specimen for an input ground motion obtained from the Northridge (1994) earthquake
(Tarzana station in the 90◦ direction) and scaled to PGA = 0.61 g.
After calibration, the FE model is artificially weakened by introducing seismically deficient
details in individual RC columns where plastic hinges are expected to form. The transverse tie
spacing in all columns is increased to 150 mm, the maximum value allowed by the 1963 ACI
building code [32]. This increases the potential for bar buckling and shear failure, especially in
columns C3 and C4 due to the formation of short columns following URM wall cracking and
Roof Horizontal Displacement [mm]

8
Shake-table experiment

OpenSees simulation
4

-4

-8
0 2 4 6 8 10 12 14 16 18 20
Time [sec]

Figure 11. Simulated and experimental time-history results of shake-table test structure.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 623

partial collapse. Lap splices are inserted at the column-footing level of columns C5 and C6 in order
to observe the resulting difference in response and avoid simultaneous, non-progressive collapse
of the otherwise symmetric bare frames. The length of the lap splice is 20 times the bar diameter
[32] which is inadequate by current standards [31]. The diagonal strut elements are assumed to
collapse and are removed upon softening to 0.75 f mo , representing brittle URM wall collapse.
Preliminary analysis of the modified FE model identified columns C3 and C4 as likely to collapse
first, imposing large deformation demands on beams B4-B7, which redistribute their loads to the
corner columns. Excessive vertical displacements lead to crushing of concrete in compression and
loss of bending moment capacity, but the continuity of the longitudinal steel bars over the beam
length forms catenaries supporting the vertical load through the action of inclined axial tension
until the longitudinal steel fails due to fracture or pull out. An independent FE simulation was
conducted for the setup in the inset of Figure 12, showing a side view of the modified FE model
with C3 removed and a displacement-controlled downward motion imposed at node 13. Beams B4
and B6 are modeled using a total of 12 displacement-based beam–column elements with five Gauss
integration points each and cross-section discretization similar to that in Figure 10. The reinforcing
steel fracture strain (pull out failure is excluded in the present study) is assumed to be 0.13. The
simulated force-displacement relationship, Figure 12, is idealized by a tri-linear envelope defined
by points (0, 0), (20, 245), (45, 150), and (490, 450); and is used to construct four nonlinear
springs connecting nodes 13 and 14 to the interior ends (at the infilled frame) of beams B4–B7.

Sensitivity to intra-event variability


A building site is chosen 43.0 km away from the epicenter of the Northridge (1994) earthquake,
Figure 13. Two orthogonal ground motion acceleration records obtained from seven stations during
the earthquake were selected in [10] to be closely located within a distance Rrup of 5.2 to 6.5 km
from the fault rupture plane, and to have nearly similar soil conditions with NEHRP soil classifi-
cation C and D [33]. The recording station designations and site information are listed in Table VI.

500
Ultimate displacement leading to
450 Idealized fracture of longitudinal steel bars
and spring element removal
OpenSees simulation
Vertical Reaction Force at Node 1 [kN]

400

350

300

250

200
B4 B6
11 15
150 13

100 C1 C5

50
1 5

0
0 50 100 150 200 250 300 350 400 450 500
Downward Displacement at Node 13 [mm]

Figure 12. OpenSees simulation of large-displacement response in transverse beams.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
624 M. TALAAT AND K. M. MOSALAM

34.5
Selected site
location

Latitude
Recording station
Epicenter
Surface projection of
fault rupture plane
34
-119 -118.5 -118
Longitude

Figure 13. Northridge (1994) earthquake locality [10] and building site.

Table VI. Summary of ground motion recording stations used in application ‘A’.

Recording station Code Rrup (km)∗ VS30 (m/s)†


Jensen Filter Plant JEN 5.4 373
Newhall Fire Station NWH 5.9 269
Newhall West Pico Canyon Road WPI 5.5 286
Rinaldi Receiving Station RRS 6.5 282
Sylmar Converter Station SCS 5.3 251
Sylmar Converter Station East SCE 5.2 371
Sylmar Olive Grove Medical Center SYL 5.3 441
∗ Closest distance from a site to the fault rupture plane.
† Average shear-wave velocity in the top 30 m of soil.

