Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of Building Engineering 49 (2022) 103992

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: www.elsevier.com/locate/jobe

Response spectrum of a reinforced concrete frame structure under


various column removal scenarios
Seweryn Kokot
Faculty of Civil Engineering and Architecture, Opole University of Technology, ul. Prószkowska 76, Opole, 45-758, Poland

A R T I C L E I N F O A B S T R A C T

Keywords: In a frame structure, sudden removal of a load-bearing member (e.g., a column) causes structural
Column removal vibrations, often leading to a partial or total progressive collapse. Since the term ”sudden column
Response spectrum removal” is imprecise, it is worth evaluating how the rate of column removal influences the
Progressive collapse maximum displacement and how this response refers to the one produced by the static column
Reinforced concrete structure removal. Both linear and nonlinear, inelastic modeling of the reinforced concrete members are
considered using the OpenSees finite element framework. The numerical simulations consist of a
gradual increase in column removal time from near zero to asymptotically approaching the static
response. Due to the asymmetry of the analyzed frame building, three column removal scenarios
are investigated. It is observed that the impact of column removal rate depends on the ability of
the remaining structure to withstand the lack of column and the natural period of the mode shape
corresponding to the downward motion of the model without the column. Also, the column
removal durations producing 95% of the maximum response are identified and their ratios to the
respective natural periods are given. These investigations can be helpful when planning numer­
ical simulations and physical experiments. Finally, a novel simplified single degree of freedom
system with support removal is proposed and its response spectrum is developed. This simplified
model helps to understand more complex multi-degree of freedom systems.

1. Introduction
The structural progressive collapse can happen when a local failure causes successive structural damage and induced vibrations (in
case of insufficient resistance and integrity) lead to the total collapse or a collapse disproportionate to the initial event. The very recent
collapse of Champlain Towers South condominium in Miami, Florida (June 24th, 2021) brought the problems of progressive collapse
to the spotlight of engineering interests. It should, however, be pointed out that the roots of this research date back to the collapse of
the Ronan Point apartment building in London, in 1968, (e.g. Ref. [1]). Later on, this subject were in focus of civil engineering in­
vestigations two more times: after the collapse of the Alfred P. Murrah Federal Building in 1995 in Oklahoma City, (e.g. Refs. [2,3]),
and the World Trade Center disaster in 2001, (e.g. Ref. [4]). The problem of progressive collapse in civil engineering can cost human
lives, cause huge economic losses and spread fear in the society. Therefore, further progress in understanding this phenomenon (in
designing, construction and maintenance of buildings) is of high importance to mitigate the negative consequences. An extensive
literature review on research and practice in the field of progressive collapse can be found in Refs. [5–8].
A deeper understanding of full-scale progressive collapse requires detailed preliminary studies of specific problems limited in their
area of analysis, for example:

E-mail address: s.kokot@po.edu.pl.

https://doi.org/10.1016/j.jobe.2022.103992
Received 21 October 2021; Received in revised form 21 December 2021; Accepted 1 January 2022
Available online 6 January 2022
2352-7102/© 2022 Published by Elsevier Ltd.
S. Kokot Journal of Building Engineering 49 (2022) 103992

a) case studies using advanced modeling of loading and structures to assess progressive collapse performance (e.g. Refs. [9–11])
b) evaluation of ultimate progressive collapse resistance of unstrengthened and strengthened RC structures through nonlinear static
analysis (e.g. Refs. [12–25])
c) evaluation of structural robustness against progressive collapse (e.g. Refs. [26–29])
d) an insight into vibration problems induced by a sudden removal of a load-bearing member (fragment of a wall, column, girder,
support, etc.) (e.g. Refs. [30–36]).
It should also be noted that the problem of a load-bearing element removal can be regarded in a broader scope as a sudden change
in constraints, restraints, or boundary conditions (e.g. Refs. [33,37]).
Progressive collapse can be triggered by many different actions. Such actions include blast loading caused by gas or bomb ex­
plosions; impacts of vehicles, ships or planes; human errors made during the design or construction stages, etc. Prediction of such
abnormal actions is very difficult and depends on many factors. Designing for example a large reinforced concrete frame building
against progressive collapse due to blast loading is a big challenge. Analysing such a building and checking if a progressive collapse
would happen or not depends on many assumptions. The major unknowns are: how large the explosive charge is? (what is the peak
pressure?), how far from the building the explosive is detonated, and whether the blast affects the corner load-bearing elements of the
building or the ones situated in the middle of the building sides? The most complex scenario is when the blast wave is modelled
explicitly by discretizing the air and the structure itself. A schematic illustration of such an approach is presented in Fig. 1a). Less
complex modelling involves obtaining an expected pressure loading from independent calculations and subsequently applying this
load to the structural model as depicted in Fig. 1b). For example many procedures and tools (e.g. Refs. [38,39] and the ConWep
LS-DYNA blast load subroutine [40,41]) can help to determine the evolution curve for pressure in time and space based on the mass of a
trinitrotoluene (TNT) equivalent and the stand-off distance from the explosion location to the structure. The difficulties and the di­
versity of possible cases mean that there are effectively no provisions in the national codes and standards to design structure to resist
blast loads caused by explosive materials detonated outside the structure. Thus instead of an explicit analysis of a structure to a specific
blast load, the current codes, standards and guidelines recommend a threat independent design, that is the design due to an unspecified
cause (Fig. 1c), or to design some elements (key elements) to resist a sufficiently high pressure (e.g. 34 kPa).
The two approaches illustrated in Fig. 1a and b) are outside the scope of this paper, instead a threat-independent approach, as
shown in Fig. 1c), is adopted in which it is assumed that a load-bearing element is removed from the structure as a consequence of its
total destruction due to an accidental action. For example, in the case of blast, the actual destroying effect equivalent to the column
removal can take from milliseconds up to fractions of a second (see e.g. Ref. [42]) depending on many factors such as explosive mass,
standoff distance, the capacity of the column itself (cross section dimensions and height of the column, concrete class, amount and
grade of reinforcement steel).
In [43], several reinforced concrete frame structures have been analyzed to simulate instantaneous column removal using various
FE packages. Both 2D and 3D models are considered to assess the contribution of spatial performance of a 3D structure. It should be
noted however that in case of the structure, analyzed in this paper, the two spans and one bay are not sufficient to contribute significant
stiffness, capacity and integrity and therefore the analysis of 2D or 3D model does not differ considerably (difference by less 10%).
Other relevant papers include [44], where the authors examined a representative four-bay three-story scaled reinforced concrete
subframe due to the quasi-static removal of a lower story column. In Ref. [45], it is introduced a new method for progressive collapse
analysis of reinforced concrete frames based on modifications (adding the non-zero initial conditions and structural damage done by
the applied blast load) in the GSA and DoD guidelines and then numerically validated on a three-story two-bay reinforced concrete
frame model.
In many guidelines on design against progressive collapse [46–48], the dynamic analysis of structures is recommended when
designing or verifying a structure under a sudden load-bearing element removal scenario. For frame structures, the sudden column
removal leads to vibrations, which in case of insufficient structural resistance can result in partial or total progressive collapse. The
term of suddenness is non-unique and therefore it is of importance to investigate how the rate of column removal influences the
structural dynamic response.

Fig. 1. Various approaches to simulate dynamic response of a building to blast loading.

