(Asce) Be 1943-5592 0000265

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Procedure for Predicting Blast Loads Acting

on Bridge Columns
G. Daniel Williams1 and Eric B. Williamson, P.E., M.ASCE2
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Historical data show that terrorist attacks against transportation assets have increased in recent years and that the vast majority of
these attacks have been bombings. As such, there is growing interest in protecting highway infrastructure from blast loads. To address this
concern, the National Cooperative Highway Research Program (NCHRP) sponsored a project to investigate the performance of highway
bridges subjected to the nearby detonation of an explosive, and this paper presents research that advances the understanding of blast loads
acting on bridge columns. Unlike large wall panels for which much of the existing knowledge about blast effects against structures has been
established, the research presented in this paper focuses on slender structural components in which the effects of cross-sectional geometry,
engulfment of blast pressures, and clearing effects strongly influence loading history. Based on the findings obtained from this study,
a simplified procedure for predicting blast loads acting against bridge columns is proposed. DOI: 10.1061/(ASCE)BE.1943-5592
.0000265. © 2012 American Society of Civil Engineers.
CE Database subject headings: Blast loads; Bridges; Columns; Computer models; Terrorism; Predictions.
Author keywords: Blast loads; Bridges; Computational modeling; Terrorism.

Introduction components, including cable-stayed bridges, suspension bridges,


and many arch bridges) are not the only transportation structures
In 2003, the Blue Ribbon Panel on Bridge and Tunnel Security targeted by terrorists. In their study, Jenkins and Gersten (2001)
(2003), organized through a joint effort of the Federal Highway indicate that between 1980 and 2006, 58% of the attacks world-
Administration (FHWA) and the American Association of State wide—and 35% of the attacks in industrialized countries—were
Highway and Transportation Officials (AASHTO), released a against highway bridges other than signature bridges. Thus, the
report that included recommendations for both near-term and development of a national standard to design blast-resistant high-
long-range research priorities in the area of transportation security. way bridges is imperative, as bridge engineers need guidelines to
This panel was formed shortly after the attacks on September 11, analyze, design, and detail bridges to resist the effects of a terrorist
2001 against the World Trade Center in New York and the Pentagon attack.
in Washington, D.C., out of concern for the vulnerability of our Several researchers have proposed methods to design and ana-
transportation infrastructure. Consistent with the recommendations lyze bridge columns subjected to nearby explosions (Holland 2008;
of the Blue Ribbon Panel on Bridge and Tunnel Security, data Fujikura et al. 2008). To utilize these procedures, it is necessary to
collected by Jenkins and Gersten (2001) of the Mineta Transpor- compute the blast loads that act on a column for a presumed threat
tation Institute demonstrate that transportation assets have been scenario. The research presented in the current manuscript utilizes
frequently targeted by terrorists in the past, and findings from this computational fluid dynamics (CFD) models of airblast scenarios
to enhance the work presented by Holland and Fujikura et al. The
study suggest that this trend will continue into the future. In their
results are used to formulate equations to account for the reduced
report, Jenkins and Gersten (2001) indicate that at least 53 terrorist
load experienced by slender circular and square members
attacks worldwide specifically targeted bridges between 1980 and
(i.e., bridge columns) relative to that of a flat wall.
2001, and 60% of those attacks were bombings. Other researchers
have obtained similar findings (Boyd and Sullivan 2000), showing
that attacks on transportation assets have increased in recent Background
years and that bridges are the most frequently attacked transporta-
tion structure. Importantly, historical data show that signature Several texts, military manuals, and computer programs include
bridges (i.e., unique bridges composed of nontypical structural various forms of equations and curves [Dept. of Defense (DoD)
2008, 2002; U.S. Army Corps of Engineers (USACE) 2001] that
1
Post-Doctoral Research Assistant, Dept. of Civil, Architectural, and can be used to calculate blast loads on flat surfaces. Some of these
Environmental Engineering, The Univ. of Texas at Austin, Austin, TX available procedures include consideration of multiple reflections
78712. E-mail: gdwilliams@mail.utexas.edu and pressure magnification, while not strictly employing computa-
2
Associate Professor, Dept. of Civil, Architectural, and Environmental tional fluid mechanics principles. BlastX [Science Applications
Engineering, The Univ. of Texas at Austin, Austin, TX 78712 (correspond- International Corporation (SAIC) 2001] is one of the most widely
ing author). E-mail: ewilliamson@mail.utexas.edu
recognized applications in this group, and Bridge Explosion Load-
Note. This manuscript was submitted on September 13, 2010; approved
on May 2, 2011; published online on May 5, 2011. Discussion period open ing (BEL) (USACE 2000) is a program developed and supported
until October 1, 2012; separate discussions must be submitted for indivi- by the U.S. Army Corps of Engineers that incorporates BlastX and
dual papers. This paper is part of the Journal of Bridge Engineering, includes a bridge-specific graphical user interface.
Vol. 17, No. 3, May 1, 2012. ©ASCE, ISSN 1084-0702/2012/3-490– Existing methods for computing blast loads are based on many
499/$25.00. years of research and test data collected from blast tests against

490 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012

J. Bridge Eng., 2012, 17(3): 490-499


structures, and, as a result, they provide accurate representations of of the reflected pressure wave form as having a linear rather than
overpressure waveforms when “clearing” effects are not a concern exponential decay from the peak value; application of the relief
(Rickman and Murrell 2007), where clearing is defined as the wave beginning at the time of incident shock arrival; and a
process by which a high reflected pressure seeks relief toward simplified determination of the clearing time. Contrary to the
lower pressure regions (free edges) through a rarefaction (or relief) assumptions that have been used historically, Rickman and Murrell
wave that propagates from the low-pressure to the high-pressure (2007) demonstrated that the measured reflected pressure rapidly
regions (DoD 2008). Because the data with which existing blast decreases to below the stagnation pressure, and the relief wave sig-
load prediction methods have been developed are from blast tests nificantly shortens the duration of the positive reflected pressure
on flat panel structures, they are not capable of modeling round phase, causing the late-time portion of the wave form to have a
geometries such as bridge columns, nor are they able to character- negative value even as Conventional Weapons Effects (CONWEP)
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

