You are on page 1of 92

Ship-like target design for underwater

explosion experiments

Malcolm J. Smith
DRDC – Atlantic Research Centre

Julian J. Lee
DRDC – Suffield Research Centre

Defence Research and Development Canada


Scientific Report
DRDC-RDDC-2017-R096
6HSWHPEHU 2017
Template in use: (2010) SR Advanced Template_EN (051115).dotm

© Her Majesty the Queen in Right of Canada, as represented by the Minister of National Defence, 2017
© Sa Majesté la Reine (en droit du Canada), telle que représentée par le ministre de la Défense nationale,
2017
Abstract

The design of floating ship-like targets is considered with the objective of reproducing at reduced
scale most of the loading and response effects that would occur in a full-scale vessel in a close
proximity underwater explosion event. Some candidate designs are produced based on simple
parameterized cross sections. Design parameters are then iteratively adjusted so as to satisfy
parameter constraints and the overall design objectives, which include alignment between the
fundamental natural period and the explosion gas bubble pulsation period. Candidate designs are
proposed, and the shock and gas bubble loading and target response in each case are extensively
studied using a 3D coupled Eulerian-Lagrangian simulation methodology. The results of these
simulations show that the details of the loading and the response among the four models are
significantly different despite the similarity of the test configurations. Local deformation and
damage are observed in some models, and a global whipping response is observed in other
models. Because of the small scale of the configurations selected, buoyancy effects are very small
and this may limit the applicability of the configurations to full-scale.

Significance to defence and security

The response of naval platforms to underwater explosion is of direct relevance to warfighting


scenarios. An assessment tool, properly validated with measurements from trials and experiments,
would enable the capability of existing vessels to be assessed, and allow the survivability of
future warship designs to be improved. The work in this report is supported by Defence Research
and Development Canada (DRDC)’s Fleet Transformation project, and is supported by the
Canada/Netherlands/Sweden Project Arrangement on Close Proximity UNDEX Effects.

DRDC-RDDC-2017-R096 i
Résumé

On envisage la conception de cibles flottantes semblables à des navires dans le but de reproduire
à une échelle réduite la plupart des effets de charge et de réaction qui se produisent sur un navire
à échelle réelle lors d’une explosion sous-marine à proximité immédiate. Des configurations
possibles ont été produites à partir de simples coupes transversales paramétrées. Les paramètres
de conception ont alors été rajustés de façon répétitive afin de satisfaire aux contraintes et aux
objectifs généraux, notamment la concordance entre la fréquence propre fondamentale et la
fréquence d’oscillation de la bulle de gaz lors de l’explosion. Dans l’étude, on propose des
configurations possibles; la charge de choc et de la bulle de gaz, ainsi que la réaction de la cible
sont étudiées en détail à l’aide de simulations 3D selon la méthode Euler-Lagrange. Les résultats
de ces simulations révèlent que les détails de la charge et de la réaction des quatre modèles
diffèrent considérablement, malgré la similitude des configurations mises à l’essai. On a observé
une déformation et des dommages locaux chez certains modèles, et une vibration de frottement
généralisée chez d’autres. En raison de l’échelle réduite de la configuration choisie, les effets de
flottabilité sont très légers et pourraient en limiter l’application à échelle réelle.

Importance pour la défense et la sécurité

La réaction des plateformes navales aux explosions sous-marines est en rapport direct avec les
scénarios de guerre. Un outil d’évaluation, validé correctement selon les mesures effectuées lors
d’essais et d’expériences, permettrait d’évaluer la capacité des navires de guerre actuels et
d’améliorer la surviabilité des futurs navires. Les travaux consignés dans ce rapport sont appuyés
par le projet de RDDC sur la transformation de la flotte et par l’entente de projet sur les effets des
explosions sous-marines (UNDEX) à proximité immédiate conclue entre le Canada, les Pays-Bas
et la Suède.

ii DRDC-RDDC-2017-R096
Table of contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Significance to defence and security . . . . . . . . . . . . . . . . . . . . . . i
Résumé . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Importance pour la défense et la sécurité . . . . . . . . . . . . . . . . . . . . ii
Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
List of figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
List of tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1 Shock interaction with structures . . . . . . . . . . . . . . . . . . . 4
2.2 Gas bubble behaviour . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.1 Analytical models for free-field bubbles . . . . . . . . . . . . . 6
2.2.2 Higher fidelity modelling of gas bubbles . . . . . . . . . . . . . 8
2.2.3 Bubble dynamics near boundaries . . . . . . . . . . . . . . . . 9
2.3 Gas bubble interaction with structures . . . . . . . . . . . . . . . . . 11
2.3.1 Far-field interactions . . . . . . . . . . . . . . . . . . . . 11
2.3.2 Close proximity interactions . . . . . . . . . . . . . . . . . . 12
2.4 Surface effects . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3 Design approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1 Design criteria and objectives . . . . . . . . . . . . . . . . . . . . 15
3.2 Design parameters . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Design steps . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.4 Estimating the natural frequency and fluid added mass . . . . . . . . . . . 20
4 Candidate designs. . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.1 Open girder designs . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2 Closed girder designs . . . . . . . . . . . . . . . . . . . . . . . 25
4.3 Segmented model . . . . . . . . . . . . . . . . . . . . . . . . 28
4.4 Barge configuration . . . . . . . . . . . . . . . . . . . . . . . . 30
5 UNDEX simulation results . . . . . . . . . . . . . . . . . . . . . . . 32
5.1 Simulation method . . . . . . . . . . . . . . . . . . . . . . . . 32
5.1.1 Fluid models . . . . . . . . . . . . . . . . . . . . . . . 32
5.1.2 Structural models . . . . . . . . . . . . . . . . . . . . . . 35
5.1.3 Coupling method . . . . . . . . . . . . . . . . . . . . . . 36
5.2 Closed rectangular girder results . . . . . . . . . . . . . . . . . . . 36
5.3 Closed semi-cylindrical girder results . . . . . . . . . . . . . . . . . 42
5.4 Segmented model results . . . . . . . . . . . . . . . . . . . . . . 49
5.5 Barge model results . . . . . . . . . . . . . . . . . . . . . . . . 55

DRDC-RDDC-2017-R096 iii
5.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
List of symbols/abbreviations/acronyms/initialisms . . . . . . . . . . . . . . . . 77

iv DRDC-RDDC-2017-R096
List of figures

Figure 1: Sequence of loading and response processes in close proximity UNDEX. . . 3


Figure 2: Open rectangular box configuration. . . . . . . . . . . . . . . . . . 17
Figure 3: Experimental configuration. . . . . . . . . . . . . . . . . . . . . . 18
Figure 4: Design process. . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Figure 5: Open semi-cylindrical configuration. . . . . . . . . . . . . . . . . . 22
Figure 6: Natural modes of the open rectangular box girder configuration. . . . . . . 23
Figure 7: Two alternative open girder designs: braced (left); and stiffened with coaming
(right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Figure 8: Two lowest natural modes for the braced open girder design. . . . . . . . 24
Figure 9: Two lowest natural modes for the open girder design with coaming.. . . . . 24
Figure 10: Lowest natural modes for the open semi-cylindrical design. . . . . . . . . 25
Figure 11: Closed rectangular box configuration: with deck removed (left); meshed FE
model (centre); section view (right). . . . . . . . . . . . . . . . . . 26
Figure 12: Closed cylindrical girder configuration: with deck removed (left); meshed FE
model (centre); section view (right). . . . . . . . . . . . . . . . . . 26
Figure 13: Closed rectangular box girder natural vibration modes. . . . . . . . . . . 27
Figure 14: Closed cylindrical box girder natural vibration modes. . . . . . . . . . . 28
Figure 15: Segmented model configuration. . . . . . . . . . . . . . . . . . . . 28
Figure 16: One-quarter finite element model of the segmented design. . . . . . . . . 29
Figure 17: Two lowest wet natural frequencies for the segmented model. . . . . . . . 30
Figure 18: Barge configuration. . . . . . . . . . . . . . . . . . . . . . . . . 30
Figure 19: Two lowest wet natural frequencies for the barge model. . . . . . . . . . 31
Figure 20: Simulation configuration. . . . . . . . . . . . . . . . . . . . . . . 32
Figure 21: 2D axisymmetric model of the fluid domain used for the Stage 1 analysis
(right); with detail showing charge geometry (left). . . . . . . . . . . . 33
Figure 22: Comparison of shock front time histories at two standoff distances. . . . . . 33
Figure 23: 3D quarter symmetric fluid domain grids used for the Stage 2 simulations: full
model (top); two views of refined region (bottom). . . . . . . . . . . . 34
Figure 24: One quarter finite element model for the close rectangular box girder. . . . . 37
Figure 25: Loading time histories at the midships centerline position, closed rectangular
model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Figure 26: Pressure time history near the first (left) and second (right) bubble collapse:
close rectangular model. . . . . . . . . . . . . . . . . . . . . . . 38

DRDC-RDDC-2017-R096 v
Figure 27: Evolution of the explosion gas bubble during the first bubble period: closed
rectangular girder model. Colours indicate density in kg/m3. . . . . . . . . 39
Figure 28: Gas bubble at full expansion. . . . . . . . . . . . . . . . . . . . . 39
Figure 29: Evolution of the explosion gas bubble during the second bubble period: closed
rectangular girder model. Colours indicate density in kg/m3. . . . . . . . . 40
Figure 30: Evolution of the explosion gas bubble during the third bubble pulse: closed
rectangular girder model. Colours indicate density in kg/m3. . . . . . . . . 41
Figure 31: Time history response of three points on the rectangular box girder. . . . . . 41
Figure 32: Displacement contour plots for the rectangular box girder. . . . . . . . . 42
Figure 33: One quarter finite element model for the semi-cylindrical girder. . . . . . . 43
Figure 34: Loading time histories at the midships keel location, semi-cylindrical model. . 43
Figure 35: Pressure time history for the first (left) and second (right) bubble pulsed at the
midships keel location: semi-cylindrical model. . . . . . . . . . . . . . 44
Figure 36: Evolution of the explosion gas bubble during the first bubble collapse:
semi-cylindrical girder model. Colours indicate density in kg/m3. . . . . . . 45
Figure 37: Gas bubble at 0.095 ms (left); and 13 ms (right). . . . . . . . . . . . . 45
Figure 38: Evolution of the explosion gas bubble during the second bubble period:
semi-cylindrical girder model. Colours indicate density in kg/m3. . . . . . . 46
Figure 39: Evolution of the explosion gas bubble during the third bubble collapse:
semi-cylindrical girder model. Colours indicate density in kg/m3. . . . . . . 47
Figure 40: Time history response at three locations of the semi-cylindrical girder design. . 48
Figure 41: Semi-cylindrical girder deformation contours. . . . . . . . . . . . . . 48
Figure 42: Loading time histories at the midships keel for the segmented model. . . . . 49
Figure 43: Pressure time history of the first (left) and second (right) bubble collapse at the
midships keel location: segmented model. . . . . . . . . . . . . . . . 50
Figure 44: Evolution of the explosion gas bubble during the first bubble collapse:
segmented model. Colours indicate density in kg/m3. . . . . . . . . . . . 51
Figure 45: Gas bubble diameter at 14 ms. . . . . . . . . . . . . . . . . . . . . 52
Figure 46: Evolution of the explosion gas bubble during the second bubble period:
segmented model. Colours indicate density in kg/m3. . . . . . . . . . . . 52
Figure 47: Evolution of the explosion gas bubble during the third bubble collapse:
segmented model. Colours indicate density in kg/m3. . . . . . . . . . . . 53
Figure 48: Time history response at three locations of the segmented model design. . . . 54
Figure 49: Segmented model deformation contours. . . . . . . . . . . . . . . . . 54
Figure 50: Structural FE model for the Barge design. . . . . . . . . . . . . . . . 55
Figure 51: Loading time histories at the midships keel position of the barge model. . . . 55

vi DRDC-RDDC-2017-R096
Figure 52: Evolution of the explosion gas bubble during the first bubble collapse: barge
model. Colours indicate density in kg/m3. . . . . . . . . . . . . . . . 56
Figure 53: Evolution of the explosion gas bubble during the second bubble collapse:
barge model. Colours indicate density in kg/m3. . . . . . . . . . . . . . 57
Figure 54: Evolution of the explosion gas bubble during the third bubble collapse: barge
model. Colours indicate density in kg/m3. . . . . . . . . . . . . . . . 58
Figure 55: Gas bubble at maximum expansion.. . . . . . . . . . . . . . . . . . 59
Figure 56: Vertical displacement time histories at several locations on the barge. . . . . 60
Figure 57: Displacement contours (in m) for the barge model.. . . . . . . . . . . . 60

