Wang 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 48

Journal of Performance of Constructed Facilities.

Submitted October 22, 2013; accepted January 29, 2014;


posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

A comparative study on the dynamic response of concrete gravity dams


subjected to underwater and air explosions
Gaohui Wang1, Sherong Zhang2, Yuan Kong3, and Hongbi LI4
Abstract: The response of dam structures subjected to explosion shock loading is a key element in
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
assessments for the dam antiknock safety and anti-terrorism applications. While the physical processes

ip
during an explosive detonated in underwater/air and the subsequent response of structures are extremely

d cr
complex, which involve lots of complex issues such as the explosion, shock wave propagation, shock
wave-structure interaction and structural response. In addition, there exists a significant contrast in wave

te s
propagation phenomena in the water and the air medium due to their different physical properties and

di nu
interface phenomena. In this paper, a fully coupled numerical approach with combined Lagrangian and
Eulerian methods is used to simulate the dynamic responses of a concrete gravity dam subjected to
ye a
underwater and air explosions. The shock wave propagation characteristics from explosions in water and
op M
air are simulated and compared. The damage profiles of concrete gravity dams subjected to underwater
and air explosions are discussed. The influence of the blast loading from explosions in water and air on
the dynamic response and the damage of the dam is also investigated. It is seen from the analysis results
C ted

that a submerged explosion causes significantly more damage to the dam in water than the same mass of
explosive in air.
Keywords: concrete gravity dam; underwater explosion; air explosion; explosion shock loading; shock
ot p

wave propagation; coupled model


N ce

Introduction
Dams are important lifeline engineering which has contributed to the development of civilization for
Ac

a long time. In order to meet the ever increasing demand for power, irrigation and drinking water et al.,
the majority of high concrete dams is being built or to be built. Dam structures are possible targets of
war or terrorist attack due to their significant political and economic benefits. The possible failure of
dams retaining large quantities of water can cause a considerable amount of devastation in the
1
Research Associate, State Key Laboratory of Hydraulic Engineering Simulation and Safety, Tianjin University, Tianjin
300072, China (corresponding author). Email: wanggaohui@whu.edu.cn.
2
Professor, State Key Laboratory of Hydraulic Engineering Simulation and Safety, Tianjin University, Tianjin 300072,
China. Email: tjudam@126.com.
3
Research Associate, State Key Laboratory of Hydraulic Engineering Simulation and Safety, Tianjin University, Tianjin
300072, China. Email: kystory0715@163.com.
4
Research Associate, State Key Laboratory of Hydraulic Engineering Simulation and Safety, Tianjin University, Tianjin
300072, China. Email: haodoudemao@163.com


Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

downstream populated area during terrorist bombing attacks or accidental explosions. Therefore it is
very important to protect dam structures against blast loads, and the antiknock safety evaluation of dam
structures is an important topic in protective engineering.
The dynamic response of dam structures under blast loading is a crucial problem to evaluate the
antiknock safety of the dam. While, the physical processes during an explosion in underwater/air and
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
shock wave propagation are extremely complex, and the subsequent response of the dam subjected to

ip
explosion shock loading is much more complicated than that under other loadings such as static and

d cr
earthquake loadings, which involves lots of complex issues such as the explosion, shock wave
propagation, shock wave-structure interaction and structural response, etc. Hence a sophisticated

te s
numerical model for the loading and material responses would be required to enable more realistic

di nu
reproduction of the underlying physical processes. Many researchers have conducted comprehensive
experimental and numerical investigations related to the blast effects on building structures (Lu and
ye a
Wang 2006, Tian and Li 2008, Jayasooriya et al. 2011, Parisi and Augenti 2012), marine structures
op M
(Zhang et al. 2011a, Jin and Ding 2011, Zhang et al. 2011b), underground structures (Lu et al. 2005,
Wang et al. 2005, Ma et al. 2009, Li et al. 2012), bridge structures (Tang and Hao. 2010, Hao and Tang
C ted

2010), and plate structures (Rajendran and Lee 2009, Wang et al. 2013, Zakrisson et al. 2011, Spranghers
et al. 2013), etc. However, the corresponding studies of concrete gravity dams subjected to blast loads
are limited. This is probably because of the large size and the interactions with the reservoir and
ot p

foundation of dams which make both numerical modeling and experimental tests very costly. In addition,
N ce

the experimental tests require the use of relatively large amounts of charges, involving potential risks
and need careful handling, which is typically not feasible in civilian research (De 2012).
Ac

In terrorist attack scenarios, dam structures may undergo air blast loading due to the attack of rockets
and missiles above the waterline and underwater explosion loading due to explosion of mines, torpedoes
and depth charges below the waterline. There exists a significant contrast in the wave propagation
phenomena between the water and the air medium due to their different physical properties and interface
phenomena. It has long been known that an underwater explosion can cause significantly more damage
to the targets in water than the same amount of explosive in air. This is because the water is considered
as incompressible and the underwater detonation products are trapped inside an expanding gas bubble
that contains significant destructive energy. Hence, attention should be paid to the shock wave
propagation characteristics from explosions in water and air and the subsequent response of structures.
Rajendran and Lee (2009) brought out a detailed review of the phenomena of air and underwater

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

explosions and their effects on plane plates. Clutter and Stahl (2004) used the hydrocode to predicting
the shock loads on facility components for vulnerability assessments produced from HE charges in air
and water. Kwak et al. (2012) studied the fire-ball expansion and subsequent shock propagation from
explosives detonation in the water and the air medium. Librescu et al. (2006) explored the problem of
the dynamic response of geometrically non-linear sandwich flat panels including initial geometric
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
imperfection and subjected to underwater and in-air explosions.

ip
Most investigations on the dynamic behavior of structures subjected to blast loading have focused

d cr
mainly on experimental works and numerical damage predictions. In October 1940 the first in a series of
trials was carried out on a scale model of the Möhne dam to see whether a big conventional bomb could

te s
destroy the dam. Further trials involving a one-fiftieth scale model of the Möhne dam and a full-size dam

di nu
in mid-Wales proved that the Möhne could be breached if 6,500lb of high explosive could be detonated
against the inner wall of the dam (Falconer 2007). On the other hand, Yu (2009) used the ALE algorithm
ye a
to study the dynamic response of the dam through establishing the fully coupled model of underwater
op M
contact explosion. Linsbauer (2009, 2011), through the establishment of coupled model of reservoir water
and the dam, studied the dynamic response, stability and failure mechanism of concrete gravity dams
C ted

(with the initial cracks at upstream surface) under the impact of blast loading at the bottom of the
reservoir. Zhang et al. (2014) discussed the possible failure modes of concrete gravity dams subjected to
underwater explosion. The influence of the dam height, standoff distance and the upstream water level on
ot p

the antiknock performance of the dam is also investigated. To the best of the authors’ knowledge, the
N ce

specialized literature addressing the dynamic response of concrete gravity dams subjected to underwater
and air explosions is rather scanty. Studying the blast response of concrete gravity dams will help
Ac

understanding and improving their blast resistance.