ize blast loads on slender members because they lack the ability to (USACE 2001), a computer program developed and supported by
accurately characterize phenomena associated with the relief wave the U.S. Army Corps of Engineers that can be used to predict blast
that travels from the free edges toward the center of a component loads on flat surfaces, indicates a positive pressure. Rickman and
(Rickman and Murrell 2007). Murrell then developed an equation and associated constants for a
Several researchers have addressed the issue of shock waves relief function to subtract the CONWEP wave form from the mea-
interacting with circular members in recent years. Ofengin and sured wave form beginning at the time of arrival of the
Drikakais (1997) studied the diffraction of shock waves over a cyl- relief wave. They also formulated a mathematical expression for
inder for three planar blast waves with short-term, long-term, and the arrival time of the relief wave. Both were combined into a com-
infinite durations. These loads were applied through a high- puter program to calculate reflected impulse curves, and predicted
pressure layer similar to the sudden rupture of a membrane within results compared favorably to both small-scale and large-scale
a shock tube. The results provided by the authors qualitatively char- experimental tests.
acterized the various shock phenomena over a cylinder, and they For the specific case of blast loads acting against highway
also demonstrated that blast wave duration significantly influences bridges, Agrawal and Yi (2009) presented the results of detailed
the pressure variation acting over a cylinder. The findings presented finite element simulations for three levels of blast loads (low,
by the authors agree well with expected results; however, they are medium, and large). While finite element programs such as
not provided in a format useful for engineering applications, LS-DYNA offer the ability to simulate shock wave propagation
especially the results for blasts of infinite duration. through a medium (e.g., air), such models can be computationally
Shi et al. (2007) used Autodyn 3D to model semispherical sur- intensive and require significant time to run. Therefore, rather than
face detonations in air and the resulting shock interaction with re- directly modeling detonation effects in their simulations, Agrawal
inforced concrete square and circular columns at a 10-m standoff. and Yi (2009) proposed a method for generating loads in their finite
The model was validated using results from a blast test reported in element models by using results computed from CONWEP. The
the literature, and the researchers then conducted a parameter study CONWEP-generated loads were then transmitted through an air
to investigate the influence of five parameters on column-shock medium to structural components, thereby making the load simu-
interaction: (1) Scaled distance (Z)—defined as the distance from lation phase of the analysis more efficient than directly modeling
the center of detonation to the target point of interest (ft. or m) detonation effects. This approach was adopted by Agrawal and Yi
divided by the charge weight (lbs. or kg) raised to the 1∕3 power (2009) for use in very detailed finite element models that were used
(i.e., Z ¼ R∕W 1∕3 ); (2) Ratio of supported mass to column mass; to parametrically evaluate the effects of various design variables
(3) The stiffness of the column; (4) The dimension of the column; (e.g., concrete strength, seismic detailing) on the performance of
and (5) The geometry of the column. The results showed that the blast-loaded bridge components. The work presented in the current
column mass ratio, the column stiffness, and the column deflection paper extends the results reported by Agrawal and Li (2009) in two
had no significant effect on the loading. The scaled distance and the significant ways. First, Agrawal and Li (2009) focused their sim-
column dimension, however, were shown to significantly influence ulations on a bridge that comprised square columns, whereas the
reflected pressure and impulse on the columns, where impulse is current work considers the effects of column cross-sectional shape
defined as the area under the pressure-time loading curve. These on the loads that must be resisted in a blast event. Second, the focus
findings are consistent with the research results reported later in of the current study is on the development of simplified load
this paper. prediction methods that can be used for design purposes rather than
Shi et al. (2007) then developed factors to account for shape detailed finite element simulations. Thus, while the research
effects when calculating reflected pressure and impulse on the front presented in the current manuscript utilizes detailed finite element
and back centerlines of a column using the positive incident pres- simulations, the emphasis is on interpreting the results so that they
sure and the positive incident impulse along with the column shape can be readily implemented in the blast-resistant design of bridge
and dimension. While these equations and factors provide insight columns.
into the interaction of blast loads on columns, they have limited Of most interest to the current work, researchers at the State
application for the design and analysis of bridge columns because University of New York investigated a design alternative and ana-
they are restricted to a standoff distance of 10 m and they only lytical approach for blast-resistant bridge columns (Fujikura et al.
consider the loads acting on the centerline of the column. Research 2008). The design utilized a concrete-filled steel tube anchored into
findings presented later in this manuscript and by others have the foundation with steel plates linked to concrete-embedded
shown a significant reduction in net impulse across the entire C-channel shapes. One-fourth-scale prototypes of the concrete-
cross-section of a column due to shape and slenderness effects filled steel tube columns were tested experimentally. The research-
(Williams 2009; Fujikura et al. 2008). ers also developed an analytical method to calculate the response of
Rickman and Murrell (2007) conducted a series of small-scale the bridge columns using an equation-based single-degree-of-
experiments to characterize the relief wave form, the clearing time, freedom (SDOF) method. The energy imparted to the structural
and the resulting loads on walls. This effort was in response to a system by the blast loads was considered to be an impulsive loading
recognition of the following errors in existing methods used to in this research because the blast load duration was much shorter
calculate reflected pressures and impulses on walls: approximation than the natural period of the columns that were tested, which is

JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012 / 491

J. Bridge Eng., 2012, 17(3): 490-499


typical for many blast-resistant design applications. The research- The parameter study was conducted in conjunction with
ers presented Eq. (1) to calculate impulse in their analytical NCHRP Project 12-72 (Williamson et al. 2010). While the charge
method: weights and standoff distances used during those tests are of inter-
est for design and research purposes, disclosure of that information
I eq ¼ βDieq ð1Þ would provide a link to the damage results in the final report for
NCHRP 12-72, and it would be irresponsible to distribute that
information publicly. Therefore, the charge weights and standoff
where β = reduction factor to account for the circular shape of the
distances are shown in Table 1 only in terms of the variables w
column; D = diameter of the column; and ieq = calculated impulse.
and z, respectively. This approach is consistent with that of other
The researchers used their predictions of column response and the
researchers reporting such data (e.g., Fujikura et al. 2008). The ma-
experimental results to back-calculate a β factor of 0.45 to account
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

jority of the charge weights and standoff distances evaluated during


for shape and slenderness effects. They found this value more
this parameter study fell within or just outside the limits of design
appropriate than the 0.80 value previously reported in the literature
Category C, as proposed in the final report for NCHRP 12-72
(Winget et al. 2005). Interestingly, as documented later in this
(Williamson et al. 2010). Category C threats are defined as those
manuscript, a value of 0.45 for β is shown to correspond extremely
having a scaled standoff less than 1:5 ft∕lbs1∕3 (0:6 m∕kg1∕3 ), and
well to the equations independently developed in the current
this design category is the only one for which a load prediction
research to account for shape and slenderness effects on bridge
technique is needed and for which a dynamic response analysis
columns.
must be conducted.
While each of the research efforts described above has signifi-
cantly advanced the understanding of shock wave interaction with Numerical Modeling of Airblast and Shock Propagation
columns, no comprehensive procedure has been developed for use
by design engineers to calculate blast loads acting on slender square Numerically modeling high-explosive burn and shock propagation
and circular bridge columns. The purpose of the research presented requires an equation-of-state to govern the expansion of gaseous
in this manuscript is to develop such a procedure utilizing computa- detonation products from a high-pressure, high-density condition
tional fluid dynamics (CFD) airblast models. The computational immediately after an explosion to a normal (i.e., ambient) pressure
models, the experimental tests used to validate these models, and density that exists after an explosion has occurred. The Jones-
and the resulting expressions that account for column shape and Wilkins-Lee (JWL) equation-of-state was utilized in this research
slenderness are presented herein. because it is commonly used and widely accepted for airblast
modeling (Sućeska 1999; Alia and Souli 2006). Eq. (2) shows
the most basic formulation of the JWL equation-of-state used in
Parametric Study to Characterize Blast Loads on this research, and Table 2 shows the input values selected for
Slender Members use with this equation.
   