DRDC-RDDC-2017-R096 vii
List of tables

Table 1: Loading processes and response effects in close proximity UNDEX. . . . . 2


Table 2: Design criteria for configuration selection. . . . . . . . . . . . . . . . 16
Table 3: Design objectives for parameter definition. . . . . . . . . . . . . . . . 16
Table 4: Design parameters for an open rectangular box configuration. . . . . . . . 17
Table 5: Parameter values for the open rectangular girder design. . . . . . . . . . 22
Table 6: Parameter values for the open semi-cylindrical design. . . . . . . . . . . 24
Table 7: Parameter values for the closed rectangular girder design. . . . . . . . . . 27
Table 8: Parameter values for the segmented configuration. . . . . . . . . . . . . 29
Table 9: Parameter values for the barge design. . . . . . . . . . . . . . . . . . 31
Table 10: Equation of state parameters. . . . . . . . . . . . . . . . . . . . . 35
Table 11: Material properties. . . . . . . . . . . . . . . . . . . . . . . . . 36
Table 12: Summary of predicted peak pressures and impulses for the shock wave and
first three bubble collapse events. . . . . . . . . . . . . . . . . . . 61
Table 13: Summary of predicted bubble periods and radii. . . . . . . . . . . . . . 62
Table 14: Summary of free-field bubble periods and radii. . . . . . . . . . . . . . 62

viii DRDC-RDDC-2017-R096
1 Introduction

Underwater explosions (UNDEX) can inflict considerable damage to naval vessels, resulting in
large loss of material and human life. Accurate prediction of UNDEX effects could help improve
hull designs to resist these loads and avoid weaknesses in designs leading to premature or
catastrophic failures. While reliable tools exist for predicting platform response to UNDEX at
medium to large standoff distances, most of these tools cannot model highly damaging close
proximity events, where the explosion occurs at standoff distances less than the diameter of the
gas bubble.

Project arrangements (Close Proximity Underwater Explosion Effects I and II), carried out under
the Canada, Netherlands, Sweden MOU, have worked toward creating validated simulation
capabilities for close proximity UNDEX and have resulted in a series of collaborative studies [1]–[4].
Objectives of the collaborations are accurate prediction of both the loading and the response
effects of targets in controlled settings, and extending this capability to full-scale scenarios.
Prediction tools used have generally been hi-fidelity modelling and simulation software, as well
as simplified tools specialized to close proximity problems.

To validate simulation tools, it is necessary to have accurate measurement data from carefully
performed tests. Full-scale test data involving destructive levels of UNDEX loading are normally
not obtainable due to the high cost of such tests and the scarcity of expendable hulls. In lieu of
full scale data, experiments are conducted at reduced scale. A wide range of reduced-scale
UNDEX testing has been performed in Canada and in other countries. Considering just close
proximity UNDEX, several series of tests have been conducted at the Suffield Research Centre to
measure the loading and response of responding and deforming targets [5]–[10]. These tests were
performed using small flat or curved target plates of varying strength or stiffness, in which the
target plate is mounted in a stiff frame. While such tests are representative of local UNDEX
loading and damage mechanisms, they do not capture global response and damage effects that
occur in a naval platform. Few tests reported in the literature have attempted to capture the full
range and complexity of global and local responses to close proximity UNDEX. This is an
important gap.

The present study aims to fill this gap by proposing a series of reduced-scale UNDEX tests
involving global and local response effects (and their interaction) in slender, floating, ship-like
targets. In Section 2 a review of the main physical processes, models, methods and previous work
in this area is provided. In Section 3, an outline of the experiments being proposed along with the
approach used to design the target specimens. In Section 4, a series of candidate tests specimens
are presented. In Section 5, the results of simulations of four proposed target specimens under
close proximity loads are presented. In Section 6, recommendations are given.

DRDC-RDDC-2017-R096 1
2 Background

The literature on underwater explosions dates back to the pioneering work done in Great Britain
and the United States during World War II [11]. In the years that followed, most research in this
area concerned standoff UNDEX effects due to their importance to the survivability design for
warships. Cole [12] provides a comprehensive treatment of the early UNDEX research. Keil [13]
and Holt [14] provide reviews of the fundamentals, and Swisdak [15] provides a useful summary
of properties and data. Relatively little work has addressed close proximity UNDEX, and modern
warships are generally not designed to survive severe close proximity loads. However, the sinking
of ROKS Cheonan in 2010 renewed interest in performance of vessels under close proximity
UNDEX.

Underwater explosions (UNDEX) damage ships or submarines through complex loading


processes. These include a short duration impulsive loading due to the shock wave and a longer
duration impulsive loading due to the pulsating motion of the explosion gas bubble. The resulting
response and damage depends on the mass and chemical composition of the explosive, its
proximity to the hull (standoff distance), the size, shape, and mass of the hull, as well as the
properties of its structure.

The main loading processes and response effects associated with this sequence of events are
summarized in Table 1, and a notional sequence of events in close proximity UNDEX is shown in
Figure 1. In this table, each response effect is labelled as Global or Local according to whether
the indicated effects would occur in a localized region of the platform (Local), or throughout the
entire platform (Global) in a typical close proximity UNDEX scenario.

The remainder of Section 2 gives an overview of the physical processes, models and methods
relevant to the close proximity UNDEX problem.

Table 1: Loading processes and response effects in close proximity UNDEX.


Loading process Response effects Global or Local Time scale
Shock wave Kickoff velocity, surface reflection, bulk Global < 0.1 s
cavitation
Equipment and machinery failure; Global or Local < 0.1 s
rupture or fracture of foundations
Dishing or rupture of hull plating; Local < 0.1 s
deformation, buckling and fracture of
members; hull cavitation
Gas bubble Lifting of the hull under added buoyancy Global 0.1–0.5 s
expansion leading to permanent deformation or
rupture
Gas bubble collapse Hull girder whipping leading to Global > 0.5 s
and high velocity permanent deformation or rupture
water jetting
Misalignment of machinery Global or local > 0.5 s

Dishing or rupture of hull plating; Local > 0.5 s


deformation, buckling and fracture of
members

2 DRDC-RDDC-2017-R096
< 0.1 s

0.1 – 0.5 s

> 0.5 s

> 0.5 s

Figure 1: Sequence of loading and response processes in close proximity UNDEX.

DRDC-RDDC-2017-R096 3
2.1 Shock interaction with structures
The fundamentals of underwater shock waves are discussed by Cole [12], Holt [14] and
Swisdak [15]. In the free field, i.e., in the absence of any surfaces, and at a sufficiently large
distance R from the detonation point, an underwater shock wave can be approximated as a step
exponential pressure wave:

ܲሺ‫ݐ‬ሻ ൌ ܲ0 ݁ ିሺ௧ି௧బ ሻȀఏ ‫ ݐ‬൒ ‫ݐ‬଴ (1)

The peak pressure, P0, and decay constant, θ, are given by the following similitude relations [13]:

஺భ
ܹ ଵȀଷ
ܲͲ ൌ ‫ܭ‬ଵ ቆ ቇ (2)
ܴ

஺మ
ܹ ଵȀଷ
ߠ ൌ ‫ܭ‬ଶ ܹ ଵȀଷ
ቆ ቇ (3)
ܴ

where W is the charge mass, and K1, K2, A1, and A2 are empirical constants that depend on the
chemical composition of the explosive.

The shock wave is spherical, but at sufficiently large distances can usually be approximated as a
plane wave such that,

ܲ ൌ ߩܿ‫ݒ‬ (4)

where ρ is the density of the fluid, c is the fluid sound speed, and ‫ ݒ‬is the fluid particle velocity.

Fundamentals of shock interactions with naval vessels are given by Keil [13] and Scavuzzo and
Pusey [16]. For a naval platform, whether floating or submerged, two processes may be
considered: a global response, often called the “kickoff;” and a local response involving dishing
of the plating and stiffened panels in the bottom or side shell, and possibly also rupture and holing
of the hull.

The kickoff effect initiates a global whipping response of the hull structure. It is determined by
summation of the fluid particle velocities that would occur in the water volume displaced by the
hull, combining the contributions from the incident wave and the surface reflections [17]. For a
surface ship, the horizontal components of the incident and reflected waves cancel each other out,
giving a net kickoff velocity in the vertical direction. For a submarine at depth, contributions from
surface reflection are negligible, and therefore the kickoff response is not restricted to vertical
motion.

For standoff explosions, the vessel can often be treated as a point mass for the purposes of
determining kickoff velocity. But for close proximity scenarios, the kickoff response occurs
progressively throughout the vessel, starting from locations closest to the charge.

4 DRDC-RDDC-2017-R096
Shock interaction with simple air-backed structures generally involves the reflection of the shock
at the fluid-structure interface (FSI), which imparts an initial velocity to the structure; the
cavitation of the water near the FSI, and the subsequent impulsive reloading the plate due to
cavitation closure; and elastic and plastic deformation of the structural material possibly resulting
in necking or rupture. The mechanisms involved are similar for standoff and non-contact close
proximity scenarios. But in close proximity UNDEX, the loading on the plate may be further
influenced by interaction of the reflected shock (rarefaction) wave with the expanding gas bubble
[18],[19]. For contact underwater detonations, different loading (e.g., fragmentation) and
structural failure modes (shear failure) influence the resulting damage, and gas bubble loading
may be less important [20]–[22].

Early work considered plane shocks at normal incidence on a flat plate target [23],[24] and these
formed the basis of subsequent design guidance [13],[25]. Many numerical and experimental studies
of flat plate, stiffened panel and cylindrical structural configurations have appeared [26]–[36].

Numerical analysis methodologies for standoff UNDEX shock interactions include the Doubly
Asymptotic Approximation (DAA) method first developed by Geers [37], which is valid for in
the early-time shock loading phase, as well as for the late-time gas bubble loading phase. The
method was extended to include cavitating fluid volumes within the DAA framework by Fellipa
and DeRuntz [38] and Sprague and Geers [39]. These methods were implemented in the
Underwater Shock Analysis (USA) code [40],[41]. Related methods include Trouwborst and
Bosman [42], who developed a simplified method for shock analysis using a plane-wave
approximation model that takes into account cavitation. Coupled Eulerian-Lagrangian and other
hydrocode methods have been used by several researchers to analyse the interaction of
underwater shock with thin-walled structures ([18],[21],[22],[43],[44]).

Two dimensional analyses of ship sections subject to underwater shock were performed by Shin
and Santiago [45], who analysed a naval vessel using the DAA method with a region with
cavitating fluid elements; and by van Aanhold, et al. [46], who used a Lagrangian methodology
for both the fluid and structure to perform underwater shock analysis on a simplified floating
cylindrical structure.

Simulations of full-scale shock trials of US destroyer designs using the DAA method were
performed by Shin and Park [47] and Shin and Schneider [48]. These studies employed detailed
3D models of the entire vessel and a fluid volume model of cavitating fluid elements extending
from the wetted hull surface out to the DAA boundary. Other shock analyses of full scale vessels
include Liang and Tai [49], who performed a DAA assessment of a 2000-tonne patrol boat
subjected to a far-field shock loading with a keel shock factor of 0.8. Sprague and Geers [50]
modelled the far-field shock response of a floating slender box structure using a DAA method
with spectral elements for the cavitating fluid. Gong and Lam [51] used an explicit finite element
method (FEM) / boundary element method (BEM) method to investigate the attenuation of
underwater shock through use of shock absorption layers. Zhang, et al. [52] analysed the shock
response of shipboard equipment using an acoustic-structure coupling method available in
commercial FEM software and a detailed whole-ship FE model of naval vessel, including detailed
modelling of shipboard equipment.