This paper presents a numerical simulation study aiming to investigate the characteristics of shock
wave propagation and compare the dynamic behavior of concrete gravity dams subjected to underwater
and air explosions. A fully coupled numerical approach with combined Lagrangian and Eulerian methods,
is adopted to allow for the incorporation of the essential processes, namely the charge detonation, shock
wave propagation, shock wave-structure interaction and structural response. The material models are
carefully selected and validated. The shock wave propagation characteristics from free-field explosions in
water and air are simulated, and the results have been compared with the published empirical formula.
The shock wave pressure propagation through a subjected medium is examined. A generic concrete
gravity dam is employed as a numerical application to illustrate the characteristics of structural effects in

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a typical underwater and air explosion scenario. The damage modes and dynamic response of the dam for
different blast loadings are also investigated.
Description of the coupled method and material models
Overview of fully coupled Lagrangian-Eulerian numerical approach
In order to incorporate the various physical processes into a single model system, many complex
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
phenomena should be considered such as large deformation at the vicinity of the charge, the

ip
fluid-structure interaction, etc. and more advanced numerical schemes beyond the capacity of commonly

d cr
used structural analysis programs are necessary. In typical hydrocodes there are two primary types of
computational physics approaches, Eulerian and Lagrangian, and each has its advantages and limitations

te s
in modelling the effects of blast loading. In the Lagrangian method, Lagrangian finite element solutions

di nu
provide the advantage of having a structural mesh that moves and deforms with the physical material,
allowing for the calculation of material and structural behaviour over large time scales. However, the
ye a
disadvantage of the Lagrange method is that the numerical grid can become severely distorted or tangled
op M
in an extremely deformed region, which can lead to adverse effects on the integration time step and
accuracy. In the Eulerian method, the numerical mesh is fixed in space and the physical material flows
C ted

through the mesh. As the mesh is fixed, there is no mesh distortion problem when large deformations or
flow occur. Eulerian solutions, however, are often limited to relatively short simulation times due to the
accumulation of advection and interface tracking errors.
ot p

A coupled method (Benson 1992, Van der Veen 2003) which combines the advantages of both
N ce

Lagrangian and Eulerian methods will facilitate the modelling of some problems involving the
fluid-structure interaction and large deformation. In this method, Eulerian meshes can be used for analysis
Ac

involving large deformation without remeshing, and Lagrangian meshes can be used for solid materials to
analysis the structural behaviour. This method has been made possible in various hydrocodes, e.g.,
LS-DYNA (2001), AUTODYN (ANSYS 2010), among others. In the present study, the coupled
Lagrangian-Eulerian method is adopted. In explicit hydrocodes such as AUTODYN(ANSYS 2010) used
in the present study, this coupling is achieved by a strategy such that, a Lagrangian interface can “cut”
through the fixed Eulerian mesh in an arbitrary manner, and Eulerian material can exert forces (pressure
boundary conditions) on the Lagrangian element causing displacement of the structure. In return, the
Lagrangian interface provides a geometric constraints (velocity boundary conditions) to Eulerian material
flow and Eulerian material cannot penetrate the Lagrangian element. Fig. 1 shows schematically a typical
coupled situation.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Concrete material models


The dynamic behaviour of the concrete material under blast loading is a complex nonlinear and
rate-dependent process. In the present study, the RHT (Riedel et al. 1999, Riedel 2000) dynamic damage
model, developed by Riedel, Hiermaier and Thoma, is used for modeling of the concrete material. This
model which can reflect the characteristics of the concrete material behavior at a high strain rate is
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
particularly useful for modeling the dynamic behavior of brittle materials such as concrete and rock under

ip
dynamic loading. The RHT model consists of three pressure-dependent surfaces in the stress space which

d cr
are different limit states for the yield surface; the elastic limit, the failure, and the residual strength, as
shown in Fig. 2.

te s
Failure surface

di nu
The failure surface, Yfail, is defined as a function of the normalized pressure p*, the lode angle θ and
strain rate İ :
ye a
( ) ( )
Yfail p ∗ , ș , İ = Yc p ∗ × r3 (ș ) × Frate (İ )
op M
(1)
( ) «¬
(
Yc p ∗ = f c × ª A × p ∗ − pspall

× Frate (İ )
»¼
)

(2)
C ted

where fc is the material uniaxial compressive strength; A is the failure surface constant; N is the failure

surface exponent; p* is the pressure normalized by fc; pspall



= f t f c , where ft is the material uniaxial

tensile strength; Frate (İ ) represents the dynamic increase factor (DIF) as a function of the strain rate; r3(θ)
ot p
N ce

defines the third invariant dependency of the model as a function of the second and third stress invariants

and a meridian ratio ȥ .


Ac

r3 (ș ) = =
( 2
) 2
( 2 2
)
r 2 1 − ȥ cos ș + (2ȥ − 1) 4 1 − ȥ cos ș + 5ȥ − 4ȥ
rc ( )
4 1 − ȥ 2 cos 2 ș + (1 − 2ȥ )2 (3)

3 3 J3
cos 3ș =
2 (J 2 ) 3 2 (4)
rt
ȥ= = Q + BQp ∗
rc (5)
where J2 and J3 are the second and the third invariants of the devaitoric stress tensor; Q is the ratio of
strength at zero pressure and the coefficient BQ is the rate at which the fracture surface transitions from
an approximately triangular in form to a circular form with increasing pressure. Fig. 3 shows the
intersections of the failure surface with different deviatoric planes.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Strain rate effects are represented through increases in fracture strength with plastic strain rate. Two
different terms can be used for compression and tension with linear interpolation being used in the
intermediate pressure regime.
­ § İ ·
Į
°1 + ¨¨ ¸¸ for p > 1 f c (compressio n )
° © İ0 3
¹
Frate =® (6)
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
į
° § İ ·
°1 + ¨¨ ¸¸ for p < 1 f t (tension )

ip
3
¯ © İ0 ¹
where į is the compression strain rate factor and Į is the tension strain rate factor.

d cr
Elastic limit surface and strain hardening

te s
The material strain hardening behavior is represented in the RHT model through the definition of an

di nu
elastic limit surface and a “hardening” slope. The elastic limit surface is scaled down from the fracture
surface.
Yelastic = Yfail × Felastic × Fcap ( p )
ye a
(7)
op M
where Felastic is the ratio of the elastic strength to failure surface strength along a radial path. Fcap(p) is a
function that limits the elastic deviatoric stresses under hydrostatic compression, varying within the range
C ted

of (0,1) for pressure between the initial compaction pressure and the solid compaction pressures.
Considering the fact that concrete initiates the inelastic behaviour under compression at around 30% and
under tension at 50-80% of the respective maximum strength, Felastic in Eq. (7) is varied linearly with
ot p

pressure between the values associated with uniaxial tension and uniaxial compression.
N ce

Linear hardening is used prior to the peak load. During hardening, the current yield surface (Y*) is
scaled between the elastic limit surface and the failure surface via
Ac

İ pl
Y ∗ = Yelastic + (Yfail − Yelastic )
İ pl(pre -softening ) (8)

where İpl and İpl(pre-softening) are the current and pre-softening plastic strain, respectively. The relationship of
İpl and İpl(pre-softening) are shown in Fig. 4.
Residual failure surface
In order to describe the strength of the completely crushed material, a residual (frictional) failure
surface is defined as