The time-history profile and intensity of the individual records show considerable differences,
representing a measure of the irreducible aleatory uncertainty due to the site location and orientation
of the building. The modified FE model is subjected to the 14 ground motion records, each scaled
to represent seven seismic hazard levels. The simulation results are used to construct fragility-like
curves for four collapse limit-states defined below. Each curve represents the conditional proba-
bility of reaching a limit-state given the specified hazard level. An empirical probability is defined
as the number of simulations in which a limit-state is reached divided by the total number of
simulations per hazard level, i.e. 14.
The directions of the horizontal ground motions recorded at each recording station are not
coincidental, and hence referred to as ‘Component 1’ and ‘Component 2’ in Table VII. Significant
differences in characteristics can be observed, with no discernible independent effects of site
location and measurement orientation. Component 1 is the direction per record in which the elastic
spectral acceleration response is higher for the considered structural system. An average value of
the 5%-damped elastic spectral acceleration Sa (T̄1 , 5%) is computed for the vibration period range
0.121–0.192 s (T̄1 ) corresponding to the FE model’s fundamental periods with the diagonal struts
included and removed, respectively. This spectral acceleration is adopted as a seismic intensity
measure. The scaling of individual components is based on the USGS uniform hazard curves [34]
via matching the system-specific Sa (T̄1 , 5%) to the hazard-specific values in Table VIII, linearly

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 625

Table VII. Unscaled seismic intensity measures of records used in application ‘A’.
Component 1 Component 2

Record PGA (g) Sa (T̄1 , 5%) (g) Record PGA (g) Sa (T̄1 , 5%) (g)
JEN292 1.023870 1.8011 JEN022 0.570622 0.9417
NWH090 0.583030 1.3155 NWH360 0.589780 1.0988
WPI046 0.454930 0.5943 WPI316 0.325431 0.4173
RRS228 0.825195 1.2973 RSS318 0.486504 1.0854
SCS142 0.897237 1.2291 SCS052 0.612467 0.8901
SCE018 0.828279 1.3825 SCE288 0.493035 1.2162
SYL360 0.843306 1.2507 SYL090 0.604490 0.8726

Table VIII. Hazard-specific seismic intensity measures at selected building site location.
Hazard Level 1/50 2/50 3/50 5/50 10/50 20/50 50/50
RP (years) 4975 2475 1642 975 475 224 72
PGA (g) 1.8407 1.5887 1.4344 1.2339 0.9676 0.6790 0.2706
Sa (0.1, 5%) (g) 3.6857 3.1366 2.7767 2.3784 1.8124 1.2484 0.5058
Sa (0.2, 5%) (g) 4.6824 3.8789 3.4726 2.9261 2.2447 1.5323 0.6389
Sa (T̄1 , 5%) (g) 4.1841 3.5078 3.1247 2.6523 2.0286 1.3904 0.5724

averaged between periods of 0.1 and 0.2 s (RP refers to return period and A/50 refers to A%
probability of being exceeded in 50 years). This procedure could have been more complex if the
mechanism of collapse were sensitive to the values of the scaling factors being computed. In the
present application, however, collapse always initiated in the URM walls due to its significantly
higher lateral stiffness. Four collapse limit-states are investigated:

(1) Partial URM wall collapse: removal of at least one diagonal strut.
(2) Complete URM wall collapse: removal of both diagonal struts.
(3) Partial structural system collapse: removal of at least two RC columns.
(4) Complete structural system collapse: removal of all six RC columns.

The conditional probability curves are shown in Figure 14 where the probabilities of the URM wall
collapse and the partial RC frame collapse are non-zero even at the lowest considered hazard level
of 50/50. The interpretation of these results deserves attention. The structural system is subjected
to ground motions recorded in similar, neighboring sites during the same earthquake event, which
nevertheless exhibited large variation in time-history profiles and response spectra. This variability
is reduced by scaling to the same elastic spectral acceleration at each hazard level. However, for
each hazard level between 20/50 and 2/50, the probability of reaching all limit-states is larger than
zero and less than unity, i.e. a simulation using at least one-scaled component reaches all limit-states
while using another similarly scaled component from the selected 14 records reaches none. This
emphasizes the paramount importance of fundamentally addressing intra-event variability within
the framework of performance-based earthquake engineering to develop consistent and unified
selection and scaling of ground motion records for simulating structural progressive collapse.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
626 M. TALAAT AND K. M. MOSALAM

1
Limit-state 1
0.9 Limit-state 2
Limit-state 3
Limit-state 4
0.8

Probability of Collapse
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
50/50 20/50 10/50 5/50 3/50 2/50 1/50
Hazard Level

Figure 14. Empirical conditional probability curves of application ‘A’.