DRDC-RDDC-2017-R096 5
2.2 Gas bubble behaviour
While the shock wave propagation and interaction occurs on a short time scale (normal
milliseconds), the gas bubble develops over a much longer time scale (seconds for full-scale
charges). The gas bubble results from explosion products remaining trapped in a higher density
medium (water). The initial high pressure in the bubble causes it to expand and displace the
surrounding water, during which time the pressure inside the bubble drops. The momentum in the
displaced water allows the bubble to over-expand, causing the pressure in the bubble to fall below
the ambient hydrostatic pressure. Eventually the momentum of the displaced fluid is overcome
and the contraction phase begins, resulting in collapse of the bubble to a minimum volume. The
following occurs near the bubble minima:
x The gas pressure inside the bubble spikes, producing a compressive disturbance in the
surrounding water. An acoustic pulse is thus radiated away at each bubble collapse, and is a
significant source of energy loss from the bubble.
x The bubble migrates: bubbles in the free field will generally migrate upwards due to
buoyancy forces; whereas, those near a solid surface migrate toward the surface due to the
Bjerknes effect.
x The bubble changes shape, becoming less spherical, and may become toroidal if a water jet
forms. This is because of uneven velocity distribution of the incoming water.

2.2.1 Analytical models for free-field bubbles

Early analytical studies assumed that bubbles maintained their spherical shape without loss of
energy throughout successive cycles of expansion and contraction. The following nonlinear
equations for a spherical, migrating bubble with constant internal energy were derived by
Herring [53] and Taylor [54]:

Ͷ ݀ܽ ଶ ͳ ݀‫ ݖ‬ଶ
ߨߩܽଷ ݃‫ ݖ‬൅ ʹߨߩܽଷ ൬ ൰ ൅ ߨߩܽଷ ൬ ൰ ൌ ‫ ܧ‬െ ‫ܩ‬ሺܽሻ (5)
͵ ݀‫ݐ‬ ͵ ݀‫ݐ‬

݀‫݃ʹ ݖ‬
െ ൌ න ܽଷ ݀‫ݐ‬ (6)
݀‫ܽ ݐ‬ଷ

where ܽ is the instantaneous bubble radius, ‫ ݖ‬is the instantaneous depth measured to the centre of
the bubble, ݃ is the acceleration of gravity; ߩ is the density of water; ‫ ܧ‬is the constant total
energy of the bubble; and ‫ܩ‬ሺܽሻ is the internal energy of the gas in the bubble.

The maximum bubble radius can be determined using the non-migrating form of (5). This is
derived by setting ݀‫ݖ‬Τ݀‫ ݐ‬ൌ Ͳ and substituting an expression for ‫ܩ‬ሺܽሻ based on the isentropic gas
law ܲ௚ ܸ ఊ ൌ ݇:


݇ Ͷ ଷ ଵିఊ
‫ܩ‬ሺܽሻ ൌ  න ܲ௚ ܸ݀ ൌ  ൬ ߨܽ ൰ (7)
௏ሺ௔ሻ ߛെͳ ͵

6 DRDC-RDDC-2017-R096
where ܲ௚ is the gas pressure inside the bubble, ܸ is the gas bubble volume, ߛ is the ratio of
specific heats, and ݇ is a constant. At maximum expansion ݀ܽΤ݀‫ ݐ‬ൌ Ͳ. Substituting this and
‫ܩ‬ሺܽሻ into (5) gives a nonlinear equation in ܽ only. Solutions for the maximum bubble radius ‫ܣ‬ெͳ
are of the form:

ܹ ଵȀଷ
‫ܣ‬ெ1 ൌ ‫଺ܭ‬ (8)
‫ܪ‬ଵଵȀଷ

where ‫ܪ‬ଵ ൌ  ܲஶ Ȁሺߩ݃ሻis the hydrostatic head at the depth of the full expanded bubble,
ܲஶ ൌ ܲ௔ ൅ ߩ݃‫ݖ‬ଵ is the static pressure at the bubble depth ‫ݖ‬ଵ , ܲ௔ is atmospheric pressure, and ‫ ଺ܭ‬is
an empirical constant that depends on the chemical composition of the explosive. Willis [55]
derived the following form for the first bubble period by assuming a non-migrating spherical
bubble in an incompressible fluid with no loss of energy:

ܹ ଵȀଷ
ܶଵ ൌ ‫ܭ‬ହ  (9)
‫ܪ‬ଵ ହȀ଺

where ‫ܭ‬ହ is an empirical constant. Equations (8) and (9) are still widely used to determine the gas
bubble properties in free field conditions. Properties for subsequent bubble cycles can be related
to the properties of the first cycle through the empirical relations [56]:

ܶ௡ ‫ݎ‬௡ ଵȀଷ ‫ܪ‬ଵ ହȀ଺


ൌ൬ ൰ ൬ ൰ (10)
ܶଵ ‫ݎ‬ଵ ‫ܪ‬௡

‫ܣ‬ெ௡ ‫ݎ‬௡ ଵȀଷ ‫ܪ‬ଵ ଵȀଷ


ൌ൬ ൰ ൬ ൰ (11)
‫ܣ‬ெଵ ‫ݎ‬ଵ ‫ܪ‬௡

where ‫ݎ‬௡ Ȁ‫ݎ‬ଵ is the ratio of bubble energies in the nth and 1st cycle, taking into account losses
associated with each bubble collapse. Snay [56] gives energy ratios of ‫ݎ‬ଶ Τ‫ݎ‬ଵ ൌ ͲǤ͵ͷand
‫ݎ‬ଷ Τ‫ݎ‬ଵ ൌ ͲǤͳͺͷ for non-migrating free-field bubbles. Thus the 2nd and 3rd bubble periods of a
non-migrating bubble will be reduced by factors of 0.705 and 0.570 relative to the first cycle
period. The steadily diminishing bubble period makes it difficult for a series of bubble pulses to
excite of a ship hull girder to resonance.

The lossless model for motion of a spherical bubble given by Equations (5) and (6) has been
found to overestimate migration in comparison to experiments. Hicks [57] remedied this by
including hydrodynamic drag in the equations of motion and obtained better predictions of
migration by adjusting the drag coefficient to match experimental observations. Vernon [58]
developed a similar model taking into account free surface effects and drag effects during
migration. Heaton [59] rederived the equations for a spherical bubble including both migration
and energy dissipation, and including hydrodynamic drag to control the migration rate.

DRDC-RDDC-2017-R096 7
However, experiments also show that although an explosion gas bubble is nearly spherical during
its first expansion cycle, it loses spherical shape during its first contraction, and is generally
non-spherical throughout remaining cycles. Penny and Price [60] first analysed non-spherical
solutions to the bubble dynamics equations by including spherical harmonic functions in the
expression for the bubble radius and the velocity potential function of the surrounding flow.
Heaton [61] developed analytical equations for an ellipsoidal shaped bubble and found that more
accurate predictions of bubble period, migration and energy loss could be obtained without
including hydrodynamic drag.

An alternative form of the bubble equation can be derived from the Lagrangian of the kinetic and
potential energy expressions in (5), neglecting migration (‫ݖ‬ሶ ൌ Ͳ), and using the isentropic model
for ‫ܩ‬ሺܽሻ in (7). This gives the Rayleigh-Plesset equation for the dilatational motion of a spherical
bubble:

͵ ܲ௚ െ ܲஶ
ܽܽሷ ൅ ܽሶ ଶ ൌ (12)
ʹ ߩ

where ܽሶ ൌ ݀ܽȀ݀‫ݐ‬. Like (5), this equation is based on incompressible theory with no energy loss.

The most important energy loss mechanism is the acoustic radiation that occurs over a short time
interval about the bubble minima. With this in mind, Herring [53], Keller and Kolodner [62], and
Prosperetti and Lezzi [63] derived equations for dilatational motion that include in different ways
the compressibility (i.e., wave propagation) in the surrounding fluid medium. These equations are
similar to (12) but include additional terms linear in the Mach number of the bubble wall (ܽሶ Ȁܿ).
Geers and Hunter [64] extended this further by considering combined dilatation and migration of
a spherical bubble with wave propagation in the gas as well as the liquid.

More accurate predictions of the dilatational and translational motion are obtained when
compressibility effects are included. Despite the success of these models, they are generally not
applicable to close proximity UNDEX scenarios where the fluid flow is constrained by the nearby
structures, thus influencing the bubble dynamics and the loading seen by the structure. Some
progress has been made modifying the Geers-Hunter model to approximate bubble behaviour
close to structures, and further developments in this direction show promise [65].

2.2.2 Higher fidelity modelling of gas bubbles

Hydrodynamic computer codes, or “hydrocodes,” are specifically intended for modelling highly
dynamic physical processes and interactions, such as shock wave propagation and ballistic
penetration [66], [67]. Mair [68] performed an extensive review of the four general classes of
hydrocode: Lagrangian, Eulerian, Coupled Eulerian-Lagrangian (CEL), and Arbitrary
Lagrangian-Eulerian (ALE). Selected naval applications of hydrocodes were discussed by
Landsberg, et al. [69] and Wardlaw, et al. [70].

Hydrocode modelling of gas bubbles in the free field has been performed using Eulerian methods
(Hsu, et al. [71] and Smith [72]) and ALE methods (Menon [73], Brett [74], Barras, et al. [75]
and Webster [76]). Eulerian hydrocode modelling of gas bubbles near a free surface was
performed by Mader [77], Li and Rong [78] and Smith [72]. Petrov and Schmidt [79] used a

8 DRDC-RDDC-2017-R096
combined Eulerian-Lagrangian approach to examine multi-phase interactions of an explosion
near the free surface.

Potential flow and boundary element methods (BEM), which are based on assumptions of
incompressible, inviscid, and irrotational flow in the surrounding fluid, were first used to model
dynamics of cavitation bubbles by Blake, et al. [80],[81]. Best and Kucera [82] extended the
method to explosion bubbles containing non-condensable explosion products. Modelling of
bubble collapse and the formation of a re-entrant water jet have proved to be formidable
difficulties for the BEM approach. A re-entrant water jet is caused by irregular rates of
contraction on the bubble surface, due to a combination of gravity/buoyancy effects and
irregularities in the flow field due to nearby surfaces (Bjerknes effect), and results in a transition
from a singly-connected (spheroidal) to a doubly-connected (toroidal) topology. Generally, BEM
methods are capable of predicting the onset of jetting, but encounter errors, invalid solutions or
numerical difficulties as the jet develops. Difficulties associated with this are numerical
instabilities resulting from the high velocity of the forward surface of the water jet, severe
distortions introduced in the bubble mesh, and difficulties with representing the change in
topology in a potential flow formulation. The main advantage of this method is that unlike most
of the analytical bubble models discussed in the previous section the bubble shape is not
constrained to remain spherical, making it useful for shallow bubbles and bubbles interacting with
surfaces. A fundamental limitation of the potential flow/BEM approach is that energy dissipation
resulting from wave propagation is excluded from the incompressible formulation.

Best [83] computed the formation of axisymmetric toroidal bubbles by introducing a cut surface
that enabled a doubly connected domain to be rendered as a singly-connected domain in the
potential flow formulation. Wang, et al. [84] introduced a vortex ring into the potential flow
formulation for axisymmetric bubbles to model the circulation of the flow resulting from the
transition from singly connected (spheroidal) to doubly connected (toroidal) topology. Zhang,
et al. [85] extended the vortex ring concept to the BEM modelling of 3D bubbles and introduced
adaptive re-meshing of the bubble surface and least square smoothing algorithm to eliminate
numerical instabilities before and after transition from single connectedness to double
connectedness. The transition itself is effected by making a surgical incision in the bubble surface
based on single-point contact of the foremost node of the jet with the opposite bubble surface.
Wang [86] used a 3D BEM method to study dynamics of a single gas bubbles near free surfaces
and inclined rigid walls, as well as a pair of bubbles near a free surface, up to the point of jet
penetration.

Best [87] compared basic parameters of free-field gas bubbles computed using an axisymmetric
BEM approach with those determined using an analytical model for a spherical bubble. Zhang,
et al. [88] used ALE and BEM techniques to study the detonation and subsequent bubble
dynamics of underwater charges, and compared results to experiments with charges of different
size and shape.

