You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/280088983

Blast resistant design of precast reinforced concrete walls for strategic


infrastructures under uncertainty

Article  in  International Journal of Critical Infrastructures · January 2015


DOI: 10.1504/IJCIS.2015.072151

CITATIONS READS

7 2,018

4 authors, including:

Pierluigi Olmati Clay J. Naito


Tokyo Polytechnic University Lehigh University
28 PUBLICATIONS   224 CITATIONS    157 PUBLICATIONS   1,100 CITATIONS   

SEE PROFILE SEE PROFILE

Franco Bontempi
Sapienza University of Rome
240 PUBLICATIONS   1,326 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Horizontal Shear Capacity of Concrete Beams View project

Tsunami Debris Project View project

All content following this page was uploaded by Clay J. Naito on 16 July 2015.

The user has requested enhancement of the downloaded file.


Int. J. Critical Infrastructures, Vol. X, No. Y, xxxx 1

Blast resistant design of precast reinforced concrete


walls for strategic infrastructures under uncertainty

Pierluigi Olmati*
Sapienza University of Rome,
Via Eudossiana 18 – 00184 Rome, Italy
E-mail: pierluigi.olmati@gmail.com
*Corresponding author

Patrick Trasborg and Clay Naito


Lehigh University,
117 IMBT Labs, ATLSS Dr, Bethlehem PA 18015, USA
E-mail: pat310@lehigh.edu
E-mail: cjn3@lehigh.edu

Franco Bontempi
Sapienza University of Rome,
Via Eudossiana 18 – 00184 Rome, Italy
E-mail: franco.bontempi@uniroma1.it

Abstract: Extreme loads can have devastating effects on civilian structures


since these buildings are not designed to withstand extreme events. Typical
buildings and other critical infrastructures are particularly prone to external
man-made attacks. This study focuses on probabilistic analyses, and
investigates the probability of exceeding a given limit state of a precast
concrete wall subjected to blast loads. The wall under investigation is a
non-load bearing precast concrete panel used as exterior cladding for buildings.
From the blast design point of view, these walls must protect people and
equipment from external detonations. The aim of the paper is to compute both
the fragility curves and the probability of exceedance of at ypical precast
concrete cladding system, considering a prescribed detonation of a vehicle
borne improvised explosive device. To this aim non-linear dynamic analyses
are carried out by the widely adopted equivalent single degree of freedom
method. The fragility curves and the probability of exceedance of the precast
concrete cladding wall are computed using Monte Carlo simulations.

Keywords: probabilistic blast engineering; fragility curves; FCs; Monte Carlo


simulation; precast wall panels; conditional approach.

Reference to this paper should be made as follows: Olmati, P., Trasborg, P.,
Naito, C. and Bontempi, F. (xxxx) ‘Blast resistant design of precast reinforced
concrete walls for strategic infrastructures under uncertainty’, Int. J. Critical
Infrastructures, Vol. X, No. Y, pp.000–000.

Copyright © 200x Inderscience Enterprises Ltd.


2 P. Olmati et al.

Biographical notes: Pierluigi Olmati graduated in Structural Engineering at


the Sapienza University of Rome (Italy). His principal research is on the blast
engineering related topics, addressed from the point of view of FE modelling
and probabilistic design. He was a Visiting Scholar at the Lehigh University
(USA) in the 2012, studying the performances of insulated wall panels
subjected to close-in detonations. Furthermore, in 2013, he was a Visiting
Scholar at the National Technical University of Athens (Greece), performing
research on the probabilistic framework of the blast design, developing the
fragility and a safety factor for steel built-up blast resistant doors.

Patrick Trasborg is in his fourth year of his PhD in Structural Engineering at


Lehigh University with advisor Professor Clay Naito. His principal research
topic is blast engineering, addressed from the point of view of analytical
modelling with experimental validation. His dissertation is on the development
of a blast and ballistic resistant insulated precast concrete wall panel. Currently,
he is characterising the performance of insulated panels with various shear ties
subjected to uniform loading. His research interests include materials subjected
to extreme demands and probabilistic analysis.

Clay Naito is an Associate Professor of Structural Engineering at Lehigh


University. His research interests include experimental and analytical
evaluation of reinforced and prestressed concrete structures subjected to
extreme dynamic events including earthquakes, impacts, and intentional blast
demands. He is the Chair of the Precast/Prestressed Concrete Institute Blast
Resistance and Structural Integrity Committee and an Associate Editor of
the ASCE Bridge Journal. He received his BS degree from the University of
Hawaii Manoa and his MS and PhD degrees from the University of California
Berkeley.

Franco Bontempi obtained a degree in Civil Engineering in 1988 and his PhD
in Structural Engineering in 1993, from the Politecnico di Milano. He is a
Professor of Structural Analysis and Design at the School of Engineering of the
Sapienza University of Rome since 2000. He has a deep research activity on
numerous themes related to structural engineering, having developed scientific
and technical publications on the topics: structural analysis and design, system
engineering, performance-based design, hazard and risk analysis, safety and
reliability engineering, dependability, structural integrity, structural dynamics,
bridges and viaducts, high-rise buildings, special structures, and offshore wind
turbines.