ϖ ϖ E
A numerical parameter study was carried out to investigate the P¼A 1 eR1 V þ B 1  eR2 V þ ϖ o ð2Þ
influence of charge weight (W), standoff distance (R), and column R1 V R2 V V
diameter (D) on the loads experienced by square and circular
columns subjected to the nearby detonation of a high explosive. where P = pressure of the gaseous detonation products; V = relative
The study included 16 sets of three analyses—one square column, specific volume (ν∕ν o ) of the gaseous detonation products; ν = cur-
one circular column, and one wall—and each analysis in a set had rent specific volume (V t ∕M) of the gaseous detonation products;
the same charge weight and standoff distance. Furthermore, the ν o = reference specific volume (V o ∕M) of the gaseous detonation
modeling parameters, including mesh density, geometry of the ex- products; V t = current volume of the gaseous detonation products;
plosive, and so on, were exactly the same for each analysis set. V o = initial volume of the gaseous detonation products; M = mass
Multiple mesh densities were considered for each set of analyses, of the gaseous detonation products; Eo = initial internal energy of
and though the mesh density varied from one set to the next, the the explosive; A, B, R1 , R2 , and ϖ are specified constants unique to
maximum mesh density computationally viable for each model was each explosive.
used to maximize accuracy based on recommendations reported in The linear polynomial equation-of-state was selected for use in
the literature (Luccioni et al. 2006; Alia and Souli 2006). Models this research among several possible alternatives for modeling the
utilized in this study included over 950,000 elements, and the response of air, and one form of this equation is shown in Eq. (3).
height and length of each air mesh was selected to be at least twice The linear polynomial equation-of-state enforces a linear relation-
the column height or the standoff distance, whichever was greatest, ship between pressure and internal energy and can be reduced to a
to allow the shock front to propagate completely past the top of the form that represents the ideal gas law. This reduction is shown
column before reaching the boundary of the mesh. In addition, all below using Eqs. (3) and (4), and the properties of air used in this
models contained at least 16 elements along the length of the ex- research are shown in Table 2.
plosive region, as this minimum number of elements is required to P ¼ C 0 þ C 1 μ þ C 2 μ2 þ C 3 μ3 þ ðC4 þ C 5 μ þ C 6 μ2 ÞE ð3Þ
adequately burn the explosive material and build the detonation
pressure (Alia and Souli 2006). The analyses containing square where
and circular columns characterized the loads acting on those mem- ρ
bers, and the wall analyses provided load data on a flat surface at μ¼ 1 ð4Þ
ρo
the same charge weights and standoff distances as the columns.
This technique was selected because analysts can use several and C 0  C 6 are constants, E = internal energy per unit
widely accepted methods to calculate accurate blast loads on flat reference volume, ρ = current density of air, ρo = initial density
panel structures when clearing is not a concern (Rickman and of air.
Murrell 2007), and the blast engineering community has high con- Fig. 1 shows an arbitrarily selected airblast model example for
fidence in these methods because they are based on extensive data which the dimensions have been adjusted intentionally to prevent
gathered from many decades of blast tests. scaling of the charge size and standoff distance. All analyses

492 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012

J. Bridge Eng., 2012, 17(3): 490-499


Table 1. Air Blast Models for Parameter Study Validation of Computational Fluid Dynamics Models
Column Column Several experimental blast tests were conducted for NCHRP 12-72
Charge diameter height Column in conjunction with the research presented in this manuscript, and
Column ID weight Standoff [m (ft)] [m(ft)] type
some of the airblast models presented in Table 1 represent columns
1.4w 1.0z 0.46 (1.5) 3.43 (11.25) Circular from that experimental test program. The objective of modeling the
Set 1 1.4w 1.0z 0.46 (1.5) 3.43 (11.25) Square experimental blast tests was to validate the numerical models prior
1.4w 1.0z Wall 3.43 (11.25) Wall to conducting a series of parameter studies for both loading and
1.4w 1.0z 0.76 (2.5) 3.43 (11.25) Circular response. The pressure-time histories derived from the CFD models
described previously were applied to numerical models of rein-
Set 2 1.4w 1.0z 0.76 (2.5) 3.43 (11.25) Square
forced concrete columns, and the response was compared qualita-
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