2.2.3 Bubble dynamics near boundaries

To extend the free-field bubble behaviour described by the scaling relations (8) and (9) to the case
of close proximity UNDEX, the behaviour of bubbles near boundaries such as the water surface
or structures has been examined. Changes in the key bubble parameters have been investigated in
the far-field for relatively shallow charges, though not as shallow as the typical depths of close

DRDC-RDDC-2017-R096 9
proximity UNDEX threats. For gas bubbles near a free surface, Snay [56] suggested a correction
factor for the bubble period given by the term in parentheses in the following:

ܹ ଵȀଷ ‫ܣ‬ெଵ
ܶଵ ൌ ‫ܭ‬ହ ହȀ଺
൬ͳ െ Ƚ ൰ (13)
‫ܪ‬ଵ ‫ݖ‬ଵ

The recommended value for the surface correction factor Ƚ is 0.2.

Chapman [89] found that even for charges at depths of 20 times the maximum bubble radius,
significant deviations from the free-field scaling laws were observed for peak pressure and
impulse of the first, second, and third bubble collapses, and modified versions of the free-field
scaling laws with an additional multiplicative factor ‫ ݖ‬ఒ (where ‫ ݖ‬is the bubble depth) were
proposed. More detailed studies on small-scale, spark-generated bubbles by Chahine and Bovis
[90] found a reduction in the first bubble period near the surface, described by the relation:

ͲǤ͹ͻ
ܶଵ ൎ ͳǤͺ͵ ൬ͳ െ ൰ (14)
Ͷܼҧ

where ܼҧ is the depth normalized to the maximum bubble radius. Although a similar expression
was also given by Cole [12], (14) agrees better with his laboratory-scale experiments and
simulations at depths of ܼҧ ൏ ̱ͳ.

Studies of the buoyancy effects on the bubble collapse dynamics near the water surface found that
the direction of jetting could be characterized by a Froude number of the form [91]:

ܲஶ െ ܲ௩
” ൌ (15)
ʹߩ݃‫ܣ‬ெଵ

where ܲ௩ is the vapour pressure of the water, and other symbols are as defined in the previous
section. Since the direction of the bubble collapse and jetting depends strongly on the buoyancy,
the Froude number should ideally be conserved in a whipping experiment to replicate the bubble
collapse loading dynamics. However, the smaller charge sizes required by subscale experiments
produce smaller bubble radii (‫ܣ‬ெଵ ), and the other parameters in the Froude number are difficult
to scale, namely gravity and atmospheric pressure. Chahine, et al. [92] successfully used Froude
number scaling in a laboratory by performing experiments in a closed tank where the atmospheric
pressure above the tank could be lowered by partially evacuating the tank ullage. When
constrained to an outdoors UNDEX pond facility, the Froude parameters cannot be easily scaled
to compensate for the reduced charge size, and a Froude number of approximately 30 or more is
achieved instead of values of less than five typically observed in full-scale close proximity
UNDEX scenarios. It was therefore expected that buoyancy effects on the bubble collapse may
not be adequately characterized in the present experiments.

Wang, et al. [86] proposed a criterion for determining jetting direction of free-field bubbles using
ܼҧߜ, where ܼҧ is the depth normalized to the maximum bubble radius and ߜ ൌ ͳȀξʹ ” is a
non-dimensional buoyancy parameter. The critical value for ܼҧߜ is 0.442 which corresponds to the

10 DRDC-RDDC-2017-R096
Kelvin impulse state at which the upward buoyancy forces are balanced with the downward
Bjerknes force. For ܼҧ ߜ ൏ ͲǤͶͶʹ, the Bjerknes forces dominates and bubbles tend to jet
downward, while for ܼҧߜ ൐ ͲǤͶͶʹ, buoyancy dominates and the jet is upward.

The other main phenomenon influencing the directionality of the bubble collapse jetting is the
attraction or repulsion from a nearby rigid or soft wall due to pressure variations caused by flow
field curvature near the wall during the expansion and contraction phases of the bubble. Previous
studies have shown that the strongest jet is formed when the charge standoff is approximately
equal to the maximum free-field bubble radius, i.e., the normalized standoff distance of the
charge, ܴത, is equal to one [91]:

ܴ
ܴത ൌ ‫ͳ׽‬
‫ܣ‬ெଵ

where R is the distance from a rigid surface. The present planned experiments are expected to
shed some light on the competition between the forces of bubble attraction to floating target and
buoyancy.

2.3 Gas bubble interaction with structures

2.3.1 Far-field interactions

Chertock [93] was the first to consider explosion-induced whipping analytically, deriving
equations of motion for an immersed elastic solid in a transient incompressible flow field
generated by a pulsating, non-migrating gas bubble. In this formulation, the bubble was assumed
to be far enough away from the solid that the bubble dynamics are not affected by the solid.
Chertock presented a modal analysis solution for proportional elastic solids, i.e., simple elastic
structures for which the mode shapes of the submerged solid are the same as those of the in vacuo
solid. Chertock [94] performed experiments on a floating, slender, transversely stiffened box
structure, and a submerged ring-stiffened cylinder. A charge equivalent to 1.2 g TNT was placed
at various positions under the floating box with a minimum standoff distance of ʹ‫ܣ‬ெଵ . For the
submerged cylinder, charge sizes were 1.2, 4.2, or 8.2 g TNT equivalent, standoff distances were
between 1 and 5 ft, and the depth of the charge and cylinder target were varied between 12 and
900 ft. The minimum standoff, in terms of bubble radii, is ʹ‫ܣ‬ெଵ , corresponding to the smallest
charge at the shallowest depth. For the 8.2 g at the 900 ft, the 5 ft standoff distance is equivalent
to ͳͶǤ͹‫ܣ‬ெଵ. The measured whipping response agreed well with Chertock’s predictions for the
response in the fundamental mode. For the floating target, the amplitude of the response varied
considerably according to the placement of the charge (from directly under the
midships/centreline position to off-axis and fore/aft positions). Furthermore, the measured
whipping response was amplified by bubble pulsation. For the submerged cylinder, the response
amplitudes varied considerably due to the variation of the frequency and phasing of bubble pulses
with depth. Using (9), the bubble frequency is found to increase by a factor of 12 when charge
depth increases from 12 ft to 900 ft.

Hicks [57] extended Chertock’s analytical method to the more practical case of lumped-mass
finite-element beam models representing immersed slender structures, and presented a numerical
calculation method incorporating loading from a migrating gas bubble, as well as elastic and

DRDC-RDDC-2017-R096 11
mildly inelastic structural response. Although Hicks’s method provided much greater flexibility
in the type of structures that could be analysed, it was still limited to far-field UNDEX of
structures represented with an equivalent beam model.

A more general far-field computational approach is the Doubly Asymptotic Approximation


(DAA) method [37]. In addition to being able to determine the early-time shock response, this
method is also valid for the late-time bubble loading phase, and can be used in concert with an
analytical model of gas bubble dynamics, such as the Taylor-Herring equations or the
Geers-Hunter model. The structural model is no longer limited to an equivalent-beam
representation, but can be any finite element discretization with a fluid-structure interface model
comprised of boundary elements on the wetted surface of the structure.

Ogilvy and Birkhead [95] performed a whipping analysis using the DAA method on 1/25 scale
model of a trimaran and compared the results with experiments. They applied the same method to
a full-scale trimaran design. Zhang, et al. [96] presented a methodology for combining far-field
whipping loads with the wave-induced bending loads, and demonstrated the methodology for a
naval surface ship design. Zhang and Zong [97] presented a whipping assessment methodology
using an equivalent beam model for the vessel, taking into account elastic and plastic
deformations of the hull girder.

Wang, et al. [98] and Zhang, et al. [99] used a structural-acoustic coupling method in combination
with analytical bubble models to study the response of a ship-like target to a close proximity gas
bubble. In this approach, as in the DAA method, the bubble influences the target but the target
does not influence the bubble dynamics.

2.3.2 Close proximity interactions

Gas bubble dynamics are influenced considerably if in close proximity to a target structure, due to
the presence of the unmoving target and possibly also due to the target motion and deformation.
The changes to the gas bubble dynamics in turn alter the loading experienced by the target.
Relatively few published studies have considered the combined effect of shock and bubble
collapse in close proximity to ship-like targets. This is likely in part due to the sensitivity of the
subject, and there may be much work not published in the open literature.

Experimental studies of explosion gas bubble interactions with fully submerged rigid or responding
targets were studied by several authors [100][104]. Bubble interactions with simple floating
structures were investigated experimentally by Imakita and Yasuda [105] and Lee, et al. [10].

Relatively few studies have considered close proximity bubble interactions with floating ship-like
targets. In the experiments by Chertock [94], the standoff distances were sufficient large that the
gas bubble was not influenced by the target. Wang, et al. [98] performed an experiment to
measure the close-in UNDEX response of a floating ship-like target in which the target response
is measured with accelerometers and strain gauges at selected points in the structure. The standoff
distances ranged from 1.1 to 3.75 times the maximum bubble radius, and therefore water jetting
was not a factor.

Chen, et al. [106] performed experiments on a floating ship-like target involving standoff
distances in the range of 0.8 to 1.2 times the maximum bubble radius. Pressure transducers are

12 DRDC-RDDC-2017-R096
used to measure loads at two locations on the bottom of the target, and accelerometers and strain
gauges are used to measure the target response. The measured response up to and slightly beyond
the first bubble collapse is provided. Experiments were also performed with a rubber coating on
the hull exterior to determine its effectiveness in mitigating UNDEX damage.

Zhang, et al. [99] performed an experiment involving a slender floating box girder structure
subjected to close proximity UNDEX from 55 g of TNT at a standoff distance of one-third of the
bubble radius, but this study focuses on the shock damage to the bottom plating and the
subsequent plastic deformation of the hull girder as it hinges at the centre. Gas bubble loading
was not investigated.

Several investigators used Eulerian CFD codes to predict bubble interactions with submerged
rigid targets [100],[107][111] and submerged responding targets [104],[112][114]. Imakita and
Yasuda [104], Kan, et al. [115] and Smith and Lee [116] used these tools to investigate bubble
interactions with responding targets floating on the surface. The multinational investigation of the
ROKS Cheonan sinking performed detailed simulations of close proximity UNDEX in an effort
to determine how the ship had been attacked [117].

Two dimensional and three dimensional BEM/FEM coupling methods were used to analyse the
interaction between bubbles and responding structures [101],[118]–[120]. Zhang, et al. [121]
applied the method to study the effect of a collapsing gas bubble on a surface ship. While coupled
BEM/FEM techniques have made considerable progress in modelling bubble collapse close to
structures, including water jet formation, they are unable to assess the combined effects of shock
and bubble collapse together given the assumption of incompressibility.

2.4 Surface effects


The proximity of an underwater explosion to the surface also leads to bulk cavitation, spray,
surface waves, and a vertical plume which may also play a role in the loading and structural
response. Although these effects are often considered secondary to the main shock and bubble
loading, their role in close proximity UNDEX damage is poorly understood, and should not be
ignored given that certain effects such as cavitation cutoff can reduce the loading impulse
significantly.

Cavitation results from the inability of water to sustain tension and can be observed as the
appearance of a uniform cloud of microscopic bubbles. Shock-induced cavitation, also known as
bulk cavitation, occurs when a rarefaction wave from the reflection off the water surface reduces
the local absolute pressure below the vapour pressure (ܲ௩ ), i.e.:

ܲ ൅ ܲஶ ൏ ܲ௩ (16)

where ܲ ൅ ܲஶ denotes the total (dynamic plus hydrostatic) pressure. The resulting cloud of
cavitated water does not sustain any pressure and abruptly drops the local pressure when it
interacts with a blast wave, resulting in a pressure “cutoff.” The extent and shape of the cavitation
cloud has been examined by Costanzo [122] who used the simple ray-tracing method of Arons

DRDC-RDDC-2017-R096 13
[123] to obtain the local pressure and applied the cavitation condition (16). He also derived an
expression to describe the pressure magnitude (ܲ௖ ) resulting from closure of the cavitation cloud:
ߩܿ
ܲ௖ ൌ ሺ‫ ݒ‬െ ‫ݒ‬௎ ሻ
ʹ ௅

where ‫ݒ‬௎ is the vertical velocity of the surface water layer above the cloud and ‫ݒ‬௅ is the vertical
velocity of the water at the lower cavitation boundary below the cloud. More sophisticated
models for the onset of cavitation incorporating a description of bubble nucleation sites have
recently been used [124], but the practical benefit of these methods is inconclusive.