This paper is a revised and expanded version of a paper entitled ‘Blast


resistance of reinforced precast concrete walls under uncertainty’ presented at
2013 Critical Infrastructure Symposium International Perspectives on Full
Spectrum Resilience, West Point, New York, 15–16 April.

1 Introduction

Designing structures to withstand blast loads is common practice for many government
and commercial buildings. Generally in design, a set of attack scenarios are selected and
the integrity of the structural members are verified using non-linear dynamic analyses
with the equivalent single degree of freedom (SDOF) method. In such a way (that is
adopting a deterministic approach) the probability of exceeding a particular limit state is
Blast resistant design of precast reinforced concrete walls 3

not evaluated, principally because there is a lack of knowledge of the hazard probability
density function. This is common for Low Probability – High Consequence (LPHC)
events; for example, Gkoumas (2010), Gentili et al. (2013), Olmati et al. (2013a, 2013b)
examine cases where the concept of LPHC events are considered concerning the threat of
progressive collapse.
In designing a structural component subjected to blast loads, the current state of
practice is to use appropriate fractile values and assume the demand is deterministic. The
DoD (2008) suggests increasing the assumed explosive by 20% for the design scenario;
however, this recommendation is limited to ammunition storage facilities where the
quantity of explosives is often well defined. In antiterrorism/force protection design, the
amount of explosive used in an attack is generally characterised by a large degree of
uncertainty.
A precast concrete cladding wall system has advantages over other traditional
non-load bearing cladding systems. The work of Davidson et al. (2005) consisted of
studies to improve traditional masonry cladding against blast loads. The first advantage
of precast concrete wall panels over traditional masonry cladding concerning the blast
loads is the increased resistance of the precast system to the impulsive demand. Precast
concrete structures have shown to provide improved resilience against blast in
comparison to traditional steel stud construction as discussed in Naito et al. (2011).
Additionally, precast concrete has many advantages over cast-in-place concrete in
general. The final condition and mechanical properties of concrete are highly sensitive to
the environmental conditions during the curing process. The more finely environmental
conditions can be controlled, such as humidity, temperature, hydration effects, etc., the
better the control over the final properties of the concrete. For this reason, precast
components are often more aesthetically pleasing than cast-in-place components.
Furthermore, precast concrete can often be more economical than cast-in-place concrete.
Formwork is generally the controlling factor in the cost of a concrete component, often
making up more than 60% of the total cost of the concrete piece (Nemati, 2007). As
precast components can be made in an assembly line fashion with the same formwork
repeatedly, the cost for constructing the component drops. Finally, precast components
will often allow for a more efficient construction process, decreasing the total
construction time. Precast components are typically fabricated offsite, transported to the
work site, and are then placed into position as opposed to cast in place concrete which
requires building formwork in the component’s final resting place and then allowing time
for curing.
Moreover, the building life-cycle should be considered when selecting the façade.
Concrete cladding can be integrated with other materials to improve the response of the
panel to environmental attack such as acid rain and chlorine ions. In Naito et al. (2012),
the behaviour of a precast concrete panel with an insulation layer to improve the thermal
resistance of the panel is investigated, focusing on the shear ties connecting the two
concrete layers confining the insulation layer. The insulated panel’s interior and exterior
surfaces are durable, not requiring expensive maintenance repairs over time. In fact, this
kind of cladding system is economically and ecologically sustainable, while maintaining
all the advantages of traditional precast components.
Considering the aforementioned, this study focuses on a probabilistic approach for
blast resistance assessment, and investigates the probability of exceedance of a precast
concrete wall subjected to blast loads (in particular far-field, surface blast demands). The
4 P. Olmati et al.

wall under investigation is a non-load bearing precast concrete wall panel used as exterior
cladding for buildings.
The non-linear dynamic analyses are carried out by the well-established method of
the equivalent non-linear SDOF system, so by definition, the precast concrete wall is
modelled by an equivalent non-linear SDOF. Both the fragility curves (FCs) and the
probability of exceeding a given limit state of the cladding wall are computed using
Monte Carlo simulations.
The aim of this paper is to apply a probabilistic approach to the blast design. To
accomplish this, the FCs for a precast concrete wall panel are computed for each
component damage level defined in a performance-based design approach [see, for
example, Ciampoli and Petrini (2012) for the performance-based design approach applied
to wind engineering]. The FCs are then used to estimate the probability of exceeding a
given limit state of the cladding panel subjected to a blast demand due to a vehicle borne
improvised explosive device (VBIED). Finally, the probability of exceedance of the wall
panel subjected to the same scenario is estimated with a Monte Carlo simulation and the
results are compared with the results determined using the FCs.