2
S. Kokot Journal of Building Engineering 49 (2022) 103992

Moreover, such influence can be characterized by the response spectrum which presents the effects of a changing quantity on the
maximum response, which can be computed using the finite element method or measured from experimental tests. In earthquake
engineering community, the classic response spectrum shows the maximum response of an oscillator of varying dynamic properties (e.
g. natural period or frequency) to a specific ground motion [49,50]. Another example of a response spectrum is when observing the
maximum response of a dynamic system to a ramp load with increasing rise time [49]. Other examples include the response spectrum
of a vibrating beams under successive moving loads [51], impulse loads on plates [52], floor vibrations due to occupant walking [53]
and in other problems (e.g. Refs. [54,55]).
Preliminary investigations on the role of column removal rate have been reported in Ref. [7], where it is concluded that the quasi
instantaneous column removal can be simulated using column removal times less than 0.005 s for selected types of structure. In
Ref. [34], the authors presented results of experimental tests against sudden column removal of the flat-slab building and in Ref. [35]
the same authors provided a parameter study of the effects of removal times on dynamic response. The simulated response covered
selected column removal times of: 0.000 1, 0.020, 0.050 and 0.100 s. The main findings are that the results for 0.000 1 and 0.020 s
gave similar response however the experimental displacements correspond between 0.050 and 0.100 s column removal times. This
specific problem is also addressed in Ref. [36] where the authors present the extensive results of a full-scale reinforced concrete
building under a sudden corner column removal and both experimental tests and numerical simulations are reported. In Ref. [56], the
instantaneous column removal is analyzed and experimentally tested on a ten-story reinforced concrete structure under explosion of an
exterior column. The paper reports that the development of Vierendeel action is observed as the main factor in the alternate load
transfer. In Ref. [57], the authors investigate a reinforced concrete substructure by contact detonation demonstrated by experimental
tests and numerical nonlinear modelling.
Since, to the author’s knowledge, there is no theoretical (or experimental) research on removal (or change) in boundary conditions
in wide range of removal rates, therefore the main objective of this paper is to examine how the column removal times, starting from
the values close to zero and asymptotically reaching the static column removal, impact the level of structural vibrations. (Although a
preliminary study in this regard on a two-span beam under support removal has been recently published by the author of this paper in
Ref. [58]). In particular, it is of interest to identify the shape of the curve representing the maximum vertical displacements in relation
to the column removal rates as well as how it influences the dynamic factors defined as the maximum peak displacement to the
displacement resulting from the static column removal. To this end, finite element simulations of have been carried out. The impact of
the column removal rate is analyzed for the three column removal scenarios of the reinforced concrete frame building. All procedures
described in this paper have been coded in Tcl and Python (importing Numpy/Matplotlib modules) with the finite element models and
analyses available in the OpenSees framework [59–61].
The main subject of this paper is a thorough examination of the influence of column removal rate on the dynamic response of a full-

Fig. 2. Computational 3D model with nodes, finite element numbers and locations of distinct cross-section types (Legend: gt = girder type, ct = column type).

3
S. Kokot Journal of Building Engineering 49 (2022) 103992

scale reinforced concrete frame building modelled as either a linear-elastic system or nonlinear, inelastic system, using distributed
inelasticity force-based finite elements and corotational geometric nonlinearity. Moreover, a characteristic pattern of the response
spectrum curve (the maximum vertical displacements against the increasing column or support removal times) has been established in
the linear and nonlinear analyses.
The three main objectives of this paper are: (1) the examination of the influence of column removal rate on the dynamic response of
a full-scale reinforced concrete frame building, (2) development of an equivalent single degree of freedom model with support
removal.
In the author’s opinion the novel findings and original contribution of this paper are:
1. Establishing a characteristic shape of a response spectrum curve for linear and nonlinear models of a 2D frame structure under
different column removal scenarios and over a wide range of column removal rates.
2. Development of a novel simplified equivalent single degree of freedom model with support removal.
The paper begins with Sections 2 and 3 which introduce and describe the computational model of the analyzed frame structure and
give the most important dynamic characteristics of the structure. Section 4 presents the main steps of the numerical column removal
procedure and then provides the results of the response spectra for both the linear and nonlinear models. Section 6 discusses the
dynamic response of the analyzed structure modelled as a plane sub-frame and the spatial frame. Based on the findings in Section 4,
Section 5 presents an equivalent single degree of freedom model with support removal which helps explain the characteristic pattern of
the response spectrum of more complex models. Finally, Section 7 summarizes and concludes this paper.

2. Description of the structure


The analyzed structure is a 3-storey 2-bay reinforced concrete frame building with a 0.24 m thick slab. The structure contains two

Fig. 3. Girders, columns and transverse beam cross-section types.

4
S. Kokot Journal of Building Engineering 49 (2022) 103992

plane frames linked by transverse beams. The girder beams are 1 m wide and 0.24 m high. The slab is composed of 0.20 m rib-and-
block floor and 0.04 m topping and has the same height as the beams. The building is supported by square columns (0.4 × 0.4 m). The
3D computational model of the structure showing the nodes, finite elements and locations of distinct cross-section types, is presented in
Fig. 2 whereas the corresponding beam and column cross-section types are shown in Fig. 3. Other details of the analyzed structures can
be found in Refs. [31,62].
The main structural materials of the building are C25/30 concrete and steel reinforcement bars of 440 MPa characteristic yield
strength. Based on the laboratory tests (details can be found in Ref. [31]) the necessary material mechanical properties have been
identified and in finite element calculations, following average values have been used: the concrete compressive strength fc = 32.77
MPa, the steel yield strength fy = 524.56 MPa, the steel ultimate strength ft = 640.0 MPa.