A final surface effect of note is the vertical jet formed at the surface above a shallow underwater
explosion. This phenomenon is sometimes called a plume, or a sultan in the Russian literature,
and appears to form from a combination of the amplification of surface waves generated by the
explosion bubble, and surface spray generated from the shock-accelerated water at the surface.
Semi-empirical expressions have been developed to describe this phenomenon such as that of
Kedrinskii [125]:

ͳ͸Ǥͷ
‫ݒ‬௝௘௧ ؄ m/s
ܼҧଵǤହ

where ‫ݒ‬௝௘௧ is the jet velocity at the surface, and ܼҧ is the charge depth normalized with the
maximum bubble radius. It is not clear what role the plume plays in ship damage from close
proximity UNDEX, but plumes of water and spray have been observed in UNDEX testing of
decommissioned ships such as the HMAS TORRENS in 1999 [126].

14 DRDC-RDDC-2017-R096
3 Design approach

The objective of the proposed series of experiments is to reproduce and measure the main global
and local response effects of a surface ship subjected to a close proximity underwater explosion.
These are as follows:

1. Lifting and subsequent whipping of the hull due to the initial shock impulse

2. Lifting of the hull due to buoyancy action of the gas bubble

3. Additional whipping of the hull due to impulsive loading from bubble collapse

4. Local deformation of the hull plating due to interaction with the shock wave

5. Local damage to the hull due to localized bubble collapse loads and water jetting

Since many experimental studies focusing exclusively on local effects have already been
performed through studies on subscale ship panels [5]–[10],[30], the first step was to design a
model to capture only the global elastic whipping response without any local effects. This model
would capture the effects 1–3 listed above. The next step was to design a model that included
some structural features to capture some local effects as well. Some of these features would
include stiffeners and panels, and would capture effects 4–5 as well as the interaction between the
global and local effects.

The design of such an experiment is a challenging task, as it involves a number of competing


design objectives and constraints. First, the tests are to be performed in the Suffield UNDEX
pond, which limits the size of the test specimen and the charge sizes that can be used. A
reasonable scale factor for the experiments is 1/40. The tests will be performed in freshwater with
a density of 1000 kg/m3.

Second, for realism, it is desirable that the experimental specimen be as ship-like as possible.
However, it is very difficult construct a test specimen in the same manner as a ship because of the
complexity (and therefore high cost) of ship construction; and because neither plate and stiffener
thicknesses, nor the welding techniques needed to assemble them, can be scaled down by as much
as a factor of 40. It is therefore necessary to simplify the design and construction while at the
same time preserving the essential ship-like characteristics of the models.

3.1 Design criteria and objectives


To better define this requirement, the design criteria in Table 2 were defined to help with the
selection of appropriate design configurations.

DRDC-RDDC-2017-R096 15
Table 2: Design criteria for configuration selection.

No. Criterion
C1 Target model with two perpendicular planes of symmetry to ease the task of simulation

C2 Global response effects measurable using digital image correlation (DIC)

C3 Local response effects measurable using DIC, if possible

C4 Fabrication of test specimens achievable at low cost

Once a configuration has been selected, the various parameters defining the design are to be
determined. To try to do this in a rational manner, a set of the design objectives to be met were
defined and these are listed in Table 3. The intention of these design objectives is to be able to
reproduce, at experimental scale, an UNDEX/target interaction comparable to what would occur
in a full-scale scenario. Setting the maximum gas bubble diameter (ʹ‫ܣ‬ெଵ ሻ to be the breadth of the
target (‫ܤ‬ሻ was based on a full-scale analogue of a 300 kg explosive charge, detonated at a depth
of 14 m. Depending on the type of explosive, the maximum bubble radius is between 8 and 9 m,
which is approximately the half-breadth of modern frigate/destroyer type vessels.

Table 3: Design objectives for parameter definition.

No. Objective
D1 Draft/breadth ratio similar to a naval surface combatant (approximately 1/3)

D2 Length/breadth ratio similar to a naval surface combatant (between 7 and 10)

D3 Diameter of gas bubble at maximum expansion approximately equal to the breadth of the
hull (ʹ‫ܣ‬ெଵ ؆ ‫)ܤ‬

D4 Frequency of the fundamental mode equal to the frequency of the first bubble cycle

D5 Fundamental natural mode is the 2-node vertical (2NV) bending mode

3.2 Design parameters


The number of design parameters available to satisfy the design objectives in Table 3 depends on
the choice of configuration. As an example of one configuration, consider the open rectangular
box configuration, with a single longitudinal stiffener at the keel shown in Figure 2. With the
simplifying assumptions of a uniform hull thickness, h, and that material and sectional properties
are uniform along the length of the hull, the complete set of available design parameters is listed
in Table 4. All of these can be varied independently except for the fluid added mass per unit
length (mf), which depends on the overall length, breadth, draft, density of the fluid, and shape of
the hull. It is included here as a parameter because an initial estimate must be provided before the
design can be attempted. Similar lists of design parameters are developed for all of the
configurations considered in this study.

16 DRDC-RDDC-2017-R096
Figure 2: Open rectangular box configuration.

Table 4: Design parameters for an open rectangular box configuration.


Symbol Parameter Symbol Parameter
W Charge mass ds Keel depth
E, U, QVy Material properties (elastic modulus, hs Keel thickness
density, Poisson’s ratio, yield stress)
h Hull thickness nb Number of transverse bulkheads
L Overall length hb Bulkhead thickness
B Section breadth P Non-structural mass per unit length
D Section depth mf Fluid-added mass per unit length

The experimental configuration to be simulated is shown in Figure 3, in which a 1.1 g RDX


(cyclotrimethylenetrinitramine) charge, which is the smallest charge that is practical to use, is
positioned below the centre of the model, at a standoff distance of 1.0‫ܣ‬ெଵ , where ‫ܣ‬ெଵ is the
maximum radius determined from (8). The empirical constant ‫ ଺ܭ‬in (8) is 3.74 for RDX in SI
units [72]. With the depth of the charge in the experiment initially estimated to be 0.3 m, the
maximum radius is determined to be 0.177 m.

DRDC-RDDC-2017-R096 17
Figure 3: Experimental configuration.

The bubble period is estimated from either (9) or (13), depending on whether a surface correction
is needed. Using ‫ܭ‬ହ = 2.232 for RDX in SI units, the bubble period is 28.96 ms and 32.84 ms,
respectively with and without the surface correction.

As the influence of the target on the bubble period is unknown, a bubble period of 31 ms (midway
between the two theoretical estimates) is assumed for design calculations. The maximum bubble
radius will be approximated as 0.18 m.

3.3 Design steps


The overall process for determining a design is illustrated in Figure 4, with the individual steps
described in the following sections. This design process illustrated is open-ended and does not
result in a unique set of design parameters for a given set of initial assumptions.

In the Initialization step, the hull form and general structural arrangement is selected. This could
be the open rectangular box configuration of Figure 2, or it could be a different configuration.
The configurations considered are discussed in Section 4. The charge size, standoff and depth are
also selected in this step. This fixes the radius and period of the gas bubble.

18 DRDC-RDDC-2017-R096
In the Design Target Values step, several of the design parameters are fixed based on the
explosion charge parameters: the gas bubble diameter is used to determine the model breadth
(‫ ܤ‬ൌ ʹ‫ܣ‬ெଵ = 0.36 m); and the gas bubble period (31 ms) determines the model’s fundamental
natural frequency (f1 =1/0.031 = 32.3 Hz). The target draft for the model is selected to be one
third of the model breadth (d = B/3). An initial L/B ratio for the model is then selected (value
between 7 and 10 is preferred).

Figure 4: Design process.

In the Parameter Iteration step, the remaining parameters of the model are determined so as to
meet the target values for the design. In varying plate thicknesses and beam dimensions, standard
sizes for which stock material were preferred.

For the Design Verification stage, a finite element model is created using the selected design
parameters, including any non-structural mass. A static balance is performed to verify the draft of
the model. Then the natural frequencies of the model are calculated to verify that the fundamental
mode is still sufficiently close to the target value, and that the fundamental mode is the two-node
vertical (2NV) hull girder mode.

If the results of the Design Verification stage are not satisfactory, the Parameter Iteration step
may be repeated using the updated values for the fluid added mass. It may also be necessary to
modify the configuration of the model, e.g., additional stiffeners to eliminate low frequency local

DRDC-RDDC-2017-R096 19
modes of vibration, or to remove low frequency torsional vibration modes. This may introduce
additional design parameters that have to be taken into account in subsequent design iterations.

Once a satisfactory design has been determined, Coupled Eulerian-Lagrangian (CEL) simulations
of the close proximity UNDEX response of the model design are performed (Design Evaluation
step). These are normally performed using one-quarter models of the structure (FE model) and the
fluid domain, provided two planes of symmetry exist in the design and that the charge is situated
directly below the centre of the model. These simulations are performed to verify that the
expected combination of loading and response processes occur in the experiment, and to
determine the deformation and damage likely to occur.

Section 4 of this report provides a summary of the design calculations performed using the above
process and Section 5 gives the details of the modelling methods used and the results obtained of
the simulations of the proposed preliminary experiment.

3.4 Estimating the natural frequency and fluid added mass


In general, natural frequencies of real ships are evaluated using finite element methods. However,
for simplified ship-like targets used in UNDEX experiments, the natural frequencies and mode
shapes can be estimated for design calculations using analytical formulas for beam vibration with
uniform elastic and inertial properties. For a free-free uniform beam, natural mode shapes in
vacuo are given by [127]:

•‹ ݇௡ ‫ ܮ‬െ •‹Š ݇௡ ‫ܮ‬


‫ݓ‬௡ ሺ‫ݔ‬ሻ ൌ ሺ•‹Š ݇௡ ‫ ݔ‬൅ •‹ ݇௡ ‫ݔ‬ሻ ൅ ሺ…‘•Š ݇௡ ‫ ݔ‬൅ …‘• ݇௡ ‫ݔ‬ሻ (17)
…‘•Š ݇௡ ‫ ܮ‬െ …‘• ݇௡ ‫ܮ‬

where L is the beam length, x is the axial coordinate, values for the wave number kn are
knL = 4.7300, 7.8532, 10.9956 for the first, second and third modes, respectively.

If we may assume that these mode shapes are unchanged when the beam is immersed in fluid,
then the natural frequencies, Zn, associated with the mode shapes are given by:

‫ܫܧ‬
߱௡ଶ ൌ ݇ସ (18)
݉௟ ൅ ݉௙ ௡

where ݉௟ is the mass per unit length of the beam structure, ݉௙ is the fluid added mass per unit
length, E is the modulus of elasticity, and I is the second moment of area for the beam section.
The above solutions are derived from the Euler-Bernoulli beam formula, and therefore neglects
shear rotations and shear lag effects which can make an important contribution to the higher hull
girder modes [57]. However, they are expected to be adequate as an estimate of the fundamental
vibration mode for uniform beams in still water.

The term ݉௙ may be estimated using strip theory, which assumes that the fluid flow about a
section of the hull is essentially two dimensional, which is valid for an infinitely long hull.

20 DRDC-RDDC-2017-R096
Lewis [128] used this approach to develop “virtual” mass coefficients for ships of finite length in
which the added mass per unit length is given by:

ͳ
݉௙ ൌ ‫ܤߩߨܬܥ‬ଶ (19)
ʹ

where ‫ ܥ‬is coefficient depending on the shape of the 2D hull sections, ‫ ܬ‬is correction factor for
the finite length of the hull, ሺͳȀʹሻߨߩ‫ܤ‬ଶ is the mass of the water displaced by a semi-circular
section of radius ‫ܤ‬. Lewis provided diagrams of various hull shapes and associated ‫ ܥ‬values, as
well as ‫ ܬ‬values for a range of ‫ܮ‬Ȁ‫ ܤ‬ratios.

For finite element calculations, the fluid added mass is determined using a boundary element
method potential flow calculation in the ADMAS5 module of VAST [129]. This is based on the
formulation of Vernon, et al. [130].