2 Blast load model

The side-on blast pressure, Ps0 [MPa], can be estimated by the formula of Mills (1987)
[equation (1)], while the side-on specific impulse, is0 [Pa sec], is estimated by the formula
of Held (1983) [equation (2)].

⎛ 2 ⎞ ⎛ 1 ⎞
⎛W⎞ ⎜ W3 ⎟ ⎜ W3 ⎟
Ps0 = 1.772 ⎜ 3 ⎟ − 0.114 ⎜ 2 ⎟ + 0.108 ⎜ ⎟ (1)
⎝R ⎠ ⎝R ⎠ ⎝ R ⎠

⎛1 ⎞
is0 = 300 ⎜ 3 W ⎟ (2)
⎝Z ⎠
where W is the explosive charge weight, usually in kgf of TNT, R is the stand-off
distance [m], and Z is the scaled distance defined as the stand-off distance divided by the
cube root of the explosive weight.
The explosive charge is commonly expressed as an equivalent TNT charge because
TNT is a very common, traditional explosive with a lot of available experimental data.
Generally, an explosive charge is adjusted to an equivalent TNT charge by the equivalent
factor (EF), which multiplies the weight of the explosive charge utilised. For more detail,
consult the DoD (2008).
Both equation (1) and equation (2) are valid for free-air explosions. In this study
detonations occurring on a surface (surface explosions) are considered, therefore the
energy of the detonation is confined by the ground surface, creating a larger demand than
that of the free-air explosion. The surface blast demand is calculated by using the same
equations for the free-air explosion but with a charge weight (W) increased by 80% (US
Army Corps of Engineers, 2008a). The reflected pressure Pr [MPa] generated on a
structure normal to the path of the expanding blast pressure is computed using equation
(3) (Mills, 1987):
Blast resistant design of precast reinforced concrete walls 5

⎛ 7P + 4Ps0 ⎞
Pr = 2Ps0 ⎜ atm ⎟ (3)
⎝ 7Patm + Ps0 ⎠
where Patm is the atmospheric pressure (0.1 MPa). For simplicity, the negative pressure
phase is neglected from the blast load time history (DoD, 2008). The duration of the blast
load (td) is computed from the side-on pressure and the specific impulse, assuming a
triangular impulse; see equation (4).
2is0
td = (4)
Ps0

The variation in reflected pressure is assumed to follow that of the Friedlander pulse
shape as shown in equation (5). Further details on the sensitivity of the structural
response due to exponential or triangular blast loading are investigated in Gantes and
Pnevmatikos (2004).
−βt
⎛ t ⎞ td
P(t) = Pr ⎜1 − ⎟e ta ≤ t ≤ td (5)
⎝ t d ⎠
In equation (5) ta [seconds] is the arrival time of the blast load, taken here as zero, and β
is the decay coefficient. In this study a value of 1.8 for β is assumed. The clearing effect
is conservatively neglected in this study since the cladding wall is generally a single part
of a large building façade; thus the conditions for clearing of the reflected shock wave are
not satisfied. More details on the clearing effect for finite width surfaces are available in
the study of Ballantyne et al. (2010). In Figure 1, the blast load time histories are
computed for different values of the explosive weight W [kgf] and stand-off distance R
[m] with the aforementioned procedure. The obtained curves are found to be in good
agreement with the curves obtained by SBEDS (US Army Corps of Engineers, 2008a).
The blast load is considered uniformly distributed on the cladding wall, which is typical
for a scaled distance higher than approximately 1.2 to 2.0 m / kg1/3 f (US Army Corps of
Engineers, 2008a).

Figure 1 Blast loads by the adopted model (broken lines) and the SBEDS model
(solid lines) (see online version for colours)
6 P. Olmati et al.

3 Cladding panel model

The cladding panel examined in the following case study is assumed to attach to a frame
structure. A 3,500 mm long by 1,500 mm wide panel is used, with a cross sectional
thickness of 150 mm. The panel is connected to a beam on the external frame of the
building. The connections are assumed to provide simple-simple boundary conditions.
Length, width and cross sectional thickness are taken as stochastic variables. Due to
the construction tolerances used in the precast concrete industry, see, for example, PCI
(2010a), the assumed coefficients of variation for the geometric variables are assumed to
be small. Both assumed mean values and coefficients of variations (COVs) are shown in
Table 1. The longitudinal reinforcement consists of ten reinforcement bars of 10 mm
diameter located in the centre of the cross section. The mean value of the reinforcement
strength and coefficient of variation are provided in Table 1. The panel under
investigation is not shear critical for the purpose of this study and consequently does not
have shear reinforcement.
Table 1 Input data

Symbol Description Mean COV Distribution


fc Concrete strength 28 MPa 0.18 Lognormal
fy Steel strength 495 MPa 0.12 Lognormal
L Panel length 3,500 mm 0.001 Lognormal
t Panel thickness 150 mm 0.001 Lognormal
b Panel width 1,500 mm 0.001 Lognormal
c Steel depth 75 mm 0.01 Lognormal
W Explosive weight 227 kgf 0.3 Lognormal
R Stand-off distance 15 m 20 m 25 m 0.05 Lognormal

