Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Structures 57 (2023) 105103

Contents lists available at ScienceDirect

Structures
journal homepage: www.elsevier.com/locate/structures

Risk-based cost-benefit analysis of structural strengthening to mitigate


disproportionate collapse of buildings under abnormal blast loading
André T. Beck a, *, Mark G. Stewart b
a
Structural Engineering Department, University of São Paulo, São Carlos, SP, Brazil
b
Centre for Built Infrastructure Resilience, School of Civil and Environmental Engineering, University of Technology Sydney, NSW 2007, Australia

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

Keywords: Strengthening building structures to mitigate their likelihood of disproportionate collapse, due to initial damage
Blast loading produced by abnormal loads, has a significant impact on construction costs. The decision to strengthen or not
Malevolent attack needs to be taken under significant epistemic uncertainty related to potential blast hazard scenarios and threat
Disproportionate collapse
probabilities. In this manuscript we present a novel risk-based cost-benefit analysis to address optimal and
Alternate path method
Risk analysis
codified design options for RC buildings subject to blast hazards. Building frames of different aspect ratios
Damage propagation (number of bays × number of stories) are addressed. Three terrorist blast hazard scenarios are considered: two
VBIEDs and one smaller suitcase bomb. The large uncertainty in threat likelihood is addressed by treating annual
threat probability as the independent parameter in the analyses. This makes the blast hazard analysis, and the
progressive collapse analysis, independent of the factors that can make a particular building a target, like
location, environment, ownership, and use. The analysis reveals the break-even hazard (threat) probabilities,
which make the cost of structural strengthening equal to the reduction in expected cost of progressive collapse.
For a particular building location, use, and blast threat, if the hazard probability is larger than the break-even
probability, strengthening is cost-effective. The break-even point between the choices of strengthening with
removal of a single or of two columns is also presented and discussed. Practical recommendations for codified
Alternate Path Method include a distinction between strengthening factors for beams and columns, much in the
same way as already done in conventional design.

1. Introduction Threat-independent risk-based cost-benefit assessment for regular


RC frames subject to localized initial damage has been considered in
Design against disproportionate collapse is an important mitigation [4–6]. In this manuscript, this is combined with the terrorist blast threat
measure against abnormal loadings such as accidental or malevolent formulation of Stewart [7,8], yielding a novel threat-dependent risk-
vehicle impact or explosions. Of particular concern are terrorist based formulation for the progressive collapse assessment of regular RC
improvised explosive device attacks on a building which may result in frames.
disproportionate collapse causing large loss of life, and significant eco­
nomic and societal losses. The Alternate Path Method [APM] [1,2] is one 2. Risk-based approaches to progressive collapse
strategy to mitigate against disproportionate collapse by ensuring that
the building is stable when one or two exterior supporting columns are Under multiple hazards, the annual probability of structural collapse
removed from the structure. However, the additional structural frame can be evaluated as [9–12]:
cost for APM design is considerable and may vary from 9 to 62% [3]. ∑∑
There is a need to assess under what threat circumstances the benefits of pCol = P[Col|LD, H]P[LD|H]pH (1)
mitigating disproportionate collapse outweigh their cost, and if design
H LD

measures may be optimized to ensure more cost-effective design solu­ where Col stands for collapse, pH is the annual probability of hazard
tions. This leads to a risk-based cost-benefit assessment which is the occurrence; P[LD|H] is the conditional probability of local damage, given
focus of this paper. hazard H; P[Col|LD, H] is the conditional probability of collapse, given

* Corresponding author.
E-mail addresses: atbeck@sc.usp.br (A.T. Beck), mark.stewart@uts.edu.au (M.G. Stewart).

https://doi.org/10.1016/j.istruc.2023.105103
Received 15 June 2023; Received in revised form 26 July 2023; Accepted 19 August 2023
Available online 9 September 2023
2352-0124/© 2023 Institution of Structural Engineers. Published by Elsevier Ltd. All rights reserved.
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

typical hazard scenarios are considered herein. Yet, from all terms in Eq.
(2), it can be argued that the term with most epistemic uncertainty, or
over which we have less control, is the hazard rate pH .
Following Ellingwood [11,12], Stewart [7,8,14], and Stewart et al.
[15], the estimated average rate of bombing attacks on large U.S.
buildings ranges from 2 × 10–6/building/year to 8 × 10–6/building/
year. These figures apply to buildings without any specific threat.
However, for key governmental and international institutions, iconic
buildings, or other critical facilities facing a specific threat, the proba­
bility of an attack may increase to 10–4/building/year [10]. The above
rates are averages and assume that attacks and target selection occur
randomly. However, this may not be the case for some terrorist and
Fig. 1. Regular building with nC columns by nS stories, subject to initial
other malicious events [11,12,16]. Also, the above numbers make no
damage leading to loss of nr,c columns of nr,s stories. Individual bays with length
L and height H. nreinf ,S is the number of stories for which beams and columns are distinction between specific threat characteristics, such as size of charge
strengthened. and standoff distance. Hence, the variable with the largest uncertainty in
Eq. (2) tends to be the likelihood of a successful attack, pH .
Considering the above, and based on the approaches developed by
local damage LD and hazard H.
Beck [4–6] and Stewart [7,8] , we consider pH as the independent var­
Local damage caused by abnormal loading can lead to multiple
iable in the risk analysis. This makes the conditional hazard and pro­
system damage states (SDS), depending on whether progressive collapse
gressive collapse analyses independent of the factors that could make a
is arrested or not. Risk is the product of probability of occurrence of a
particular building a potential target for a terrorist plot. By varying pH
damage state, and a measure of the consequences: R = P[SDS]C(SDS).
from a very small pH = 10− 8 to a large pH = 10− 1 , we find the break-
Total risk is given by the sum of direct (RD ) and indirect (RID ) risks,
associated to direct and indirect consequences [13]: even annual hazard probability (or threshold value pth H ) which makes

∑∑[∑ the strengthening cost equal to the reduction in expected costs of pro­
R = RD + RID = CID (SDS, LD)P[SDS|LD, H] gressive collapse failure. The decision to strengthen a particular building
H LD SDS (2) should then be taken based on a risk assessment [17], considering
+ CD (LD) ]P[LD|H]pH
ownership, use and surrounding environment of the structure, to find
where CID (SDS, LD) are the indirect costs, for each system damage state out if for a particular hazard, pH is smaller or greater than pthH.

and initial damage, and CD (LD) are the direct costs of local damage. When the hazard probability is considered the independent param­
This manuscript addresses optimal design of regular reinforced eter, conventional design under normal loading conditions and the
concrete (RC) frame buildings, as illustrated in Fig. 1. For such regular strengthening for load bridging under initial damage need to be com­
buildings, the initial loss of load-bearing elements due to abnormal bined in the same formulation. In practice, this means that normal
loading can propagate (see Fig. 2): vertically by a plastic beam bending loading condition is considered as one of the ‘hazards’ in Eq. (2), with
mechanism (denoted as B); or horizontally, either by local crushing or pH = pNLC = 1. For abnormal load hazards leading to very small pH ,
instability failure of the adjacent columns (denoted as PL) or by a global conventional design controls optimal frame configurations; for large pH ,
column crushing mechanism involving all ground floor columns strengthening for alternate load paths controls optimal configurations.
(denoted as PG) leading to what is known as a pancake-type collapse. The threshold hazard probability pth H divides the two solution spaces.
Local column crushing may also develop into a beam bending mecha­ Since the abnormal load hazard pH is the independent variable, it is
nism involving two additional bays. Hence, the basic or primary system not possible to consider the sum over hazards in Eq. (2).1 For each
damage states form the set SDS = {B, PL, PG}. abnormal load hazard acting in isolation, combined with the normal
The analysis of system behavior is addressed by terms CID (SDS, loading condition, Eq. (2) becomes:
LD)P[SDS|LD, H] in Eq. (2). CID (SDS, LD) measures the consequences of ∑
R = RD + RID = CID (SDS, NLC)P[SDS|NLC]pNLC
each type of progressive collapse failure, and P[SDS|LD, H] accounts for SDS
[ ] (3)
the probabilities of occurrence of each type of progressive collapse event ∑ ∑
conditional on the hazard. In this manuscript we address optimal design + CID (SDS, LD)P[SDS|LD, H] + CD (LD) P[LD|H]pH
of the structural system focusing on initial damage caused by malevolent LD SDS

explosive blast loading events. The Alternate Path Method (APM)


where NLC is for Normal Loading Condition, and pNLC = 1.
considered herein consists of strengthening elements adjacent to po­
tential target areas, to allow the structure to bridge over the initial
3. Structural configurations
damage, thus avoiding disproportionate collapse.
Herein, we dont specifically address the Enhanced Local Resistance
Herein we address the optimal design of regular frame buildings
Method (ELRM), which looks at blast-resistant strengthening of poten­
subject to initial damage by blast loading, as illustrated in Figs. 1 and 2.
tial target elements based on specific explosive threats to reduce term
The optimal design against progressive collapse changes significantly
P[LD|H]. Yet, column strengthening in APM has a compounding effect: it
with the aspect ratio of the building frames, as shown by [4–6]. Hence,
increases resistance against collapse propagation, specifically in the
we consider seven frame configurations, with similar total floorspace,
local column crushing failure mode, but it also increases column resis­
and aspect ratios described in Table 1. Herein, the aspect ratio of indi­
tance against blast loading (ELRM approach). This compounding effect
vidual bays is fixed, with beam length (span between columns) L =
is considered herein.
2H = 6m. Variations of bay aspect ratio have been considered in [4,5].
In this manuscript, three specific threats are considered: a bench­
The frames are designed following ASCE 7 (2016) and ACI 318–19
mark Vehicle-Borne Improvised Explosive Device (VBIED) case, a VBIED
(American Concrete Institute 2019), considering mean concrete strength
with reduced charge, and a suitcase bomb. The term P[LD|H] in Eq. (2)
accounts for the local strength of columns subject to different blast
loads. As will be shown, local strength is strongly dependent on the
assumed hazard scenarios. Hence, term P[LD|H] is characterized by 1
Considering a multi-hazard environment (such as prior earthquake damage
significant epistemic uncertainty. To address this uncertainty, three
or deterioration) can be relevant and will be subject of future studies.

2
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 2. System damage states for regular buildings subject to initial damage.

Table 1
Variants of aspect ratio (number of stories × bays) of frames studied.
Aspect ratio Number of stories × bays (ns × Axial strength of columns, Column cost factor (w.r.t. cost of Whole frame cost factor (w.r.t. cost of
(ar) (nc − 1) ) RD [kN] conventional beams) conventional 8x8)

Tall frame (16 × 4) 9, 359 3.33 1.39


Intermediate (13 × 5) 7, 900 3.09 1.25
Intermediate (11 × 6) 6, 639 2.86 1.14
“Square” frame (8 × 8) 5, 624 2.45 1
Intermediate (6 × 11) 4, 709 2.04 0.93
Intermediate (5 × 13) 3, 895 1.70 0.84
Low frame (4 × 16) 3, 181 1.37 0.76

Fig. 3. Frame aspect ratios, column cross-sections and detailing.

3
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 4. Overview of the blast hazard progressive collapse analyses (design-verify-strengthen loop).