DRDC-RDDC-2017-R096 21
4 Candidate designs

4.1 Open girder designs


Some initial design configurations were needed to begin the design process, and the first two that
were tried were the open rectangular box girder of Figure 2, and the open semi-cylindrical girder
of Figure 5. Both of these configurations satisfy the design criteria set out in Table 2.

The design steps described in Section 3 were performed, resulting in the design parameters shown
in Table 5 for the open rectangular girder. Finite element analysis of this design gave the wet
natural vibration modes shown in Figure 6.

Figure 5: Open semi-cylindrical configuration.

Table 5: Parameter values for the open rectangular girder design.

Parameter Value Parameter Value


W 1.1 g (RDX) ds 20 mm
Material Aluminum hs 5 mm
h 5 mm nb 12
L 3.6 m hb 5 mm
B 0.36 m P 28.61 kg/m
D 0.226 m mf 76.9 kg/m

Figure 6 shows that the fundamental mode is not the 2NV mode and therefore this design
objective is not met. The open girder design is torsionally very flexible, and therefore not very
representative of a surface combatant. To correct this, two variants of this configuration were
attempted, as shown in Figure 7. One has cross bracing to control the shear deformation in the
openings, while the other has a coaming around the deck openings. The finite element analysis

22 DRDC-RDDC-2017-R096
results in Figure 8 and Figure 9 show that with these modifications, the 2NV mode is the
fundamental mode. The penalty for this is that the structures are more complex to build, and the
cross bracing and coamings will interfere with the DIC imaging of the bottom plate.

Figure 6: Natural modes of the open rectangular box girder configuration.

Figure 7: Two alternative open girder designs: braced (left); and stiffened with coaming (right).

DRDC-RDDC-2017-R096 23
Figure 8: Two lowest natural modes for the braced open girder design.

Figure 9: Two lowest natural modes for the open girder design with coaming.

The design parameters determined for the open semi-cylindrical design are given in Table 6. This
was also an aluminum design with a shell plate thickness of 4 mm. The calculated draft for the
model was 0.119 m.

Table 6: Parameter values for the open semi-cylindrical design.

Parameter Value Parameter Value


W 1.1 g (RDX) D 0.230 m
Material Aluminum nb 8
h 4 mm hb 4 mm
L 3.5 m P 20.86 kg/m
B 0.36 m mf 39.66 kg/m

24 DRDC-RDDC-2017-R096
Finite element analysis of this design gave the wet vibration modes shown in Figure 10. These
show a similar excess of torsional flexibility as was found for the open rectangular box design.
While it might be possible to correct for this with bracing or flanges, the same issues remain
regarding their interference with the DIC imaging. Furthermore, the semi-cylindrical bottom of
this design is not considered to be ideal for the DIC system, and there is a risk that the response
may not be accurately measured. This design is therefore not pursued further.

Figure 10: Lowest natural modes for the open semi-cylindrical design.

4.2 Closed girder designs


The next set of configurations attempted were closed girder designs of Figure 11 and Figure 12.
Both of these designs have longitudinal stiffeners running down the centre of the deck. The
rectangular box girder also has a keel member. This is not needed for the cylindrical hull as the
curvature of the hull provides sufficient flexural stiffness. The rectangular girder has ten internal
bulkheads, whereas as the cylindrical hull only has six. This is again because the cylindrical hull
requires less transverse support to prevent lateral distortions.

Both of these configurations satisfy criteria C1 and C2 in Table 2. However, because only the
upper surface (deck) of the models can be imaged with DIC, the local response of the bottom
plating will be inaccessible to DIC imaging. Therefore criterion C3 is not satisfied. Although
these configurations would be difficult to fabricate using only welded connections, it was thought
that by attaching transverse bulkheads to the plating using machine screws, the models could be
adequately assembled while still welding all of the longitudinal connections. Therefore the
buildability criterion C4 could be satisfied. Although these models are not well suited for
measuring the local response, they could still produce an effective ship-like design for measuring
the global response.

DRDC-RDDC-2017-R096 25
Figure 11: Closed rectangular box configuration: with deck removed (left);
meshed FE model (centre); section view (right).

Figure 12: Closed cylindrical girder configuration: with deck removed (left);
meshed FE model (centre); section view (right).

The design parameters determined are shown in Table 7 for both designs, and the natural modes
for these designs are shown in Figure 13 and Figure 14. Overall, these designs can be made to
satisfy the design objectives in Table 3 in a satisfactory manner: with the indicated additional
mass per unit length, a draft of 0.12 m can be achieved; the length to breadth requirement is
achieved; and the fundamental vibration mode is the 2NV mode. Apart from the C3 criterion,
these designs appear to be viable.

26 DRDC-RDDC-2017-R096
Table 7: Parameter values for the closed rectangular girder design.

Parameter Closed rectangular girder Closed cylindrical girder


W 1.1 g (RDX) 1.1 g (RDX)
Material Steel Steel
h 0.912 mm 0.912 mm
L 3.6 m 3.6 m
B 0.36 m 0.36 m
D 0.2 m 0.19 m
ds 10 mm 10 mm
hs 0.912 mm 0.912 mm
nb 12 8
hb 10 mm 10 mm
P 14.47 kg/m 13.36 kg/m
mf 75.52 kg/m 39.7 kg/m

Figure 13: Closed rectangular box girder natural vibration modes.

DRDC-RDDC-2017-R096 27
Figure 14: Closed cylindrical box girder natural vibration modes.

4.3 Segmented model


The proposed configuration in Figure 15 is similar to those used in hydroelastic model testing.
The central beam (in this case a hollow rectangular section) is the only longitudinally continuous
structure and is therefore entirely responsible for the global structural stiffness. The
semi-cylindrical sections determine the hydromechanical behavior of the model, and contribute to
the global mass distribution. The segments are separated, and therefore have no longitudinal
continuity. Each segment on its own will respond locally to shock and bubble loads, and segments
could be designed to deform and fail in realistic manner. However the local deformation
processes cannot influence the global response, as it can with the other girder designs discussed
above.

Figure 15: Segmented model configuration.

DIC could be used to measure the global response of the segmented configuration by imaging the
upper surface of the continuous central beam. Measuring the local response will be more difficult
because of the curved hull surface and the obstruction of the central beam. Therefore, criterion C3
cannot easily be met. However, as the other criteria are met, it was decided to proceed with the
design.

28 DRDC-RDDC-2017-R096
The design values determined are given in Table 8, and a one-quarter finite element model is
shown in Figure 16. In the latter it can be seen how the central hull segment can be designed so as
to allow for a more realistic local response effects. The values of D and h used elsewhere in the
model can be selected somewhat arbitrarily since they do not affect the global stiffness of the
structure. The total mass of the structure can be adjusted subsequent to selection of the other
design variables using the non-structural mass parameter, μ. The two lowest wet vibration modes,
shown in Figure 17, satisfy design objectives D4 and D5.

Table 8: Parameter values for the segmented configuration.

Parameter Value Parameter Value


W 1.1 g (RDX) dg 0.1 m
Material Steel beam, Aluminum shell bg 0.12 m
h 3 mm hg 3 mm
L 3.24 m ns 11
B 0.36 m hb 3 mm
D 0.24 m P 9.88 kg/m
d 0.12 m mf 36.9 kg/m

Figure 16: One-quarter finite element model of the segmented design.

DRDC-RDDC-2017-R096 29
Figure 17: Two lowest wet natural frequencies for the segmented model.

4.4 Barge configuration


Given the complexities of the designs given above, it was of interest to produce a simpler
configuration for preliminary studies. For this reason, the “barge” configuration shown in Figure 18
was considered. This configuration is structurally simple, allows for DIC imaging of the global
and local response, and has two planes of symmetry. Therefore all of the design criteria are
satisfied. The configuration is not very ship-like, however, in that a realistic draft/breadth ratio
will not be achievable (objective D1).

The design process was carried out, resulting in the design parameters shown in Table 9. Note
that the length/breadth ratio is 5.56, which is somewhat lower the minimal value of 7. This was
done so as to ease the handling of the model during the experiments, and to accommodate a
standard size of aluminum hollow section for the two gunwales. The draft is determined to be
25 mm with no additional weight added to the model. This gives a draft/breadth ratio of 0.0694.
The two lowest vibration modes are shown in Figure 19 and indicate that the design objectives
D4 and D5 are satisfied.

Figure 18: Barge configuration.

30 DRDC-RDDC-2017-R096
Table 9: Parameter values for the barge design.

Parameter Value Parameter Value


W 1.1 g (RDX) dg 50 mm
Material 6061 aluminum bg 40 mm
hp 6 mm hg 3 mm
L 2.0 m nb 2
B 0.36 m hb 3 mm
D 56 mm P -
d 25 mm mf 54.8 kg/m

Figure 19: Two lowest wet natural frequencies for the barge model.

DRDC-RDDC-2017-R096 31
5 UNDEX simulation results

Design evaluation simulations for four of the designs proposed in Section 4 were conducted to
determine if the desired response characteristics are produced at model scale. The general
configuration for the simulations is shown in Figure 20, and is the same for all four simulations,
except for differences in the target geometries and drafts.

Figure 20: Simulation configuration.

5.1 Simulation method

5.1.1 Fluid models

A two-stage simulation procedure was used as described in Smith and Lee [116]. The first stage is
a standalone analysis using the Eulerian CFD code Chinook [131] of the early-time shock
propagation and initial bubble expansion phase, and is performed using a two dimensional
axisymmetric model of the fluid domain. The second stage is a coupled analysis involving
parallel execution of Chinook and the structural finite element analysis code LS-Dyna [132]. This
stage is performed using three dimensional quarter symmetry models of the fluid domain and the
target structure. The transition between the first and second stages is effected just before the
shock wave arrives at the target, at which time the 3D fluid grid is initialized using the field
variables from the 2D grid using Chinook’s native mapping capability. In addition, the grids are
initialized to the hydrostatic pressure distribution, and a gravity load is applied to the structural
models.

The 2D axisymmetric grid used for the Stage 1 analysis is shown in Figure 21. The 2D grid has a
cell size of 0.5 mm within the footprint of the charge, increasing to 1 mm at a radius of 0.25 m
from the charge. At the time of mapping (0.095 ms), the shock front has travelled a distance
slightly less than the standoff distance of 0.179 m, and therefore remains in the refined region of
the grid during the Stage 1 analysis. The 1.1 g RDX charge is modelled as a small cylinder of
detonated explosion products with the same dimensions as the undetonated charge. Because of
the similarity in the configurations, the same Stage 1 simulation is used in all cases.

The shock pressures predicted with this 2D axisymmetric model are shown in Figure 22 for two
standoff distances: 0.0895 m (0.5R) and 0.179 m (1.0R). The pressure time-histories are compared

32 DRDC-RDDC-2017-R096
for two other methods as well: the similitude Equations (1)–(3), and a 1D Chinook analysis using a
fluid grid with 0.025 mm cell size (grid 4d in [72]). The shocks arrive 0.01–0.02 ms earlier in the
Chinook simulations due to nonlinearities arising from the large amplitude of the shock at early
time, causing the propagation speed to be initially greater than the sound speed; by contrast, a
uniform sound speed is assumed for the similitude equation.

Figure 21: 2D axisymmetric model of the fluid domain used for the Stage 1 analysis (right);
with detail showing charge geometry (left).

Figure 22: Comparison of shock front time histories at two standoff distances.

DRDC-RDDC-2017-R096 33
Another difference is the height and shape of the shocks. The numerical dispersion in the
Chinook simulations results in the non-vertical shock front, the blunting of the shock peak, and
the spreading of the wave along the time axis. This is especially noticeable in the 2D simulations.
The peak pressure at 1.0R in the 2D model is slightly less than 50% of the peak of the similitude
equation. However, the impulse carried in the shock (equal to the area under the curve) is actually
greater in the Chinook analysis due to the long spreading of the tail. The 1D simulation shows a
somewhat higher peak pressure and a more vertical shock front but is otherwise similar to the 2D
simulation. The 2D simulation was selected for the initial shock propagation phase because
(a) the shape of explosive charge can be more realistically modelled, and (b) a previous
benchmark study [72] showed that more accurate gas bubble dynamics is achieved in the 3D
stage when the initial stage is performed in 2D rather than in 1D.