3.1 Concrete model


The concrete compressive strength fc is taken as a stochastic variable, while the Young’s
modulus of the concrete Ec and the concrete density ρ are expressed as functions of
fc. The mean value of fc is 28 MPa, with a COV of 0.18, as adopted in Enright and
Frangopol (1998) for a lognormal probability density function (see Table 1). The
Young’s modulus is computed by equation (6) (ECS, 2005) while the concrete density is
computed by equation (7) (ACI, 2011). Both Ec and fc are expressed in MPa while ρ is
expressed in kg/m3.
0.3
⎛f ⎞
E c = 22, 000 ⎜ c ⎟ (6)
⎝ 10 ⎠
1
⎛ Ec ⎞1.5
ρ=⎜ (7)
⎜ 0.043 ( f 0.5 ) ⎟⎟
⎝ c ⎠
Blast resistant design of precast reinforced concrete walls 7

The compressive strength enhancement of the concrete due to high strain rate is
considered in this study. This increase is taken into account by means of the dynamic
increase factor (DIF), a multiplicative coefficient of the concrete compressive resistance.
Since the compressive strength enhancement of the concrete varies marginally for a
ductile flexural response over the range of the strain velocity, the DIF can be assumed
constant and equal to 1.19 times the static compressive strength. This hypothesis is in
accordance with the compressive strength enhancement proposed in US Army Corps of
Engineers (2008a), and increases computational efficiency by avoiding cyclic iterations
in the algorithm of the SDOF equation solver. However, cyclic iterations are necessary
for computing the strength enhancement of the reinforcement, which is very sensitive to
ductile flexural response as opposed to the compressive strength enhancement of the
concrete.

3.2 Reinforcing steel model


For this case study, grade 450 MPa steel is used. Due to material standard requirements
the average yield strength is higher than the specified yield. For estimating the mean
value of the yield strength, a static strength increase factor equal to 1.1 is adopted as
indicated in US Army Corps of Engineers (2008a). A COV of 0.12 is used as proposed in
Enright and Frangopol (1998) for a lognormal probability density function (see Table 1).
Young’s modulus is taken as a deterministic value and assumed to be equal to
210 GPa. The steel strength enhancement due to the strain velocity is taken into account
by the Cowper and Symonds (1957) model. Thus, the DIF is provided by equation (8):
1
⎛ d ∈ /dt ⎞ p
DIF = 1 + ⎜ ⎟ (8)
⎝ C ⎠
where dε/dt is the strain-rate demand of the reinforcement, C is taken equal to 500 sec–1
and p is taken equal to 6. Both C and p are estimated by fitting the strength increase
versus the strain rate in US Army Corps of Engineers (2008a). By solving the SDOF
equation of motion, the DIF is iteratively updated until a convergence threshold is
reached. The strain rate of the steel reinforcement (dε/dt) in equation (8) is calculated
approximately by equation (9).
dε dS L ⎛ d ⎞
= (9)
dt dt 8 ⎜⎝ 2E c Ic ⎟⎠

where L is the length of the cladding panel, Ec is the Young’s modulus of the concrete, Ic
is the moment of inertia of the cracked cross section, d is the distance from the extreme
compression fibre of the cross section to the centroid of the tensile reinforcement, and
dR/dt is the rate of the resistance force developed by the panel (S) when subjected to the
demand. Equation (9) is valid for simply supported elements when the response is
governed by the flexural behaviour without shear failure.
8 P. Olmati et al.

4 The SDOF model and the component damage levels

The load considered in this paper is due to a far-field detonation that tends to create a
uniform pressure distribution across the face of the component and cause a flexural
failure mode as opposed to a close-in detonation that tends to cause localised effects such
as spall and breach. Typical blast design methodology utilises time stepping methods to
solve the differential equation of motion in order to predict the response of the reinforced
concrete component to the blast loading. In many cases, structural components subjected
to blast load can be modelled as an equivalent SDOF mass-spring system with a
non-linear spring.
An equivalent SDOF system is created by developing appropriate transformation
factors for the system’s mass, damping, load and resistance. Furthermore, inherit with an
SDOF analysis is the assumption that the system behaves only in a single deflected
shape. As the system begins to deflect under the blast load, it eventually yields and forms
plastic hinges at various locations depending on the applied boundary conditions. Thus in
reality, the system’s deflected shape changes with the progression of plastic hinges.
Therefore, the transformation factors are adjusted accordingly to account for the change
in deflected shape. For a more detailed review on the development of transformation
factors, see the PCI (2010b).
For a simply supported, one way panel under uniform loading, it is assumed that a
single plastic hinge will form at centre span of the panel. The resistance-deflection
relationship for such a panel is assumed to act in an elastic-perfectly plastic manner.
Thus, at a certain deflection, the component will continuously yield at near constant
resistance until an ultimate deflection limit is reached, at which point the component will
fail. This resistance-deflection relationship (resistance function) serves as the property for
the non-linear spring in the equation of motion.
Once the time-history deflection of the component is known, it is desirable to
determine the amount of damage that the component has undergone. The US Army Corps
has developed component damage levels (CDLs), based on the building level of
protection and component type, which are correlated with two response parameters: the
support rotation angle (θ) and the ductility ratio (μ). These parameters are defined in
equation (10) and equation (11) for a simply supported component:

⎛ 2δ ⎞
θ = arctan ⎜ max ⎟ (10)
⎝ L ⎠
δ max
μ= (11)
δe

where δmax is the maximum deflection and δe is the equivalent yield deflection of the
panel. A structural component subjected to a blast load is generally expected to yield
(ductility greater than 1), as it is economically impractical to design a member to remain
in the elastic range. While other significant response parameters can be defined, this
Blast resistant design of precast reinforced concrete walls 9

study focuses on the response parameters usually adopted for antiterrorism design (US
Army Corps of Engineers, 2008b). In the USA anti-terrorism performance-based blast
design approach, there are five CDLs considered, listed in order of decreasing damage:
blowout (BO), hazardous failure (HF), heavy damage (HD), moderate damage (MD), and
superficial damage (SD). The thresholds corresponding to these CDLs are defined in
terms of the response parameters θ and μ. For a non-structural concrete cladding wall
without shear reinforcement, neglecting tension membrane effects, the CDL thresholds
are those reported in Table 2. The FCs computed in this paper are for the CDLs used in
design.
Table 2 Component damage levels, and the associated thresholds in terms of response
parameters

Component damage levels θ [degree] μ [–]


Blowout > 10° None
Hazardous failure ≤ 10° None
Heavy damage ≤ 5° None
Moderate damage ≤ 2° None
Superficial damage None 1
Source: US Army Corps of Engineers (2008b)

5 Fragility curves

The blast load on the panel depends on both the peak pressure and the impulse. The
pressure depends only on the scaled distance, while the impulse density depends on
both the scaled distance and the explosive weight [equation (1) and equation (2)].
Consequently, two detonations with the same scaled distance can have different impulse
densities, depending on the amount of explosive. Thus, two explosions may have the
same peak pressure but different durations. Therefore, for the blast evaluations both the
scaled distance and the explosive weight are considered as demand inputs.
The choice of the opportune intensity measure (IM) for computing the FCs is crucial.
Discussions on the opportune IM are published by several authors. The approach of using
IM and FCs was developed in the field of earthquake engineering. The works of Cornell
and Krawinkler (2000), Vamvatsikos and Dolšek (2011), Luco and Cornell (2000) are
some of the original references for the conditional approach. Additionally, the works of
Ciampoli et al. (2011), and Barbato et al. (2013) in wind engineering are useful
references for the proposed approach; as the approach originally made for earthquake
engineering is adapted for performance-based wind engineering (Petrini and Ciampoli,
2010). In this study, the scaled distance is taken as the IM. The FCs are consequently
plotted as a function of the scaled distance and the stand-off distance (R = Z 3 W ). This
takes into account the dependence on the blast demand, and consequently the sensitivity,
to both the scaled distance and the amount of explosive. The complete procedure is
shown in Figure 2.
10 P. Olmati et al.

Figure 2 FC computation flowchart

j=1  i=1 k=1 • CDL: Component Damage Level
• R: Stand‐off distance
• Z: Scaled distance
R=k • FC‐CDL: numerical Fragility Curve 
of the Component Damage Level
CDL (j) • i: the i‐th point, of the j‐th FC‐CDL 
corresponding to the k‐th R
• j: the j‐th CDL
Z=i
• k: the k‐th stand‐off distance
• MC analysis: Monte Carlo analysis
MC analysis • N: number of FC‐CDL points, or 

i=i+1
number of the Zs
FC‐CDL (i, j, k) • M: number of the CDLs
• L: number of the stand‐off 
distances
NO i=N ? • Interpolated FC‐CDL: lognormal 
interpolated Fragility Curve of the 
YES Component Damage Level
j=j+1

FC‐CDL (j,k)

NO j=M ?

YES
k=k+1

FC‐CDL (k)

NO k=L ?