Yresidual = B × p∗M
(9)
where B is the residual failure surface constant, M is the residual failure surface exponent.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Damage evolution
When hardening states reach the ultimate strength of the concrete on the failure surface Yfailure
damage is accumulated during further inelastic loading controlled by plastic strain. The model by
Holmquist and Johnson (1993) is taken as a basis for the evolution law. Damage is assumed to
accumulate due to inelastic deviatoric straining (shear induced cracking) using the relationships
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
ǻİ pl

ip
D=¦
İpfailure (10)

( )
D2

d cr
İpfailure = D1 p ∗ − pspall

≥ İfmin
(11)
where D1 and D2 are damage constants used to describe the effect strain to fracture as a function of

te s
pressure, is the minimum strain to reach failure.

di nu
İ fmin

The current fracture surface (for a given level of damage) is scaled down from the intact surface
using the expression
ye a

Yfracture = (1 − D ) × Yfailure
∗ ∗
+ D × Yresidual
op M
(12)
The current shear modulus is defined through the expression
Gfracture = (1 − D ) × Ginitial + D × Gresidual
C ted

(13)
where Gfracture, Ginitial, Gresidual are the shear moduli.
The material constants (ANSYS 2010, Tu and Lu 2009) adopted in the present work are based on the
ot p

typical data for concrete material, as shown in Table 1.


N ce

Rock material models


The Johnson-Holmquist (JH-2) constitutive model, developed by Johnson and Holmquist (1999), is
Ac

used for modeling of the rock material. This model was originally formulated for description of the brittle
response of ceramics (Johnson and Holmquist 1999, Holmquist and Johnson 2002, Johnson and
Holmquist 2003, Holmquist and Johnson 2005). A summary of these constants for granite is listed in
Table 2.
Material model for explosive, water and air
The explosive charges are typically modeled using standard Jones, Wilkins, and Lee (JWL) equation
of state. The values of the constants for many common explosives have been determined from dynamic
experiments and are available in AUTODYN (ANSYS 2010). In the present simulation, for a TNT
explosive charge, C1, C2, R1, R2, and Ȧ are 3.7377e11Pa, 3.747e9Pa, 4.15, 0.9, and 0.35, respectively.
The water included in the analyses is modeled using the Two Phase expansion EOS with a

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Polynomial EOS. In the simulation, the standard constants of air from the AUTODYN material library
(ANSYS 2010) are used. The material parameters are: initial density, ȡ0=999kg/m3; the constants, A1, A2,
A3, B0, B1 and T1 are 2.2e9Pa, 9.54e9Pa, 1.457e10Pa, 0.28, 0.28, 2.2e9Pa, and 0, respectively.
Air is modeled by the ideal gas equation of state, which is defined by the ideal-gas gamma-law
relation as:
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
ȡ
p = (Ȗ − 1) E

ip
ȡ0 (14)
3
where E is the specific energy (253.4kJ/m ), Ȗ is the constant-pressure to constant-volume specific heat

d cr
ratio (=1. 4 for a diatomic gas like air), ȡ0 is the initial air density (with a reference mass density of
1.225×10-3 g/cm3), and ȡ is the current density.

te s
di nu
Shock wave propagation characteristics from explosions in water and air
The shock wave propagation from explosions in water and air
In underwater explosion, the peak overpressure is of several orders of magnitude greater than the
ye a
hydrostatic pressure. Therefore, the hydrostatic pressure is ignored and the peak overpressure is simply
op M
called as peak pressure. The pressure history of the shock wave at a given point starts with an
instantaneous pressure increase to a peak pressure, followed by a decaying exponential function, which is
C ted

given by (Cole 1948, Bjørnø and Levin 1976)

P(t ) = Pme − t ș (15)


ot p

where Pm is the peak pressure of the shock wave; ș is the exponential decay time constant; t is the time
N ce

since the shock wave front arrived at the target point. For trinitrotoluene (TNT) or Pentolite, the peak
pressure Pm and the time decay constant ș can be expressed as follows:
Į1
Ac

§W1 3 ·
Pm = k1 ⋅ ¨¨ ¸
¸ (16)
© R ¹
Į2
§ W1 3 ·
ș = k2 ⋅ W 13
⋅¨ ¸ (17)
¨ R ¸
© ¹

where Pm is in MPa, ș is in microseconds, W is the weight of the explosive charge in kilograms, R is the
standoff distance in meters, k1, Į1, k2 and Į2 are constants which depend on explosive charge type the
shock parameters of the explosive.
The effectiveness of the shock wave depends on the time integral of the pressure, or impulse, which
is more significant than on the detailed form of pressure versus time. The free field impulse per unit area,
I, in N·s/m2 is given as

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

9
ȕ
§W1 3 ·
I = l ⋅W 13
⋅ ¨¨ ¸ (18)
¸
© R ¹
where, l and ȕ are the shock parameters of the explosive. For TNT, the shock parameters are as follows:
k1=52.16MPa, Į1=1.13ˈk2=96.5×10-6, Į2=-0.22, l=5760, ȕ=0.89.
The typical pressure-time history of the positive phase of an ideal air blast wave is given by the
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
Friedlander equation as (Kinnery and Graham 1985)

ip
ª § t · -Įt º
p (t ) = Po + Pm «1 − ¨¨ ¸¸e td
»
«¬ © t d ¹ »¼ (19)

d cr
where Po is the reference ambient pressure, Pm is the peak overpressure of the shock wave, t is the
instantaneous time, td is the positive phase duration, and Į is called the waveform parameter that depends

te s
upon the Pm. The overpressure-distance relation for chemical explosions can be written as

di nu
ª § R′ · 2 º
808«1 + ¨ ¸ »
Pm «¬ © 4.5 ¹ »¼
=
Po 2 2 2 (20)
ye a § R′ ·
1+ ¨ ¸
§ R′ ·
1+ ¨ ¸
§ R′ ·
1+ ¨ ¸
© 0.048 ¹ © 0.32 ¹ © 1.35 ¹
op M
R
R′ =
W13 (21)
where R′ is the scaled distance, R is the standoff distance from the explosion in m, and W is the TNT
C ted

equivalent of the explosive charge weight in kg.


The impulse of the shock wave I in kN·s/m2 is given as
4
§ R′ ·
0.067 1 + ¨ ¸
ot p

© 0.23 ¹
I=
§ R′ ·
3 (22)
R′ 2 3 1 + ¨
N ce

¸
© 1.55 ¹

Comparative analysis of shock wave propagation between underwater explosion and air explosion
Ac

To verify the effectiveness of the numerical model and investigate the shock wave propagation
characteristics, due to lack of experimental data at present, a numerical model of free-field explosions in
water and air is established, and the results have been compared with the published empirical formulas
(Cole 1948, Kinnery and Graham 1985). In the numerical simulation, the initial detonation and blast wave
propagation are modeled by 1/8th of cube with three symmetric planes. The three-dimensional model
with an element size of 100mm is shown in Fig. 5. Exposed area of the numerical model is 10×10×10m3.
The weight of the charge is 1000kg TNT. The water, air and high explosive are modeled by the Euler
subgrid, in which the grid is fixed and material flows through it. In order to simulate a free field
underwater explosion, the outflow boundary condition is applied at all the artificial boundaries to
minimize the stress wave reflection at these computational boundaries.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