APPLICATION ‘B’: FIVE-STORY STRUCTURAL SYSTEM

The structural system is selected to represent complex multi-story behavior, where the collapse of
stiff URM walls may lead to the formation of a soft-story mechanism. The objectives of application
‘B’ are to investigate the collapse pattern and its sensitivity to uncertainties in selected system
parameters and in seismic hazard.

Description and modeling of test-bed structural system


A three-bay five-story planar RC frame infilled with a URM wall in the middle bay of each
story representing a typical model of a 1960s building structure and located at the same site as
application ‘A’ is adopted for application ‘B’. The RC frames and URM walls, as well as meshing
and fiber-discretization of the FE model, are similar in dimensions and reinforcement to application
‘A’, except where mentioned here. For the purpose of estimating the dead and live loads on the
analyzed planar frame, the spacing between the individual frames is assumed to be 3500 mm with
floor and roof dead and live loads assumed as 5.5 and 2.5 kPa, respectively, [28]. Longitudinal
lap splices are inserted in columns C21 and C41 just above the second and fourth floor levels,
respectively. The corresponding FE model consists of displacement-based beam–column elements
for the RC frame members, compression-only truss members for the URM wall, and coupled
shear-axial springs at the bases of the intermediate bay columns in all stories, Figure 15. The
springs are shown only at the foundation level for clarity. Nodes 1-4 are fully restrained. Masses
and gravity loads are lumped at the nodes. The FE model has simulated vibration periods of
0.446, 0.146, 0.086, 0.081, and 0.064 s for the first five modes of free vibration. Rayleigh viscous
damping ratio  = 0.05 is assumed for the first and third vibration modes by combining mass- and
stiffness-proportional damping.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 627

B51 52 B52 53 B53


51 54
251 S53

C51 C52 C53 C54


S52

151 B41 B42 B43


41 44
241 S43

C41 C42 C43 C44


S42

141 B31 B32 B33


31 34
231 S33

C31 C32 S32 C33 C34

131 B21 B22 B23


21
221 S23

C21 C22 C23 C24


S22

121 B11 B12 B13


11 14
211 S13

C11 C12 C13 C14


S12 Typical detail at base of
Typical detail at base of
all C?2 columns above all C?3 columns above
111
1 20 30 4
2 3

Figure 15. OpenSees FE model of the five-story test-bed structural system.

Reference ground motion record and hazard level


The ground motions recorded during Düzce (1999) earthquake are characterized by two intervals of
high-amplitude shaking. Therefore, a structural system, which has been seriously damaged during
the first interval, may as a result completely collapse during the second interval. The Lamont
station record (North direction) used in the shake-table tests in [28] has a scaled PGA of 1.26 g.
It is used as the reference ground motion record for this application. Scaling is conducted taking
into consideration the significant change that takes place in the structural system’s stiffness after
the collapse of the URM wall panels. The simulated fundamental vibration period increases from
0.446 to 0.591 s upon elimination of the struts in the first story, which were observed to collapse
first during preliminary simulations. Hence, an average value of the 5%-damped elastic spectral

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
628 M. TALAAT AND K. M. MOSALAM

Figure 16. Elastic 5%-damped response spectrum used in application ‘B’.

acceleration of the FE model is computed as Sa (T̄1 , 5%) = 1.038 g, Figure 16. The target spectral
acceleration value, corresponding to the reference hazard level (2/50) at the selected building site,
is computed for the average first-mode period as 2.493 g [34].