1.4w 1.0z Wall 3.43 (11.25) Wall


tively and quantitatively to the results of the experimental blast
1.4w 1.0z 1.22 (4.0) 3.43 (11.25) Circular tests. Due to length restrictions, a detailed discussion of the exper-
Set 3 1.4w 1.0z 1.22 (4.0) 3.43 (11.25) Square imental blast tests and corresponding test results is beyond the
1.4w 1.0z Wall 3.43 (11.25) Wall scope of the current paper, and only a brief description is provided
1.4w 2.0z 0.46 (1.5) 3.43 (11.25) Circular to show the procedure by which the CFD models were validated.
Set 4 1.4w 2.0z 0.46 (1.5) 3.43 (11.25) Square Readers are directed to Holland (2008), Williams (2009), and
1.4w 2.0z Wall 3.43 (11.25) Wall Williamson et al. (2010) for further information on the experimen-
1.4w 2.0z 0.76 (2.5) 3.43 (11.25) Circular tal tests and validation procedure.
Set 5 1.4w 2.0z 0.76 (2.5) 3.43 (11.25) Square
Experimental Test Setup
1.4w 2.0z Wall 3.43 (11.25) Wall
1.4w 2.0z 1.22 (4.0) 3.43 (11.25) Circular The research program included the evaluation of large-scale rein-
Set 6 1.4w 2.0z 1.22 (4.0) 3.43 (11.25) Square forced concrete bridge columns subjected to severe (i.e., close-in)
1.4w 2.0z Wall 3.43 (11.25) Wall threat scenarios. The boundary conditions used in the test program
1.4w 3.0z 0.46 (1.5) 3.43 (11.25) Circular were those of a propped cantilever. These boundary conditions
Set 7 1.4w 3.0z 0.46 (1.5) 3.43 (11.25) Square modeled a scenario in which the location of an explosion is directly
beneath the deck and directly to the side of a column, assuming the
1.4w 3.0z Wall 3.43 (11.25) Wall
superstructure provides lateral restraint. Axial loads experienced by
1.4w 3.9z 0.46 (1.5) 3.43 (11.25) Circular
bridge columns in service typically do not exceed the balanced
Set 8 1.4w 3.9z 0.46 (1.5) 3.43 (11.25) Square failure point, and any applied axial load tends to increase the
1.4w 3.9z Wall 3.43 (11.25) Wall strength and resistance of a column (Williamson et al. 2010). Thus,
1.4w 3.9z 0.76 (2.5) 3.43 (11.25) Circular testing specimens as propped-cantilevers with no applied axial load
Set 9 1.4w 3.9z 0.76 (2.5) 3.43 (11.25) Square was conservative and logistically desirable. Furthermore, reflected
1.4w 3.9z Wall 3.43 (11.25) Wall loads from the deck occur sufficiently late in time so as to have
1.4w 3.9z 1.22 (4.0) 3.43 (11.25) Circular limited impact on peak column displacement in most cases (Winget
Set 10 1.4w 3.9z 1.22 (4.0) 3.43 (11.25) Square et al. 2005), and therefore testing a column as shown in Fig. 2
1.4w 3.9z Wall 3.43 (11.25) Wall provides a convenient and effective setup for evaluating bridge
2.9w 2.8z 0.76 (2.5) 3.43 (11.25) Circular column performance under severe blast loads.
Set 11 2.9w 2.8z 0.76 (2.5) 3.43 (11.25) Square
Numerical Models of Experimentally Tested Reinforced
2.9w 2.8z Wall 3.43 (11.25) Wall Concrete Columns
2.9w 3.1z 0.46 (1.5) 3.43 (11.25) Circular
Set 12 2.9w 3.1z 0.46 (1.5) 3.43 (11.25) Square Each of the computational models representing column response
2.9w 3.1z Wall 3.43 (11.25) Wall
had multiple analyses that evaluated the effectiveness of different
mesh sizes and modeling approximations for the transverse
4.3w 2.0z 0.76 (2.5) 3.43 (11.25) Circular
reinforcement, and two of the final models employed erosion tech-
Set 13 4.3w 2.0z 0.76 (2.5) 3.43 (11.25) Square
niques to visually represent spalling of cover concrete. The adequacy
4.3w 2.0z Wall 3.43 (11.25) Wall of the models was evaluated using the peak residual displacements,
8.6w 4.1z 0.76 (2.5) 3.43 (11.25) Circular shapes of global response, and cracking and damage patterns.
Set 14 8.6w 4.1z 0.76 (2.5) 3.43 (11.25) Square Accurate finite element modeling of concrete requires a com-
8.6w 4.1z Wall 3.43 (11.25) Wall plex constitutive model that accurately represents the nonlinear
9.8w 3.1z 0.76 (2.5) 3.43 (11.25) Circular behavior of concrete. The Karagozian and Case concrete model
Set 15 9.8w 3.1z 0.76 (2.5) 3.43 (11.25) Square used in this research is a three-invariant model that includes
9.8w 3.1z Wall 3.43 (11.25) Wall confinement effects, strain-rate effects, and brittle and/or ductile
23.0w 5.0z 0.76 (2.5) 3.43 (11.25) Circular damage (LSTC 2007; Schwer and Malvar 2005; Malvar et al.
Set 16 23.0w 5.0z 0.76 (2.5) 3.43 (11.25) Square
2000; Williams 2009). The Simplified Johnson-Cook material
model (LSTC 2007) was selected as the constitutive model for
23.0w 5.0z Wall 3.43 (11.25) Wall
the reinforcing steel in this research (Williams 2009).
All column analyses utilized one-half symmetry along the
longitudinal centerline to reduce computational demand.
Constant-stress solid elements and multiple mesh densities were
assumed the detonation point was at the center of the explosive. considered to ensure computational efficiency. The models utilized
Pressure-time histories were recorded at 220 and 240 intervals truss elements to represent the longitudinal and transverse
along the height and width of the circular and square columns, reinforcement in order to maintain displacement compatibility with
respectively, on both the front and back faces. the solid elements. The truss elements that were used to model the

JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012 / 493

J. Bridge Eng., 2012, 17(3): 490-499


Table 2. Equation-of-State Input Parameters for Air and TNT
Material A (MPa) B (MPa) R1 R2 Eo (MPa) Ω V d (cm∕m sec) pCJ (MPa) ρo (g∕cm3 ) γ
TNT 371,200 3,230 4.15 0.95 7,000 0.30 0.693 21,000 1.630 —
Air — — — — 0.25 — — — 1:293 × 103 1.4
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Example of numerical air blast models

reinforcing steel shared nodes with the concrete elements, and as vertical base constraints (i.e., no horizontal base constraints in ei-
such the model did not explicitly consider bar slip. This approxi- ther direction) on base nodes outside the transverse reinforcement
mation was acceptable because highly distorted concrete elements (i.e., in the region of cover concrete) to allow erosion of cover con-
could be eroded from the model, and as such rebar elements did not crete. The pressure-time histories applied to the numerical models
necessarily remain connected to concrete elements for the entire of the blast-tested columns were computed using the computational
analysis duration, allowing them to move independently from fluid dynamics (CFD) numerical techniques discussed above
the concrete elements under large deformations when rebar slip (Williams 2009). Readers are directed to Williams (2009) for addi-
tional information on the finite element models used to simulate the
is most pronounced. All nodal translational displacements were
experimental blast tests carried out under NCHRP 12-72.
constrained at the base of the model to be consistent with the fixed
boundary condition provided by the column foundation and reac-
tion structure. The models in which spalling occurred had only Validation Efforts