10

Fig. 6 shows typical pressure time histories at 5m from charge center in water and air. In air
explosion, the shock wave is characterized by a sudden pressure rise to the peak value at the shock front,
and followed by an exponential decrease back to ambient value and then a negative phase. The negative
phase of the blast wave is generally ignored. In underwater explosion, the shock wave also has an
instantaneous rise and an exponential fall, which is similar to the air blast at the positive phase. The shock
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
wave peak pressure from underwater explosion is 39.62 times that from air blast.

ip
In Fig. 7, the numerical peak pressures and impulses from air blast are compared with that from

d cr
underwater explosion. It can be seen from Fig. 7 that both the peak pressures and the impulses from
underwater explosion are significantly higher than that from air blast. More rapid decrease of pressure at

te s
the center of explosion can be expected in air because the shock radius increased more rapidly in air. The

di nu
attenuation of the shock waves from underwater explosion and air blast can be also clearly seen in this
figure in which the explosive wave pressure is high in the vicinity of the charge and they decrease with
ye a
the increase of the distance from the charge center. The shock wave peak pressures and impulses from
op M
underwater explosion are 42.35 and 71.99 times that from air blast on average, respectively.
In order to examine the mesh sensitivity of the model and calibrate the accuracy of the numerical
C ted

results, two meshes with different element sizes, namely, 100mm and 200mm for the 3D model are
considered in the present study. Fig. 8 shows the peak pressure vs. distance plots from the present
numerical analysis and those from empirical formulas (Cole 1948, Kinnery and Graham 1985). Compare
ot p

the numerical result with the empirical result at the same concentration of different measured points, it is
N ce

not so hard to be seen from Fig. 8 that there are some difference between simulation results and
empirical results. In addition to the vicinity of the charge, the empirical peak pressures from explosions
Ac

in water and air are marginally higher than the present numerical results. This may be caused owing to
insufficiently small mesh size used in the simulation. Further reduction of the element size will result in
a better estimation of peak pressures, but will substantially increase the computational time and will
cause computer memory overflows. Fig. 9 shows the comparison of the present numerical results and the
prediction from empirical formulas (Cole 1948, Kinnery and Graham 1985) for the impulse from
explosions in water and air. The comparison of impulses presented in Fig. 9 shows that the impulse
agrees well with the empirical results except the vicinity of the charge. These observations indicated that
the numerical model fails to capture the very sharp peak of the blast wave. However on the whole, the
prediction of the peak magnitude by the present method shown in the graph represents a good
approximation for this problem, and the impulse agrees well with the empirical results. Further when the

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

11

actual distance is far, the calculation efficiency can be improved through appropriately increasing the
mesh size without reducing the accuracy of the results.
Dynamic response of concrete gravity dams under blast loadings
Numerical model
A typical non-overflow monolith of concrete gravity dams, which is 75m high with a 70m deep
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
reservoir, is chosen for the analysis to characterize the structural effects from explosions in water and air.

ip
In order to compare the dynamic response of concrete gravity dams subjected to underwater and air

d cr
explosions, the conditions of full reservoir and empty reservoir are assumed. It should be noted that it is
not realistic scenario set for an air blast loading acting on the whole front surface of the dam in this study.

te s
The maximum reservoir water level of 70.0m is considered. The width of the non-overflow monolith of

di nu
the dam is 15m. Considering the symmetries of the geometrical model as shown in Fig.10 (a), to save
computation time, a half of the dam, rock mass, water, air, and explosive are modeled as shown in Fig.10
ye a
(b) which consists of 3,022,575 elements. In the central part of the charge, the element size is 100mm for
op M
the high explosive and water. The mesh size increases gradually away from the charge center. In the upper
part of the dam, the element size is 200mm. The standoff distance and detonation depth are all assumed to
C ted

be 10m. A cubical explosive charge of 1000kg of TNT is used for this calculation.
In numerical simulation, a fully coupled approach combining the Lagrangian and Eulerian methods
is adopted to allow for the incorporation of the essential processes. The solid concrete and rock are
ot p

modelled using a Lagrange subgrid, in which the coordinates move with the material, while the explosive
N ce

charge, the air and the water are modeled using an Euler subgrid, in which the grid is fixed and material
flows through it. At the Euler-Lagrange interface, interaction is considered. The bottom of the mesh
Ac

which represents the bedrock is considered as fixed in all directions. Along the truncated boundary for the
entire computational domain, the boundary condition of the Euler subgrid is set as an outflow boundary,
the symmetric boundaries are applied on the symmetric plane, and the transmission boundary conditions
are applied to allow for free passage of shock/stress waves.
The mesh sensitivity of the fully couple model for concrete gravity dams
In order to examine the mesh sensitivity of the fully couple model for concrete gravity dams
subjected to blast loads, two meshes with different densities, namely Mesh I and Mesh II for the couple
model are utilized in this study. For Mesh I, the element size is 100mm for the high explosive and water
at the central part of the charge. Then the mesh size increases gradually away from the charge center. The
element size is 200mm for the upper part of the dam. It should be noted that the Mesh I is the same as in

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

12

Fig. 10, and the size of the elements in the dam of the Mesh II is half of the size of the correspondent
elements of the Mesh I in each direction.The charge weight, exposed area, and boundary conditions are
assumed as aforementioned. A fully coupled approach combining the Lagrangian and Eulerian methods is
adopted to consider the shock wave-structure interaction. Because the peak pressures and the impulses
from underwater explosion are significantly higher than that from air blast, only the damage analysis of
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
concrete gravity dams subjected to underwater explosion is considered for the mesh sensitivity

ip
investigation.

d cr
The damage profiles of concrete gravity dams subjected to underwater explosion with two mesh size
cases using the fully coupled numerical model is shown in Fig. 11 and Fig. 12. As observed by comparing

te s
Fig.11 and Fig. 12, the damage profiles of the dam under blast loads is very similar to each other. The

di nu
small mesh sensitivity to the damage growth path is observed. As seen from Fig. 11(b) and Fig. 12(b), the
final damage profiles for the two meshes are similar in orientation and localization on the whole. Fig. 13
ye a
shows the computed velocity time histories at the dam crest. As can be seen, the results are very similar
op M
between the Mesh I and Mesh II models. Marked difference is observed in in subsequent dynamic
response, where the MeshII results exhibit larger velocity response than from the Mesh I model. These
C ted

point out that it is feasible to adopt the couple model with the aforementioned mesh to analyze the
dynamic response and damage profile of concrete gravity dams subjected to blast loads qualitatively.
Comparative analysis of the dam subjected to underwater and air explosions
ot p

Subjecting to the explosion load, the model of the concrete gravity dam is analyzed numerically. The
N ce

shock wave loads from explosions in water and air act on the dam structure and generate dynamic
response. It should be noted that underwater explosion will damage the dam structure by the shock wave
Ac

and the bubble pulsation pressure. This study, however, concentrates on the comparative study on the
structure responses and damage to direct blast shock wave from explosions in water and air. Further
studies of the combined effect of hydrostatic pressure, underwater blast shock wave and bubble pulsation
pressure on dam structure responses and damage are deemed necessary.
Immediately after detonation of the charges, spherical shock waves radiate out from the charge
locations. Through the simulated analysis for the whole process of the dam subjected to underwater and
air explosions, the explosion, shock wave propagation, shock wave-structure interaction and structural
response the dam are obtained. The contours of the pressure wave propagation from explosions in water
and air at several time instants are shown in Fig. 14 and Fig. 15, respectively. These graphs show clearly
the propagation of the shock wave. The incident and reflected waves can also be clearly observed. It can