Modeling uncertainty in parameters for deterministic sensitivity study


The collapse of the relatively stiff URM wall in any story increases the demands on the deficient
RC columns and changes the system’s dynamic characteristics. The resulting localization of further
damage may lead to a soft-story collapse. The progressive collapse response is thus dependant on
the URM wall stiffness, story mass and gravity loads, damping in the system, and intensity of
the earthquake; all uncertain random variables. The effect of uncertainty in system parameters is
investigated and compared with that of uncertainty in seismic intensity to determine the relative
importance using the Tornado diagram method [35]. In this method, the response quantities of
interest are computed using a set of reference values of the uncertain input parameters. Next,
input parameters, assumed uncorrelated, are individually varied within a given range parameterized
by their coefficient of variation (COV). The ‘swings’, i.e. the resulting changes in the response
quantity, as a measure of relative importance, are sorted with the largest at top and the smallest
at bottom forming a tornado-like shape. Uncertainty in a parameter with larger swing has more
influence on the reliability and accuracy of the simulated response.
The reference mass and gravity loads are related to the occupancy of the structure and are
equal to the full dead load in addition to a live load factor LL = 0.25 [36]. The maximum and
minimum gravity loads correspond to full and no occupancy. The URM wall stiffness is calculated,
assuming uncertain strut strength f mo and a deterministic corresponding strain εmo = 0.0028, as
E mi = 2 f mo /εmo . The non-negative f mo is assigned a lognormal distribution. The tested sample
mean and COV were respectively reported in [28] as 17.0 MPa and 0.1, and taken as estimates
of the median and COV. The mean, maximum and minimum values are estimated for a 0.997
confidence interval; since the gravity load is varied over a near-certain interval. For the lognormal

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 629

Table IX. Statistics of uncertain system parameters in deterministic sensitivity analysis.


Parameter Distribution Reference COV Minimum Maximum
E mi (GPa) Lognormal 6.07 0.1 4.46 8.14
 (%) Lognormal 5.0 0.4 2.0 10.0
LL (%) Not applicable 25.0 Not applicable 0.0 100.0
Sa (T̄1 , 5%) (g) Not applicable 2.493 Not applicable 2.205 2.915
(Hazard level) (2/50) (3/50) (1/50)

distribution, these quantities are,



ˆf¯ = fˆ ˆ2
mo mo(0.5) 1+  (4)
2
min,max
f mo = fˆmo(0.5) exp[(−1)1,2 3(ln(1+ ˆ ))0.5 ] (5)

where fˆ¯mo , fˆmo(0.5) , and ˆ are the estimated mean, median, and COV. For the viscous damping
ratio, mean, maximum and minimum values of 0.05, 0.10 and 0.02 are commonly assumed in
practice. For a 0.997 confidence interval and lognormal distribution, this corresponds to a median
value of 0.5 = 0.045 [28] and COV of 0.4. The reference seismic hazard level is set to 2/50 [36].
The higher and lower hazard levels of 1/50 and 3/50 are investigated. The values of the corre-
sponding elastic spectral accelerations are listed along with the uncertain system parameter ranges in
Table IX.

Progressive collapse response of reference structural system


The simulated system response eventually exhibits global collapse due to the formation of a soft
first story mechanism after collapse of the first story URM wall. Figure 17 shows the progression
of collapse via snapshots of the deformed FE model geometry (no displacement magnification)
at time intervals of significant events. In Figure 18, the story drift (‘ratio’ hereafter is omitted) is
defined as the average drift calculated from the four pairs of nodes defining story columns, Figure
15, and normalized by the story height. All story drift time-histories are similar before removal
of the diagonal strut S12 at 5.061 s, indicating that the dynamic response is dominated by the
first mode of vibration [28]. Diagonal strut S13 collapses at 5.581 s. This leads to shear failure
in columns C12 and C13, and the first story drift increases disproportionately to the fifth story
due to the localization of damage in the weakened first story. At 6.761 s, shear-damaged column
C13 loses its axial load capacity and is removed, closely followed by C12 removal at 6.826 s.
Subsequently, the first story beams act as catenaries and pull the two remaining columns at the
first story laterally inwards. Excessive lap-splice slip in C21 results in its collapse at 7.690 s. The
subsequent increases in first-story drift result in severe bar buckling and ultimately collapse of
C11 at 7.891 s. The structural system starts to undergo almost-rigid body rotation around node 14,
where the first story drift values start to exhibit unrestrained increase in the positive direction while
the fifth story and overall drifts exhibit unrestrained increase in the negative direction. Ultimately,
the imposed curvature at the top of column C14 results in its flexure-axial collapse at 8.229 s,
and the structure undergoes unrestrained free fall. The time at incipient collapse is defined as
the time when unrestrained motion is initiated at 7.891 s (removal of C11), and the maximum
‘meaningful’ drifts are defined for the time period preceding it. The maximum first story, fifth