√̅̅̅̅
illustrated in Fig. 4. The frames are initially designed considering the
fc′ = 40 MPa, tangent elastic modulus Ec = 4.73 f ′c GPa, reinforcing
usual dead and live loads, as shown in Section 5. First-floor columns are
steel strength of 500 MPa, and total distributed design loads on all exposed to blast hazards, where column damage and their failure
beams of all floors of q = 50 kN/m, roof included. All beams have the probabilities are evaluated (Section 5). At this point, columns could be
same design, with a cross-section of 200 × 500 mm2 (width × depth); strengthened to reduce damage and failure probabilities, but this would
three bars of 20 mm diameter (3Φ20) on top side (924 mm2 of steel); characterize the Enhanced Local Resistance Method (not our target).
2Φ16 + 1Φ12 at the bottom (515 mm2). The ratio of nominal live-to- Instead, we proceed with a risk-based progressive collapse analysis,
dead loads is DLnn = 0.6 for all frames. Floor and roof loads are assumed where expected progressive collapse costs are evaluated. Based on the
the same, as while different design loads will affect member sizing, they risk analysis, for given hazard probability pH and for each blast hazard,
do not affect the optimal design factors. we decide whether each frame should be strengthened or not. If not, the
The conventional design resulted in the four column cross-sections conventional design is deemed sufficient, and the design-verify-
and detailing shown in Fig. 3, for four of the seven frames. The hazard strengthen loop is exited. If frame strengthening is found to be
analysis addresses conditional failure probabilities P[LD|H] and damage required, this leads to strengthening of beams and columns, typically by
to these four columns, with results for frames of intermediate aspect first considering discretionary removal of a single column. Column
ratios being obtained by quadratic regression. Same regression curves removal leads to load redistribution, which controls the amount of
are employed later for column strengthening. Naturally, the tallest strengthening required. Strengthening columns has a dual effect: it re­
buildings have more sturdy columns, which suffer smaller damage under duces the probabilities of cascading progressive collapse failures, but it
the same blast loading. also increases the local resistance against the blast loading. Therefore,
the hazard analysis needs to be updated for the strengthened columns,
4. Overview of the blast hazard progressive collapse analyses characterizing the loop. The strengthened frame is then subject to pro­
gressive collapse analysis. If no further strengthening is required, the
The blast-hazard and progressive collapse analyses performed herein loop is exited. If further strengthening is required, this will involve for
follow a typical design-verify-strengthen cycle of structural analysis, as instance discretionary removal of two columns.

Table 2
Hazard scenarios considered.
Hazard case Range R(m) Charge Wmass APM strengthening options studied
(kg) of ANFO

Benchmark VBIED 10 1000 nAPM


r,c = 1 and nAPM
r,c =2
VBIED with Reduced W 10 500 nAPM
r,c =1
Suitcase bomb 3.5 50 nAPM
r,c =1

4
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

While the terrorists may, for example, plan for a 1000 kg charge, in
reality there may be weighing and mixing difficulties resulting in a mass
tolerance of ± 10%. If Wmass is normally distributed, then the COV of
Wmass is 0.05.
The effectiveness (energetic output) of explosives is measured in
terms of TNT equivalency. As ANFO manufactured by terrorists is
typically home-made (as there are restrictions on purchasing materials
for commercial grade ANFO in many countries), the TNT equivalency of
home-made explosives can be quite low, especially as terrorists in
Western countries tend to be self-starters or lone-wolves (e.g. [16,18])
with no access to commercial or military explosives, or to high grade
bomb making materials, and often lack any training, and lack the op­
portunity to test their product [19]. Hence, it is assumed herein that
WTNT is modelled as a triangular distribution with lower bound of 0.6 (a
poorly built IED), upper limit and mode equal to 0.82 (a well-built IED).
This results in a mean value of WTNT = 0.75, i.e., 1000 kg of ANFO has
Fig. 5. Layout of IED and building RC columns. Dashed line around the lower an average TNT equivalency of 750 kg of TNT.
3rd column indicates influence area for this column. For a more detailed discussion of terrorist threats and the quantifi­
cation of R and W see Stewart [20–22].
The design-verify-strengthen loop in Fig. 4 is done throughout the
manuscript, treating in a batch approach the seven frames described in 5.2. Conditional probabilities of column loss, and damage indexes
Section 5, the three blast hazards described in Section 5, and considering
different hazard probabilities pH . The effect of explosive blast loading on damage and failure proba­
bilities of structural elements is a complex phenomenon. Damage to RC
5. Blast hazard analysis columns can be modelled by pressure-impulse (P-I) curves developed by
Mutalib and Hao [23]. The level of damage is represented by a damage
5.1. Terrorist blast hazard scenarios index defined as:
Rdesign − Rdamage
For sake of illustration, three hypothetical blast hazard scenarios are DI = (5)
Rdesign
assumed, as summarized in Table 2. These represent a range of terrorist
threats. A VBIED being a favoured mode of attack on large buildings or
where Rdesign is the axial load carrying capacity of the undamaged col­
public spaces where the minimum distance the vehicle may get to a
umn, and Rdamage is the axial load carrying capacity of the column due to
building is assumed as 10 m (due to fencing, bollards, or other protective
blast damage. If the column is undamaged by the blast then DI = 0. The
security). A suitcase IED represents a smaller IED, but since it is person
derived P-I curves are based on numerical modelling that considered
delivered it may be able to get closer to the building, in this case, 3.5 m.
shear and flexural damage. The probability of damage DI given an
The range (R) and explosive mass (W) are taken as random variables.
explosive hazard is a function of the random variability of column width,
The probabilistic distributions of these variables are heuristic and
column depth, column height, concrete compressive strength, reinforc­
selected to help illustrate the uncertainty and variability of predicting
ing yield strength, longitudinal and transverse reinforcement ratio, and
the placement, mass, and quality of IEDs.
P-I curve model errors MEPo and MEIo (see Table 3). Airblast variability
for peak reflected pressure and impulse is dependent on the random
5.1.1. Range
variability of explosive mass and range (see Section 5.1). Added to this
The VBIED location will be variable in x and y directions (see Fig. 5).
are blast loading modelling (air blast) errors MEP and MEI developed by
The variability of range is measured along the x-axis. The range is
Stewart et al. [24] where the predictive model for blast loading is taken
assumed to be a uniform distribution − 0.5 m to + 0.5 m from mean.
from the Kingery and Bulmash [25] model widely used for structural
This represents the fact that even though a terrorist will aim to get the
design and damage assessment (e.g., [26]. More recently, Stewart [27]
IED as close to a perimeter fence or bollard, it may be ± 0.5 m from the
showed that blast loading model errors are not statistically independent
desired location. Moreover, in a car the IED may be placed on one side of
for adjacent building columns. There is air blast spatial variability with
a vehicle, or maybe in the middle, etc. There is also uncertainty about
correlation coefficients ρP = 0.45 and ρI = 0.58 for pressure and im­
where along the building the IED will be placed. A reasonable assump­
pulse, respectively. For more details see Mutalib and Hao [23] and
tion is it will be desired to have it centrally placed (i.e., directly in front
Stewart [27].
of the middle column) with a 95% tolerance of ± 2 m leading to a
standard deviation of σy = 1.0 m assumed normally distributed.
Table 3
Statistical parameters for analysis of conditional column loss probabilities, P[LD|
5.1.2. Explosive mass
H] (adapted from [27]).
The equivalent explosive mass is estimated as
Mean COV Distribution
W = WTNT × Wmass (4)
Column width and breadth 1.005 0.04 Normal
Column height 1.00 0.03 Normal
where WTNT is the TNT effectiveness factor, and Wmass is the actual Longitudinal reinforcement 1.02 0.05 Modified Lognormal
(charge) mass of explosives. Transverse reinforcement 1.03 0.06 Modified Lognormal
The terrorist blast scenarios involve IEDs containing home-made Concrete compressive strength 1.00 0.15 Normal
ANFO (ammonium nitrate fuel oil). For the benchmark case, Wmass is Yield stress 1.15 0.05 Normal
Model Error for axial capacity (ME) 1.01 0.08 Normal
assumed as 1,000 kg of ANFO, 500 kg for a smaller VBIED, and 50 kg for Model Error for P-I curves (MEPo , MEIo ) 1.00 0.05 Normal
a suitcase bomb. The latter may be seen as a large amount of ANFO for a Dead load - D 1.05 Dn 0.10 Normal
suitcase; however, ANFO has a bulk density of 840 kg/m3 meaning a Arbitrary point-in-time live load -Lapt 0.5 kPa 0.72 Gamma
large mass can be concealed inside a typical 60 L suitcase. Airblast Model Error (MEP , MEI ): [24] Lognormal

5
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Table 4 failure probability is a high 91% and 39% for these columns, respec­
Computed failure probabilities P[LD|H] for frame columns and three hazards. tively. For the (4 × 16) frame, with significantly smaller RC cross-
Hazard scenario Frame 1st Column 2nd Column 3rd Column sections, the DI is 1.0 and 0.96 for the 1st and 2nd columns respec­
tively. The corresponding failure probabilities are both 100%. However,
(16 × 4) 0.91 0.39 7.5 × 10− 6