The experimental blast tests on reinforced concrete columns pro-


vided important data to verify the fidelity of the simulation models
both qualitatively and quantitatively. Strain gauges were employed
during the tests; however, only one strain gauge data set out of eight
tests survived with sufficient data to compare to the strains recorded
from the numerical models, and one data set does not provide a
statistically significant comparison for eight total tests. Therefore,
the experimental data available for validating the computational
models were limited to photographs, hand-drawn cracking dia-
grams, and hand-measured residual displacements. Using that data,
the qualitative performance of the numerical column models was
assessed using the global response shapes, the observed damage
patterns, and the recorded cracking patterns, and the quantitative
performance of the numerical column models was assessed using
peak residual displacements. The use of peak residual displacement
as a means of evaluating computational model accuracy for col-
umns subjected to blast loads has been utilized previously by other
researchers (Fujikura et al. 2008). Because of the extensive quantity
Fig. 2. Experimental test setup
of numerical data collected during these verification analyses, it is

494 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012

J. Bridge Eng., 2012, 17(3): 490-499


Table 3. Numerical Column Model Performance Criteria pressure-time histories calculated from the CFD models, represent
Experimental Computational the cracking and damage patterns quite well, as shown in Fig. 3.
Column ID displacement mm (in) displacement mm (in) The performance of all final column models was judged based
on peak residual displacement, response shape, and damage and
1A-2 127 (5.0) 109 (4.3) cracking patterns. The response shape and cracking patterns of each
1B 19.1 (0.75) 25.4 (1.0) numerical column model were assessed qualitatively relative to the
2A-2 12.7 (0.5) 12.7 (0.5) other columns modeled during the research program. Table 3 com-
2B 6.35 (0.25) 2.54 (0.1) pares the recorded and computed residual displacements. It is
2-Seismic 5.08 (0.2) 2.54 (0.1) notable that the numerical models for the three columns that sus-
2-Blast 12.7 (0.5) 10.2 (0.4) tained the most displacement and damage—Columns 1A-2, 3A,
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

3A 76.2 (3.0) 66.0 (2.6) and 3Blast—generated slightly lower displacements relative to
3-Blast 114 (4.5) 78.7 (3.1) the experimental results. The main contributor to the small pre-
dicted displacement for these columns is the inability of the con-
crete constitutive model to accurately represent degradation of
not practical to provide a detailed discussion of each individual concrete in compression and shear. Stress versus strain curves from
numerical column model used to validate the CFD models. The single element stress analyses of unconfined axial compression
results of all models are summarized in Table 3, and one numerical tests at various loading rates using the Karagozian and Case con-
column model is briefly presented herein. crete model showed that the material model exhibits an unrealisti-
Fig. 3 shows both the experimental and numerical damage and cally long unloading phase after reaching the peak concrete
cracking patterns for Column 3A. The damage is shown using a compressive strength with a strain well beyond 0.008, which is sig-
normalized scale of 0 to 2. A value of 1 indicates that the material nificantly larger than the typical range of 0.003 to 0.004
has just reached its peak strength, and a value between 1 and 2 (MacGregor 1997). Concrete constitutive modeling of blast-loaded
indicates that the material has passed its peak strength and is soft- structures remains a computational challenge (DoD 2002), and this
ening. Accordingly, for tension, values close to 2 indicate that the field continues to be an active area of research. Despite this limi-
concrete is cracking, and for compression, such values indicate that tation, the numerical models used in this study performed quite well
the concrete is crushing (i.e., softening) and nearing failure (or per- with the pressure-time history data derived from the CFD analyses
haps has already reached failure). The appearance of failure on the described above.
compression side of the column must be evaluated carefully, as an
Characterization of Reflected Impulse on Columns
indication of large plastic strains in the numerical models do not
always correlate to similar damage in the experimental models Eqs. (5) and (6) were employed to compute the net resultant im-
due to limitations in the constitutive models used to represent pulse (i.e., the difference in impulse between the front and back
the concrete behavior [as described further below and in Williams faces) at a given cross-section using the pressure-time history data
(2009)]. Overall, the numerical models, which were loaded using and the tributary areas shown in Fig. 4.

Fig. 3. Comparison of response shapes and damage patterns for Column 3A: (a) cracking and damage pattern from analytical model; (b) analytical
model with overlay of real cracking pattern; and (c) real cracking and damage pattern

JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012 / 495

J. Bridge Eng., 2012, 17(3): 490-499


ðT1  T11Þ þ 2ðT2  T10Þ þ 2ðT3  T9Þ þ 2ðT4  T8Þ þ 2ðT5  T7Þ þ ðT6  T6Þ
I nr:circular ¼ ð5Þ
10

ðT1  T12Þ þ 2ðT2  T11Þ þ 2ðT3  T10Þ þ 2ðT4  T9Þ þ 2ðT5  T8Þ þ ðT6  T7Þ
I nr:square ¼ ð6Þ
10

The designation T# in Eqs. (5) and (6) represents numerically at a position above the ground than at the base of the column.
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

generated pressure-time history data from the locations labeled Shi et al. (2007) reported similar findings for small scaled distan-
by the corresponding numbers in Fig. 4. For example, considering ces, albeit at a slightly higher position along the column. This phe-
the first term in the numerator of Eq. (6), Fig. 4 shows that locations nomenon is likely due to pressure magnification from reflections of
“1” and “12” are directly opposite each other, and it is the differ- the shock wave off the ground, the nature of which is a function of
ence in loading between these locations that is used to compute the standoff, charge size, and shape.
net resultant impulse. Impulse values are considered to be the most Several attempts were made to normalize these data into equa-
important load parameter for this study because the net resultant tions that could be used to calculate equivalent uniform impulse
impulse typically governs the global response of blast-loaded mem- values acting against circular and square columns for blast-resistant
bers when the duration of loading is much less than the natural design purposes. In this context, an equivalent loading that has a
period of the structure being analyzed (Biggs 1964; DoD 2008; uniform distribution rather than one that varies along the column
Fujikura et al. 2008), which was the case for all scenarios height can be derived using principles of work and energy conser-
considered during this research program. Curves that indicate vation, and this approach is used in various existing software pack-
how blast intensity varies with position along the height of a ages that compute blast loads, including CONWEP (USACE 2001)
column were developed based on the net resultant impulse values and BEL (USACE 2000). Given the uncertainty associated with
computed at 20 equally spaced cross-sections along the height of potential threat scenarios, an approach that utilizes an equivalent
each column. uniform load is often preferred to simplify the design process
Because the numerical CFD parameter study included a variety (DOD 2008; Conrath et al. 1999). The effort to develop equivalent
of charge sizes and standoff distances, actual pressures and uniform impulse values encountered significant difficulty because
impulses acting on the analyzed columns varied considerably. of the highly nonlinear nature of shock propagation and the
To allow for consistency in interpreting the computed results, nor- complex interaction between a shock wave and a column. The best
malized net resultant impulse curves (termed “load curves” in this solution yielded results that could differ up to 35% from the results
paper) were developed by calculating and plotting the net resultant of the numerical models. Thus, this approach was abandoned in
impulse at various vertical locations along the height of the col- favor of an alternative method for computing equivalent blast
umns and normalizing those values to the computed peak net result- impulse acting on a column.
ant impulse value for each column (Fig. 5). This figure is difficult to Methods already exist to calculate pressure wave forms and the
understand with the numerous curves shown on the plot, but all resulting impulses acting against flat panel structures (DoD 2008;
curves are shown to allow a detailed independent investigation USACE 2001; SAIC 2001; USACE 2000), and recent research
of the results if desired. The peak reflected impulse on each column confirms their applicability and accuracy when clearing from free
was chosen as the basis for normalization because peak reflected edges is not a concern (Rickman and Murrell 2007). Therefore, a
impulse can be readily calculated with high confidence using revised method was developed for computing equivalent uniform
several widely accepted methods (e.g., USACE 2000; DoD impulse values for use in analyzing blast-loaded bridge columns.
2008). The curves in Fig. 5 show a higher net resultant impulse The proposed method utilizes existing procedures for computing
equivalent uniform impulse values against flat panels (USACE
2001; SAIC 2001; USACE 2000), coupled with column shape
factors that account for the effects of cross-sectional shape and
slenderness on load history. The procedure used to develop the
column shape factors for slender circular and square members
subjected to blast loads is described in the following section. After-
ward, a simplified design procedure based on these factors is
presented.