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
630 M. TALAAT AND K. M. MOSALAM

14 14
Vertical Co-ordinate [m] 12 12

Vertical Co-ordinate [m]


10 10

8 8

6 6

4 4

2 2

0 0

0 2 4 6 8 10 12 0 2 4 6 8 10 12
(a) Horizontal Co-ordinate [m] (b) Horizontal Co-ordinate [m]

14 14

12 12
Vertical Co-ordinate [m]
Vertical Co-ordinate [m]

10 10

8 8

6 6

4 4

2 2

0 0

0 2 4 6 8 10 12 0 2 4 6 8 10 12
(c) Horizontal Co-ordinate [m] (d) Horizontal Co-ordinate [m]

Figure 17. Collapse progression in reference five-story structural system: (a) time = 7.7 s
(removal of C21); (b) time = 7.9 s (removal of C11); (c) time = 8.2 s (collapse mechanism);
and (d) time = 8.3 s (free-fall).

story, and overall drifts prior to collapse are 0.037, 0.013, and 0.017, respectively, with mode and
sequence of collapse given in Table X.

Deterministic sensitivity to uncertainty in system and hazard parameters


The effect of changing several parameters results either in further localizing the structural damage
in one story, decreasing the time to collapse; or spreading the damage into other stories, delaying
collapse or avoiding it altogether. Hence, the change in the simulated time at incipient collapse is
found to be an effective measure for the influence of an uncertain parameter value on the system’s
vulnerability to collapse. On the other hand, the simulated story drifts exhibit a decrease when
global collapse is avoided. However, when global collapse takes place, drifts may not exhibit a

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 631

Figure 18. Simulated drift ratio time-histories of reference five-story structural system.

Table X. Simulated sequence of collapse for reference five-story structural system.


Collapse mode Wall strut Shear-axial Confinement loss
Collapsed FE S12 S13 C13 C12 C21 C11 C14
Time (s) 5.061 5.581 6.761 6.826 7.690 7.891 8.229

direct correlation with the induced change in uncertain system parameters. Instead, the maximum
drift is controlled by local variables such as the axial load and the corresponding deformation
ductility of RC columns. Accordingly, Tornado diagram results of story drifts are found less
valuable by comparison.
The plot of the swings resulting from the deterministic sensitivity analysis in ascending order is
the graphical output of a Tornado diagram analysis. In the Tornado diagram shown in Figure 19,
the largest swing is associated with the earthquake intensity. Increasing the hazard level to 1/50
results in a stronger earthquake and earlier collapse at 6.92 s, while decreasing it to 3/50 results in
no collapse other than the first story URM wall. Next in importance is the URM wall stiffness. The
higher stiffness value results in earlier collapse at 7.469 s. The lower stiffness results in lower force
redistribution to the neighboring columns. In addition, the low-stiffness URM walls exhibit damage
in the upper stories, which reduces the subsequent localization of damage and story drift demand
in any one story and enables the structural system to survive the earthquake without gravity load
collapse in any RC column. Next in importance is the live load on the structure. Full occupancy
results in higher gravity-induced axial loads on the RC columns and the URM walls. This leads
to the collapse of at least one diagonal strut in each of the first four stories and to slightly earlier
collapse at 7.681 s. On the other hand, no occupancy results in lower gravity-induced axial loads.
As a result, the structure survives the first interval of high-amplitude shaking but collapses during
the second high-amplitude shaking interval at 16.37 s. Finally, the least important parameter is the

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
632 M. TALAAT AND K. M. MOSALAM

No incipient collapse
Reference value

Sa (T1 5,%)

Uncertain Parameter
Emi

LL

6 8 10 12 14 16 18 20 22 24
Time at Incipient Collapse [sec]

Figure 19. Tornado diagram for simulated sensitivity of time at incipient collapse.

viscous damping ratio of the undamaged structure, unless the increase in frictional damping due
to damage accumulation is explicitly modeled, which is not the case in the present study.
The swings for the top two uncertain parameters in Figure 19 are close in magnitude. However,
for the lower stiffness URM wall, the survival of the structural system comes at the expense of more
extensive damage propagation into the upper stories of the structure and a larger maximum first
story drift. Moreover, the minimum URM wall stiffness corresponds to a 0.997 confidence interval,
while the minimum ground motion intensity corresponds to decreasing the reference hazard level
by only one percentage point of being exceeded in 50 years. Thus, the most important uncertain
parameter in deciding the possibility of progressive collapse is the seismic intensity. However,
the most important system parameter when subjected to a given ground motion is the URM wall
stiffness.