3. Computational model and input parameters


The computational model of the analyzed structure is presented in Fig. 2, however, due to the symmetry and regularity, as well as
preliminary numerical tests reported in Section 6, for the subsequent studies, the plane frame model has been considered, which
constitutes the front plane frame presented in Fig. 2.
For the purpose of this study, three column removal scenarios have been considered (labeled A, B and C) as shown in Fig. 4, in
which separately the column: A, B or C is suddenly removed from the frame. The computational model consists of 30 finite elements
(21 for girders and 9 for columns).
For both linear elastic and nonlinear inelastic analyses, the OpenSees ’forcedBeamColumn’ elements have been used with the so
called fiber section approach. The nonlinear force-based formulation is described in e.g. Refs. [63–66]. The idea is that the actual
reinforced concrete cross-section is divided into both concrete and steel fibers to which the constitutive relationships are assigned. In
the linear elastic case, to concrete fibers the elastic uniaxial material model is applied with the concrete modulus of elasticity Ec =
24.86 GPa and similarly Es = 200 GPa for steel fibers. Alternatively, for the linear elastic studies, the OpenSees ’elasticBeamColumn’
Euler-Bernoulli finite elements can be used, however to include both the concrete and steel material the force-based finite elements
provide more accurate results without cross-section homogenization technique.
The inelastic material behaviour can occur at five pre-specified Gauss-Lobatto integration points along the member length and at
these points, the cross-section are modelled using fibers. To each fiber, a nonlinear inelastic constitutive model of steel or concrete is
assigned as indicate below. The number of concrete fibers have been set to 120 and 4, along the depth (local y-axis) and width (local z-
axis) direction of the cross-section, respectively. The steel fibers are located according to the cross-section types as shown in Fig. 3,
where the concrete cover is equal to 0.04 m.
It should be noted that the nonlinear inelastic modeling of RC frame members is performed using the guidelines recommended in
Refs. [9,32,67] and is additionally validated through the reproduction of the vertical pushover curves in Ref. [9] where the authors
compared the behaviour of reinforced concrete subassemblies modelled with micro- and macro-elements.
The constitutive relationship for concrete used in this study is based on the Kent-Park model [68] which satisfy the observations by
Ref. [69]. In compression, in the initial loading stage, the stress-strain curve is parabolic [70] ascending up to the compressive strength
fc0 at strain ϵc0. This point is related with the onset of concrete crushing so then it is observed a descending line. After reaching the
ultimate strain ϵcu, the concrete is regarded as completely crushed. At this point, the unconfined concrete in cover and beam core is
assumed to carry no stresses while the confined concrete in column core is capable of carrying the residual stress fcu. The linear
descending branch depends on the amount and distance of transverse reinforcement [71]. The confinement effect of ties in column
core results also in greater values of compressive strength as opposed to the unconfined concrete in beams and in the cover of columns.
Also in this study it is assumed that the tensile concrete strength is negligible.
The most common uniaxial constitutive model for steel is bilinear elasto-plastic with kinematic or isotropic hardening with or
without the Bauschinger effect. The main parameters of a bilinear model are the modulus of elasticity Es, yield strength fy, hardening
ratio b = Eh/Es, ultimate stress fu and strain ϵu. If the Bauschinger effect is required, the constitutive model introduced by Menegotto-
Pinto [72] represents accurately the steel rebar material behaviour. A modified version of the Menegotto-Pinto model is presented in

Fig. 4. Three A, B and C column removal scenarios.

5
S. Kokot Journal of Building Engineering 49 (2022) 103992

Refs. [73,74] and is implemented in OpenSees [59–61]. For the smooth transition between the two lines of the initial steel modulus of
elasticity Es and the hardening modulus Eh, a curve can be controlled by additional parameters: R0, cR1, cR2.
It should be noted that these constitutive models take into account the strength and stiffness degradation under cyclic loading as
shown in Fig. 5 for concrete and Fig. 6 for steel.
In the inelastic finite element simulations, the following nominal material parameters have been used for the nonlinear analyses
using the OpenSees ’Concrete02’ material: σc0 = 37.22 MPa, ϵc0 = 0.002, σ cu = 6 MPa, ϵcu = 0.003 5, σsy = 524.6 MPa, Es = 200 GPa,
σsu = 640 MPa and ϵsu = 0.10.
It should also be noted that the corotational formulation of the geometric nonlinearity which accounts for large displacements has
been used according to formulation reported in Refs. [75–78].

3.1. Modal properties


In structural dynamic investigations it is important to perform the modal analysis to obtain dynamic characteristics of the structure.
To this end, the eigenvalue problem is solved and the first two natural periods are provided in Table 1. For each column removal
scenario, in the first row, the values refer to the linear elastic model, in the second row – to the nonlinear inelastic model at the
beginning of the analysis, while the values in the third row (marked with asterisk ,*”) refer to the structure at the end of the time history
calculations where plastic hinges developed in a few cross-sections. Relevant, from the column removal point of view, are the mode
shapes showing movement in the vertical direction and are illustrated in Fig. 7. These mode shapes refer to the natural periods T2 in
Table 1, because the first mode shape is typically related to movement in the horizontal direction.

4. Computational procedure and results


To evaluate the response spectrum of the analyzed structure under column removal, the numerical procedure requires calculations
in a loop, where individual iterations refer to the increasing duration times tr of the column removal. Inside the loop, for a given value
of tr, the calculations are performed according to the following steps:
● Create the finite element model.
● Apply the nominal values of self-weight.
● Record the nodal sectional forces (bending moment, shear and axial forces) in the upper end of the column to be removed (see
Fig. 8a).
● Remove the column from the finite element model (see Fig. 8b).
● At the node where the upper end of the column was connected, apply the set of forces recorded in the previous step (this set of forces
is applied as constant in time) (see Fig. 8b).
● Apply at the same node, another set of forces acting in the opposite direction and increasing in time tr to the reference value, so as to
simulate the column removal (see Fig. 8c).
● Perform the dynamic analysis using the finite element method software.
● As output, the time histories of dynamic response are obtained from which the maximum value of the vertical displacement needs
to be found.
Using the finite element method formulation, for the nonlinear inelastic case the following matrix equation of motion needs to be
iteratively solved

MÜ + Pr (U, U̇) = P(t) (1)

and for the linear elastic case the equation simplifies to

Fig. 5. Example of uniaxial cyclic test for concrete.

6
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 6. Example of uniaxial cyclic test for steel.

Table 1
First two natural periods of free vibrations for three column removal scenarios.

Mode T1 [s] T2 [s]

A)lin 0.387 0 0.212 5


A)nlin 0.342 2 0.188 3
A)nlin* 0.751 3 0.377 7
B)lin 0.310 5 0.138 8
B)nlin 0.274 6 0.122 7
B)nlin* 0.509 6 0.251 2
C)lin 0.391 5 0.130 1
C)nlin 0.346 9 0.115 1
C)nlin* 0.622 9 0.211 9

Fig. 7. Natural mode shapes: mode 1 (solid line) and mode 2 (dashed line) for A), B) and C) column removal scenarios (the corresponding natural periods are listed
in Table 1).

MÜ + CU̇ + KU = P(t) (2)

where M, C and K are the mass, damping and stiffness matrices, while Ü, U̇ and U are the vectors of accelerations, velocities and
displacements, respectively. P(t) is the vector of external time-dependent loads, while Pr (U, U̇) is the vector of nodal resisting forces
which can be linearized using the Newton-Raphson method.

4.1. Linear elastic response spectra


After completing the whole loop of calculations for discrete values of the column removal time tr, the results are presented in plots.
The total number of tr discrete values is in the range of 140–200 to obtain smooth curves in the response spectra. The tr list starts from
0.001 s up to 3 s with variable increments.
In the dynamic finite element calculations, the average-acceleration Newmark method has been used to solve either Eq. (1) or Eq.
(2) with the specified time interval (e.g. 0.001 s). As far as damping is concerned, the Rayleigh damping matrix has been determined
based on the 5% damping ratio for the first two natural mode shapes.
The procedure described above can be implemented in any finite element software which enables us the dynamic analysis.

7
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 8. Procedure steps in the simulation of sudden column removal: (a) the column to be removed is present – record the internal forces N, V, M, (b) remove the
column and apply the N, V, M forces accordingly to reproduce the state in the first step, (c) cancel the N, V, M forces by applying the opposite − N, − V, − M forces
acting linearly in a short time tr.