YES

FC‐CDL

Lognormal 
Interpolation

Interpolated 
FC‐CDL

6 Results

This section presents the results on:


1 the FCs of the cladding panel case study
2 the probability of exceeding a particular limit state of the cladding panel.
The fragility analysis results are compared with the results obtained by using a Monte
Carlo simulation. In Figure 3, the computed FCs are plotted as functions of the scaled
distance. For the specific panel case study, the FC of the SD component damage level has
a different slope compared to that of the other three CDLs (HF, HD, and MD). It should
be noted that the SD component damage level is based on the ductility (μ) of the
component while HF, HD, and MD CDLs are based on the support rotation (θ). The SD
component damage level for a concrete cladding panel thus requires the component to
respond elastically, and for the case study conducted, the panel appears to be more
sensitive to the considered uncertainties when compared to the HF, HD, and MD damage
levels.
Blast resistant design of precast reinforced concrete walls 11

Figure 3 From top left clockwise, FCs for the HF, HD, SD, MD component damage levels

Hazardous Failure Heavy Damage


100 100

80 80
P (X> x0|Z) [%]

P (X> x0|Z) [%]


60 60

40 40

20 20

0 0
2.4 2.6 2.8 3.0 3.2 3.4 2.8 3.0 3.2 3.4 3.6 3.8 4.0
Z Z

Moderate Damage Superficial Damage


100 100

80 80
P (X> x0|Z) [%]
P (X> x0|Z) [%]

60 60

40 40

20 20

0 0
3.0 3.5 4.0 4.5 5.0 5 6 7 8 9 10 11
Z Z

By varying the number of samples, the maximum COV obtained for the lower probability
of exceedance (close to zero) is about 9%. This value is considered acceptable for the
specific case, and it is consistent with other studies on blast applications.
These FCs are the probabilistic representation of the mechanical properties of the
concrete cladding panel and are useful for computing the probability of exceedance for a
set of blast scenarios with different CDLs. If the stochastic characterisation of the blast
scenario is neglected, assuming for example the conventional performance demand, the
FCs of Figure 3 can be used directly as a design tool. By selecting a blast scenario
characterised by both the amount of explosive and the stand-off distance, it is possible to
establish the probability of exceedance of a particular limit state of the panel and verify
that this probability is lower than an acceptable threshold. Different cladding panels can
be certified using these FCs, expediting the design process of each panel against the
12 P. Olmati et al.

corresponding blast demand. Thus, the FCs (Figure 3) can be considered as design charts
when the load is conventionally defined. As an alternate approach the FCs can be
computed for variations of explosive size (i.e., car bomb, truck bomb, etc…) rather than
standoff distance. This would allow for direct determination of the probability of
exceedance at a given standoff distance and would be useful for determining the location
of protective measures, such as bollards or fences, required to achieve the tolerable
probability of failure. But generally for computing the probability of exceedance for a
given limit state from the FCs, it is necessary to stochastically characterise the blast
scenario and integrate equation (12).
In this study a VBIED is considered. The amount of explosive in the vehicle depends,
among other parameters, on the security measures in place, and for each level of security
a different mean value of explosive weight is expected. The mean value of the explosive
charge in the scenario decreases with a decreasing stand-off distance from the target,
assuming a security system progressively reduces the severity of the possible attack with
a decrease in stand-off. In the specific case a VBIED (with about 27 to 454 kgf of TNT or
equivalent) is prescribed. The mean amount of TNT equivalent explosive in the vehicle is
227 kgf with a coefficient of variation equal to 0.3 (see Table 1). This estimation is in line
with the data provided in FEMA (2005). A set of stand-off distances are considered (15,
20 and 25 m) each with a coefficient of variation equal to 0.05, assuming that the vehicle
could impact a fence barrier but move no further. Equation (12) computes the probability
of exceedance of the CDL (P(X > x0)). X is the union of the two response parameters
characterising the component damage level (see Table 2). The probability density
function of the scaled distance (p(Z)) is computed by fitting both the explosive weight
and the stand-off distance to a lognormal distribution. The fragility curve (P(X > x0|Z))
for equation (12) corresponds to the mean value of the stand-off distance of the blast
scenario (Table 1).
+∞ ∞
P ( X > x0 ) = ∫ P ( X > x 0 Z ) p(Z)dz ≅ ∑ P ( X > x Z) p(Z) ΔZ
i=0
0 i i i (12)
−∞

P(X > x0) is estimated for three scenarios. The amount of explosive is the same but the
mean value of the stand-off distance varies 15, 20 and 25 metres respectively. As a
consequence, the considered FC for each scenario is different. Results are provided in
Table 4, where the first column reports the CDLs, and the second and third report the
probability of exceedance for each blast scenario obtained by equation (12) and by the
Monte Carlo analysis respectively.
The first consideration is the sensitivity of the probability of exceedance to the
amount of explosives in the vehicle. The standard deviation of the explosive weight is 0.3
times the mean value which is predominant for computing the probability of exceeding a
CDL. If both the mean value explosive weight and the mean value of the stand-off
distance result in a scaled distance on the range of the FC, the probability of exceedance
is close to 50%, as the explosive weight has a higher standard of deviation than the other
parameters.
Blast resistant design of precast reinforced concrete walls 13

Table 4 Comparisons of the probability of exceedance for a sample charge size and standoff
using MC and FC analysis