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

13

be seen from Fig. 14 and Fig. 15 that there is some difference between the pressure distribution for
explosions in water and air, and pressures from underwater explosion are significantly higher than that
from air blast.
Underwater explosion, a bulk cavitation phenomenon occurs in water caused by the reflection of a
shock wave at a free surface. As shown in Fig. 14(b) (at t=5ms), the initial shock generated by the
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
underwater explosion has impacted on the free surface and the upstream surface of the dam structure.

ip
When the incident shock wave impacts on the free surface, the compression shock wave reflecting from

d cr
the free surface results in a tensile reflected wave (Prandtl-Meyer rarefaction wave), and a transmitted
shock is propagating into the air medium. Due to the much lower acoustic impedance of the air medium,

te s
the transmitted shock is much weaker than the reflected rarefaction waves. Since the water cannot sustain

di nu
a significant amount of tension, the rarefaction waves which propagate in the opposite directions along
the water-air interface will cause the water pressure just below the free surface to drop very rapidly, and a
ye a
cavitation “surface cutoff” effect occurs just below the free surface as shown in Fig.14(c) at t= 7ms. The
op M
pressures of the shock and reflected wave cancel each other out reducing the pressure at that point to a
negative pressure causing cavitation which reflects the influence of the free surface on the propagation of
C ted

the blast wave. When the incident shock wave impacts on the dam upstream face, the shock-structure
interaction will firstly result in a reflected shock wave which causes the water pressure to rise near the
vicinity of the upstream face of the dam facing the explosion center. It should be noted that a rarefaction
ot p

wave will also generate when a reflected shock resulted from the shock-structure interaction interacts
N ce

with the gas bubble. This rarefaction wave will next interact with water-structure interface and create a
cavitation near the structural surface.
Ac

In order to study the propagation of the shock wave in water and air towards to the dam structure, as
well as the distribution of the pressure loading (reflected pressure) on the dam structure, a number of
target points are arranged in the numerical model to record the pressure histories, as shown in Fig. 16.
The pressure histories from underwater explosion and the overpressure histories from air blast at
selected targets along the straight path from the charge center to the dam structure are shown in Figs. 17
and 18. As can be seen from Figs. 17 and 18, both in underwater explosion and air blast, the attenuation
of the shock wave peak pressure is rapid (exponential) with the increase of the distance from the charge
center. But the peak overpressure from air blast decreases more rapidly than from underwater explosion.
Due to the impedance of the water and air being lower than that of the concrete structure, there will be a
reflected shock wave as the original shock wave propagates through the interface of them. This reflected

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

14

shock wave will cause the pressure to be higher than that of the free field explosion. So there are two
peaks of the curve at some selected points, the first peak is caused by the direct wave of the explosion,
and the second peak is caused by the shock wave reflected. When the target points are closer to the dam
structure, the reflected wave becomes more and more significant, the peak reflected pressure/overpressure
appears to be higher than the incident wave pressure/overpressure.
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
Fig. 19 shows the blast pressure/overpressure on the front face of the dam structure. As expected, the

ip
peak overpressure from air blast tends to decrease gradually at lower positions on the front face of the

d cr
dam. But unlike the air blast, the peak pressure from underwater explosion does not exhibit very sensible
attenuation among different target points due to the reflected wave. The arrival time of the shock front do

te s
not exhibit sensible difference among underwater explosion and air blast for Target 1 and Target 2. As the

di nu
height decreases, the arrival time of the shock front for air blast delays more than underwater explosion.
This is because the shock wave propagates with rapid expansion of the products in the vicinity of the
ye a
charge. The shock wave travels radially from the burst point with diminishing velocity. With the increase
op M
of the distance from the charge center, the shock wave gradually propagates at the sound velocity of the
surrounding medium. The velocity of sound in air is 340m/s, and the velocity of sound in water is
C ted

1483m/s (approximately 4.36 times the velocity of sound in air).


Figs. 20 and 21 show the comparison of recorded pressure time histories at two different target
points for the underwater explosion and air blast. As can be seen from Figs. 20 and 21 that when the
ot p

shock wave arrives at the downstream face, the compression shock wave reflecting from the downstream
N ce

surface will result in a tensile reflected wave, and the reflected wave becomes more and more significant.
At the downstream face (Target 4), the peak reflected pressure appears to be much higher than the
Ac

incident compression wave pressure. It can also be found from Figs. 20 and 21 that the shock wave peak
pressure from underwater explosion is higher than that from air blast with the same mass of charge. The
maximum peak pressure of the compression shock wave in the dam for air blast is 1.52MPa, and for
underwater explosion is 34.86MPa (approximately 30.02 times the peak pressure of the compression
shock wave in the dam for air blast). The maximum peak pressure of the reflected shock wave in the dam
for air blast is 0.45MPa, and for underwater explosion is 5.52MPa (approximately 12.22 times the peak
pressure of the reflected shock wave in the dam for air blast).
Fig. 22 displays the velocity time response curve of the dam crest between underwater explosion and
air blast, and Fig. 23 shows the corresponding acceleration time curve. It can be seen from Figs. 22 and
23 that the response velocity and the corresponding acceleration of the dam from underwater explosion is

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

15

significantly higher than that from air blast with the same mass of charge. In air blast, the maximum
velocity and acceleration occur at the dam crest (represented by Target 17), which are 0.45m/s and
474.02m/s2, respectively. In underwater explosion, the highest velocity of about 4.21 m/s occurs at the
upstream face of the dam near the explosion center (represented by Target 14), followed by the dam crest
about 1.79 m/s (represented by Target 17, approximately 3.98 times the highest velocity of the dam crest
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
for air blast).

ip
Damage analyses of the dam subjected to underwater and air explosions

d cr
The accumulated damage process of the dam for 1000kg of TNT which exploded at a standoff
distance of 10m is shown in Fig. 24, and Fig. 25 illustrates the damage contours of the dam subjected to

te s
air blast at the same detonation location as that in underwater explosion. It should be noted that the

di nu
contour value between 0 and 1 indicates the concrete element damage states in the range from undamaged
material to fully damaged material.
ye a
As shown in the damage process in Fig. 24, the blast pressure generated from underwater explosion
op M
causes a highly localized damage to the upper zone of the dam. It can be found that the damage is initially
observed to occur mainly on the upstream face of the dam near the explosion center due apparently to the
C ted

strike of direct shock wave, and the vicinity of the free water surface due to the effect of the bulk
cavitation. The crack damage in the upstream face of the dam near the free water surface extends deeper
inside of the dam and propagates horizontally toward the downstream face. Therefore the bulk cavitation
ot p

effect on the dam in the vicinity of the free water surface has to be taken into consideration in the overall
N ce

evaluation of the effect of an underwater explosion. Moreover, owing to the stress wave transmitted from
the upstream face, the downstream face of the dam also suffers intensive damage although the blast
Ac

pressures acting on the downstream face are smaller. This is because of the lower tensile strength of
concrete material than its compressive strength. When the shock wave arrives at the downstream face, the
compression shock wave reflecting from the downstream surface will result in a tensile reflected wave,
and the reflected wave becomes more and more significant. The remaining part of the dam structure
exhibit more or less a uniform distribution of damage among different elements. It can also be found that
the tensile damage appears at the dam heel through the global dynamic system.
It can be seen from Fig. 25 that the concrete gravity dam is not as significantly damaged because the
pressures acting on the dam structure from air blast are substantially smaller as compared to underwater
explosion. Only some localized damage is observed at the upstream and downstream faces due to the
strike of direct air shock wave and reflected shock wave.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