However, due to multiple reiterated calculations with modified parameters it is convenient to use a finite element framework which
allows us to automate the finite element calculations in a loop. Here, the OpenSees framework has been used [60], which allows users
to create applications using the Tcl [59] or Python [61] programming languages.
A practical and technical aspect can be mentioned here that the computational time of about 120 linear elastic dynamic analyses
performed on a personal notebook (processor: Intel Core i7-8650U 8th Gen quad-core 1.9 GHz–4.2 GHz, 8 MB cache, 32 GB RAM) for
increasing column removal times take about 140 s (∼ 1 s per one analysis), whereas in the case of the nonlinear inelastic analyses, one
analysis takes about 60 s–120 s depending on the plastic deformation and issues in achieving numerical convergence. Other aspects
which impact the computational time for a given structural model are: number of finite elements per member, number of Gauss
integration points, number of concrete and steel fibers, computational complexity of the constitutive material models (number of
internal variables, etc.).
If we look at the left column removal scenario (see Fig. 4a), a selected number of displacement time histories is shown in Fig. 9, and
the maximum vertical displacement at node 1 for incremented values of tr ∈ 〈0.001 s − 3.0 s〉 are given in Fig. 10.
The time history curve in Fig. 9 for the shortest tr = 0.001 s shows that the maximum vertical displacement occurs in the second
downward cycle of vibration, which is caused by the coupling of the two first natural mode shapes (the first mode shape moves the
structure to the left and at the same time the second mode moves the structure upwards, thus reducing the overall downward
movement), while in the second cycle of vibration, the second mode shape synchronises with the first mode and produces the
maximum deflection of max u1y = 0.098 4 m. Similarly, for tr = 0.10 s the maximum deflection happens in the second cycle, while for tr
= 0.18 s and longer column removal duration times, the maximum vertical displacements take place in the first cycle. For all cases of tr,
the vibrations stabilize at the static deflection level of 0.060 4 m. In Fig. 10 we can observe that the maximum value u1y = 0.098 4 m
among all discrete values of tr occurs at the shortest tr = 0.001 s. Next, the curve decreases smoothly down to the point of the second
natural period tr ≈ T2 = 0.21 s. Crossing this peculiar point, the curve decreases to point (tr = 0.430 s, max u1y = 0.061 4 m), and after
that a few waves with alternate increasing and decreasing segments follow to eventually converge to the static deflection. The 95% of

Fig. 9. Time history plots of the vertical displacement at node 1 for selected removal times tr – left column removal – linear analysis.

8
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 10. Plot of removal time tr versus maximum vertical displacement at node 1 – left column removal – linear analysis.

the maximum overall displacement is found to take place for tr95 = 0.09 s which gives the ratio of 0.42 with respect to the second
natural period T2. This 95% notional threshold can help decide on setting the time interval used in the Newmark method for solving the
equations of motion. On the right hand side of Fig. 10, the dynamic factor is related to the static stabilized deflection. For the left
column removal scenario, the maximum dynamic factor is obtained, as expected, for the shortest column removal time tr = 0.001 s and
is equal to 1.63.
When it comes to the central column removal scenario a few displacement time histories are illustrated in Fig. 11, while the
associated maximum vertical displacements at node 4 are shown in Fig. 12.
The vibration curves in Fig. 11 seems simpler than in the first scenario. Here, the second mode shape is dominant and clearly
separated from other modes. The time history of u4y for the column removal duration time tr = 0.001 s gives the maximum deflection in
the first downward vibration cycle and is equal to 0.029 m. Then for tr = 0.125 s (which is close to the second natural period T2 = 0.138
8 s), the dynamic effects are less pronounced giving the max u4y equal to 0.018 8 m which is 35% smaller than for the shortest tr =
0.001 s. This trend is confirmed in Fig. 12 where the curve of max u4y decreases steeply with increasing values of tr. However after
reaching 0.017 5 m at tr = 0.145 s, the max u4y increases to 0.019 6 m for tr = 0.205 s and then again decreases to the minimum point.
This wave repeats, but the amplitude keeps decreasing and converges to the static deflection 0.016 7 m. The 95% threshold is achieved
for tr95 = 0.04 s which can be related by the ratio to the second natural period tr95/T2 = 0.29 and the maximum dynamic factor equals
1.7.
For the right column removal scenario, a few of the 116 individual displacement time histories are shown in Fig. 13, and the
maximum vertical displacements at node 7 are plotted in Fig. 14.
This scenario seems similar to the left column removal one (indeed the mode shapes in Fig. 7 are similar), however, here the span of
the second bay is shorter and also the ratio between T C1 /T C2 = 3.0 is different to the similar ratio of T A1 /TA2 = 1.8 for the left column
removal scenario. Looking at Fig. 13, this fact leads to the situation when for the shortest tr = 0.001 s, the vibration curve u7y of the first
natural period (related to T1 = 0.39 s) is disrupted by the second mode of vibration and after adds up together to reach the maximum
vertical displacement at the level of 0.029 5 m. The close coupling is also visible in two vibration curves for tr = 0.10 s and tr = 0.18 s in
Fig. 13. Fig. 14 shows similar pattern of the maximum vertical deflections as in Fig. 10, however, here the 95% of the overall maximum
deflection occurs for tr95 = 0.05 s, for which the ratio tr95/T2 = 0.38 and the maximum dynamic factor is equal to 1.77.

4.2. Nonlinear inelastic response spectra


Similar analyses have been carried out for the inelastic modelling of reinforced concrete members. In order to facilitate the
comparison between the linear and nonlinear analysis the individual time history plots for the A, B and C column removal scenarios are
shown in Figs. 15, 17 and 19, whereas the response spectra (the collection of maximum vertical displacement for various increasing

Fig. 11. Time history plots of the vertical displacement at node 4 for selected removal times tr – central column removal – linear analysis.

9
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 12. Plot of removal time tr versus maximum vertical displacement at node 4 – central column removal – linear analysis.

Fig. 13. Time history plots of the vertical displacement at node 7 for selected removal times tr – right column removal – linear analysis.

Fig. 14. Plot of removal time tr versus maximum vertical displacement at node 7 – right column removal – linear analysis.

values of column removal times tr) are presented in Figs. 16, 18 and 20. The summary of results for both linear and nonlinear analyses
and for all column removal scenarios are presented in Table 2. Since for nonlinear analysis, two values of the natural frequencies are
identified (see 1), the ratio tr95/T2 refers to the natural frequency at the beginning of the analysis, whereas the ratio tr95 /T∗2 in
parenthesis is for the natural frequency when decreased tangent stiffness is caused by material yielding.
The main findings of the nonlinear studies can be summarized as follows:
● the maximum vertical displacement at node 4 (see Fig. 4) for the shortest time of column removal (0.001 s) is 2–3 times greater than
the displacement obtained from the quasi-static column removal (low rate of column removal)
● the time duration of the column removal causing the vertical displacement at the level of 95% of the maximum possible vertical
displacement (for tr = 0.001 s) are of the order of 0.03 − 0.07 s (see Figs. 16, 18 and 20), which means that, both the time interval
and the time duration of the column removal at most can be equal to 0.01 s. However shorter time intervals are required at the first
stage of response (up to reaching the maximum vertical displacement), and longer time intervals at the second state of free vi­
brations, what suggests the use of the Newmark integration schemes where the time interval can be adapted (variable time

10
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 15. Time history plots of the vertical displacement at node 1 for selected removal times tr – left column removal – nonlinear analysis.

Fig. 16. Plot of removal time tr versus maximum vertical displacement at node 1 – left column removal – nonlinear analysis.