The 3D fluid grid used for all Stage 2 simulations is shown in Figure 23. The dimensions of the
gridded region is 24u24u9.5 m. While the dimensions are roughly similar to those of the Suffield
test pond, it is not necessary to accurately represent the pond geometry in the fluid model. Instead
it is only necessary that a sufficiently large fluid domain be modelled so that reflections from the
boundaries have no effect on the loading and response of the target, as is the case with real
experiments in the pond. The 2D and 3D unstructured grids were all created using Pointwise
V17.3R2 [133]. The 3D grid shown in Figure 23 has 496,020 grid points and 2,897,686 cells and
a cell size that varies from a minimum of 4 mm near the detonation point up to a maximum of
2 m at the extreme boundaries of the domain. Smith [72] determined that this grid cell size is
adequate for 3D simulations of explosion gas bubbles for 1.1 g charges.

Figure 23: 3D quarter symmetric fluid domain grids used for the Stage 2 simulations:
full model (top); two views of refined region (bottom).

At the time of mapping, the 3D model is initialized with the field variables of the 2D
axisymmetric model, including the water and air volumes that fill the grid. At the same time,

34 DRDC-RDDC-2017-R096
additional air volumes are introduced to account for the displacement of the floating targets. The
Tait equation of state (EOS) [134] is used for the water, the Jones-Wilkins-Lee (JWL) EOS
[135]–[137] is used for the RDX explosive, and an ideal gas EOS is used for the air. The
parameter values for the EOS models are listed in Table 10.

Table 10: Equation of state parameters.

Tait equation (water) JWL (RDX explosive) Ideal gas (air)


ρ0 1000.0 kg/m3 ρ 1659 kg/m3 ρ0 1.177 kg/m3
P0 101,325 Pa A 495.1 GPa γ 1.4
B 3000 atm B 7.21 GPa E0 215 kJ/kg
n 7.14 C 1.62 GPa
m 5.00 R1 4.387
αc 0.05 R2 0.9954
E0 354 kJ/kg ω 0.3469
E0 5877.9 kJ/kg

5.1.2 Structural models

Finite element meshes of the target structures were created with SubSAS (Trident
Modeller) [138] and subsequently translated to LS-Dyna. One-quarter of the structure is modelled
with symmetry constraints applied to nodes in the two symmetry planes. For the steel models,
A569 steel (mild steel) is assumed. The material properties used are listed in Table 11 and are
applied to the *MAT_PLASTIC_KINEMATIC material model in LS-Dyna in which high strain
rate effects are introduced via the Cowper-Symonds coefficients. By this material model, the
dynamic yield stress is obtained with the formula:

ߝሶ ଵȀ௣
ߪ௬ ൌ  ߪ௬0 ቆͳ ൅ ൬ ൰ ቇ (20)
‫ܥ‬

where ߪ௬Ͳ is the static yield stress of the material, and C and p are the Cowper-Symonds
coefficients.

For the aluminium models, 6061 aluminum is used with the material properties given in
Table 11. The Johnson Cook parameters were taken from [139] and input to the
*MAT_SIMPLIFIED_-JOHNSON_COOK model in LS-Dyna in which the yield stress of the
material is determined using the formula:

ߪ› ൌ ൫‫ ܣ‬൅ ‫߳ܤ‬௣ ௡ ൯ሺͳ ൅ ‫߳ Ž ܥ‬ሶ ‫ כ‬ሻ (21)

Where A, B, C and n and empirical derived constants, ߳௣ is the effective plastic strain, and ߳ሶ ‫ כ‬is
the normalized effective strain rate.

DRDC-RDDC-2017-R096 35
Table 11: Material properties.

A569 Steel 6061 Aluminum


ρ 7850 kg/m3 ρ 2700 kg/m3
E 207 GPa E 68.9 GPa
ν 0.3 ν 0.33
σy 425 MPa Johnson-Cook parameters
Et 2.0 GPa A 236 MPa
Cowper-Symonds parameters B 443 MPa
C 40 n 0.376
p 5.0 C 0.024

5.1.3 Coupling method

Coupling between the fluid model and the structural model is invoked by running the LS-Dyna
executable in parallel with Chinook, and activating the *USER_LOADING keyword in LS-Dyna.
A small deformation two-way coupling scheme is used in which pressures on the fluid-structure
interface (FSI) are transferred from Chinook to LS-Dyna, and the normal velocities on the FSI are
transferred from LS-Dyna to Chinook. The transfer of data occurs at fixed coupling times (a 2 μs
coupling interval is used for all simulations). The small deformation coupling scheme is based on
the undeformed position of the target, i.e., the FSI position is not updated when determining the
pressure loads to be transferred to the target. A large deformation coupling scheme is also
available in Chinook but has been found to be unreliable [72]. The small deformation scheme was
previously found to give results that agreed well with measurements on a simple floating box
structure for the charge size and standoff distance considered here [116].

5.2 Closed rectangular girder results


The quarter-model of the closed rectangular girder target is shown in Figure 24, in which
symmetry constraints are applied to all nodes on the XZ and YZ planes. The duration of the
simulation was 76 ms. Three distinct bubble collapse events are visible in pressure time history at
the midship centerline position of the target, shown in Figure 25. After the initial shock pulse, a
second pulse at about 1.5 ms and lower peaks following it are the result of the closure of the
cavitation zone generated by the initial shock response. The peak near 28 ms coincides with the
first bubble collapse, while the much larger peak near 49 ms is due to the second bubble collapse.
The third bubble collapse produces the small peak near 68 ms.

A detailed view of the pressure time history near the first bubble collapse is shown in Figure 26.
Here it can be seen that the pressure pulse is a series of three peaks, the first occurring at
28.28 ms, and the highest occurring at 28.38 ms. Between each of the peaks are short intervals at
zero pressure, indicating cavitation resulting from the target response to the pressure pulse. The
pressure time history for the second bubble collapse is also shown in Figure 26. It shows a single
large pressure pulse at 49.57 ms, with a peak pressure about seven times higher than the first
bubble pulse.

36 DRDC-RDDC-2017-R096
Figure 24: One quarter finite element model for the close rectangular box girder.

Figure 25: Loading time histories at the midships centerline position, closed rectangular model.

DRDC-RDDC-2017-R096 37
Figure 26: Pressure time history near the first (left) and second (right)
bubble collapse: close rectangular model.

The sequence of images in Figure 27 shows the evolution of the gas bubble as predicted by
Chinook during the first bubble period. The bubble is near maximum expansion at 13 ms.
Although the standoff distance is 1.0‫ܣ‬ெଵ , the top surface of the bubble never reaches the target.
The bubble is close to its minimum at 28 ms, and during contraction has lost its spherical shape.
Also, the bubble does not at this stage appear to be collapsing on to the target surface or forming
a water jet as expected.

Figure 28 provides another view of the full expanded gas bubble. The edge of the bubble is
indistinct because of the multi-material formulation used in the Chinook. This allows a mixture of
explosion products and water to exist within the same cell, as occurs with cells near the edge of
the bubble and near the air-water interface. Using the 500 kg/m3 density contour (the average of
the densities of the expanded explosion products and water), the diameter in the vertical direction
is 0.304 m and the horizontal radius is 0.171 m.

The sequence of images in Figure 29 shows the bubble evolution during the second pulse. The
re-expanded bubble has now lost much of its spheroidal shape, and is now prolate instead of
oblate. It reaches a second minimum at close to 50 ms with the remnants of the bubble now in
contact with the target. This may explain why the pressure peak and change in impulse associated
with the second collapse is larger than that of the first collapse.

38 DRDC-RDDC-2017-R096
Figure 27: Evolution of the explosion gas bubble during the first bubble period: closed
rectangular girder model. Colours indicate density in kg/m3.

Figure 28: Gas bubble at full expansion.

DRDC-RDDC-2017-R096 39
Figure 29: Evolution of the explosion gas bubble during the second bubble period:
closed rectangular girder model. Colours indicate density in kg/m3.

The sequence of images in Figure 30 shows the remnants of the gas bubble during the third
period. The explosion products retain a prolate form, and are now more quiescent. There is less
obvious pulsing of the bubble, and the weak minimum near 68 ms is not apparent in these images.

The vertical displacement response at three locations in the structure is given in Figure 31. The
closed girder structure exhibits a weak overall whipping response with a period of 30 ms, as seen
in the response at the end location. Oscillations with a period in the range of 2.5–15 ms are visible
at the midships locations, and these are associated with local vibration of the deck and bottom
plating. A 40 mm drop in the displacement at the midships top location occurs after 18 ms,
resulting from the buckling of the deck structure. Furthermore, both midships locations show a
sharp rise in displacement of 15–20 mm starting at 28 ms as a result of the first bubble collapse.
Views of the box girder response in Figure 32 show the localized deformation in the deck and
bottom near midships, with the permanent deformation being about 10 mm in each. No
permanent hinging of the box girder is evident.

40 DRDC-RDDC-2017-R096
Figure 30: Evolution of the explosion gas bubble during the third bubble pulse: closed
rectangular girder model. Colours indicate density in kg/m3.

Figure 31: Time history response of three points on the rectangular box girder.

DRDC-RDDC-2017-R096 41
Figure 32: Displacement contour plots for the rectangular box girder.

5.3 Closed semi-cylindrical girder results


The simulation of the semi-cylindrical girder design was performed using the quarter-symmetric
finite element model shown in Figure 33, in which symmetry constraints are applied to nodes on
the XZ and YZ planes. The model is made of A516 steel with materials given in Table 11.

The duration of the simulation was 76 ms, long enough to capture the three bubble collapse
events. The pressure time history at the midships keel location, shown in Figure 34, has features
similar to those seen for the closed box girder. The first, second and third bubble collapses occur
at 28.98, 48.53 and 65.73 ms, respectively. The second bubble collapse has a higher peak pressure
than that of the first collapse, whereas the third collapse is weak. Less cavitation is evident in the
first bubble collapse pressure pulse than was seen for the closed box. Local cavitation appears to
be less of a factor with the curved bottom than it was for a structure with a flat bottom. The
detailed view of the two bubble collapse events in Figure 35 shows that one brief interval at the
cutoff pressure occurs during the first collapse.

42 DRDC-RDDC-2017-R096
Figure 33: One quarter finite element model for the semi-cylindrical girder.

Figure 34: Loading time histories at the midships keel location, semi-cylindrical model.

DRDC-RDDC-2017-R096 43
Figure 35: Pressure time history for the first (left) and second (right) bubble pulsed at the
midships keel location: semi-cylindrical model.

The sequence of images in Figure 36 shows the evolution of the bubble during its first period. It
can be seen that the bubble collapses without contacting the target. Another view of the bubble at
the time of mapping and at full expansion is shown in Figure 37, showing the horizontal radius to
be 0.173 m and the vertical diameter to be 0.324 m. Immediately after collapse, at 30 ms,
(see Figure 38), the re-expanding bubble is now in contact with the target, indicating that upward
migration occurred at the first minimum. The re-expanded bubble now remains in contact with
the target until the second minimum at 49 ms. After the second collapse, the shape of the bubble
becomes less distinct (Figure 39). The third minimum corresponding to the small peak at 65.7 ms
in Figure 34 is not readily identifiable in the bubble contours.

44 DRDC-RDDC-2017-R096
Figure 36: Evolution of the explosion gas bubble during the first bubble collapse:
semi-cylindrical girder model. Colours indicate density in kg/m3.

Figure 37: Gas bubble at 0.095 ms (left); and 13 ms (right).

DRDC-RDDC-2017-R096 45
Figure 38: Evolution of the explosion gas bubble during the second bubble period:
semi-cylindrical girder model. Colours indicate density in kg/m3.

46 DRDC-RDDC-2017-R096
Figure 39: Evolution of the explosion gas bubble during the third bubble collapse:
semi-cylindrical girder model. Colours indicate density in kg/m3.

The time histories of the vertical displacement at three locations on the target are shown in Figure 40.
The overall whipping motion of the model is only weakly evident in the displacement at the End
location. The displacements at the midships locations show that the hull bottom undergoes a
35 mm upward deformation as a result of the first bubble, while the second bubble collapse adds
a further 10 mm. The permanent deformation in the bottom appears to be greater than 20 mm.