16

By comparing Fig. 24 (underwater explosion) with Fig. 25 (air blast), it can be noted that the damage
area of the dam subjected to air blast is smaller in comparison to that subjected to underwater explosion.
An underwater explosion can cause significantly more damage to the dam in water than the same amount
of explosive in air due to their different physical properties and the interface phenomena between the
explosive product gases and the surrounding medium. Hence, attention should be paid to the dynamic
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
response and damage process of concrete gravity dams subjected to underwater explosion. As usual, the

ip
modern war is predictable. This means that lowering the water level in the reservoir before the

d cr
foreseeable war can effectively provide the better antiknock performance for the dam, which is an
effective defense measure to reduce the risk of the dam failure.

te s
Conclusions

di nu
There exists a significant contrast in the wave propagation characteristics between the water and the
air medium. In underwater explosion, the shock wave has an instantaneous rise and an exponential fall,
ye a
which is similar to the air blast at the positive phase. The shock wave peak pressures and impulses from
op M
underwater explosion are significantly higher than that from air blast, which are 42.35 and 71.99 times
that from air blast on average, respectively. More rapid decrease of pressure at the center of explosion can
C ted

be expected in air because the shock radius increased more rapidly in air. There are some difference
between simulation results and empirical results. However on the whole, the prediction of the peak
magnitude by the present method shown in the graph represents a good approximation for this problem
ot p

and the impulse agrees well with the empirical results.


N ce

Underwater explosion, a bulk cavitation phenomenon occurs in water caused by the reflection of a
shock wave at a free surface. The damage to the dam structure is initially observed to occur mainly on the
Ac

upstream face of the dam due apparently to the strike of direct shock wave, and the vicinity of the free
water surface due to the effect of the bulk cavitation. The crack damage in the upstream face of the dam
near the free water surface extends deeper inside of the dam. The dynamic evolution of the bulk
cavitation has significant influence on the structures in the vicinity of the free water surface. Moreover,
owing to the stress wave transmitted from the upstream face, the downstream face of the dam also suffers
tensile damage.
By comparing analyses of concrete gravity dams subjected to underwater explosion and air blast, it
can be found that the response velocity and the corresponding acceleration of the dam from underwater
explosion is significantly higher than that from air blast with the same mass of charge. An underwater
explosion can cause significantly more damage to the dam than the same amount of explosive in air due

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

17

to their different physical properties and interface phenomena. Hence, attention should be paid to the
shock wave propagation characteristics from underwater explosion and the subsequent response of dam
structures.
Acknowledgements
The authors gratefully appreciate the supports from the State Key Laboratory of Hydraulic
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
Engineering Simulation and Safety (Tianjin University), the Foundation for Innovative Research Groups

ip
of the National Natural Science Foundation of China (No. 51021004), the National Natural Science

d cr
Foundation of China (No. 51379141), and Tianjin Research Program of Application Foundation and
Advanced Technology (No. 13JCYBJC19400).

te s
References

di nu
ANSYS Inc. (2010) AUTODYN user manual version 13.

Bjørnø, L., and Levin, P. (1976). “Underwater explosion research using small amount of chemical explosives.” Ultrasonics, 14(6), 263–267.
ye a
Benson, D. J. (1992). “Computational methods in Lagrangian and Eulerian hydrocodes.” Computer methods in Applied mechanics and
op M
Engineering, 99(2), 235-394.

Cole, R. H. (1948). “Underwater explosions.” New York: Dover Publications Inc.

Cluttera, J. K., and Stahlb, M. (2004). “Hydrocode simulations of air and water shocks for facility vulnerability assessments.” Journal of
C ted

Hazardous Materials, 106(1), 9–24.

De, A. (2012). “Numerical simulation of surface explosions over dry, cohesionless soil.” Computers and Geotechnics, 43,72–79.
ot p

Falconer, J. (2007). “The Dam Busters Story.” United Kingdom: Sutton Publishing Ltd.
N ce

Holmquist, T. J., and Johnson, G. R. (1993). “A computational constitutive model for concrete subjected to large strains, high strain rates,

and high pressures.” In: Fourteenth International Symposium on Ballistics, Québec.


Ac

Holmquist, T. J., and Johnson, G. R. (2002). Response of silicon carbide to high velocity impact. Journal of Applied Physics, 91, 5858–5866.

Holmquist, T. J., and Johnson, G. R. (2005). “Characterization and evaluation of silicon carbide for high-velocity impact.” Journal of

Applied Physics, 97, 93502-1–93502-12.

Hao, H., and Tang, E. K. C. (2010). “Numerical simulation of a cable-stayed bridge response to blast loads, Part II: Damage prediction and

FRP strengthening.” Engineering Structures, 32(10), 3193–3205.

Johnson, G. R., and Holmquist, T. J. (1999). “Response of boron carbide subjected to large strains, high strain rates, and high pressures.”

Journal of Applied Physics, 85, 8060–8073.

Johnson, G. R., Holmquist, T. J., and Beissel, S. R. (2003). “Response of aluminum nitride (including a phase change) to large strains, high

strain rates, and high pressures.” Journal of Applied Physics, 94, 1639–1646.

Jin, Q., and Ding, G. (2011). “A finite element analysis of ship sections subjected to underwater explosion.” International Journal of Impact

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

18

Engineering, 38(7), 558–566.

Jayasooriya R., Thambiratnam D. P., Perera N. J., Kosse V. (2011). “Blast and residual capacity analysis of reinforced concrete framed

buildings.” Engineering Structures, 33(12): 3438–3492.

Kinnery, G. F., and Graham, K. J. (1985). “Explosive shocks in air.” Berlin: Springer Verlag.

Kwak, H. Y., Kang, K. M., Ko, I., and Kang, J. H. (2012). “Fire-ball expansion and subsequent shock wave propagation from explosives
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
detonation.” International Journal of Thermal Sciences, 59, 9–16.

ip
LS-DYNA. (2001). Livermore Software Technology Corporation, LS-DYNA user’s manual, version 960, vols. 1–2, Livermore, CA, USA.

d cr
Lu, Y., Wang, Z., and Chong, K. (2005). “A comparative study of buried structure in soil subjected to blast load using 2D and 3D numerical

simulations.” Soil Dynamics and Earthquake Engineering, 25(4), 275–288.

te s
Librescu, L., Oh, S. Y., and Hohe, J. (2006). “Dynamic response of anisotropic sandwich flat panels to underwater and in-air explosions.”

di nu
International Journal of Solids and Structures, 43(13), 3794–3816.

Lu, Y., and Wang, Z. (2006) “Characterization of structural effects from above-ground explosion using coupled numerical simulation.”
ye a
Computers & Structures, 84(28), 1729–1742.
op M
Linsbauer, H. N. (2009). “Damage potential of an upstream-side crack in a gravity dam subjected to an impact loading in the reservoir.”