Fig. 17. Time history plots of the vertical displacement at node 4 for selected removal times tr – central column removal – nonlinear analysis.

interval). On the other hand, the nonlinear inelastic iterations can require subdivision into shorter time intervals to achieve the
convergence criteria as in the Newton-Raphson method,
● the solution of the eigenvalue problem indicates, that the first mode refers to vibrations of the frame in the horizontal direction, and
in the second mode, vertical motion is dominant. The second mode is also dominant after the removal of column A and B (Fig. 4 A
and B), while in the case of the C scenario, additionally it is visible the contribution of the next mode describing the vertical vi­
brations (Fig. 4C). For the structure analyzed in the intact state (no material yielding) the natural periods of free vibrations T2 are in
the range of 0.11 − 0.19 s (see Table 1),
● when the structure deforms inelastically due to the formation of plastic hinges at specified cross-sections, the stiffness of the
structure decreases which results in longer natural periods of vibrations (see Table 1, values marked with asterisks ,*”)
On the other hand, if comparing the linear and nonlinear analyses, it should be noted that the characteristic patterns in the response
spectra for the linear analysis (see Figs. 10, 12 and 14) have been ,flatten” in the nonlinear analyses (compared to Figs. 16, 18 and 20).

11
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 18. Plot of removal time tr versus maximum vertical displacement at node 4 – central column removal – nonlinear analysis.

Fig. 19. Time history plots of the vertical displacement at node 7 for selected removal times tr – right column removal – nonlinear analysis.

Fig. 20. Plot of removal time tr versus maximum vertical displacement at node 7 – right column removal – nonlinear analysis.

Table 2
Summary of results for all column removal scenarios: the maximum displacement, the dynamic amplification factor (DAF), the 95% of maximum displacement time and
the ratio of the 95% to the downward natural frequency T2.

Case max uy [m] DAF tr95 [s] tr95/T2 (tr95 /T∗2 )

A)lin 0.098 1.63 0.09 0.42


A)nlin 0.53 2.8 0.07 0.37 (0.18)
B)lin 0.029 1.70 0.04 0.29
B)nlin 0.114 2.24 0.03 0.24 (0.12)
C)lin 0.030 1.77 0.05 0.38
C)nlin 0.06 1.85 0.04 0.35 (0.19)

12
S. Kokot Journal of Building Engineering 49 (2022) 103992

This can be explained by the fact that in the linear analysis, the stiffness and natural frequencies remain constant during the whole
individual response while in the nonlinear analysis, both the stiffness and natural frequencies change: first, due to the nonlinear
constitutive models and second, due to the formation of plastic hinges in the first downward movement of the structure.
It should also be noted that using the nonlinear inelastic model of reinforced concrete, the analyzed frame structure withstands the
assumed sudden column removal scenarios; however, the structure is only partially damaged and can be further assessed using various
damage identification methods (e.g. Refs. [79–84]).

5. A novel simplified single degree of freedom model


Observing the repeatable pattern in plots representing the maximum values of vertical displacement against the increasing column/
support removal times; made the author of this paper further explore the topic and found that in Refs. [49,51], a similar pattern in the
response spectrum is observed. In Ref. [49] the response spectrum of a single degree of freedom system for an arbitrary dynamic
loading using the Duhamel’s integral is analyzed, while in Ref. [51], the response spectrum and dynamic factors of vibrating beams
under successive moving loads are investigated.
The term of response spectrum is widely used in earthquake engineering design, where the maximum response values are extracted
from individual time history responses for increasing values of natural periods. However, other examples of adopting the term of
response spectrum can be found in Refs. [49,50,52–55].
To illustrate the problem in a analogous way as in Ref. [49], let us consider a single degree of freedom system with the following
data: the natural period T = 1.0 s, the vibrating mass: m = 0.101 3 kg, the stiffness: k = 4.0 N/m, the circular frequency: ω = 6.28 rad/s.
To compare, the systems without and with ξ = 0.05 damping ratio are considered. A mass m suspended on a spring k is supported by an
element below and loaded with a force P as illustrated in Fig. 21a. Next the support element is removed and replaced by the reaction
force R (see Fig. 21b). Finally the reaction R is removed by the counter-reaction, − R(t), applied in removal time tr (see Fig. 21c). If such
a formulated problem is solved numerically for a range of increasing tr values and for no (or 5% damping), the response spectrum is
depicted in Fig. 22. It should be noted that the minima visible in this plot, which coincide exactly with the static displacement ustat, take
place exactly at the multiple values of the natural period T = 1.0 s. Also, the maximum dynamic factor (for the shortest tr = 0.001 s) is
equal to 2.0 (no damping) and 1.976 (5% damping), similarly to the single degree of freedom system with an external ramp force.
Identifying the characteristic shape of the resulted curve similar to the response spectrum found in Ref. [49], the adapted formula
for the undamped single degree of freedom under support removal can be expressed as:
⃒ ⃒
⃒ sin π(tr /T0 )⃒
max u2 = ustat + ⃒⃒ustat ⃒ (3)
π(tr /T0 ) ⃒

and its plot in Fig. 22 matches exactly the finite element computations if no damping is included.
However, the studies in this paper concern the multi-degree of freedom system, and based on Figs. 10, 12 and 14, it seems that the
response spectrum pattern, presented here, is also valid to a great extent for multi-degree of freedom discrete systems as resulting from
the finite element discretization. Although, unlike the single degree of freedom system, the multi-degree of freedom systems are
characterized by the finite number of frequencies and associated mode shapes, where the first subsequent modes are dominant and
their contribution with respect to the total number of modes can be measured by modal participation factors (see e.g. Ref. [49]).
To obtain modal participation factor, first the generalized mass matrix can be calculated as follows

M = ΦT MΦ (4)

Fig. 21. Sudden support removal of a single degree of freedom system: a) the single degree of freedom system with support and applied load P, b) the single degree of
freedom system with the reaction R replacing the support, c) modelling a sudden reaction removal by cancelling the reaction R in time tr.

13
S. Kokot Journal of Building Engineering 49 (2022) 103992

Fig. 22. The response spectrum representing the support removal time tr versus the maximum vertical displacement for a single degree of freedom system – linear
analysis – FEM no damping, FEM 5% damping, Eq. (3) no damping.

where M and Φ are the global mass matrix and the matrix of natural mode shapes, respectively.
The coefficient vector L is calculated using the influence vector r composed of unit values in the direction (either in x or y) of the
coordinates of interest

L = ΦT Mr (5)

In progressive collapse analysis, the vertical (downward motion) y direction is more important than the horizontal x one, because the
dominant motion of the structure under column removal is the direction of gravity acceleration.
Finally, the modal participation factor Γi for the i-th mode can be obtained from
Li
Γi = (6)
M ii

The modal participation factors give information on relative importance of a particular mode shape in the dynamic response in the
specified direction.
In the analyzed cases of the left, central and right column removal scenarios, the normalized modal participation factors expressed
in percent for the first mode only are equal to 32%, 1.4% and 11%, and for the second mode, the respective normalized modal
participation factors are 68%, 98.6% and 89%. This can explain, why for the left and right column removal scenarios, the response
spectra in Figs. 10 and 14 are not as pronounced as for the response spectra of the central column removal scenario in Fig. 12. It should
be noted that for the frame model considered in this paper, the first mode shows the movement in horizontal x direction and the second
mode represents the motion in the vertical y direction.

Fig. 23. Comparison of displacement time histories for the 3D frame model with rigid connection to transverse beams, the 3D frame model with pinned connection to
transverse beams and the 2D single plane frame model.