Mean W = 227 kgf COV = 0.30 lognormal distribution R, COV = 0.05 lognormal distribution
CDL FC analysis MC analysis Percentage difference Δ%
Mean R = 25 m
SD 100.0% 100.0% 0.0%
MD 74.6% 77.3% 3.5%
HD 14.2% 12.6% 11.2%
HF 1.02% 1.02% 0.0%
Mean R = 20 m
SD 100.0% 100.0% 0.0%
MD 96.6% 97.5% 0.9%
HD 55.7% 55.5% 0.3%
HF 13.6% 12.1% 11.0%
Mean R = 15 m
SD 100.0% 100.0% 0.0%
MD 97.9% 99.9% 2.0%
HD 93.6% 96.9% 3.4%
HF 67.8% 72.6% 6.6%

From these results the maximum percentage difference between the probability of
exceedance computed by the FCs and the Monte Carlo simulation is 11%. Further studies
are necessary to confirm whether this percentage difference is acceptable or not.
However, it is necessary to consider that the amount of explosive in the vehicle has an
elevated variability, something that amplifies the difference in the fragility analysis and
the Monte Carlo simulation due to the dependence of the impulse density to both the
scaled distance and the explosive weight. Thus, being that the impulse density is
proportional to both the explosive weight and the scaled distance, the difference between
the probability of exceedance computed by the fragility analysis method and by the
Monte Carlo simulation method increases with the increase in the difference between the
stand-off distance with which the FC is computed and the stand-off distance of the blast
scenario estimated by Z 3 W. A future study could focus on the quantification of this
difference and on developing a corrective method that leads to using a FC independently
from its stand-off from which it is computed. Also the deterioration of the materials
during the entire life cycle could be taken into account in computing the FCs, see, for
example, Biondini et al. (2006) and Sgambi et al. (2012a). Moreover, soft-computing
methods could be implemented for the more efficient treatment of the uncertainties
Sgambi et al. (2012b).

7 Conclusions

In this paper, a probabilistic approach for designing and assessing a cladding panel
subjected to a blast load has been proposed. The blast load model has been adopted on
14 P. Olmati et al.

the basis of empirical laws, and the geometry and mechanical properties of the panel are
assumed as stochastic. A mechanical model equivalent to a SDOF has been adopted for
describing the motion of the panel under the blast load and the FCs for each component
damage level are computed. These FCs are used in computing the probability of
exceedance of the panel subjected to three different blast scenarios for each component
damage level. Furthermore, results of the fragility analysis were compared to Monte
Carlo simulations.
This study highlights the feasibility and effectiveness of using a probabilistic
approach for the design and assessment of protective structures, having as a principal
limitation, that the computed probability of exceedance of the component damage level is
conditional on the event occurring. Computing the probability of exceedance in time is
difficult as the probability density function of the event in time is unknown. Even if the
probability of occurring in time was known, it would be difficult to implement in design,
since a terroristic attack is a low probability event. The solution presented in this paper is
posed as a basis for the design based on the probability of exceeding a CDL for a specific
event. However, the probability of exceeding a component damage level has to be fixed
on the basis of some criteria.
In blast engineering the fragility curve analysis maintains the same advantages over
the Monte Carlo analysis as in earthquake engineering. In application, FCs can be
computed by the precast fabricator with the proposed method. This would allow for the
reliability analysis to be conducted during the design process for selected blast scenarios.

Acknowledgements

The authors gratefully acknowledge the precious scientific contribution of Dr. Francesco
Petrini, and Dr. Konstantinos Gkoumas of the Sapienza University of Rome.

References
American Concrete Institute (ACI) (2011) Building Code Requirements for Reinforced Concrete.
Ballantyne, G.J., Whittaker, A.S., Dargush, G.F. and Aref, A.J. (2010) ‘Air-blast effects on
structural shapes of finite width’, Journal of Structural Engineering, Vol. 136, No. 2,
pp.152–159.
Barbato, M., Petrini, F., Unnikrishnan, V.U. and Ciampoli, M. (2013) ‘Probabilistic
performance-based hurricane engineering (PBHE) frame work’, Structural Safety, Vol. 45,
No. 1, pp.24–35.
Biondini, F., Bontempi, F., Frangopol, D.M. and Malerba, P.G. (2006) ‘Probabilistic service life
assessment and maintenance planning of concrete structures’, Journal of Structural
Engineering, Vol. 132, No. 5, pp.810–825.
Ciampoli, M. and Petrini, F. (2012) ‘Performance-based aeolian risk assessment and reduction for
tall buildings’, Probabilistic Engineering Mechanics, Vol. 28, No. 1, pp.75–84.
Ciampoli, M., Petrini, F. and Augusti, G. (2011) ‘Performance-based wind engineering: towards a
general procedure’, Structural Safety, Vol. 33, No. 6, pp.367–378.
Cornell, C.A. and Krawinkler, H. (2000) ‘Progress and challenges in seismic performance
assessment’, PEER Center News, Vol. 3, No. 2.
Cowper, G.R. and Symonds, P.S. (1957) Strain Hardening and Strain Rate Effects in the Impact
Loading of Cantilever Beams, Applied Mathematics Report No. 28, Brown University,
Providence, Rhode Island, USA.
Blast resistant design of precast reinforced concrete walls 15