Proceedings of the 2nd International Conference on Long Term Behavior of Dams (LTBD09), Graz, 817–822.
C ted

Linsbauer, H. N. (2011). Hazard potential of zones of weakness in gravity dams under impact loading conditions. FRONTIERS OF

ARCHITECTURE AND CIVIL ENGINEERING, 5(1), 90–97.

Li, J. C., Li, H. B., Ma, G. W., and Zhou, Y. X. (2012). “Assessment of underground tunnel stability to adjacent tunnel explosion.” Tunnelling
ot p

and Underground Space Technology, 35, 227–234.


N ce

Ma, G. W., Huang, X., and Li, J. C. (2009). “Simplified damage assessment method for buried structures against external blast load.” Journal

of structural engineering, 136(5), 603–612.


Ac

Parisi, F., and Augenti, N. (2012). “Influence of seismic design criteria on blast resistance of RC framed buildings: A case study.”

Engineering Structures, 44, 78–93.

Riedel, W., Thoma, K., and Hiermaier, S. (1999). “Penetration of reinforced concrete by BETA-B-500 numerical analysis using a new

macroscopic concrete model for hydrocodes.” In: Proceedings of 9th international symposium on interaction of the effects of munitions

with structures. Berlin-Strausberg, Germany, 315–322.

Riedel, W. (2000). “Beton unter dynamischen Lasten: Meso-und makromechanische Modelle und ihre Parameter.” Doctoral Thesis, Institut

Kurzzeitdynamik, Ernst-Mach-Institut, der Bundeswehr Munchen, Freiburg, 210. [In German]

Rajendran, R., and Lee, J. M. (2009). “Blast loaded plates.” Marine Structures, 22(2), 99–127.

Spranghers, K., Vasilakos, I., Lecompte, D., Sol, H., and Vantomme, J. (2013). “Numerical simulation and experimental validation of the

dynamic response of aluminum plates under free air explosions.” International Journal of Impact Engineering, 54, 83–95.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

19

Tian L., Li Z. X. (2008). “Dynamic response analysis of a building structure subjected to ground shock from a tunnel explosion.”

International Journal of Impact Engineering, 35(10): 1164–1178.

Tu, Z., and Lu, Y. (2009). “Evaluation of typical concrete material models used in hydrocodes for high dynamic response simulations.”

International Journal of Impact Engineering, 36(1), 132-146.

Tang, E. K. C., and Hao, H. (2010). “Numerical simulation of a cable-stayed bridge response to blast loads, Part I: Model development and
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
ip
response calculations.” Engineering Structures, 32(10), 3180–3192.

Van der Veen, W. A. (2003) “Simulation of a compartmented airbag deployment using an explicit, coupled Euler/Lagranginan method with

d cr
adaptive Euler domains.” In: NAFEMS.

te s
Wang, Z., Lu, Y., Hao, H., and Chong, K. (2005). “A full coupled numerical analysis approach for buried structures subjected to subsurface

di nu
blast.” Computers & Structures, 83(4-5), 339–356.

Wang, W., Zhang, D., Lu, F., Wang, S. C., and Tang, F. (2013). “Experimental study and numerical simulation of the damage mode of a
ye a
square reinforced concrete slab under close-in explosion.” Engineering Failure Analysis, 27, 41–51.
op M
Yu, T. (2009). “Dynamical Response Simulation of Concrete Dam Subjected to Underwater Contact Explosion Load.” Computer Science

and Information Engineering, 769–774.


C ted

Zhang, A., Zeng, L., Cheng, X., Wang, S., and Chen, Y. (2011a). “The evaluation method of total damage to ship in underwater explosion.”

Applied Ocean Research, 33(4), 240–251.

Zhang, A., Zhou, W., Wang, S., Feng, L. (2011b). “Dynamic response of the non-contact underwater explosions on naval equipment.” Marine
ot p

Structures, 24(4), 396–411.


N ce

Zakrisson, B., Wikman, B., and Häggblad, H. A. (2011). “Numerical simulations of blast loads and structural deformation from near-field

explosions in air.” International Journal of Impact Engineering, 38(7), 597–612.


Ac

Zhang, S., Wang, G., Wang, C., Pang, B., Du, C. (2014). “Numerical simulation of failure modes of concrete gravity dams subjected to

underwater explosion.” Engineering Failure Analysis,36: 49-64.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

20

Table 1

Parameters used in the RHT model for concrete

Equation of state P alpha Strength RHT concrete


3
Reference density 2.750e3 (kg/m ) Shear modulus 1.670e10 (Pa)
Porous density 2.314e3 (kg/m3) Compressive strength (fc) 3.500e7 (Pa)
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Porous soundspeed 2.920e3 (m/s) Tensile strength (ft/fc) 0.100


Initial compaction pressure 2.330e7 (Pa) Shear strength (fs/fc) 0.180
Solid compaction pressure 5.999e9 (Pa) Intact failure surface constant A 1.600
Compaction pressure 3.000 Intact failure surface exponent N 0.610
Solid EOS Polynomial Tens./Comp. meridian ratio (Q) 0.6805
Bulk modulus A1 3.527e10 (Pa) Brittle to ductile transition 0.0105
Parameter A2 3.958e10 (Pa) G (elastic)/(elastic-plastic) 2.000
Parameter A3 9.040e9 (Pa) Elastic strength/ft 0.700
Parameter B0 1.220 Elastic strength/fc 0.530
Parameter B1 1.220 Fractured strength constant B 0.700
Parameter T1 3.527e10 (Pa) Fractured strength exponent M 0.800
Parameter T2 0 (Pa) Compressive strain-rate exponent Į 0.032
Reference temperature 300.000 (K) Tensile strain-rate exponent į 0.036
Specific heat 654.000 (J/kgK) Max. fracture strength ratio 1.000e20
Failure RHT concrete Erosion strain Geometric strain
Damage constant D1 0.015 Erosion strain 2.000
Damage constant D2 1.000 Type of geometric strain Instantaneous
Minimum strain to failure 8e-4
Residual shear modulus
0.130
fraction

Accepted Manuscript
Not Copyedited
Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

21

Table 2

Parameters used in theJH2 model for rock

Equation of state
Reference density ȡ0 2.660e3 (kg/m3)
Bulk modulus K1 2.570e10 (Pa)
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Polynomial EOS constant K2 -4.500e12(Pa)


Polynomial EOS constant K3 3.000e14 (Pa)
Johnson-Holmquist strength model
Shear modulus G 2.190e10 (Pa)
Hugoniot elastic limit HEL 4.500e9 (Pa)
Intact strength constant A 0.760
Intact strength exponent N 0.620
Strain rate constant C 0.005
Fractured strength constant B 0.250
Fractured strength exponent M 0.620
Max. fractured strength ratio 0.250
Johnson-Holmquist failure model
Hydro tensile limit HTL -5.400e7 (Pa)
Damage constant D1 0.005
Damage constant D2 0.700
Bulking constant ȕ 0.500
Type of tensile failure Hydro

Accepted Manuscript
Not Copyedited
Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJXUH&DSWLRQV/LVW

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Fig. 1. Schematic of the coupling Lagrangian-Eulerian approach.