Shape Factors to Account for Column Shape and


Slenderness

Several factors contribute to the fact that a circular column will


experience less impulse than a similarly sized square column.
Reflected pressures, and more importantly reflected impulses, de-
pend upon the peak incident overpressure and the angle of inci-
(a) (b) dence of a shock front’s direction of travel relative to the target
it impacts. The angle of incidence at a given location around the
Fig. 4. Tributary areas for net resultant impulse and geometric differ-
front face of a circular column is greater than that of a similar
ences between circular and square column blast cases: (a) circular
position along the front face of a square column (i.e., locations with
column; and (b) square column
the same horizontal distance from the column centerline), as shown

496 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012

J. Bridge Eng., 2012, 17(3): 490-499


1.0
W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 1.5 ft; circular
All charges are on ground. W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 1.5 ft; square
W = 1.4w sph.; R = 1.0z; Z = 0.9; D = 2.5 ft; circular
0.9 W = 1.4w sph.; R = 1.0z; Z = 0.9; D = 2.5 ft; square
W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 2.5 ft; circular
W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 2.5 ft; square
W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 4.0 ft; circular
0.8
W = 1.4w cyl.; R = 1.0z; Z = 0.9; D = 4.0 ft; square
W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 1.5 ft; circular
W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 1.5 ft; square
0.7 W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 2.5 ft; circular
W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 2.5 ft; square
Column Location, y/Lo

W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 4.0 ft; square


W = 1.4w cyl.; R = 2.0z; Z = 1.7; D = 4.0 ft; circular
0.6
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

W = 1.4w cyl.; R = 3.1z; Z = 2.7; D = 1.5 ft; circular


W = 1.4w cyl.; R = 3.1z; Z = 2.7; D = 1.5 ft; square
W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 1.5 ft; circular
0.5 W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 1.5 ft; square
W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 2.5 ft; circular
W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 2.5 ft; square
W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 4.0 ft; circular
0.4
W = 1.4w cyl.; R = 3.9z; Z = 3.5; D = 4.0 ft; square
W = 2.9w cyl.; R = 2.8z; Z = 2.0; D = 2.5 ft; circular
W = 2.9w cyl.; R = 2.8z; Z = 2.0; D = 2.5 ft; square
0.3 W = 2.9w cyl.; R = 3.1z; Z = 2.2; D = 1.5 ft; circular
W = 2.9w cyl.; R = 3.1z; Z = 2.2; D = 1.5 ft; square
W = 2.9w cyl.; R = 4.7z; Z = 3.3; D = 2.5 ft; circular
W = 4.3w cyl.; R = 2.0z; Z = 1.2; D = 2.5 ft; square
0.2
W = 4.3w cyl.; R = 2.0z; Z = 1.2; D = 2.5 ft; circular
W = 8.6w cyl.; R = 4.1z; Z = 2.0; D = 2.5 ft; circular
W = 8.6w cyl.; R = 4.1z; Z = 2.0; D = 2.5 ft; square
0.1 W = 9.8w cyl.; R = 3.1z; Z = 1.5; D = 2.5 ft; circular
W = 9.8w cyl.; R = 3.1z; Z = 1.5; D = 2.5 ft; square
W = 23.0w cyl.; R = 5.0z; Z = 1.7; D = 2.5 ft; circular
W = 23.0w cyl.; R = 5.0z; Z = 1.7; D = 2.5 ft; square
0.0
0.00 0.20 0.40 0.60 0.80 1.00 1.20
Normalized Impulse, Ir/Ir.peak

Fig. 5. Normalized impulse curves for cylindrical bursts against bridge columns

in Fig. 4. Moreover, the peak incident overpressure is less at a given other factors contribute to the nonlinearity of the data, including
location around the circumference of a circular column than at a varying charge weights, charge sizes, standoff distances, column
similar position along the face of a square column due to the sizes, clearing times, angles of incidence, heights-of-burst, and
increased standoff distance to the position on the circular column ground reflections. Therefore, rather than normalizing the data,
relative to that on a square column. Both an increase in angle of it is more appropriate for design purposes to determine suitable
incidence and a decrease in peak incident overpressure individually upper bounds as shown in Fig. 6. The column shape factors are
produce reduced reflected pressures and impulses, and thus with found using Eq. (7) for a square column and Eq. (8) for a circular
the combination of the two a circular column will experience less column:
net load than a square column. Furthermore, it is important to note  
that load clearing around circular columns is different than around R
Ss ¼ 0:013 þ 0:49 ð7Þ
square columns. Shock waves engulf circular columns more easily D
than square columns, making the resultant impulse on the back face
of circular columns greater than that on the back face of similarly
sized square columns, further reducing the net resultant load a
circular column must resist relative to a square column. This find- 0.60
ing is consistent with those of Shi et al. (2007) as described above.
Fig. 6 shows the column shape factors derived from the airblast 0.55
models analyzed for this research. These values were determined
0.50
Column Shape Factor, S

by normalizing the net reflected impulse on each column to the


peak reflected impulse acting on a wall with the same charge 0.45
weight and standoff distance. As described previously, blast loads
acting on large, flat walls have been studied extensively in the past, 0.40
allowing peak reflected impulses and pressures to be readily calcu- 0.35
lated for these components (USACE 2001; SAIC 2001; USACE
Square Column Shape Factors
2000). Thus, blast loading parameters computed for a flat wall 0.30
Circular Column Shape Factors
provide a convenient basis for normalizing the loads acting on slen-
0.25 Square Upper Bound
der circular and square columns of finite width. The numerical data Circular Upper Bound
clearly show that circular columns experience less net resultant 0.20
impulse than similarly sized square columns; however, there is 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
noticeable scatter in the data. Scatter in the computed results exists Standoff Distance/Column Diameter, R /D (ft./ft.)
because of the highly nonlinear relationship between scaled stand-
Fig. 6. Column shape factors as a function of the ratio of the standoff
off and peak reflected impulse, which are only two of the factors
distance to the column diameter
that contribute to the net resultant impulse (DoD 2008). Multiple

JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012 / 497

J. Bridge Eng., 2012, 17(3): 490-499


 
R that occurs relative to a large, flat surface for slender structural
SC ¼ 0:019 þ 0:39 ð8Þ members of finite width.
D
The first step in the load prediction procedure is the calculation
of the peak equivalent uniform pressure and an equivalent uniform
impulse acting along the centerline of the column being analyzed.
where SS = column shape factor for a square column (unitless);
This calculation can be achieved by assuming an ideal reflected
SC = column shape factor for a circular column (unitless);
pressure wave form utilizing widely accepted software for comput-
R = standoff distance (ft. or m); D = column diameter or edge width
ing blast loads on flat panel surfaces, such as CONWEP (USACE
(ft. or m).
2001), BlastX (SAIC 2001), or BEL (USACE 2000). The second
These upper-bound equations are valid for R∕D ratios less than
step in the proposed load prediction procedure is the calculation of
4.5. Most importantly, the numerical models used to generate the
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

a column shape factor, SS or SC , using Eq. (7) for square columns


data on which these equations are based use scaled standoffs that
and Eq. (8) for circular columns. The final step in the proposed
correspond to Blast Design Category C, as outlined by Williamson
procedure to predict blast loads on bridge columns is the determi-
et al. (2010). As described earlier in this paper, Blast Design
nation of the load duration, t d . This value is determined using
Category C is the only category that requires a dynamic analysis,
Eq. (9).
and thus these equations are valid over the range of scaled standoffs
for which they will be used for design. Eqs. (7) and (8) may over- 2Sie
estimate the net resultant impulses for some scenarios because they td ¼ ð9Þ
pe
are upper bounds, but they provide much more accurate and less
conservative loads than those predicted using experimental data where pe = peak equivalent uniform pressure on the column (lbs∕in
from blasts against flat walls, which is the method currently or kN∕m); S = appropriate column shape factor found by Eq. (7)
employed by software used to predict blast loads acting on bridge or (8) (unitless); ie = equivalent uniform impulse (psi-msec or
columns (USACE 2000). kPa-msec); t d = duration of the positive phase of blast
It is notable that the values calculated by these equations com- loading (msec).
pare extremely well to the value of 0.45 presented for the load Eq. (9) preserves the total equivalent uniform impulse on a col-
reduction factor β by Fujikura et al. (2008). The intent of the β umn for an assumed linear variation in the blast pressure as a func-
factor in that work was the same as the column shape factors in tion of time, which is a common assumption for blast-resistant
the current research; the constant value of 0.45 was an average design (e.g., DoD 2008). The peak equivalent uniform pressure
of several values found using a single-degree-of-freedom (SDOF) leads to a resultant load for an SDOF analysis by multiplying
response analysis and the results of experimental blast tests on by the height and dimension (i.e., edge width or diameter) of
concrete-filled steel tube columns. While the values computed the column being analyzed. The resultant load, and the duration
using Eqs. (7) and (8) compare favorably to the work reported of the positive phase, td , are the parameters typically needed to
by Fujikura et al. (2008), it is important to recognize that the define an idealized triangular blast pulse (i.e., one that varies lin-
approach taken to arrive at these values for the current study is quite early in time) for an SDOF dynamic analysis. The procedure
different from that of Fujikura et al. In their research, a load reduc- described in this section does not compute the resultant load di-
tion factor was computed based on response analyses using an rectly to allow bridge designers to use the peak equivalent uniform
SDOF model, while the values obtained from the current study load with a different analysis procedure if desired (i.e., frame
have been computed solely from CFD analyses of loads and thus analysis).
do not depend on the subsequent column response. Although the
exact standoffs used by Fujikura et al. are undisclosed for security
purposes, the column and blast scenarios presented appear to have Evaluation of Simplified Load Prediction Method
R∕D ratios ranging from 1.7 to 5.0. These R∕D ratios yield column
shape factors calculated by Eq. (8) that range from 0.423 to 0.485. Table 4 shows a comparison among the impulse values computed
These values are consistent with the range of 0.417 to 0.472 for β as using the column shape factors, the method currently employed by
back-calculated by Fujikura et al. (2008) based on experimental existing blast-prediction software such as BEL (USACE 2000)
data. Furthermore, the actual data presented for β do show a general (i.e., no adjustment for column shape and slenderness), and a
increase as the standoff increases, confirming the results and computational fluid dynamics model. Percentage difference values
equations presented in this paper. Therefore, Eqs. (7) and (8) cal- are computed assuming the CFD solution is the correct result. The
culate column shape factors validated both by computational fluid proposed method generates a load value for design that is only 7%
dynamics and experimental data. different than the value assumed to be correct, whereas the method
currently used in existing software drastically overpredicts the
loads on the slender circular member by 120%. While the proposed
Load Prediction Method method does calculate a slightly unconservative value compared to
the numerical airblast model, this level of inaccuracy is considered
The column shape factors can be used with any method that to be acceptable given the high amount of uncertainty typically
calculates load variation along the height of a column based on
the ideal reflected impulse. For the present discussion, one example
procedure is provided to demonstrate how the column shape factors
can be used in the analysis and design of bridge columns subjected Table 4. Final Comparison of Impulse Values
to blast loads when an equivalent uniform loading is desired. Impulse load Percent
Because the column shape factors are based only on shock inter- value ðN∕mmÞ-msec difference (%)
action with columns, they are applicable to design procedures that Proposed method 3,424.6 7:3
utilize SDOF methods as well as more complex methods, such as
Regular method 8,153.9 120.7
multiple-degree-of-freedom (MDOF) analyses. Thus, the proposed
LS-DYNA 3,694.8 0
column shape factors simply account for the reduction in loading