Benchmark VBIED (11 × 6) 0.99 0.61 3 for a lesser 500 kg VBIED the DI and P[LD|H] both reduce as expected.
1.0 × 10−
(6 × 11) 1 0.67 1.1 × 10− 3 Similar trends are also observed for the 50 kg suitcase bomb. The trends
(4 × 16) 1 0.87 4.4 × 10− 3 in mean DI are shown in Fig. 7. As expected, damage indices are smaller
(16 × 4) 0.16 3.1 × 10− 3 0 for sturdier columns of the taller frames and for columns at a greater
VBIED Reduced W (11 × 6) 0.37 1.8 × 10− 2 0 distance from the IED (Fig. 7).
(6 × 11) 0.44 2.0 × 10− 2 0 Conditional column loss probabilities (Table 4 and Fig. 6) and
(4 × 16) 0.75 7.8 × 10− 2
2.0 × 10− 6
damage indices (Table 5 and Fig. 7) show that the increasing order of
(16 × 4) 3.7 × 10− 2 1.5 × 10− 7 0
Suitcase bomb (11 × 6) 0.25 5 0
blast load damage is Suitcase bomb, Reduced Charge W VBIED and
1.7 × 10−
(6 × 11) 0.42 7.5 × 10− 5 0 Benchmark VBIED, with one exception. For the weaker columns of the
(4 × 16) 0.8 7.8 × 10− 4 0 lower (4 × 16) building, the suitcase bomb results in larger conditional
column loss probabilities and larger damage index, in comparison to the
Reduced Charge W VBIED. This occurs because the standoff distance for
The probability of failure for a column i conditional on a hazard H is: the suitcase bomb is smaller (3.5 m against 10 m of the VBIED), and this
[ ( ) ]
P[LD|H] = P[Gi (X)⩽0] = P Rdesign (1 − DI) − D + Lapt ⩽0 (6) appears to have a greater effect on weaker columns.
As illustrated in Fig. 4, strengthening the columns against damage
where Rdesign is the undamaged axial strength for a column ([28]), DI is propagation also influences their direct resistance against blast loading.
the damage index, D is the actual dead load effect, Lapt is the arbitrary This leads to a reduction of conditional failure probabilities and of
point-in-time live load effect for a single column, and Gi (X) is the limit damage indices. This strengthening is addressed in Section 8.5, as it
state function equal to resistance minus load where failure occurs if applies only to those frame and hazard combinations for which
Gi (X) ≤ 0. For more details on the structural reliability analysis for blast-
induced damaged columns see [27].
Table 5
Conditional column loss probabilities P[LD|H] computed for the three
Computed mean and C.O.V. of Damage Index for frame columns and three
hazards, for up to three adjacent columns (Fig. 5), are presented in hazards.
Table 4. As illustrated in Fig. 6, failure probability trends are as ex­
Hazard scenario Frame 1st Column 2nd Column 3rd Column
pected: they are smaller for the sturdier columns of the taller frames; for
the less intense blast hazards, and for the 2nd and 3rd columns, due to (16 × 4) (0.92,0.06) (0.75, 0.20) (0.098, 0.83)
Benchmark VBIED (11 × 6) (0.98,0.034) (0.85,0.12) (0.13,0.79)
larger stand-off distance and angle of incidence from the IED. Note that
(6 × 11) (0.996, 0.014) (0.91,0.07) (0.21,0.7)
failure probabilities for the seven frames of Fig. 6 are evaluated by (4 × 16) (1.0,0.0) (0.96,0.050) (0.34,0.57)
quadratic regression, from the four columns of Fig. 3 (values in Table 4). (16 × 4) (0.61,0.30) (0.22,0.57) (0,0)
The same regression curves are later employed for the strengthened VBIED Reduced W (11 × 6) (0.78, 0.18) (0.31,0.58) (0,0)
columns (Section 8.5). (6 × 11) (0.87,0.083) (0.51,0.44) (0,0)
(4 × 16) (0.93, 0.058) (0.69,0.26) (0.06,0.95)
The damage index statistics for the frame columns and three hazards
(16 × 4) (0.32, 0.65) (0,0) (0,0)
are shown in Table 5. It is observed that for the benchmark VBIED Suitcase bomb (11 × 6) (0.59,0.45) (0,0) (0,0)
located centrally in front of column 1 for the (16 × 4) frame the damage (6 × 11) (0.83,0.13) (0,0) (0,0)
index is a very high 0.92 (i.e., loss of 92% of capacity), and a high 0.75 (4 × 16) (0.96, 0.04) (0.024, 3.10) (0,0)
for the adjacent columns. As expected, Table 4 then shows that the

Fig. 6. Conditional column loss probabilities P[LD|H] for all buildings (e.g., columns of Fig. 3) for the three threats considered (1st CL = first column lost, …).

6
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 7. Mean damage index (DI) for all buildings (e.g., columns of Fig. 3) for the three threats considered (1st CL = first column lost, …).

strengthening is found to be cost-effective. 4. Failure of beam-column connections not considered, and


5. Geometric non-linearities are not considered.
6. Progressive collapse and system damage analysis
Slab and infill effects [29–32], compressive arch [33,34] and cate­
6.1. Mechanical model nary actions [35–37] significantly contribute to progressive collapse
resistance. The model of [38] does allow consideration of catenary ac­
The progressive collapse analysis is made using the simplified tion, within nonlinear static APM [39] yet, within the optimization
analytical model of [61]. The model is applicable for steel or RC frames, problem this leads to beams of very small bending strength which do not
and considers plastic bending failure of beams, and (brittle) crushing respect vertical displacement service limit states [4] . Such combined
failure of columns. The model considers regular frames with nc columns effects should be studied with more explicit numerical models
and ns storeys, and initial damage events leading to failure of nr,c col­ [2,40–42]. In summary, due to the above limitations, this study provides
umns of nr,s stories, where subscript r is for ‘removed’. Here, nr,s refers to conceptual insight, and specific numerical results should be considered
the vertical extent of the initial damage. as indicative. Numerical results obtained recently by the authors [42],
The analytical model has some known limitations: considering geometrical and material non-linear FE modelling, have
been confirming the conceptual insight obtained with the simplified
1. Only gravity loads are considered. No lateral loads nor second-order analytical model.
effects; Epistemic uncertainty in the mechanical model was not considered,
2. Slab, infill, and other three-dimensional effects not considered; mainly because hazard epistemic uncertainty will dominate the out­
3. Compressive arch and catenary action effects not considered; comes of the risk assessment. However, this is a topic for future research.

Fig. 8. Cascading progressive collapse events, triggered by removal of n0r,c columns, indicating two sequential local column failures involving four addi­
tional columns.

7
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

In the following, superscript I is employed for the intact structure. 6.3. Strengthening for alternate load paths
Superscripts B and P refer to bending and column crushing (pancake)
collapse. Local column crushing (PL) refers to failure of the columns Three situations need to be accounted for in the formulation for
adjacent to the damaged area, including cascading column failures alternate load paths:
(Fig. 8). Global column crushing (PG) refers to the instantaneous failure
of all columns of the first floor, leading to a pancake-type collapse a. The discretionary initial damage against which the frames are
(Figs. 2 and 8). Furthermore, perfect brittleness is denoted el, whereas strengthened in design stage (APM strengthening);
ideal plastic behavior is written as pl. This notation follows [38] for easy b. The initial damage scenarios the structure will be subject to during
cross-referencing. The entire formulation is presented in terms of ver­ normal service, and;
tical distributed load capacities qu . The plastic hinge moment for beams c. The potential for cascading failure of the structure, in case of initial
is By , and crushing strength of columns is Rc . damage.
For the intact structure, the ultimate strength in bending collapse is:
Equations (8), (10) and (11) can be employed in the three situations
16By
qI,B,pl
u = (7) above. Yet, to differentiate between these situations, nAPM
r,c is used for the
L2
discretionary number of columns removed in APM strengthening (a.); n0r,c
This plastic solution is derived from the kinematic theorem,
is used for the number of lost (collapsed) columns (initial damage sce­
considering a triple-hinge plastic mechanism. In case of loss of nr,c col­
narios) resulting from blast loading (b.); and nr,c is used for the cascading
umns, bending collapse strength is reduced leading to:
failure of columns in damage propagation phase (c.).
4By In the following, it is assumed that any column of the first (ground)
qB,pl
c (nr,c ) = (8)
L2 nr,c floor could be lost due to a blast damage; hence, all columns and all
beams of the first two floors need to be strengthened.
For the intact structure, static crushing of all ground floor columns
Beams and columns are strengthened in such a way as to provide an
occurs when maximum compressive force at base equals total axial
alternative path to the loads supported by the removed elements [1]. For
strength, leading to the global pancake collapse strength:
the beams to bridge over nAPM r,c removed columns, the required bending
qI,PG =
Rc nc
(9) strength is [1,44]:
u
Lns (nc − 1)
L2 nAPM L2 nAPM
In case of initial damage, local and global column crushing (pancake) (14)
r,c r,c
BAPM
y (nAPM
r,c ) = qB,pl
c = (1.2Dn + 0.5Ln )
4ϕ 4ϕ
collapses are possible. In case of local crushing, loads are redistributed to
the two nearest columns. Brittle failure occurs for: Considering load redistribution, required strength for adjacent col­
umns is:
Rc 1
qPL,el (nr,c , nr,s ) = ( ( )) (10) [ ( ( )) ]
c
Lns 2 − nc − 1
+ nr,c 1 −
nr,s
APM APM NLC Lns nc − 1 APM nAPM
r,s
nc ns RAPM
c (nr,c ,n ) = Max R c , 2− +n r,c 1− qP,loc,el
c
r,s ϕ nc ns
In case of loss of nr,c columns, brittle global crushing collapse occurs (15)
for:
with qP,loc,el
c = (1.2Dn +0.5Ln ). In Eq. (15), operator Max[] is used to also
Rc nc (nc − nr,c )
qPG,el (nr,c , nr,s ) = ( ) (11) consider the required column strength under normal loading conditions.
c
Lns (nc − 1)( nc + nr,c ) − 2 nr,s nr,c nc
ns This operator is not required in Eq. (14) because the impact of a lost
column in bending moments is large for L ≥ H.
As shown above, herein we consider ductile failure of beams, and In codified APM strengthening, ϕ = 1 is recommended [45], with no
brittle failure of columns, which is representative of RC frame building distinction for beams and columns. To address optimal design within an
behavior. The Dynamic load Amplification Factors (DAF) as a result of APM framework, herein we introduce independent strengthening fac­
sudden column removal included in Eqs. (8), (10) and (11) are DAF = 1
tors for beams (λB = 1/ϕAPM B ) and columns (λC = 1/ϕAPM C ), such that
for beams, and DAF = 2 for columns, following [38]. This is closely
λB = λC = 1 recovers the code-recommended strengthening [1,45], with
related to the recommended DAF values for deformation-controlled
ϕAPM = ϕAPM = ϕ = 1.
plastic failure of beams, and load-controlled brittle failure of columns B C

[1]. No uncertainty is considered in DAF values [43]. Other damage In the following risk-based optimization, strength equations (Eqs. (7)
modes may also be important, such as to beam-column connections, but to (11) are computed with By replaced by By = λB BAPM y , and with Rc
this is a topic for further research. replaced by Rc = λC RAPM
c , where BAPM
y and RAPM
c are the initial values of
beam plastic hinge moment and column crushing capacities, respec­
6.2. Conventional design under normal loading conditions tively, resulting from APM strengthening with discretionary removal of
nAPM
r,c columns.
For conventional design under normal loading conditions, the Strengthening beams and columns of a frame to support removal of
required beam strength is [28,44] : one or two columns leads to increase in bending moment and axial load
L2 I,B,pl L2 capacities. The APM design strengthening factors are obtained by
BNLC
y = q = (1.2Dn + 1.6Ln ). (12) dividing the design capacity of the strengthened element, by the design
16ϕB u 16ϕB
capacity from conventional design. For the frames addressed herein, and
where superscript (•)NLC refers to the normal loading condition, and for λB = 1, the resulting strengthening factors for beams, with removal
ϕB = 0.85. The required axial column strength is: of a single (nAPM
r,c = 1) or of two columns (nAPM
r,c = 2), are:

Lns (nc − 1) I,P Lns (nc − 1) ⃒
RNLC = qu = (1.2Dn + 1.6Ln ) (13) BAPM
y (nAPM ⃒
r,c )⃒
c
ϕC nc ϕC nc SF B = NLC ⃒ = 2.06
By ⃒
nAPM
r,c =1
where ϕC = 0.65, due to the brittle crushing failure mode of columns.

8
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Table 6
Required percentage increase in column axial strength, for updated hazard analysis.
Aspect ratio (ar) Number of stories × bays (ns × (nc − 1) ) *Required % increase in RD with:
nAPM
r,c = 1, DAF = 2 nAPM
r,c = 2, DAF = 2

Tall frame (16 × 4) 28.0 116


Intermediate (13 × 5) 40.0 120
Intermediate (11 × 6) 48.0 125
“Square” frame (8 × 8) 53.0 131
Intermediate (6 × 11) 55.0 139
Intermediate (5 × 13) 58.5 147
Low frame (4 × 16) 62.0 157

* Evaluated by FE analysis for damaged frames, with DAF × (1.2Dn + 0.5Ln ), w.r.t. conventional design (1.2Dn + 1.6Ln ), assuming the floors above have not collapsed
(Fig. 2).