Fig. 2. Three strength surface.

Fig. 3. Typical failure curves in a deviatoric plane for different hydrostatic pressures.

Fig. 4. Illustration of strain hardening in RHT model.

Fig. 5. Computational model of free-field explosion.


Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
ip
Fig. 6. Typical free pressure time history at 5m from charge center. (a) Underwater explosion pressure time history; (b) Air blast pressure

time history

d cr
Fig. 7. Comparison of peak pressures and impulses from explosions in water and air. (a) Peak pressure; (b) Impulse

te s
di nu
Fig. 8. Comparison of numerical and experimental values of peak pressures (a) in underwater and (b) in air.

Fig. 9. Comparison of numerical and experimental values of impulse (a) in underwater and (b) in air.
ye a
Fig. 10. Configuration of numerical model and coupled numerical model. (a) Schematic description of model configurations; (b) Coupled
op M

model for numerical simulation.


C ted

Fig. 11. Damage profiles of the dam under Mesh I at two selected times. (a) t=10.0ms; (b) t=35.0ms.

Fig. 12. Damage profiles of the dam under Mesh II at two selected times. (a) t=10.0ms; (b) t=35.0ms.
ot p

Fig. 13. Computed velocity time histories at the dam crest.


N ce

Fig. 14. Underwater explosion pressure wave propagation at different times (a) t=3ms, (b) t=5ms, (c) t=6ms, (d) t=7ms.
Ac

Fig. 15. Air blast pressure wave propagation at different times (a) t=3ms, (b) t=5ms, (c) t=6ms, (d) t=7ms.

Fig. 16. Arrangement of target points. (a) Target point in water or air; (b) Target point in dam.

Fig. 17. Pressure histories at targets from explosion in water.

Fig. 18. Overpressure histories at targets from explosion in air.

Fig. 19. Comparison of blast pressure load on the front face of the dam structure: (a) underwater explosion, (b) air blast.

Fig. 20. Comparison of stress wave in the dam at Target 12: (a) underwater explosion, (b) air blast.

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

Fig. 21. Comparison of stress wave in the dam at Target 13: (a) underwater explosion, (b) air blast.

Fig. 22. Comparison of horizontal velocity time histories at selected target points: (a) underwater explosion, (b) air blast.

Fig. 23. Comparison of horizontal acceleration time histories at selected target points: (a) Target 14; (b) Target 15; (b) Target 16; (b) Target

Fig. 24. The accumulated damage propagation process of the dam subjected to underwater explosion.

Fig. 25. Distribution of cumulative damage in the dam subjected to air blast.
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

t
ip
d cr
te s
di nu
ye a
op M
C ted
ot p
N ce
Ac

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 2.4
b
80
Peak pressre 2.0

1.6
60
Pressure (MPa)

Pressure (MPa)
1.2 Peak overpressre

40 0.8
Ambient pressure
0.4
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

20
0.0
Positive phase Negative phase
duration duration
0 -0.4
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 1200 b 500
1000 Underwater explosion 400 Underwater explosion
Peak pressure (MPa)

Air blast Air blast

Impulse (KPa-s)
800
300
600
200
400
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

100
200

0 0

1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
Distance from charge center (m) Distance from charge center (m)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 1200 b
80
1000 100mm 100mm
200mm 200mm

Pressure (MPa)
60
Pressure (MPa)

800 Empirical formula Empirical formula

600 40

400
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

20
200
0
0
2 4 6 8 10 2 4 6 8 10
Distance from charge center (m) Distance from charge center (m)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 500 b 20
100mm 100mm
400
200mm 15 200mm
Empirical formula

Impulse (KPa-s)
Empirical formula
Impulse (KPa-s)

300
10
200
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

5
100

0 0
2 4 6 8 10 2 4 6 8 10
Distance from charge center (m) Distance from charge center (m)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited
Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

b
a
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

2.0

1.5

Velocity (m/s)
1.0

0.5 Meshĉ
MeshĊ

0.0
0 50 100 150 200 250 300
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

c d

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

c d

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

1200 200 80
Target 1 Target 2 Target 3
1000
150 60
Pressure (MPa)

Pressure (MPa)

Pressure (MPa)
800
600 100 40
400
200 50 20

0
0 0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms) Time (ms)

50 50
Target 4 Target 5 Target 6
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

30
40 40
Pressure (MPa)

Pressure (MPa)

Pressure (MPa)
30 30
20
20 20
10
10
10
0 0 0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

100 4 2.5
Target 1 Target 2 Target 3

Overpressure (MPa)
Overpressure (MPa)
Overpressure (MPa)

80 2.0
3
60 1.5
2
40 1.0
1 0.5
20

0 0.0
0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms) Time (ms)

1.2 2.5
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Target 4 Target 5 Target 6


Overpressure (MPa)

0.8
0.9 Overpressure (MPa)

Overpressure (MPa)
2.0
0.6
0.6 1.5
0.4
1.0
0.3
0.2
0.5
0.0
0.0
0.0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

50
a b T6
T6 2.5
Target 6 Target 6
40 Target 7 Target 7
2.0

Overpressure (MPa)
Target 8 Target 8
Pressure (MPa)

Target 9 Target 9
30 1.5
T7 Target 10 Target 10
T10 Target 11 Target 11
T9
20 1.0
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

T8 T7
T11
10 0.5
T8
T9 T10 T11
0.0
0
0 20 40 60 80 100 0 20 40 60 80 100
Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 40 b 1.5
Target 12
Target 12
1.0
30

0.5
Pressure (MPa)

Pressure (MPa)
20
0.0
10
-0.5

0
-1.0
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

-10 -1.5
0 20 40 60 80 100 0 20 40 60 80 100
Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

4 b 0.6
a Target 13 Target 13
2
0.3
Pressure (MPa)

Pressure (MPa)
0
0.0
-2

-0.3
-4
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

-6 -0.6
0 20 40 60 80 100 0 20 40 60 80 100
Time (ms) Time (ms) 

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b 0.6 Target 14
4 Target 14 Target 15
Target 15 Target 16
Target 16 0.4 Target 17
3
Target 17

Velocity (m/s)
Velocity (m/s)

2 0.2

1
0.0
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

0
-0.2
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a 3000 Underwater explosion b 1500 Underwater explosion


Acceleration (m/s ) Air blast Air blast

Acceleration (m/s )
2

2
1500
0

-1500
-1500

-3000 -3000
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (ms) Time (ms)
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

2000 2000
c Underwater explosion
d Underwater explosion
Air blast Air blast
1000 1000
Acceleration (m/s )

Acceleration (m/s )
2

2
0 0

-1000 -1000

-2000 -2000
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (ms) Time (ms)

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589

a b
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

c d

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.


)LJSGI

Journal of Performance of Constructed Facilities. Submitted October 22, 2013; accepted January 29, 2014;
posted ahead of print January 31, 2014. doi:10.1061/(ASCE)CF.1943-5509.0000589
Downloaded from ascelibrary.org by PENN STATE UNIV on 08/12/14. Copyright ASCE. For personal use only; all rights reserved.

Accepted Manuscript
Not Copyedited

Copyright 2014 by the American Society of Civil Engineers

J. Perform. Constr. Facil.

You might also like