14
S. Kokot Journal of Building Engineering 49 (2022) 103992

6. 3D versus 2D modelling
Real reinforced concrete buildings are 3D continuous structures which can be modelled as a 3D frame or a 3D frame with slabs and
walls. In the cases of regular structures it can be justified to model and analyze only 2D frames which can be virtually extracted from
the entire structure. Analysing the 2D model can substantially reduce the time devoted to modelling the structure as well as decrease
the numerical calculation time. However, whether a 2D frame model leads to credible results depends on many aspects such as the
spatial integration of girder and slab elements or the load pattern.
In this section it is intended to assess the difference and possible inaccuracy when the reinforced concrete frame building is
modelled as full 3D spatial frame model or a 2D plane frame model. The 3D model is shown in Fig. 2 while the 2D model is the front
plane frame extracted from the 3D model.
Both 3D and 2D models are analyzed under sudden central column removal. Note that in this case the story level consists of
transverse beams connecting two plane frames, therefore if the story levels are made of reinforced concrete slabs, the results can differ.
Since the connection stiffness between transverse beams and columns falls in the range between a fully rigid connection and a pinned
one, hence only the borderline cases are considered in the comparison.
Fig. 23 shows the time history of the displacement at the point connecting the missing column. The stabilized displacements for the
3D model with rigid connection, 3D model with pinned connection and 2D plane model are 0.013 8 m, 0.014 4 m and 0.013 8 m,
respectively. Therefore it can be observed that the relative error between the 3D model with rigid connections and 2D model is of 7.2%,
while for the comparison between the 3D models, the relative difference decreases to 4.3%. If comparing the pinned connection 3D
model with the 2D plane frame model then the relative error further drops to 2.7%. These relatively small differences justify the use of
the simpler 2D model in selected further analyses.
However, it should be noted that when a structure is moderately or highly irregular in plan and/or elevation, then only 3D models
of such structures can correctly provide prediction of a structural global behaviour (see e.g. Refs. [62,85]).

7. Conclusions
A parametric study of the influence of a sudden column removal of a reinforced concrete frame structure is presented. Linear and
nonlinear dynamic responses are analyzed in detail. Response spectra of the maximum vertical displacement and dynamic factors for
three column removal scenarios are obtained.
The linear analysis reveals that the time during which the support is removed could be in the range of 29%–42% of the downward-
motion natural period, yet the column removal can still be regarded as ”sudden”, giving only a 5% relative difference. This conclusion
agrees with the investigations of Russel et al. [35] on other types of structures. Specifically, 95% of the overall maximum deflection for
a wide range of column removal durations times tr could be related to the relevant natural periods by ratios: 0.42, 0.29, and 0.38 for the
left, central and right column removal scenarios, respectively.
A characteristic pattern of the response spectrum curve (the maximum vertical displacements against the increasing column or
support removal times) is revealed in the linear analysis. An analogous characteristic pattern can be found in a single degree of freedom
system excited by an external ramp-constant force or in beam vibrations under moving loads.
When the material nonlinearity is included in the analysis, the characteristic pattern in the response spectrum is less pronounced
and can be explained by the following facts: (a) in the nonlinear analysis, the stiffness contributing to the mode shape is calculated
according to the current tangent stiffness, and thus, it is responsible for the constant change in natural frequencies, which in turn
desynchronize the contribution of mode shapes in the total structural dynamic response; (b) in the nonlinear cyclic analysis, at each
deformation and loading cycle, the energy is dissipated. It provides additional damping, which further influences frequencies of os­
cillations and respective mode shapes (see, e.g. Ref. [86]).
Moreover, the response spectrum resulting from the analysis of a structure under column removal can help plan experiments of
sudden support or column removal, where slower removal times can be applied and still the response is close to the quasi instantaneous
removal time. This response spectrum may also be useful in specifying the optimum time interval in dynamic analyses based on
implicit step-by-step numerical integration using, e.g., the Newmark family methods.
Finally, a novel simplified single degree of freedom system with support removal is proposed and its response spectrum is
developed. This simplified model together with modal participation factors can help to understand more complex multi-degree of
freedom systems.

Author agreement statement


I, the undersigned declare that this manuscript is original, has not been published before and is not currently being considered for
publication elsewhere.
I confirm that the manuscript has been read and approved by all named authors and that there are no other persons who satisfied
the criteria for authorship but are not listed.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

15
S. Kokot Journal of Building Engineering 49 (2022) 103992

Acknowledgements
The reported research would not be possible without a three-year (2009–2012) post-doc stay at the European Laboratory for
Structural Assessment, Joint Research Centre, European Commission, Ispra, Italy. The time spent at the ELSA laboratory brought the
necessary experience to further research career of the author. The respective European Commission support is gratefully
acknowledged.
The author would also like to gratefully acknowledge professor Zbigniew Zembaty for reviewing the final manuscript and many
helpful remarks and suggestions.
The numerical research studies presented in this paper were funded by the Ministry of Education and Science of Poland: TETA 004/
21.