CONCLUSIONS

An analytical formulation is developed based on the dynamic equilibrium of forces to account


for abrupt removal of collapsed elements from a structural model. The algorithm simulates the
dynamic redistribution of forces in addition to a simplified modeling of impact and recently
developed criteria for element removal. Two applications are presented using structural systems
of seismically deficient RC frames with URM infill walls. First, a probabilistic study is performed
on the influence of intra-event variability on the outcome of collapse simulation in a one-story
system. Conditional probability curves are developed for four limit-states characterizing the extent
of system damage and eventual collapse. The results demonstrate that intra-event variability is
an important factor in determining the outcome of progressive collapse simulation. Second, a
deterministic simulation and sensitivity study are performed for a five-story system subjected to
the 1999 Düzce earthquake. The system collapsed in a soft-story mechanism after collapse and
element removal of the first story URM wall. Time to collapse is shown to correlate well with
the influence of uncertain parameters on the system’s vulnerability to collapse. The sensitivity
study identifies the uncertainty in earthquake intensity as the most important factor affecting the

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
MODELING PROGRESSIVE COLLAPSE 633

outcome, followed by URM wall stiffness for a given ground motion. Finally, uncertainties in live
load (occupancy level) and viscous damping ratio (excluding frictional damping due to damage)
have moderate and minor influences, respectively.

ACKNOWLEDGEMENTS
This study was supported by the Earthquake Engineering Research Centers Program of the NSF under
Award No. EEC-9701568 to PEER at University of California, Berkeley. Financial support from the
research sponsor is gratefully acknowledged. Opinions and findings presented are those of the authors
and do not reflect views of the sponsor.

REFERENCES
1. Virdi KS, Beshara FB. Numerical simulation of building decommission as a progressive collapse process. Third
International Conference on Decommissioning and Demolition. UMIST: U.K., 1992.
2. Gilmour JR, Virdi KS. Numerical modeling of the progressive collapse of framed structures as a result of impact
or explosion. Second International PhD Symposium in Civil Engineering, Budapest, 1998.
3. Krauthammer T, Hall RL, Woodson SC, Baylot JT, Hayes JR, Sohn Y. Development of progressive collapse analysis
procedure and condition assessment for structures. National Workshop—Prevention of Progressive Collapse,
Rosemont, IL, 2002.
4. Houghton DL, Karns JE. Post 9-11 multi-hazard mitigation in steel frame structures as a function of connection
geometry. Seventy-first International Convention of SEAOC, Santa Barbara, CA, 2002.
5. Crawford JE. Retrofit methods to mitigate progressive collapse. Technical Report TR-02-7, Karagozian and Case,
Glendale, CA, 2002.
6. Starossek U. Progressive collapse study of a multi-span bridge. Structural Engineering International, IABSE
1999; 16(2):121–125.
7. Ellingwood BR. Mitigating risk from abnormal loads and progressive collapse. Journal of Performance of
Constructed Facilities 2006; 20(4):315–323.
8. Kim Y, Kabeyasawa T. Dynamic test and analysis of an eccentric reinforced concrete frame to collapse. Thirteenth
World Conference on Earthquake Engineering, Vancouver, BC, Canada, 2004.
9. Wu C-L, Loh C-H, Yang Y-S. Shake table tests on gravity load collapse of low-ductility RC frames under near-
fault earthquake excitation. Proceedings, 1st International Conference on Advances in Experimental Structural
Engineering, Nagoya, Japan, Itoh Y, Aoki T (eds), 2005; 725–732.
10. Ghannoum W. Experimental and analytical dynamic collapse study of a reinforced concrete frame with light
transverse reinforcements. Ph.D. Dissertation, University of California, Berkeley, 2007.
11. Grierson DE, Xu L, Liu Y. Progressive-failure analysis of buildings subjected to abnormal loading. Computer-aided
Civil and Infrastructure Engineering 2005; 20(3):155–171.
12. Kaewkulchai G, Williamson EB. Modeling the impact of failed members for progressive collapse analysis of
frame structures. Journal of Performance of Constructed Facilities 2006; 20(4):375–383.
13. Kaewkulchai G, Williamson EB. Beam element formulation and solution procedure for dynamic progressive
collapse analysis. Computers and Structures 2004; 82(7–8):639–651.
14. Pretlove AJ, Ramsden M, Atkins AG. Dynamic effects in progressive failure of structures. International Journal
of Impact Engineering 1991; 11(4):539–546.
15. Ramsden M. Dynamic effects in the progressive failure of lattice structures. Ph.D. Dissertation, University of
Reading, U.K., 1987.
16. Powell G. Progressive collapse: case studies using nonlinear analysis. Structures Congress and Forensic
Engineering Symposium, New York, 2005; 2185–2198.
17. Haselton CB, Deierlein GG. Benchmarking seismic performance of reinforced concrete frame buildings. Structures
Congress and Forensic Engineering Symposium, New York, 2005; 1891–1902.
18. Zareian F, Krawinkler H. Assessment of probability of collapse and design for collapse safety. Earthquake
Engineering and Structural Dynamics 2007; 36(16):1901–1914.
19. Elwood KJ, Moehle JP. Axial capacity model for shear-damaged columns. ACI Structural Journal 2005;
102(4):578–587.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe
634 M. TALAAT AND K. M. MOSALAM