⃒ CCF (ar) is the cost factor for columns normalized to conventional beam
BAPM (nAPM ⃒
r,c )⃒
SF B =
y
⃒ = 4.13 (16) construction cost. Values of CCF (ar) for the different frames are shown in
ByNLC

nAPM
Table 1. Hence, the total construction cost for each frame is:
r,c =2

CREF (ar) = Lns (nc − 1) + Hns nc CCF (ar) (20)


where SF refers to design strengthening factor, and subscript (•)B refers
to beams. The resulting design strengthening factor for columns is given As suggested in Eq. (16), the cost of strengthening beams to bridge
by Eq. (17), with numerical values given in Table 6 for λC = 1. over a lost column is higher than the cost of strengthening columns.
Strengthening beams usually involves increasing the reinforcement
RAPM (nAPM
r,c ) ratio, or the amount of steel. Hence, the additional cost of strengthening
(17)
c
SF C =
one meter of beams for bridging over nAPM removed columns, is pro­
RcNLC
r,c
portional to:
6.4. Limit states for probabilistic progressive collapse analysis
CstB (λB ) = Max[0, λB αB SF B + (1 − αB ) − 1] (21)
For the intact structure, the limit state function for bending and
where αB is participation factor of steel cost in total construction costs of
global pancake collapse is:
a beam. Following [4], herein we consider αB = 0.7, assuming the cost
gI (λB , λC , X) = RI rI (λB , λC , ⋯) − D − L50 (18) of steel to be 70% of a RC beam cost, with the other 30% given by cost of
concrete, formwork, workmanship, etc. In Eq. (21), operator Max[⋅] is
where rI (⋅) is a deterministic strength function for the intact structure, RI used such that strengthening costs are not smaller than zero for λB →0, so
is a non-dimensional random variable describing uncertainty in strength that conventional design is not bypassed.
of the intact structure, including model error, D is the dead load, L50 is Accordingly, the additional cost of strengthening one meter of frame
the fifty-year peak live load, and X is the vector of random system pa­ columns is proportional to:
rameters. The strength rI (⋅) for bending failure is given by Eq. (7), and
for global pancake by Eq. (9). CstC (λC ) = Max[0, λC αC SFC + (1 − αC ) − 1] (22)
In case of localized initial damage, the limit state function is given
where αC is the participation factor of steel cost in strengthening a frame
by:
column where αC depends on concrete strength, load eccentricity and
( )
gLD (λB , λC , X) = RD rD λB , λC , nr,c , nr,s ⋯ − D − Lapt (19) other factors. Based on [4], herein we consider αC = 0.5 in all analyses,
as we understand there is greater contribution of concrete to strength of
where rD (⋅) is a deterministic strength for the damaged structure, RD is a columns. The impact of varying strengthening cost factors α is investi­
non-dimensional random variable describing uncertainty in strength of gated in [4]. Cost factors α affect the threshold column loss probabilities
the damaged structure, including model error, and Lapt is the sustained which are discussed in [4]. Larger strengthening cost factors imply in
component of live load. In Eq. (19), the strength rD ( • ) for bending is Eq. larger threat probabilities to justify APM strengthening of the frames to
(8), for local crushing is Eq. (10), and for global pancake failure is Eq. resist local damage.
(11). The above equations are valid for any number of removed columns, The total construction cost is:
which includes the initial triggering damage (n0r,c ), as well as cascading [ { } ]
1 L(nc − 1) ns + nreinf ,s CstB (λB )
progressive failure of columns. Cconst. (λB , λC ) = (23)
CREF (8 × 8) +Hnc CCF (ar){ns + nreinf ,s CstC (λC )}
Herein, approximate failure probabilities are computed based on the
Cornell reliability index, as the limit state functions are linear in the where nreinf,s is the number of stories for which beams and columns will
random variables, but also by approximating the distributions of L50 and be strengthened.
Lapt as Gaussian. Formulation for Cornell reliability indexes and data for Eq. (23) is normalized by CREF (8 × 8) such that the “square” frame is
the probabilistic progressive collapse evaluation are presented in [4]. the reference for frame construction cost with no strengthening (nreinf,s =
0). Table 1 shows the cost factors for all frames addressed herein. By
6.5. Cost functions: construction, strengthening and expected costs of considering the (8 × 8) frame construction cost as unitary, collapse costs
failure can be related to construction costs by a simple multiplication factor k.
Herein, we assume that any column of the first floor of the buildings
6.5.1. Construction and strengthening costs is a potential target for the bombing. Therefore, when the decision is to
Since all beams of all frames are the same size, the elementary strengthen the frames to produce alternate load paths, we require
building block for construction costs is the unitary cost of a RC beam strengthening of all columns and all beams of the first two floors; hence,
(CB = 1, per meter), in arbitrary units (Dollar, Euro, etc.). The cost of nreinf,s = 2. This goes inline with recent findings of [41], that optimal
columns depends on frame aspect ratio, as taller frames have more strengthening reinforcement should be limited and placed in the first
expensive columns. It is given by CC = CB CCF (ar) (per meter), where

9
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

story above the lost column. Other strengthening decisions are studied design factor for beams (0.85 0.65 = 1.3), which suggests that the brittle/
in [4]. ductile cost ratio considered is realistic. Following the arguments in [4],
In Eq. (21), cost of strengthening beams depends on factor λB αB SFB . in Eq. (25) we consider kductile = 40 for plastic beam bending failures
Similarly, cost of strengthening columns depends on factor λC αC SFC , Eq. (j ∈ {B}), and kbrittle = 80 for brittle crushing failure of columns (j ∈ {PL,
(22). In the risk optimization, a continuous transition is obtained be­ PG}). For the direct consequences of local damage in Eq. (24), we
tween the decisions to strengthen, or not to strengthen the frames, by consider kdirect = kbrittle , since blast loading produces immediate load-
considering λB and λC as design variables. If a particular frame should controlled failure of beams and columns. With the above figures, if for
not be strengthened, the algorithm will reduce λB and λC , compensating example building cost is $50 million, and there is global pancake
for the strengthening factors SFB and SFC . Eventually, for λB = 1/SFB and (brittle) collapse, then total failure cost is expected to be $4 billion (for
λC = 1/SFC , the cost of strengthening is zero (see Eqs. (21) and (22)), kbrittle = 80). Yet, if only 30% of the building suffers progressive collapse,
which corresponds to conventional design. Progressive collapse strength then failure costs reduce to $1.2 billion. This roughly accounts for the
equations (Eqs. (7) to (11)) are updated accordingly. Any values λB , λC < fraction of building occupancy directly affected by collapse, plus the cost
1 indicate that codified APM strengthening for load bridging over nAPM r,c of reconstructing the whole building.
removed columns is not economical. By codified strengthening, we
mean use of Eqs. (14) and (15) with λB = λC = 1.
6.6. Evaluation of system damage probabilities
6.5.2. Cost of initial localized damage
A blast event leading to loss of n0r,c columns of a frame produces
The ith local damage state LDi is characterized by loss (collapse) of n0r,c
localized damage, which may propagate upwards due to a plastic beam
columns of n0r,s stories. Herein, we consider n0r,s = 1, and n0r,c ∈ {1, 2, 3}.
bending mechanism, or horizontally, due to column crushing mecha­
For the strengthened frame, and considering that collapsed (lost) col­
nisms. The basic set of damage propagation events is given by {B, PG}.
umns are from strengthened floors, the cost of the ith local damage is: Yet, the local crushing failure of adjacent columns leads to nr,c = n0r,c +
⎡ ( ){ } ⎤ 2, and this may again propagate upwards by beam bending, affecting
L n0r,c + 1 n0r,s CstB (λB )
CD (LDi ) =
kdirect ⎢
⎣ {

}⎦ (24) two additional bays, or lead to global crushing, with nr,c = n0r,c + 2, and
CREF (8 × 8) +Hn0 C (ar) n0 C (λ ) so on (Fig. 8). Hence, the total indirect risk of damage propagation is
r,c CF r,s stC C
given by:
where kdirect is a consequence multiplication factor. For kdirect = 1, the P[LDi |H]
direct consequence of local damage CD (LDi ) is a fraction of whole frame CID (LDi )P[LDi |H] =
CREF (8 × 8)
construction cost, corresponding to the fraction of the frame cross- [ ] (26)
( )
section area affected by local damage. For the frames considered here­ ×P ∪j,n0r,c aj pj kj ˝ + ˝∪n0r,c +2→nFr,c pPL nr,c ∪j aj pj kj , j ∈ {B, PG}
in, and for n0r,c = n0r,s = 1, this fraction varies between 0.025 and 0.033.
where pPL (nr,c ) is the probability of crushing failure of the adjacent
6.5.3. Expected costs of progressive collapse failure columns, which is the main parameter controlling the cascading prop­
The cost of progressive collapse failure is accounted for by means of agation of failure. This term is directly affected by blast damage, by way
cost multiplication factors kj , which relate to the unitary cost of the of the damage indexes shown in Fig. 7. The ˝ +˝ sign in Eq. (26) also
structure, and by the frame frontal area aj affected by each failure mode represents a union of events. In Eq. (26), nFr,c is the maximum number of
(Fig. 8). Hence, the expected cost of progressive collapse failure, for the symmetric cascading local column failure events, given by nFr,c =
ith initial damage state, is given by: floor[(nc − n0r,c + 1)]/2. This corresponds to initial damage at the central
1 [⋃ ] column of the frame.
CEFi = CID (LDi )P[LDi |H] = P[LDi |H]P pj aj kj ,j ∈ {B,PL,PG} The unions of events in Eq. (26) correspond to series systems, with
CREF (8×8) j

(25) failure consequences accounted for each event. Evaluation of the prob­
ability in Eq. (26) involves the probability of occurrence of individual
where {B, PL, PG} represents the set of possible damage propagation events, as well as dual and multiple intersections [46]. As damage
events (bending failure, local and global column crushing). Cascading propagates by means of local column crushing, the impacted frame area
damage propagation is addressed in the next section. By considering increases, but the event probabilities decrease, especially due to the
collapse failure costs proportional to the affected frontal area (aj in Eq. conditioning on local column failure, controlled by conditional proba­
(25)), we assume failure costs to be proportional to building occupancy bility pPL (nr,c ). Importantly, the failure events in Eq. (26) involve a su­
(equipment, people), occupancy to be uniform, and total losses to be perposition of frame areas, but each frame element can fail only once. If
significantly higher than building construction costs. the individual event probabilities were added, accurate intersections
Failure cost factors kj in Eq. (25) have been investigated in [4], based would have to be discounted. As it is difficult to accurately compute such
on results by [3] which relate the cost of the structural frame, to total intersections, we employ the max[.] operator to evaluate the probabilities
construction cost of different buildings. The analysis in [4] compares of event unions. Hence, among all terms in Eq. (26), only the system
cost of construction to the cost of collapses for the Alfred P. Murrah damage state leading to maximum risk is considered.
Federal Building, 9/11 World Trade Center and Pentagon, and considers
cost multipliers considered in studies such as [8]. The analysis by [4] 7. Risk-based cost-benefit analysis
results in failure cost factors in the range kj ∈ (40, 80).
One relevant point in a structural risk analysis is to distinguish be­ The risk-based cost-benefit analysis is composed of the initial dam­
tween the more severe consequences of brittle failure of columns, from age and damage propagation events, their occurrence probabilities, and
the ductile failure of beams. [4] use kbrittle = 2 kductile , and they arrive at the cost terms discussed in Section 6. The solution leading to minimal
very realistic relation between the design factor for columns and the total expected cost is found by solving the following optimization
problem:

10
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Given pH and for a given blast loading scenario, probability of losing the first column.
find:(λ*B , λ*C ) In Section 8.4 we investigate optimal reliability indexes, in terms of
which minimizes: hazard probabilities pH . This helps to interpret how the optimum solu­
tions adapt to increasing hazard probabilities. In this section we also

CTE (λB , λC ) = Cconst. (λB , λC ) (a)



+ SDS CID (SDS, NLC)P[SDS|NLC]pNLC (b) (27)
⋃3
+ i=1 [USDS CID (SDS, LDi )P[SDS|LDi , H] + CD (LDi ) ]P[LDi |H]pH50 , (c)

where pH50 = 1 − (1 − pH )50 is the fifty-year probability that the struc­ observe that the break-even hazard probability (pth H ) is associated with
ture is exposed to a threat with annual hazard rate pH . In Eq. (27), line beam bending reliability indexes becoming zero, which aids in inter­
(b) corresponds to the expected cost of failure under normal loading preting optimal strengthening factors discussed in Section 8.7.
conditions. Bending and global pancake collapse failures are considered From Section 8.5 onwards we consider a fixed hazard probability
for the intact structure, such that line (b) becomes: large enough to justify strengthening all frames addressed herein, for the
three blast hazards. Hence, in Section 8.5 we address the increased local
Lkductile [ ] [ ]
Φ − β50 50
B (λB ) + CPG Φ − βPG (λC ) (28) resistance of the strengthened columns against the blast loads. In Section
CREF (8 × 8)
8.6 we study how optimal reliability indexes vary in terms of frame
aspect ratio, and in Section 8.7 we discuss optimal strengthening factors,
where β50 B is the 50-year reliability index for plastic collapse of beams,
with relevance to codified APM strengthening.
and β50PG is the 50-year reliability index for global pancake collapse, both
calculated from Eq. (18). Although all beams are identical, it is assumed
that under normal loading conditions only one span of one floor would 8.2. Frame strengthening for alternate path method
fail, due to (partial) independence of live loads in different spans and
floors. The cost of a global pancake failure is CPG = Cconst. (λB , λC )kbrittle . As observed in Fig. 6, a terrorist bombing blast hazard can lead to
3
The ∪i=1 [⋅] union in line (c) of Eq. (27) corresponds to the different loss of one, two or three columns. For the benchmark VBIED case, the
initial damage states which may be caused by blast loading. Considering probability of losing a second column, as direct result of the blast, is
the conditional failure probabilities in Fig. 6, and the mean Damage significant, and of the same order of magnitude as the probability of
Indexes in Fig. 7, herein we consider initial damage leading to loss of one losing the first column. Hence, for this case it is important to consider
to three columns. Hence, the ith local damage state LDi is characterized the options of removing one or two columns in the APM strengthening.
This choice relates to use of design codes (Eqs. (14) and (15)), as well as
by loss of n0r,c columns of n0r,s stories, where n0r,c ∈ {1, 2, 3} and n0r,s = 1.
to the risk optimization. As will be shown in the sequence, the risk
For all blast scenarios considered herein, the conditional probability of
analysis can “partially compensate” for an inappropriate choice of nAPM
r,c .
losing four columns as a direct result of the blast is insignificant and can
be neglected. Yet, the optimal beam and column design factors, λ*B and λ*C , are related
Note that the sums over system damage states in Eq. (3), SDS [⋅],
∑ to the discretionary column removal option considered.
have been replaced by unions of events in Eq. (27), since each element of Column strengthening has a dual role in the progressive collapse
the frame can fail only once. Also in Eq. (27), the union of events is analysis under blast loading. The first and immediate role discussed in
handled by means of the max[⋅] operator. Section 6.3 is the axial strengthening required for the adjacent columns
to absorb the load redistributed from failed column(s). The second role is
8. Numerical results to increase column resistance against blast loadings, reducing the term
P[LD|H] addressed in the Hazard Analysis (Section 5). Yet, this increase
8.1. Summary of results and problem variants investigated in column strength against blast loading does not apply when the
optimal decision is not to strengthen the frames. Hence, an interactive
When considering blast hazard threats, the first relevant design analysis (Fig. 4) would be required to find the break-even hazard
question is whether a particular threat is significant to justify the probability pth th
H , while considering the strengthened columns for pH > pH ,
spending in strengthening beams and columns of a particular structure. and the unstrengthened columns for pH < pth H . Herein, to avoid the
By considering the hazard probability pH as the independent parameter interactive analysis, this secondary column strengthening effect is
in the risk-based formulation, this question is addressed by finding the initially neglected in the risk analysis for pth
H (Sections 8.3 and 8.4). This
break-even point (pth H ), or the point where the two alternatives effect is considered, however, in the analysis of optimal reliability in­
(strengthening for APM or conventional design) have the same total dexes and optimal safety factors resulting from the risk analysis (Sec­
expected cost. This decision also relates to the choice on the number of tions 8.6 and 8.7), for a hypothetical threat above the break-even point,
columns to be removed in the APM strengthening, given by nAPM r,c . As
that is, for a hazard probability which justifies strengthening all frames
observed in Fig. 6, for the Benchmark VBIED blast the probabilities of studied herein.
losing one or two columns are significant. Hence, the options nAPM
r,c =1
and nAPM
r,c = 2 are considered for the Benchmark VBIED threat, starting in 8.3. Objective function and hazard probability thresholds
Section 8.2. This is followed by a discussion of the break-even hazard
probabilities pth
H in Section 8.3. For the Reduced W and Suitcase hazard
8.3.1. Objective function
cases, the intuitive choice is nAPM = 1, as the probabilities of losing a Figs. 9 and 10 illustrate the total expected cost (objective functions)
r,c
second column in the blast are one order of magnitude smaller than the in Eq. (27) for the Benchmark VBIED case, for the “square” frame (ar =
8 × 8), considering strengthening options with a single (nAPM r,c = 1,

11
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 9. Total expected cost (objective function) for the benchmark VBIED case, nAPM
r,c = 1, ar = (8 × 8), showing an annual hazard threshold of (pth
H )OPT = 3.9 ×
10− 4 , for which two alternative optimal solutions are obtained: one with λB ≈ 0.48, corresponding to conventional design (no strengthening), and one for λB ≈ 1.43,
corresponding to APM strengthening of beams to bridge over a single lost column.

Fig. 10. Total expected cost (objective function) for the benchmark VBIED case, nAPM
r,c = 2, ar = (8 × 8), showing an annual hazard threshold of (pth
H )OPT = 3.3 ×
10− 4 , for which two alternative optimal solutions are obtained: one with λB ≈ 0.24, corresponding to conventional design (no strengthening), and one for λB ≈ 0.73,
corresponding to APM strengthening of beams to bridge over two lost columns.

Fig. 9) and with two (nAPMr,c = 2, Fig. 10) columns removed. The figures (Fig. 10) relate to APM strengthening, following Section 6.3. The hazard
show total expected cost as a function of λB , for λC at the optimal value probability value characterizing this transition, or cost break-even point,
λ*C . Four total expected cost functions are shown, for pH going from small is called the hazard probability threshold, pth H . In Figs. 9 and 10, we use
to large. (pth
H )OPT to refer to numerical values evaluated from the risk-based
For one particular value of pH in each figure, two points of local optimization.
minima are observed, with almost the same total expected cost. This It can be observed in Figs. 9 and 10 that the risk-based optimization
occurs due to a transition, from the optimal solution corresponding to can “compensate” for an inappropriate decision on the discretionary
conventional design (no strengthening) to the optimal solution being an number of removed columns (nAPM r,c ) against which the frames are
APM strengthened design. The beam strengthening factors λB ≈ 0.48 ≈ strengthened. Due to the large probability of losing two columns in the
1/SFB (Fig. 9) and λB ≈ 0.24 ≈ 1/SFB (Fig. 10) correspond to conven­ Benchmark VBIED blast event (see Fig. 6), strengthening the frame for
tional design. The strengthening factors λB ≈ 1.43 (Fig. 9) and λB ≈ 0.73 nAPM
r,c = 2 seems to be a more intuitive decision, which leads to λB ≈ 0.73.

12
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 11. Total expected cost for the benchmark VBIED case, showing three cost break-even points, for three code-based solutions: conventional design, and APM
strengthening for nAPM
r,c = 1 and nAPM
r,c = 2, frame with ar = (8 × 8).

Fig. 12. Annual hazard thresholds (pth APM


H )CD , in terms of frame aspect ratio, for the three blast threats, nr,c = 1 unless stated otherwise.

Yet, if the frame is strengthened for nAPM


r,c = 1 only, this is compensated collapse would be justified for hazards leading to pH > pth
H.
by increasing the (optimal) beam strengthening factor to λB ≈ 1.43. This Conventionally, the decision about strengthening a frame or not is
represents a 43% increase in the APM strength factor for beams, leading based on use of design codes, such as [1,44,45]. Yet, the hazard
to an overall SFB = 2.06 × 1.43 = 2.95 (see Eq. (16)). These values also threshold values observed in Figs. 9 and 10 are computed from a risk-
show that the optimal solution is an intermediate one, with nAPM
r,c = 1 but
based optimization; hence, these values do not directly relate to codi­
43% higher beam strength, or nAPM = 2 but 27% smaller beam strength. fied design. Fig. 11 compares the total expected costs obtained from use
r,c
of codes, for conventional design and for APM strengthening with λB =
This intermediate solution is addressed in Section 8.5.
λC = 1, with those resulting from the optimization. Three cost break-
even points are shown in Fig. 11. Point A is the break-even between
8.3.2. Hazard probability break-even point from codified design
costs of conventional design and codified APM strengthening consid­
The hazard threshold values pth
H shown in Figs. 9 and 10 correspond
ering nAPM
r,c = 1. Point B is the break-even between costs of conventional
to the break-even points, for which a conventional design and an
alternate load path design would have roughly the same total expected design and codified APM strengthening considering nAPM
r,c = 2. Point C is
costs. The conventional design would be cheaper to construct, but with the break-even between costs of APM strengthening considering nAPM
r,c =
higher associated expected costs of progressive collapse failure. The 1 and nAPM
r,c = 2. Point A corresponds to Fig. 9, and Point B corresponds
APM strengthened design would be more costly to construct, but with a to Fig. 10. Points A and B have been identified and studied previously
corresponding reduction in expected costs of failure. Hence, frame [4–6], in hazard-independent analyses. Herein, points A and B are dis­
strengthening as a measure to reduce the likelihood of disproportionate cussed within a hazard-dependent analysis. Point C is identified for the