498 / JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012

J. Bridge Eng., 2012, 17(3): 490-499


associated with potential blast threats. Most often, the exact charge Structural design for physical security: State of practice, ASCE,
weight, shape, size, and standoff distances are unknown, and only a Reston, VA.
general level of blast-resistance can be provided based on a reason- Dept. of Defense (DoD). (2002). Design and analysis of hardened struc-
able estimate of the loads. The proposed method provides an tures to conventional weapons effects, unified facilities criteria (UFC)
acceptably accurate estimate of the loads on a bridge column 3-340-01, United States Army Corps of Engineers, Washington, DC.
Dept. of Defense (DoD). (2008). Structures to resist the effects of acciden-
due to airblast and provides a significant advancement over the
tal explosions, UFC 3-340-02, United States Army Corps of Engineers,
loads computed by currently available methods.
Washington, DC.
Fujikura, S., Bruneau, M., and Lopez-Garcia, D. (2008). “Experimental
investigation of multihazard resistant bridge piers having concrete-filled
Conclusions
steel tube under blast loading.” J. Bridge Eng., 13(6), 586–594.
Downloaded from ascelibrary.org by "Indian Institute of Technology, Gandhinagar" on 08/07/23. Copyright ASCE. For personal use only; all rights reserved.

This paper presents research results obtained from computational Holland, C. E. (2008). “Blast-resistant design of highway bridge columns.”
Master’s thesis, Dept. of Civil, Architectural, and Environmental Engi-
fluid dynamics models of blast loads acting against circular and
neering at the Univ. of Texas at Austin, Austin, TX.
square bridge columns, although the findings apply in general to
Jenkins, B. M., and Gersten, L. N. (2001). “Protecting public surface
slender structural members directly exposed to airblast. The com- transportation against terrorism and serious crime.” MTI Report
putational models developed for this research were validated using 01-14, Mineta Transportation Institute, San Jose, CA.
data from experimental blast tests and results reported in the liter- Livermore Software Technology Corporation (LSTC). (2007). LS-DYNA
ature. The numerical results gathered from the computational keyword user’s manual, version 971, LSTC, Livermore, CA.
models showed the influence of column cross-sectional shape Luccioni, B., Ambrosini, D., and Danesi, R. (2006). “Blast load assessment
on net resultant impulse, and this information was used to develop using hydrocodes.” Eng. Struct., 28(12), 1736–1744.
equations for column shape factors that accurately account for the MacGregor, J. G. (1997). Reinforced concrete: Mechanics and design, 3rd
reduced impulses experienced by slender square and circular Ed., Prentice Hall, Upper Saddle River, NJ.
members relative to those acting on a wall subjected to the same Malvar, L. J., Crawford, J. E., and Morrill, K. B. (2000). “K&C concrete
threat scenario. These equations can be combined with an SDOF material model release III—automated generation of material model
analysis model, which provides a simple method for predicting input.” Karagozian and Case Report TR-99-24.3, Karagozian & Case
response, to ensure that bridge columns have adequate capacity Structural Engineers, Burbank, CA.
to resist blast loads. Ofengein, D. K., and Drikakais, D. (1997). “Simulation of blast wave
propagation over a cylinder.” Shock Waves, 7(5), 305–317.
Rickman, D. D., and Murrell, D. W. (2007). “Development of an improved
Acknowledgments methodology for predicting airblast pressure relief on a directly loaded
wall.” J. Pressure Vessel Technol., 129(1), 195–204.
The authors gratefully acknowledge the support of the National Schwer, L. E., and Malvar, L. J. (2005). “Simplified concrete modeling
Cooperative Highway Research Program for providing the finan- with *MAT_concrete_damage_REL3.” JRI LS−DYNA USER WEEK
cial resources that made this project possible. The authors also wish 2005, Karagozian & Case Structural Engineers, Burbank, CA.
Science Applications International Corporation (SAIC). (2001). BlastX
to thank the Southwest Research Institute and Protection Engineer-
version 4.2.3.0, Science Applications International Corporation. San
ing Consultants for their assistance in completing the experimental
Diego, CA.
tests. The opinions expressed in this paper are those of the authors Shi, Y., Hao, H., and Li, Z. (2007). “Numerical simulation of blast wave
and do not necessarily reflect those of the sponsor. interaction with structure columns.” Shock Waves, 17(1–2), 113–133.
Sućeska, M. (1999). “Evaluation of detonation energy from EXPLO5 com-
puter code results.” Propellants Explos. Pyrotech., 24(5), 280–285.
References U.S. Army Corps of Engineers (USACE). (2000). Bridge Explosive
Agrawal, A. K., and Yi, Z. (2009). “Blast load effects on highway bridges.” Loading (BEL), version 1.1.0.3, Engineer Research and Development
Univ. Transportation Research Center, City College of New York, NY. Center. Vicksburg, MS.
〈http://www.utrc2.org/research/assets/123/FR_Blast_bridge1.pdf〉 U.S. Army Corps of Engineers (USACE). (2001). CONWEP, Version
(Apr. 3, 2012). 2.0.6.0, U.S. Army Engineer Research and Development Center,
Alia, A., and Souli, M. (2006). “High explosive simulation using Vicksburg, MS.
multi-material formulations.” Appl. Therm. Eng., 26(10), 1032–1042. Williams, G. D. (2009). “Analysis and response mechanisms of
Biggs, J. M. (1964). Introduction to structural dynamics, McGraw-Hill, blast-resistant bridge columns.” Ph.D. dissertation, Dept. of Civil, Arch.
Inc, New York. and Env. Eng., Univ. of Texas at Austin, Austin, TX.
The Blue Ribbon Panel on Bridge and Tunnel Security (BRPBTS). (2003). Williamson, E. B. et al. (2010). “Blast-resistant highway bridges: Design
“Recommendations for bridge and tunnel security.” Special report and detailing guidelines.” National Cooperative Highway Research
prepared for FHWA and AASHTO, Washington, DC. Program (NCHRP) Report 645, Transportation Research Board of
Boyd, A., and Sullivan, J. P. (2000). “Emergency preparedness for transit the National Academies, Washington, DC.
terrorism.” TR News. No. 208. 12–17. 〈 http://onlinepubs.trb.org/ Winget, D. G., Marchand, K. A., and Williamson, E. B. (2005). “Analysis
onlinepubs/trnews/trnews208_transit_security.pdf〉 (Apr. 6, 2012). and design of critical bridges subjected to blast loads.” J. Struct. Eng.,
Conrath, E. J., Krauthammer, T., Marchand, K. A., and Mlakar, P. F. (1999). 131(8), 1243–1255.

JOURNAL OF BRIDGE ENGINEERING © ASCE / MAY/JUNE 2012 / 499

J. Bridge Eng., 2012, 17(3): 490-499

You might also like