References
[1] C. Pearson, N. Delatte, Lessons from the progressive collapse of the ronan point apartment tower, in: Third Forensic Engineering Congress, San Diego, California,
United States, 2003.
[2] J.D. Osteraas, Murrah Building bombing revisited: a qualitative assessment of blast damage and collapse patterns, J. Perform. Constr. Facil. 20 (4) (2006)
330–335.
[3] A. Kazemi-Moghaddam, M. Sasani, Progressive collapse evaluation of murrah federal building following sudden loss of column g20, Eng. Struct. 89 (2015)
162–171.
[4] Z.P. Bazant, M. Verdure, Mechanics of progressive collapse: learning from World trade center and building demolitions, J. Eng. Mech. 133 (3) (2007) 308–319.
[5] O.A. Mohamed, Progressive collapse of structures: annotated bibliography and comparison of codes and standards, J. Perform. Constr. Facil. 20 (4) (2006)
418–425.
[6] R.S. Nair, Preventing disproportionate collapse, J. Perform. Constr. Facil. 20 (4) (2006) 309–314.
[7] S. Kokot, Literature Survey on Current Methodologies of Assessment of Building Robustness and Avoidance of Progressive Collapse, JRC Scientific and Technical
Reports JRC 5598, European Commission, Joint Research Centre, Ispra, Italy, 2009.
[8] J.M. Adam, F. Parisi, J. Sagaseta, X. Lu, Research and practice on progressive collapse and robustness of building structures in the 21st century, Eng. Struct. 173
(2018) 122–149.
[9] Y. Bao, S.K. Kunnath, S. El-Tawil, H.S. Lew, Macromodel-based simulation of progressive collapse: RC frame structures, J. Struct. Eng. 134 (7) (2008)
1079–1091.
[10] L. Kwaśniewski, Nonlinear dynamic simulations of progressive collapse for a multistory building, Eng. Struct. 32 (5) (2010) 1223–1235.
[11] K. Qian, J. Cheng, Y. Weng, F. Fu, Effect of loading methods on progressive collapse behavior of rc beam-slab substructures under corner column removal
scenario, J. Build. Eng. 44 (2021) 103258.
[12] Y. Bao, S.K. Kunnath, Simplified progressive collapse simulation of rc frame-wall structures, Eng. Struct. 32 (2010) 3153–3162.
[13] H. Lew, Y. Bao, F. Sadek, J.A. Main, S. Pujol, M.A. Sozen, An Experimental and Computational Study of Reinforced Concrete Assemblies under a Column
Removal Scenario, Tech. Rep. NIST Technical Note 1720, National Institute of Standards and Technology, 2011.
[14] H.S. Lew, Y. Bao, S. Pujol, M.A. Sozen, Experimental study of RC assemblies under a column removal scenario, ACI Struct. J. 111 (4) (2014) 881–892.
[15] F. Wang, J. Yang, S. Nyunn, I. Azim, Effect of concrete infill walls on the progressive collapse performance of precast concrete framed substructures, J. Build.
Eng. 32 (2020) 101461.
[16] Y.-H. Weng, K. Qian, F. Fu, Q. Fang, Numerical investigation on load redistribution capacity of flat slab substructures to resist progressive collapse, J. Build. Eng.
29 (2020) 101109.
[17] I. Azim, J. Yang, S. Bhatta, F. Wang, Q. feng Liu, Factors influencing the progressive collapse resistance of rc frame structures, J. Build. Eng. 27 (2020) 100986.
[18] K. Qian, Y.-H. Weng, F. Fu, X.-F. Deng, Numerical evaluation of the reliability of using single-story substructures to study progressive collapse behaviour of
multi-story RC frames, J. Build. Eng. 33 (2021) 101636.
[19] M. Husain, J. Yu, B.H. Osman, J. Jid, Progressive collapse resistance of post-tensioned concrete beam-column assemblies under a middle column removal
scenario, J. Build. Eng. 34 (2021) 101945.
[20] X. Gu, B. Zhanga, Y. Wang, X. Wang, Experimental investigation and numerical simulation on progressive collapse resistance of RC frame structures considering
beam flange effects, J. Build. Eng. 42 (2021) 102797.
[21] H. Huang, M. Huang, W. Zhang, M. Guo, Z. Chen, M. Li, Progressive collapse resistance of multistory rc frame strengthened with hpfl-bsp, J. Build. Eng. 43
(2021) 103123.
[22] X. Deng, S.-L. Liang, F. Fu, K. Qian, Progressive collapse resistance of emulative precast concrete frames with various reinforcing details, J. Struct. Eng. 146 (6)
(2020), 04020078.
[23] K. Qian, S. Liang, F. Fu, Y. Li, Progressive collapse resistance of emulative precast concrete frames with various reinforcing details, J. Struct. Eng. 147 (8) (2021),
04021107.
[24] L. Qiu, F. Lin, K. Wu, X. Gu, Progressive collapse resistance of rc t-beam cable subassemblages under a middle-column-removal scenario, J. Build. Eng. 42 (2021)
102814.
[25] Y. Huang, Y. Tao, W. Yi, Y. Zhou, L. Deng, Numerical investigation on compressive arch action of prestressed concrete beam-column assemblies against
progressive collapse, J. Build. Eng. 44 (2021) 102991.
[26] U. Starossek, M. Haberland, Disproportionate collapse: terminology and procedures, J. Perform. Constr. Facil. 24 (6) (2010) 519–528.
[27] U. Starossek, M. Haberland, Measures of structural robustness - requirements and applicaitons, in: ASCE SEI 2008 Structures Congress - Crossing Borders,
Vancouver, Canada, 2008.
[28] J. W. Baker, M. Schubert, M. H. Faber, On the assessment of robustness, Struct. Saf. 30 (253-267).
[29] D.-C. Feng, S.-C. Xie, J. Xu, K. Qian, Robustness quantification of reinforced concrete structures subjected to progressive collapse via the probability density
evolution method, Eng. Struct. 202 (2020) 109877.
[30] S. Marjanishvili, E. Agnew, Comparison of various procedures for progressive collapse analysis, J. Perform. Constr. Facil. 20 (4) (2006) 365–374.
[31] S. Kokot, A. Anthoine, P. Negro, G. Solomos, Static and dynamic analysis of a reinforced concrete flat slab frame building for progressive collapse, Eng. Struct.
40 (2012) 205–217.
[32] Y. Bao, H.S. Lew, S.K. Kunnath, Modeling of reinforced concrete assemblies under column-removal scenario, J. Struct. Eng. 140 (1) (2012), 04013026.
[33] A. Alsahlani, F.B. Mathis, R. Mukherjee, Vibration suppression of a string through cyclic application and removal of constraints, J. Sound Vib. 331 (20) (2012)
4395–4405.
[34] J.M. Russell, J. Owen, I. Hajirasouliha, Experimental investigation on the dynamic response rc flat slabs after a sudden column loss, Eng. Struct. 99 (2015)
28–41.
[35] J.M. Russell, J. Owen, I. Hajirasouliha, Dynamic column loss analysis of reinforced concrete flat slabs, Eng. Struct. 198 (2019) 109453.
[36] J.M. Adam, M. Buitrago, J. Sagaseta, J.J. Moragues, Dynamic performance of a real-scale reinforced concrete building test under a corner-column failure
scenario, Eng. Struct. 210 (2020) 110414.
[37] A. Humer, Dynamic modeling of beams with non-material, deformation-dependent boundary conditions, J. Sound Vib. 332 (2013) 622–641.
[38] C. Kingery, G. Bulmash, Air Blast Parameters from TNT Spherical Air Burst and Hemispherical Burst, Tech. Rep. ARBRL-TR-02555, U.S. Army Ballistic Research
Laboratory, Aberdeen Proving Ground, MD, 1984.