13
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

first time in this manuscript and is discussed later in the paper (Section hazard probabilities pH . For large hazard probabilities (left in Fig. 13),
8.5 onwards). optimal β*B′s are quite large, around β*B ≈ 3.7; this occurs because beams
The concept of a break-even point is the same for codified design and were strengthened considering removal of two columns (nAPM = 2), but
r,c
in risk optimization. Yet, the numerical values are not the same, but of Fig. 13 shows reliability indexes conditional on loss of one column.
similar magnitude. Hence, for hazard thresholds computed by cost Around pH = 7 × 10− 4 , for low frames, and around pH = 10− 4 , for
comparison from codified design (as in Fig. 11), we use label (⋅)CD for tall frames, the optimal reliability index for beams drops significantly,
Codified Design. Yet, when referring to the concept, we continue to use eventually becoming zero. Since this is the optimal bending reliability
simply pthH . Note that the total expected cost for the optimal solutions for beams strengthened to bridge over two removed columns, this sharp
(blue dashed line in Fig. 11) is always below the cost functions for the drop corresponds to the two equivalent solutions shown in Fig. 10,
codified designs. This occurs because the optimization can fully repre­ which is also related to the threshold hazard probabilities shown in
sent the entire spectrum of possible design factors, leading to the Figs. 11 and 12. For smaller pH , the drop in bending reliability reveals
optimal design factors for any hazard probability. that it is not optimal to strengthen the frames to bridge over two
Fig. 12 illustrates the annual hazard thresholds, (pth
H )CD , for the three removed columns. The optimal reliabilities for the column crushing
blast threats studied herein, computed by comparing costs resulting failure modes are more uniform in terms of hazard probabilities, with
from codified design (conventional, and APM with λB = λC = 1, nAPM r,c as some difference for low and tall buildings. Perhaps surprisingly, the
indicated). Fig. 12 also includes a constant pH = 10− 3 per building per optimal reliability of columns of lower buildings is higher than for
year, an arbitrary value chosen for discussion purposes. As shown in higher buildings. This is a consequence of horizontal propagation of the
Fig. 12, for any hazard leading to pH = 10− 3 , frame strengthening column crushing failure modes, which affects a greater frame area for
against the VBIED (Benchmark and Lower W) bombing cases is only the wider buildings. This effect could be partially limited by including
cost-effective for the frames taller than 6 stories (ar ≥ (6 × 11), left in structural fuses (not considered herein) or by considering the natural
Fig. 12), since for these frames, pH = 10− 3 > pth arrest of horizontal failure propagation, which occurs when slabs of
H . As expected, for less
collapsing floors rest on the ground, no longer exerting horizontal pull of
damaging hazards like the suitcase blast, pth H increases as strengthening
adjacent elements (also not considered).
becomes less cost-effective.
In Fig. 12, the cases identified as Lower W w’DI and B. VBIED w’DI
8.4.2. VBIED with Reduced charge W threat
include the damage to adjacent columns, in the progressive collapse
Fig. 14 illustrates optimal reliability indexes conditional on one lost
analysis. As observed in Fig. 6, a blast with Lower W characteristics
column (n0r,c = 1) for beam bending β*B , local β*PL and global β*PG column
(Table 4) produces loss of a first column, with probabilities varying from
0.16 for the stronger to 0.75 for the weaker columns (Fig. 6). The blast failures, in terms of annual hazard probabilities, for the VBIED with
also produces damage to the adjacent columns, which are the most likely Reduced W threat. In comparison to the Benchmark case (Fig. 13), one
to fail, in a cascading collapse. Following Fig. 7 and Table 5, mean observes that optimal bending reliability indexes are smaller (APM
damage index for the second column varies from 0.22 for stronger col­ strengthening is for a single removed column, nAPM r,c = 1), and the drop
umns, to 0.69 for the weaker columns. As observed in Fig. 12, this for smaller pH is not as sharp; hence, the transition between the optimal
additional damage has a significant effect on the optimal decision on solutions involving strengthening or not is also not as clear-cut, as for the
strengthening the frame or not. For a hazard leading to pH = 10− 3 , for Benchmark VBIED case. Optimal reliability indexes for column crushing
instance, strengthening of all frames in Fig. 12 would be cost-effective are uniform for different threat probabilities, with slightly smaller
for Lower W w’DI, if the damage to adjacent columns is considered, as values for larger pH′s (left in Fig. 14).
it should. Results for the Benchmark VBIED w’DI case for the taller Fig. 15 illustrates the impact of considering blast damage to the
buildings with stronger columns are similar. Yet, for the lower buildings adjacent columns, quantified through the Damage Index DI (Fig. 7).
with weaker columns, an unusual result is obtained: strengthening has Fig. 15 addresses the Reduced W threat; hence, it can be directly
an impact on construction costs, which is not enough to reduce the compared to Fig. 14. It is noted that optimal column crushing reliability
probability of progressive collapse, due to high damage index of the indexes increase significantly, indicating the need to strengthen col­
second and third columns. The threshold probability increases, sug­ umns, to compensate the initial damage to adjacent columns, and the
gesting that strengthening is not cost-effective for any hazard proba­ greater probability of horizontal damage propagation by means of local
bility, but this trend may change if the effect of column strengthening on column crushing. In comparison, the impact on optimal bending re­
direct damage is considered, as shown in Section 8.5. For the Suitcase liabilities is opposite: these are reduced, compensating the increase in
blast case, damage to the second column makes no difference to results, optimal column crushing reliabilities, as a result of competition for the
as this damage is very small. strengthening budget. The optimal bending reliability for lower frames
is zero, because it is not worth strengthening those frames (cost of
strengthening to mitigate the probability of progressive collapse is larger
8.4. Optimal reliability indexes × hazard probabilities
than the reduction in expected costs of progressive collapse). This sce­
nario changes, however, when the columns are strengthened against the
8.4.1. Benchmark VBIED case
blast hazard, as shown in Section 8.5. Similar results are obtained for the
Optimal reliability indexes for beam bending β*B (λ*B ), local β*PL (λ*C ) Benchmark VBIED case with damage index, but are not shown for
and global β*PG (λ*C ) column failures are obtained as by-product of the brevity.
risk-based optimization stated in Eq. (27). These indices correspond to
the optimum reliability for each primary failure mode, which depends 8.4.3. Suitcase bomb case
on threat probabilities and frame aspect ratios, as explored herein. Results for the Suitcase bomb case are similar to those obtained for
Optimal reliability indexes reveal how optimal frame designs adapt Reduced W case (Fig. 14) and are omitted for brevity.
to initial damage, considering its occurrence probability. The behavior
of the bending reliability index, in particular, is illustrative of the
transitions observed in Figs. 9 and 10. Fig. 13 shows optimal reliability
indexes conditional on initial damage with one column loss only (n0r,c =
1) for beam bending β*B , local β*PL and global β*PG column failures, in
terms of annual hazard probabilities, for the Benchmark VBIED threat.
The optimal reliability against beam bending is most affected by

14
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 13. Optimal reliability indexes conditional on n0r,c = 1 for beam bending β*B , local β*PL and global β*PG column failures, in terms of annual hazard probabilities, for
the Benchmark VBIED threat, nAPM
r,c = 2.

Fig. 14. Optimal reliability indexes conditional on n0r,c = 1 for beam bending β*B , local β*PL and global β*PG column failures, in terms of annual hazard probabilities, for
the VBIED with Reduced W threat, nAPM
r,c = 1.

Fig. 15. Optimal reliability indexes conditional on n0r,c = 1 for beam bending β*B , local β*PL and global β*PG column failures, in terms of annual hazard probabilities, for
the VBIED with Reduced W threat, nAPM
r,c = 1, considering blast damage to adjacent columns (DI).

15
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 16. Conditional column loss probabilities in terms of (required % increase in) column axial strength (interpolated RD × Log10 (P[LD|H]) curves).

Fig. 17. Optimal beam bending reliability indexes β*B , and those obtained by APM strengthening, conditional on n0r,c = 1, in terms of frame aspect ratio, nAPM
r,c =1
unless stated otherwise.

8.5. Strengthening columns against the blast hard for cross-referencing. For pH = 6 × 10− 3 , APM strengthening would be
economically justified for all frames and blast hazards; hence, it is useful
In the following analyses, we consider a fixed hazard probability to consider the effect of column strengthening in their increased resis­
pH = 6 × 10− 3 , which is above the thresholds2 for all frames and blast tance against the blast hazard (iteractive solution in Fig. 4). This local
hazards considered herein, but otherwise arbitrarily chosen. This hazard strengthening is based on the load redistribution, considering two col­
probability is identified as a dashed green vertical line in Figs. 13 to 15 umn removal scenarios (nAPMr,c = 1 and nAPM
r,c = 2). Since by assumption
the target column can be any column of the first floor, all columns of this
floor are strengthened. Yet, since local column strengthening has an
2 impact on progressive collapse, the numerical results presented in Figs. 9
This is the (pth
H )OPT value computed from risk optimization, which cannot be
to 15 become “outdated”, in the sense of the design-verify-strengthen
directly compared to values in Figs. 11 and 12.

16
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

cycles of Fig. 4. Specifically, local strengthening of columns leads to strengthen a frame for nAPM
r,c = 2 based on large failure probabilities for
smaller conditional failure probabilities due to blast, and smaller dam­ initial conventional column design, but the (over-) strengthened col­
age indices to adjacent columns. This reduces the probabilities of initial umns make this choice inappropriate. This can happen for all blast
damage, given a successful attack, and reduces the probabilities of hazards and frames considered herein. One reasonable decision, and the
progressive collapse. Hence, this changes the risk analysis, producing for one first considered herein, is to strengthen all columns for nAPM = 1,
r,c
instance a translation of threshold hazard probabilities (upwards in
which leads to the 28% increase in axial capacity points identified in
Fig. 12, to the left in Figs. 13 to 15). As a result, for some strengthened
Fig. 16 for the stronger columns. One then updates the progressive
frames pH = 6 × 10− 3 ≈ pth H . This leads to very interesting results, collapse analysis, checking whether any further strengthening is
related to the elementary decision on whether the frames should be required.
strengthened or not. In particular, this will challenge the choice of APM A preliminary risk analysis with the choice above (not shown) re­
strengthening for two removed columns (nAPM r,c = 2), considered in the veals that this local column strengthening is not sufficient for the
Benchmark VBIED scenario. Benchmark VBIED case. Yet, local column strengthening for nAPM r,c =2
Table 6 shows the required increase in codified APM design axial leads to overshooting the optimal solutions. Note that the λB ≈ 0.73
strength of all columns, for nAPM
r,c = 1 and nAPM
r,c = 2, considering dynamic (nAPM = 2) and λB ≈ 1.43 (nAPM = 1) results of section 8.3 were already
r,c r,c
load amplification DAF = 2. These values are obtained considering the pointing to an optimal intermediate solution, although those results
influence area for the lost column(s), (dashed square in Fig. 5), and the were obtained before the local column strengthening. Hence, to avoid
ratio: overshooting, we consider an intermediate local column strengthening
RAPM (DAF × (1.2Dn + 0.5Ln )) for the Benchmark VBIED case, midway between nAPM r,c = 1 and nr,c
APM
=
c
(29)
RNLC
c (1.2Dn + 1.6Ln ) 2. For the taller frame, this leads to a 28%+116%
2 = 72% increase in axial
capacity, as indicated in Fig. 16, with corresponding reductions in P[LD|
Using results in Table 6, the blast hazard analysis (Section 5) is
H] and DI. The progressive collapse risk optimization will still be eval­
repeated for columns of higher axial strength. Quadratic regression
uated for nAPM APM
r,c = 1 and nr,c = 2, as these are the codified strengthening
curves RD × Log10 (P[LD|H]) and RD × DI are constructed, using results in
Tables 4 and 5, and the required additional results to avoid extrapola­ options. Yet, the resulting optimal design factors will reveal, for
tion (up to 116% for the stronger columns of the taller frames). The same instance, that APM strengthening should be made with nAPM r,c = 1 but
curves are also employed to estimate P[LD|H] and DI for those columns with larger strengthening factors for the columns, or for nAPMr,c = 2 but
not investigated in the original blast hazard analyses (Section 5). with smaller strengthening factors. Importantly, herein we are not
Resulting conditional column loss probabilities are shown in Fig. 16. aiming at solving a real structural design problem; instead, we are trying
As observed in Table 6, for an extreme case with nAPM
r,c = 2, axial load to conceptually understand the tradeoffs of structural strengthening to
in adjacent columns can increase between 116 and 157%. This leads to a mitigate progressive collapse.
significant reduction in conditional failure probabilities and in damage
indices. The round markers in Fig. 16 and their accompanying labels 8.6. Optimal reliability indexes × frame aspect ratio
identify the percentage increase in axial strength of the columns at
different points along each conditional failure probability curve. Figs. 17 to 19 show optimal reliability indexes conditional on first
As noted in Fig. 16, for the Benchmark VBIED blast, P[LD|H] for the
column failure (n0r,c = 1) for beam bending β*B (Fig. 17), local column
first lost column (Fig. 6) reduces from 0.91 to around 10− 2 , for an APM
crushing β*PL (Fig. 18) and global column failures β*PG (Fig. 19), in terms
strengthened column for nAPM r,c = 2. For the second lost column, P[LD|H]
of frame aspect ratio. The three figures also show, for comparison pur­
reduces from 0.39 to 5 × 10− 5 . In this scenario, APM strengthening for
poses, the reliability indexes for the same failure modes for λB = λC = 1,
nAPM
r,c = 2 would not be cost-effective, as conditional failure probabilities which corresponds to codified APM strengthening [1,45].
are very low. This is an example of overshooting: one chooses to Results in Fig. 17 may appear surprising at first. However, recall that