20. Talaat M, Mosalam KM. Computational modeling of progressive collapse in reinforced concrete frame structures.
PEER Technical Report 2007/10, 2008.
21. Mosalam KM, Talaat M, Binici B. A computational model for reinforced concrete members confined with fiber
reinforced polymer lamina: implementation and experimental validation. Composites Part B: Engineering 2007;
38(5–6):598–613.
22. Mazzoni S, McKenna F, Scott MH, Fenves G. Opensees user’s manual, 2004. Available from:
www.opensees.berkeley.edu.
23. Zineddin M, Krauthammer T. Dynamic response and behavior of reinforced concrete slabs under impact loading.
International Journal of Impact Engineering 2007; 34(9):1517–1534.
24. Schonberg WP, Keer LM, Woo TK. Low velocity impact of transversely isotropic beams and plates. International
Journal of Solids and Structures 1987; 23(7):871–896.
25. The MathWorks, Inc. Getting Started with Matlab 7 (Release 2007b), Natick, MA, 2007.
26. Crisfield MA, Moita GF. Unified co-rotational framework for solids, shells and beams. International Journal of
Solids and Structures 1996; 33(20–22):2969–2992.
27. Brown J, Kunnath SK. Low-cycle fatigue failure of reinforcing steel bars. ACI Materials Journal 2004; 101(6):
457–466.
28. Hashemi A, Mosalam KM. Seismic evaluation of reinforced concrete buildings including effects of masonry
infill walls. PEER Technical Report 2007/100, 2007.
29. Menegotto M, Pinto P. Method of analysis for cyclically loaded reinforced concrete plane frames including
changes in geometry and non-elastic behavior of elements under combined normal force and bending. Proceedings,
Symposium on Resistance and Ultimate Deformability of Structures Acted on by Well Defined Repeated Loads,
International Association for Bridge and Structural Engineering (IABSE), Lisbon, 1973; 15–22.
30. Mander JB, Priestley MJN, Park R. Theoretical stress–strain model for confined concrete. Journal of Structural
Engineering 1988; 114(8):1804–1826.
31. ACI. Standard building code requirements for reinforced concrete. ACI 318-02, Detroit, MI, 2002.
32. ACI. Standard Building Code Requirements for Reinforced Concrete. ACI 318-63, Detroit, MI, 1963.
33. ICC. International Building Code. International Code Council, Falls Church, VA, 2003.
34. USGS Probabilistic Hazard Curves, 2007. Available from: http://earthquake.usgs.gov/research/hazmaps/design/.
35. Lee T-H, Mosalam KM. Seismic demand sensitivity of reinforced concrete shear-wall building using FOSM
method. Earthquake Engineering and Structural Dynamics 2005; 34(14):1719–1736.
36. NEHRP. Recommended Provisions for Seismic Regulations for New Buildings and Other Structures. Building
Seismic Safety Council, Washington, DC, 2001.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:609–634
DOI: 10.1002/eqe

S-ar putea să vă placă și