Davidson, J.S., Fisher, J.W., Hammons, M.I., Porter, J.R. and Dinan, J. (2005) ‘Failure mechanisms
of polymer-reinforced concrete masonry walls subjected to blast’, Structural Engineering,
Vol. 131, No. 8, pp.1194–1205.
Department of Defense (DoD) (2008) Structures to Resist the Effects of Accidental Explosions,
USA.
Enright, M.P. and Frangopol, D.M. (1998) ‘Probabilistic analysis of resistance degradation of
reinforced concrete bridge beams under corrosion’, Engineering Structures, Vol. 20, No. 11,
pp.960–971.
European Committee for Standardization (ECS) (2005) Eurocode 2 – Design of Concrete
Structures – Part 1-1: General Rules and Rules for Buildings.
Gantes, C.J. and Pnevmatikos, N.G. (2004) ‘Elastic-plastic response spectra for exponential blast
loading’, International Journal of Impact Engineering, Vol. 30, No. 3, pp.323–343.
Gentili, F., Giuliani, L. and Bontempi, F. (2013) ‘Structural response of steel high rise buildings to
fire: system characteristics and failure mechanisms’, Journal of Structural Fire Engineering,
Vol. 4, No. 1, pp.9–26.
Gkoumas, K. (2010) ‘A risk analysis framework for offshore wind turbines’, Proceeding of the
12th International Conference on Engineering, Science, Construction, and Operations in
Challenging Environments – Earth and Space 2010, 14–17 March, Honolulu, HI, USA.
Held, M. (1983) ‘Blast waves in free air’, Propellants, Explosives, Pyrotechnics, Vol. 8, No. 1,
pp.1–7.
Luco, N. and Cornell, C.A. (2000) ‘Structure-specific scalar intensity measures for near-source and
ordinary earthquake ground motions’, Earthquake Spectra, Vol. 23, No. 2, pp.357–392.
Mills, C.A. (1987) ‘The design of concrete structures to resist explosions and weapon effects’,
Proceedings of the 1st International Conference for Hazard Protection, Edinburgh, 27–30
September.
Naito, C.J., Dinan, R. and Bewick, B. (2011) ‘Use of precast concrete walls for blast protection of
steel stud construction’, Journal of Performance of Constructed Facilities, Vol. 25, No. 5,
pp.454–463.
Naito, C.J., Hoemann, J., Beacraft, M. and Bewick, B. (2012) ‘Performance and characterization of
shear ties for use in insulated precast concrete sandwich wall panels’, Journal of Structural
Engineering, Vol. 138, No. 1, pp.1–11.
Nemati, K.N. (2007) Formwork for Concrete, CM 420 temporary structures, University of
Washington.
Olmati, P., Petrini, F. and Bontempi, F. (2013a) ‘Numerical analyses for the structural assessment
of steel buildings under explosions’, Structural Engineering and Mechanics, Vol. 45, No. 6,
pp.803–819.
Olmati, P., Gkoumas, K., Brando, F. and Cao, L. (2013b) ‘Consequence-based robustness
assessment of a steel truss bridge’, Steel and Composite Structures, Vol. 14, No. 4, pp.379–395.
Petrini, F. and Ciampoli, M. (2010) ‘Performance-based wind design of tall buildings’, Structure
and Infrastructure Engineering, Vol. 8, No. 10, pp.954–966.
Precast Prestressed Concrete Institute (PCI) (2010a) PCI Design Handbook, 7th ed.
Precast Prestressed Concrete Institute (PCI) (2010b) Blast-Resistant Design of Precast/Prestressed
Concrete Components, PCI Blast Resistance and Structural Integrity Committee.
Sgambi, L., Malerba, P.G., Gotti, G. and Ielmini, D. (2012a) ‘The influence of degradation
phenomena on collapse modes in prestressed concrete beams’, International Journal of
Lifecycle Performance Engineering, Vol. 1, No. 1, pp.41–63.
Sgambi, L., Gkoumas, K. and Bontempi, F. (2012b) ‘Genetic algorithms for the dependability
assurance in the design of a long-span suspension bridge’, Computer-Aided Civil and
Infrastructure Engineering, Vol. 27, No. 9, pp.655–675.
The Federal Emergency Management Agency (FEMA) (2005) Risk assessment: A How-To-Guide
to Mitigate Potential Terrorist Attacks against Buildings, Providing protection to people and
buildings, FEMA 452.
US Army Corps of Engineers (2008a) Methodology Manual for the Single-Degree-of-Freedom
Blast Effects Design Spreadsheets (SBEDS).
US Army Corps of Engineers (2008b) Single Degree of Freedom Structural Response Limits for
Antiterrorism Design.
16 P. Olmati et al.

Vamvatsikos, D. and Dolšek, M. (2011) ‘Equivalent constant rates for performance-based seismic
assessment of ageing structures’, Structural Safety, Vol. 33, No. 1, pp.8–18.

View publication stats

You might also like