Fig. 18. Optimal local column crushing reliability indexes β*PL , and those obtained by APM strengthening, conditional on n0r,c = 1, in terms of frame aspect ratio,
nAPM
r,c = 1 unless stated otherwise.

17
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

Fig. 19. Optimal global column crushing reliability indexes β*PG , and those obtained by APM strengthening, conditional on n0r,c = 1, in terms of frame aspect ratio,
nAPM
r,c = 1 unless stated otherwise.

zero bending reliability indices are associated to the threshold, or break- effective for the stronger columns of the taller frames. For the lower
even pth
H , as observed in Figs. 13 to 15. Fig. 17 shows that for the taller frames, the optimal solutions are obtained for nAPM
r,c = 2. Hence, the
frames exposed to the Suitcase bomb blast, we have overshot the optimal jumps in β*B , observed in Fig. 17 around aspect ratio of (6 × 11), corre­
solutions. Apparently, this overshooting was modest because it did not spond to point C in Fig. 11, i.e., this is the break-even point for solutions
occur for the lower frames. Our interpretation is that for the Suitcase with nAPM APM
r,c = 1 and nr,c = 2. In fact, Fig. 17 shows this break-even point
bomb case, the threshold probability has increased to around
in terms of frame aspect ratio, for a fixed pH = 6 × 10− 3 . Yet, a corre­
pth − 3
H ≈ 6 × 10 , because of column strengthening (Section 8.5). The sponding point exists in terms of hazard probabilities pH , where the
overshooting for taller buildings occurs because the effect of strength­
break-even point for the (6 × 11) frame is pth
H ≈ 6 × 10
− 3
(this would be
ening the sturdier columns (right in Fig. 16) is larger than the effect of
point C in Fig. 11, but for the (6 × 11) frame).
strengthening the weaker columns (left in Fig. 16), for the same nAPM =
r,c In Fig. 17, the horizontal lines correspond to the bending reliability
1. indexes resulting from codified APM strengthening with λB = λC = 1.
For the Benchmark VBIED blast scenario, we observe jumps in
The optimal β*B ’s are always below the values resulting from codified
optimal bending reliability indices β*B . Recall that local column design, suggesting that the resistance partial factor for strengthening of
strengthening was made for the midway 72% increase in axial strength, beams (ϕ in Eq. (14)) could be greater than one (ϕ > 1). Note that the
hence the nAPM
r,c = 1 and nAPM
r,c = 2 risk-based solutions are built from the jump of Benchmark VBIED solutions, for nAPM = 1 and nAPM = 2, is
r,c r,c
same initial condition. For the taller frames, the curves for nAPM
r,c = 1 and compatible with the corresponding codified solutions.
nAPM
r,c = 2 both reveal that strengthening for nAPM
r,c = 1 is a better solution. The Benchmark VBIED scenario considering damage to adjacent
Again, this occurs because the local strengthening of columns is more columns (B. VBIED w’DI in legend of Fig. 17) leads to optimal bending

Fig. 20. Optimal strengthening factors for beams and columns (λ*B and λ*C ), in terms of frame aspect ratio, for the three blast threat scenarios, for pH = 6 × 10− 3 ,
nAPM
r,c = 1 unless stated otherwise.

18
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

reliabilities which are close to zero, and effectively zero for lower 1. columns should be strengthened beyond APM recommendations, to
frames. For this loading scenario the intermediate 72% strengthening avoid cascading horizontal propagation of initial damage.
(see Section 8.5) is not sufficient: the increase in construction costs is not Strengthening columns also reduces conditional failure probabilities
matched by a corresponding reduction in expected cost of progressive and damage to adjacent columns, conditional on a successful attack,
collapse failure, for pH = 6 × 10− 3 . For this case, strengthening for leading to a two-fold benefit.
nAPM
r,c = 2 is probably a better solution. 2. Beam strengthening in codified APM strengthening is slightly
For column crushing, Figs. 18 and 19 show similar patterns between conservative.
local and global crushing failure modes, only with optimal reliability
indexes against global crushing failure significantly higher. For codified These results should be taken as indicative, as they strictly apply to
APM strengthening, column crushing reliabilities change significantly the frames and hazard scenarios investigated herein, for a relevant
with frame aspect ratio, for nAPM = 1. Results are more uniform when threat probability, and with the model limitations acknowledged
r,c
throughout the paper. Nevertheless, to incorporate such a result in
nAPM = 2 is considered. All solutions are in near agreement for low
r,c codified APM strengthening, the resistance factor ϕ of [45] could be
frames. Interestingly, optimal column crushing reliabilities are closer to
individualized as ϕAPM
C for columns and ϕAPM
B for beams, as already done
those obtained in codified APM for nAPM = 2. This suggests that the
r,c in [1] with conventional design values. This is a direct result of a risk
higher load redistribution promoted by removal of two columns, and analysis, balancing the costs of strengthening beams and columns to
which leads to increase in required column strength, produces columns reduce the likelihood of disproportional collapse, with the expected
with a reliability closer to the optimal. This suggests that the resistance costs of failure. An indicative value found herein for taller buildings, for
partial factor for strengthening of columns (ϕ in Eq. (15)) could be one column removal, and subject to further studies, is given by the
smaller than one (ϕ < 1). λ*C 1.2
Optimal reliabilities against column crushing are larger for lower (optimum) ratio of strengthening factors: λ*B
= 0.8 = 1.5. Distinct resis­
frames to avoid horizontal propagation of damage, which would affect tance factors are already considered in conventional design of beams
larger portions of the lower (and wider) frames. Large column crushing and columns, with use of ϕB = 0.85 for beams with ductile failure, and
β* ’s for low frames could be avoided by adopting structural fuses and ϕC = 0.65 for columns with brittle failure, leading to a ratio ϕϕB = 0.85
0.65 =
C
compartmentalization, which is not considered herein. 1.3. Results presented herein suggest that this ratio could be increased.
For the Benchmark VBIED scenario considering damage to adjacent Fixing ϕAPM at the conventional ϕC = 0.65, for instance, one would
C
columns, optimal crushing reliabilities are significantly larger, except
obtain ϕAPM
B = 0.65 × 1.5 ≈ 1. This corresponds to an intermediate sit­
for the tallest frame.
uation between the ϕB = ϕC = 1 of [45], and the ϕC = 0.65 with ϕB =
0.85 of [1].
8.7. Optimal strengthening design factors
9. Concluding remarks
Fig. 20 shows optimal strengthening factors for beams and columns
In this manuscript we addressed the optimal risk-based cost-benefit
(λ*B and λ*C ), in terms of frame aspect ratio, for the three blast threat
analysis of strengthening RC frames to mitigate the likelihood of
scenarios. Recall that these strengthening factors correspond to λB = 1/ϕ
disproportionate collapse under terrorist blast loading. The large
and λC = 1/ϕ, with ϕ = 1 being the strength reduction factors of codi­
epistemic uncertainty in evaluating the probability of initial damage,
fied APM strengthening (Eqs. (14) and (15)). Values of λ*B and λ*C above
given a successful attack, was addressed by considering three blast
one correspond to additional required strength, in the abnormal load
hazard scenarios, and by considering the annual threat probability as the
scenarios of [1,45], considering discretionary removal of one or two
independent parameter. The cost-benefit analysis confronted the costs of
columns. As observed in Fig. 20, the column crushing strengths could be
strengthening the frames, with the expected reduction in cost of pro­
increased by a factor between 1.0 and 1.25, with higher values applying
gressive collapse failure. A break-even point was identified, corre­
to the taller frames considered herein, in case of significant threats with
sponding to the annual threat probability for which conventional design
annual hazard rates of the order of pH = 6 × 10− 3 . and Alternate Path Method (APM) strengthening lead to same total ex­
The optimal column strengthening for the Benchmark VBIED cases pected costs. A risk analysis involving actual building location, envi­
are near unitary for nAPM
r,c = 2, and between 1.0 and 1.25 for nAPM
r,c = 1. ronment, ownership and use, should be conducted to evaluate if the
These results reflect the intermediate 72% local strengthening discussed hazard probability is above or below the break-even point. If the hazard
in Section 8.5, meaning that optimal solutions are in-between, corrob­ probability for a particular building and blast threat is above the break-
orating results observed in Figs. 18 and 19. For the tall building, for even point, APM strengthening is more cost-effective than conventional
instance, the optimal solution would be obtained with nAPM r,c = 1, but design. Herein, we also identified for the first time the break-even point
with columns of 25% larger axial strengths. From Table 6, one obtains between the decisions to strengthen a given frame for it to bridge over a
1.28 × 1.25 = 1.60, a 60% increase which is close to the intermediate single, or over two removed columns. We have shown how the
72% axial strength already computed in Section 8.5. Hence, this solution strengthening of columns to avoid horizontal propagation of initial
is very close to the optimal. damage has a dual effect in the direct resistance of these columns against
Optimal strengthening factors for beam bending confirm results the blast loading, which again changes the probabilities of damage
observed in Fig. 17: for the Benchmark VBIED case and taller frames, propagation. This eventually leads to an overshooting of the optimal
APM strengthening with nAPM r,c = 1 is closer to optimal; for lower frames,
solutions, which was discussed in the paper. Finaly, results shown herein
nAPM = 2 is closer to optimal. For a frame with aspect ratio (6 × 11) provide useful insight towards the revision of guidelines and codified
r,c
strengthening factors for resistance against disproportionate collapse of
there is a transition, where both solutions are equivalent (green circle in
buildings. In particular, for buildings that are potential targets to ma­
Fig. 20). This corresponds to a break-even hazard probability of pth H ≈ levolent blast threats, a distinction should be considered between the
6 × 10− 3 for this frame and hazard, given by point C in Fig. 11. In fact, resistance factors for beams and columns, much in the same way as is
the two risk-based solutions discussed here are the same, since as noted current practice in conventional design: strengthening of columns with
earlier, the risk optimization covers the whole range of design factors. brittle failure should be slightly increased, while strengthening could be
Hence, as noted in Fig. 20, all beam strengthening factors could be reduced for beams with ductile failure. Specific values computed herein
reduced by at least 20%. are valid only for the buildings, parameters, models, and hazard
For all scenarios investigated herein, two results are noteworthy:

19
A.T. Beck and M.G. Stewart Structures 57 (2023) 105103

scenarios considered herein, but the conclusions can be generalized with [18] Mueller J, Stewart MG. The Policing of Terrorism. New York: Oxford University
Press; 2016.
further research.
[19] Stewart MG, Mueller J. Terrorism risks chasing ghosts, and infrastructure
resilience. Sustain Resilient Infrastruct 2020;5(1-2):78–89.
10. Replication of results [20] Stewart MG. Reliability-Based Load Factor Design Model for Explosive Blast
Loading. Struct Saf 2018;71:13–23.
[21] Stewart MG. Reliability-Based Load Factors for Airblast and Structural Reliability
All information required to reproduce results presented herein are of RC Columns for Protective Structures. Struct Infrastruct Eng 2019;15:634–46.
included in the manuscript, and in the cited references. [22] Stewart MG. Reliability-Based Design and Robustness for Blast-Resistant Design of
RC Buildings. Adv Struct Eng 2022;25(7):1402–12.
[23] Mutalib AA, Hao H. Development of P-I diagrams for FRP strengthened RC
Declaration of Competing Interest columns. Int J Impact Eng 2011;38(5):290–304.
[24] Stewart MG, Netherton MD, Baldacchino H. Observed Airblast Variability and
The authors declare that they have no known competing financial Model Error from Repeatable Explosive Field Trials. Int J Prot Struct 2020;11(2):
235–57.
interests or personal relationships that could have appeared to influence [25] Kingery CN, Bulmash G (1984) Airblast Parameters From TNT Spherical Air Burst
the work reported in this paper. and Hemispherical Surface Burst.
[26] UFC 3-340-02 (2014) Structures To Resist the Effects of Accidental Explosions,
Department of Defense, Washington D.C.
Acknowledgements [27] Stewart MG. Spatial variability of explosive blast loading and its effect on damage
risks to reinforced concrete buildings. Eng Struct 2023;285:115650. https://doi.
Funding of this research project by Brazilian agencies CNPq (Na­ org/10.1016/j.engstruct.2023.115650.
[28] ACI 318-19 (2019) Requirements for Structural Concrete. Am. Concr. Inst. 988.
tional Council for Scientific and Technological Development, grant n.
[29] Bredean LA, Botez MD. The influence of beams design and the slabs effect on the
309107/2020-2) and joint FAPESP-ANID (São Paulo State Foundation progressive collapse resisting mechanisms development for RC framed structures.
for Research - Chilean National Agency for Research and Development, Eng Fail Anal 2018;91:527–42. https://doi.org/10.1016/j.
grant n. 2019/13080-9) are cheerfully acknowledged. The support from engfailanal.2018.04.052.
[30] Yu J, Tang J-H, Luo L-Z, Fang Q. Effect of boundary conditions on progressive
the Australian Research Council Discovery Project DP210101487 is also collapse resistance of RC beam-slab assemblies under edge column removal
gratefully acknowledged. scenario. Eng Struct 2020;225:111272. https://doi.org/10.1016/j.
engstruct.2020.111272.
[31] Qian K, Cheng J-F, Weng Y-H, Fu F. Effect of loading methods on progressive
References collapse behavior of RC beam-slab substructures under corner column removal
scenario. J Build Eng 2021;44:103258. https://doi.org/10.1016/j.
[1] GSA (2016) Alternate path analysis and design guidelines for progressive collapse jobe.2021.103258.
resistance. [32] Shan S, Wang H, Li S, Wang B. Evaluation of progressive collapse resistances of RC
[2] Adam JM, Parisi F, Sagaseta J, Lu X. Research and practice on progressive collapse frame with contributions of beam, slab and infill wall. Structures 2023;53:
and robustness of building structures in the 21st century. Eng Struct 2018;173(15): 1463–75. https://doi.org/10.1016/j.istruc.2023.04.114.
122–49. https://doi.org/10.1016/j.engstruct.2018.06.082. [33] Huang Y, Tao Y, Yi W, Zhou Y, Deng L. Numerical investigation on compressive
[3] Marchand KA, Stevens DJ. Progressive collapse criteria and design approaches arch action of prestressed concrete beam-column assemblies against progressive
improvement. J Perform Constr Facil 2015;29:B4015004. https://doi.org/ collapse. J Build Eng 2021;44:102991. https://doi.org/10.1016/j.
10.1061/(ASCE)CF.1943-5509.0000706. jobe.2021.102991.
[4] Beck AT, da Rosa RL, Valdebenito M, Jensen H. Risk-Based Design of Regular Plane [34] Xi Z, Zhang Z, Qin W, Zhang P. Experiments and a reverse-curved compressive arch
Frames Subject to Damage by Abnormal Events: A Conceptual Study. J Struct Eng model for the progressive collapse resistance of reinforced concrete frames. Eng
2022;148. https://doi.org/10.1061/(asce)st.1943-541x.0003196. Fail Anal 2022;135:106054. https://doi.org/10.1016/j.engfailanal.2022.106054.
[5] Beck AT, Ribeiro LdR, Valdebenito M. Risk-based cost-benefit analysis of frame [35] Long X, Wang S, Huang X-J, Li C, Kang S-B. Progressive collapse resistance of
structures considering progressive collapse under column removal scenarios. Eng exterior reinforced concrete frames and simplified method for catenary action. Eng
Struct 2020;225:111295. Struct 2021;249:113316. https://doi.org/10.1016/j.engstruct.2021.113316.
[6] Beck AT, Ribeiro LR, Valdebenito MA. Cost-Benefit Analysis of Design for [36] Subki NEA, Mansor H, Hamid YS, Parke GAR. The development of a moment-
Progressive Collapse Under Accidental or Malevolent Extreme Events. In: rotation model for progressive collapse analysis under the influence of tensile
Stewart MG, Rosowsky DV, editors. Engineering for Extremes, Decision-Making in catenary action. J Constr Steel Res 2021;187:106960. https://doi.org/10.1016/j.
an Uncertain World; 2022. jcsr.2021.106960.
[7] Stewart MG. Life-safety risks and optimisation of protective measures against [37] Yang B, Kang S-B, Tan KH, Zhou X-H, (eds), 2022: Analytical model for
terrorist threats to infrastructure. Struct Infrastruct Eng 2011;7:431–40. https:// compressive arch action and catenary action in reinforced concrete beams against
doi.org/10.1080/15732470902726023. progressive collapse, Chapter 7 in Behaviour of Building Structures Subjected to
[8] Stewart MG. Risk of Progressive Collapse of Buildings from Terrorist Attacks: Are Progressive Collapse, Woodhead Publishing, 315-367, https://doi.org/10.1016/
the Benefits of Protection Worth the Cost? J Perform Constr Facil 2017;31:1–9. B978-0-12-822267-6.00006-1.
https://doi.org/10.1061/(asce)cf.1943-5509.0000954. [38] Masoero E, Darò P, Chiaia BM. Progressive collapse of 2D framed structures: An
[9] Ellingwood BR, Dusenberry DO. Building design for abnormal loads and analytical model. Eng Struct 2013;54:94–102. https://doi.org/10.1016/j.
progressive collapse. Comput Civ Infrastruct Eng 2005;20:194–205. https://doi. engstruct.2013.03.053.
org/10.1111/j.1467-8667.2005.00387.x. [39] Department of Defense (DoD), UFC 4-023-03. Design of Buildings to Resist
[10] Ellingwood BR. Mitigating risk from abnormal loads and progressive collapse. Progressive Collapse. (2016) 34-37.
J Perform Constr Facil 2006;20:315–23. https://doi.org/10.1061/(ASCE)0887- [40] Praxedes C, Yuan X-X. Robustness Assessment of Reinforced Concrete Frames
3828(2006)20:4(315). under Progressive Collapse Hazards: Novel Risk-Based Framework. J Struct Eng
[11] Ellingwood BR, Smilowitz R, Dusenberry DO, et al (2007) Best practices for 2021;147:1–14. https://doi.org/10.1061/(asce)st.1943-541x.0003075.
reducing the potential for progressive collapse in buildings. US Natl Inst Stand [41] Praxedes C, Yuan XX. Robustness-oriented optimal design for reinforced concrete
Technol (NIST) 216. frames considering the large uncertainty of progressive collapse threats. Struct Saf
[12] Ellingwood BR. Strategies for Mitigating Risk to Buildings from Abnormal Load 2022;94:102139. https://doi.org/10.1016/j.strusafe.2021.102139.
Events. Int J Risk Assess Manag 2007;7:828–45. [42] Ribeiro LR, Kroetz HM, Beck AT, Parisi F (202?): Optimal risk-based design of
[13] Faber MH. Risk Assessment in Engineering - Principles. System Representation & reinforced concrete beams against progressive collapse, Engineering Structures
Risk Criteria; 2008. (under review).
[14] Stewart MG. Risk-informed decision support for assessing the costs and benefits of [43] Marchand K, McKay A, Stevens DJ (2009): Development and Application of Linear
counter-terrorism protective measures for infrastructure. Int J Crit Infrastruct Prot and Non-Linear Static Approaches in UFC 4-023-03, in: Struct. Congr. 2009,
2010;3:29–40. https://doi.org/10.1016/j.ijcip.2009.09.001. American Society of Civil Engineers, Reston, VA, 2009: pp. 1-10. https://doi.org/
[15] Stewart MG, Thöns S, Beck AT. Assessment of risk reduction strategies for terrorist 10.1061/41031(341)191.
attacks on structures. Struct Saf 2023;104. https://doi.org/10.1016/j. [44] ASCE 7 (2017) Minimum design loads for buildings and other structures. ASCE
strusafe.2023.102353. Stand. 1–456.
[16] Mueller J, Stewart MG. Terror, Security, and Money: Balancing the Risks, Benefits, [45] ASCE/SEI 41-13 (2017) Seismic Evaluation and Retrofit of Existing Buildings.
and Costs of Homeland Security. New York: Oxford University Press; 2011. [46] Melchers RE, Beck AT. Structural Reliability Analysis and Prediction3rd ed.. Wiley;
[17] Beck AT, Rodrigues da Silva LA, Miguel LFF. The latent failure probability: A 2018. p. 528. ISBN-10 1119265991, ISBN-13 9781119265993.
conceptual basis for robust, reliability-based and risk-based design optimization.
Reliab Eng Syst Saf 2023;233:109127. https://doi.org/10.1016/j.
ress.2023.109127.

20

You might also like