16
S. Kokot Journal of Building Engineering 49 (2022) 103992

[39] M. Larcher, Pressure-time Functions for the Description of Air Blast Waves, JRC Scientific and Technical Reports JRC 46829, European Commission, Joint
Research Centre, Ispra, Italy, 2008.
[40] G. Randers-Pehrson, K.A. Bannister, Airblast Loading Model for Dyna2d and Dyna3d, Tech. Rep. ARL-TR-1310, Army Research Laboratory, Aberdeen Proving
Ground, MD, 1997.
[41] J.O. Hallquist, LS-DYNA, THEORY MANUAL, 2006.
[42] Q. Ma, P. Yuan, J. Zhang, R. Ma, B. Han, Blast-induced damage on millisecond blasting model test with multicircle vertical blastholes, Shock Vib. (2015) 1–6,
2015.
[43] E. Brunesi, R. Nascimbene, Extreme response of reinforced concrete buildings through fiber force-based finite element analysis, Eng. Struct. 69 (2014) 206–215.
[44] W.-J. Yi, Q.-F. He, Y. Xiao, S.K. Kunnath, Experimental study on progressive collapse-resistant behavior of reinforced concrete frame structures, ACI Struct. J.
105 (4) (2008) 433–439.
[45] Y. Shi, Z.-X. Li, H. Hao, A new method for progressive collapse analysis of rc frames under blast loading, Eng. Struct. 32 (2010) 1691–1703.
[46] DoD UFC Guidelines, Design of Buildings to Resist Progressive Collapse, Unified Facilities Criteria (UFC) 4-023-03, Department of Defence (DoD), 2010.
[47] EN 1991-1-7, Eurocode 1 - EN 1991-1-7: Actions on Structures - Part 1-7: General Actions - Accidental Actions, 2006.
[48] GSA Guidelines, GSA Progressive Collapse Analysis and Design Guidelines for New Federal Office Buildings and Major Modernizations Projects, General Services
Administration (GSA), 2003.
[49] A.K. Chopra, Dynamics of Structures, Prentice Hall, Englewood Cliffs, New Jersey, 1995.
[50] R.W. Clough, J. Penzien, Dynamics of Structures, Computers & Structures, Inc., Berkeley, USA, 2003.
[51] E. Savin, Dynamic amplification factor and response spectrum for the evaluation of vibrations of beams under successive moving loads, J. Sound Vib. 248 (2)
(2001) 267–288.
[52] F. Botta, G. Cerri, Shock response spectrum in plates under impulse loads, J. Sound Vib. 308 (2007) 563–578.
[53] J. Chen, R. Xu, M. Zhang, Acceleration response spectrum for predicting floor vibration due to occupant walking, J. Sound Vib. 333 (2014) 3564–3579.
[54] S.J. Ahn, W.B. Jeong, W.S. Yoo, Improvement of impulse response spectrum and its application, J. Sound Vib. 288 (2005) 1223–1239.
[55] S. Pathak, V.K. Gupta, On nonstationarity-related errors in modal combination rules of the response spectrum method, J. Sound Vib. 407 (2017) 106–127.
[56] M. Sasani, M. Bazan, S. Sagiroglu, Experimental and analytical progressive collapse evaluation of actual reinforced concrete structure, ACI Struct. J. 104 (6)
(2007) 731–739.
[57] J. Yu, T. Rinder, A. Stolz, K.H. Tan, W. Riedel, Dynamic progressive collapse of an RC assemblage induced by contact detonation, J. Struct. Eng. 140 (6) (2014),
04014014.
[58] S. Kokot, Reinforced concrete beam under support removal—parametric analysis, Materials 14 (20) (2021) 5917.
[59] F. McKenna, G.L. Fenves, M.H. Scott, OpenSees - Open System for Earthquake Engineering Simulation, Pacific Earthquake Engineering Research Center,
University of California, Berkeley, CA, 2003. Available from, http://opensees.berkeley.edu/.
[60] F. McKenna, M.H. Scott, G.L. Fenves, Nonlinear finite-element analysis software architecture using object composition, J. Comput. Civ. Eng. 24 (1) (2010)
95–107.
[61] M. Zhu, F. McKenna, M.H. Scott, OpenSeesPy: Python library for the OpenSees finite element framework, Software X 7 (2018) 6–11.
[62] P. Negro, E. Mola, Current assessment procedures: application to regular and irregular structures compared to experimental results, in: Third European
Workshop on the Seismic Behaviour of Irregular and Complex Structures, Florence, 2002.
[63] F. F. Taucer, E. Spacone, F. C. Filippou, A Fiber Beam-Column Element for Seismic Response Analysis or Reinforced Concrete Structures, Tech. Rep., University
of California, Berkeley (12 1991)..
[64] A. Neuenhofer, F.C. Filippou, Evaluation of nonlinear frame finite element models, J. Struct. Eng. 123 (7) (1997) 958–966.
[65] M. Petrangeli, P.E. Pinto, V. Ciampi, Fiber element for cyclic bending and shear of RC structures. I: Theory, J. Eng. Mech. 125 (9) (1999) 994–1001.
[66] R.L. Taylor, R.C. Filippou, A. Saritas, F. Aurcchio, A mixed finite element method for beam and frame problems, Comput. Mech. 31 (2003) 192–203.
[67] Y. Alashker, H. Li, S. El-Tawil, Approximations in progressive collapse modeling, J. Struct. Eng. 137 (9) (2011) 914–924.
[68] D.C. Kent, R. Park, Flexural members with confined concrete, J. Struct. Div., ASCE 97 (1971) 1969–1990.
[69] I.D. Karsan, J.O. Jirsa, Behavior of concrete under compressive loading, J. Struct. Div., ASCE 95 (12) (1969) 2543–2563.
[70] E. Hognestad, A Study of Combined Bending and Axial Load in Reinforced Concrete Members, Tech. Rep. Bulletin Series No. 399, University of Illinois, College
of Engineering. Engineering Experiment Station, Urbana Champaign, USA, 1951.
[71] B.D. Scott, R. Park, M.J.N. Priestley, Stress-strain behavior of concrete confined by overlapping hoops at low and high strain rates, ACI Journal 79 (1) (1982)
13–27.
[72] M. Menegotto, P.E. Pinto, 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, in: IABSE Symposium on Resistance and Ultimate Deformability of Structures Acted on by Well Defined
Repeated Loads, Lisbon, 1973, pp. 15–22.
[73] F. C. Filippou, E. P. Popov, V. V. Bertero, Effects of Bond Deterioration on Hysteretic Behavior of Reinforced Concrete Joints, Tech. rep., University of California,
Berkeley (8 1983).
[74] M. Bosco, E. Ferrara, A. Ghersi, E.M. Marino, P.P. Rossi, Improvement of the model proposed by Menegotto and Pinto for steel, Eng. Struct. 124 (2016) 442–456.
[75] T. Belytschko, L. Schwer, M. Klein, Large displacement, transient analysis of space frames, Int. J. Numer. Methods Eng. 11 (1977) 65–84.
[76] M. Crisfield, Nonlinear Finite Element Analysis of Solids and Structures, vol. 1, Wiley, New York, 1991.
[77] C. Felippa, B. Haugen, A unified formulation of small-strain corotational finite elements: I. theory, Comput. Methods Appl. Mech. Eng. 194 (2005) 2285–2355.
[78] R.M. De Souza, Force-based Finite Element for Large Displacement Inelastic Analysis of Frames, Ph.D. thesis, University of California, Berkeley, California, 2000.
[79] S. Fang, R. Perera, G.D. Roeck, Damage identification of a reinforced concrete frame by finite element model updating using damage parameterization, J. Sound
Vib. 313 (2008) 544–559.
[80] M. Pircher, B. Lechner, O. Mariani, A. Kammersberger, Damage due to heavy traffic on three rc road bridges, J. Sound Vib. 33 (2011) 3755–3761.
[81] W. Hong, Z. Wu, C. Yang, C. Wan, G. Wu, Y. Zhang, Condition assessment of reinforced concrete beams using dynamic data measured with distributed long-gage
macro-strain sensors, J. Sound Vib. 331 (2012) 2764–2782.
[82] Z. Nie, H. Hao, H. Ma, Structural damage detection based on the reconstructed phase space for reinforced concrete slab: experimental study, J. Sound Vib. 332
(2013) 1061–1078.
[83] B.L. Wahalathantri, D.P. Thambiratnam, T.H. Chan, S. Fawzia, Vibration based baseline updating method to localize crack formation and propagation in
reinforced concrete members, J. Sound Vib. 344 (2015) 258–276.
[84] T. Xu, A. Castel, Modeling the dynamic stiffness of cracked reinforced concrete beams under low-amplitude vibration loads, J. Sound Vib. 368 (2016) 135–147.
[85] Z. Zembaty, M. De Stefano (Eds.), Seismic Behaviour and Design of Irregular and Complex Civil Structures II, Springer, 2012.
[86] H. Rodrigues, A. Arede, H. Varum, A.G. Costa, Energy dissipation and equivalent damping of rc columns subjected to biaxial bending: an investigation based in
experimental results, in: Proceedings of the Fifthteenth World Conference on Earthquake Engineering, Lisbon, Portugal, 2012.

17

You might also like