Figure 41 shows the displacement contours of the model at six different times during the
simulation. The deformation damage in the bottom near midships is visible from 32 ms on. A
small negative permanent deformation of about 2 mm appears to have developed at the End
location, indicating some slight downward hinging of the girder.

DRDC-RDDC-2017-R096 47
Figure 40: Time history response at three locations of the semi-cylindrical girder design.

Figure 41: Semi-cylindrical girder deformation contours.

48 DRDC-RDDC-2017-R096
5.4 Segmented model results
The simulation of the segmented design was performed using the quarter-symmetric finite
element model shown in Figure 16. The central hull compartment and the rectangular box girder
are made of A516 steel, while the rest of the model is made from 6061 aluminum. Properties of
the materials are given in Table 11.

The duration of the simulation was 76 ms. The pressure time history at the midships keel location
is shown in Figure 42. The first, second and third bubble collapses occur at 29.42 ms, 47.65 ms,
and 70.87 ms, respectively. Note that the third period is longer than the second period. This is
contrary to expectation, as bubble periods normally progressively decrease due to the energy
losses that occur at each collapse. The peak pressures and impulses per unit area fall with each
successive bubble collapse. Pressure cutoff due to cavitation occurs repeatedly between the
arrival of the shock and the first bubble collapse but not after the first bubble collapse (Figure 43).
The peak shock pressure is of similar magnitude as for the other simulations.

Figure 42: Loading time histories at the midships keel for the segmented model.

DRDC-RDDC-2017-R096 49
Figure 43: Pressure time history of the first (left) and second (right) bubble collapse at the
midships keel location: segmented model.

The sequence of images in Figure 44 shows the evolution of the bubble during its first period,
which is similar to the previous cases. The bubble at the time of full expansion is shown in
Figure 45. After the first collapse, the re-expanding bubble is now in contact with the target and is
prolate (Figure 46). It remains in contact with the target through the second collapse and into the
third cycle (Figure 47). The bubble is more coherent in the second and third cycles than it is with
the two previous models, and a distinct third bubble minimum is visible in the contours.

50 DRDC-RDDC-2017-R096
Figure 44: Evolution of the explosion gas bubble during the first bubble collapse:
segmented model. Colours indicate density in kg/m3.

DRDC-RDDC-2017-R096 51
Figure 45: Gas bubble diameter at 14 ms.

Figure 46: Evolution of the explosion gas bubble during the second bubble period:
segmented model. Colours indicate density in kg/m3.

52 DRDC-RDDC-2017-R096
The time histories of the vertical displacement at three locations on the target are shown in Figure 48.
A strong whipping motion with a maximum amplitude of about 20 mm and a period of 32 ms is
visible in the displacement at the midships and end locations. Relatively little local deformation
occurs, as can be seen from the similarity in the responses at the three midships locations. A
permanent deformation of about 3 mm at the midships keel locations mostly occurs at about
48 ms, after the second bubble collapse. The first bubble collapse also produces a small amount
of local deformation at this location. A small higher frequency oscillation with amplitude of about
1 mm is visible at the midships girder location. This is due to local vibration in the central box
girder.

Figure 49 shows the displacement contours of the model at six different times during the
simulation, which confirms that the motion is predominantly hull girder whipping. Additional
local deformation and damage could be produced by redesigning the midships segment to have
less stiffness (fewer or smaller stiffeners) or by using a weaker material.

Figure 47: Evolution of the explosion gas bubble during the third bubble collapse:
segmented model. Colours indicate density in kg/m3.

DRDC-RDDC-2017-R096 53
Figure 48: Time history response at three locations of the segmented model design.

Figure 49: Segmented model deformation contours.

54 DRDC-RDDC-2017-R096
5.5 Barge model results
A simulation of the barge response was conducted using the quarter symmetric model shown in
Figure 50. The draft of the model is 2.5 cm, instead of the 12 cm used for the other models.
Otherwise, the fluid grid and other model parameters are identical to those used for the other
models. The barge structure material is 6061 aluminum with the properties given in Table 11.

The duration of the simulation was 76 ms. In the pressure time history shown (Figure 51), the
incident shock pressure, cavitation closure pressure and first bubble collapse are clearly apparent.
The first bubble collapse occurs at 26.76 ms. The second bubble pulse occurs at 46.2 ms and is
quite weak.

Figure 50: Structural FE model for the Barge design.

Figure 51: Loading time histories at the midships keel position of the barge model.

DRDC-RDDC-2017-R096 55
The sequence of images showing the evolution of the explosion gas bubble is given in Figure 52
through Figure 54. These show three distinct bubble collapse events, the second and third
occurring at approximately 46 ms and 65 ms, respectively. The latter is not identifiable in the
pressure time history in Figure 51. The bubble behaviour is different from the other models in that
the gas bubble migrates and jets downward during all three bubble collapses. This accounts for
why the bubble pressures and impulses at the target in Figure 51 are smaller than those seen in
previous models.

Figure 55 shows the bubble at close to full expansion, from which the vertical diameter of the
bubble was determined to be 0.311 m and the horizontal radius is 0.172 m.

Figure 52: Evolution of the explosion gas bubble during the first bubble collapse:
barge model. Colours indicate density in kg/m3.

56 DRDC-RDDC-2017-R096
Figure 53: Evolution of the explosion gas bubble during the second bubble collapse: barge
model. Colours indicate density in kg/m3.

DRDC-RDDC-2017-R096 57
Figure 54: Evolution of the explosion gas bubble during the third bubble collapse:
barge model. Colours indicate density in kg/m3.

58 DRDC-RDDC-2017-R096
Figure 55: Gas bubble at maximum expansion.

The time history response of four locations on the barge model is shown in Figure 56 and shows a
strong hull girder whipping motion. These clearly show two oscillations in the fundamental mode
with a period of approximately 28 ms. The amplitude of oscillation at the end location is slightly
greater than 30 mm. Relatively little local deformation occurs at midships, as is indicated by the
small difference between lines A and B. In addition to the whipping motion, the model as a whole
undergoes a downward drift. This is likely due to the applied gravity load not being well balanced
with the hydrostatic pressure for this model. The deformation of the barge at six different times is
shown in Figure 57. This further confirms that a global whipping motion is the predominant
motion with little local deformation occurring. No permanent deformation of the model was
observed in the simulation.

DRDC-RDDC-2017-R096 59
Figure 56: Vertical displacement time histories at several locations on the barge.

Figure 57: Displacement contours (in m) for the barge model.

60 DRDC-RDDC-2017-R096
5.6 Discussion
The simulations of the four candidate designs show that the segmented model and the barge
model exhibit a much stronger hull-girder whipping response than the two closed girder models,
whereas the latter models exhibit much more local deformation and damage than the former. The
thin steel plating used in the two closed girder models appears to promote local response
behaviour at the expensive of global whipping response. This is especially noticeable in the
closed rectangular box model, in spite of the stiffening introduced in the deck and bottom.
Notwithstanding these differences, the maximum target displacements (neglecting rigid body drift
of the models) were comparable for all four models, the highest being approximately 40 mm for
the closed rectangular model and the lowest being 25 mm for the segmented model).

None of the models exhibited failure of the hull girder (i.e., breaking the back of the hull),
although the semi-cylindrical girder showed slight hinging at the mid-plane. Hull girder failures
evidently require a larger charge size.

The simulations were all performed with identical fluid grids, the same charge size and standoff
distance, and with all other modelling parameters being the same. Nonetheless significant
differences in the loading time history (pressures and impulse/area) are predicted for the four
models. At the midships keel location (directly above the detonation), the loading is greatly
influenced by the flexibility of target structure, as seen by the large difference in the peak
pressures and changes in impulse associated with each bubble pulse. These are summarized in
Table 12. The change in impulse is estimated as the maximum change in the impulse/area over a
time period starting immediately before one event and ending immediately before the next event.

Table 12: Summary of predicted peak pressures and impulses for the shock wave
and first three bubble collapse events.
Pressure (MPa) Impulse/area (Pa-s)
st nd rd
Model Shock 1 2 3 Shock 1st 2nd 3rd
Closed Rect. 29.0 12.9 89.0 3.52 1115 1442 3448 1429
Semi-cylinder 25.3 18.0 41.9 0.37 1166 2144 1947 882
Segmented 25.7 14.7 4.91 2.58 1324 3445 2213 2112
Barge 23.2 16.5 1.23 - 1050 1596 273 694

The simulations for the two closed girder models and the segmented model exhibit similar evolution
of the gas bubble in that after the first collapse the remnants of the gas bubble become prolate in shape
and attach to the structure. The gas bubble evolution with the barge model is different in that the
bubble migrates and jets downward. To explore why this is so, consider the criterion for bubble jetting
direction proposed by Wang, et al. [84]. The Froude numbers (15) for these experimental
configurations are determined to be 28.8 for the two closed girder and segmented models, and 28.5 for
the barge model. To get this result it was assumed that ሺܲ௔ െ ܲ௩ ሻȀߩ݃ ൌ ͳͲǤ The term ܼҧߜ is then
evaluated as 0.22 for the first three models and 0.15 for the barge model. The lower value for ܼҧߜ for
the barge model is due to the shallower depth of the charge with that model, and indicates a greater
likelihood for downward jetting. Both values are significantly lower than the critical value of 0.442
(the Kelvin impulse state), but that value is based on a bubble near a free surface with no target.

DRDC-RDDC-2017-R096 61
The predicted periods of the gas bubble pulsation are summarized in Table 13. For the two closed
girder and segmented models the charge depth is 0.3 m, and the first cycle periods range between
28.38 and 29.42 ms. These differences in the periods are due solely to the differences in the target
geometry and stiffness, as all other aspects of the simulations are identical. For the barge model, the
charge depth is 0.204 m, resulting in a bubble period of 26.76, more than 1.5 ms shorter than for the
0.3 m charge depth simulations. With this model the bubble was seen to jet downward as well, and
therefore its behaviour is very similar to that of a bubble near a free surface. The bubble does not
appear to be influenced by the target in the case of the barge model, with its shallow draft.

In Table 14, bubble periods predicted by the empirical formula including the surface correction
factor are given for the two depths. These are in good agreement with the simulation results, the
largest differences being about 0.6 ms. These predicted periods are significantly smaller than the
target design natural period of 31 ms assumed in Sections 3 and 4, and the periods for the 2nd and
3rd cycles are even smaller. Indeed there is little evidence in the predicted target response that the
bubble collapses magnify the initial whipping response, as in a resonant excitation condition.

Table 13: Summary of predicted bubble periods and radii.

Period (ms) Average radius (m)


Model st nd rd
1 2 3 1st
Box girder 28.38 21.19 18.51 0.165
Semi-cylinder 28.98 19.55 17.20 0.169
Segmented 29.42 18.23 23.22 0.170
Barge 26.76 19.43 - 0.167

Table 14: Summary of free-field bubble periods and radii.

Charge depth (m) 1st bubble period (ms)a Max. radius (m)b
0.2 27.22 0.178
0.3 28.96 0.177
a
Eqn (13)
b
Eqn (8)

The predicted average maximum radii in the first bubble cycle are also summarized in Table 13,
while the empirical values are given in Table 14. The average maximum radius here means the
radius averaged in the three coordinate directions with the bubble at maximum expansion. Radii
for the 2nd and 3rd bubble cycles are not given due to the shape distortions after the first collapse.
The predicted first cycle radius is consistently smaller than the empirical values, the largest
difference being 6.7%. Smith [72] showed that gas bubble radius predictions with Chinook were
accurate in the free field and near free surfaces, but that bubble radius is underestimated in FSI
simulations with a floating target.

All the simulations ran successfully to the termination time without the numerical stability
problems previously encountered with Chinook in coupled simulations [72].

62 DRDC-RDDC-2017-R096
6 Conclusion

Several candidate designs of ship-like targets for UNDEX experiments have been considered in
this study. Design verification simulations for four of these were performed to determine if a
realistic ship-like response could be produced at model scale. The simulations showed that the
two closed girder designs exhibit considerable local response behaviour, including partial
collapse of the hull bottom and deck structure, but showed relatively weak hull girder whipping
response. Conversely, the segmented and barge designs exhibited a strong whipping response but
showed very little local response. The simulation results therefore illustrate the tradeoffs to be
encountered with different experimental designs. It is possible that by altering these designs a
model could be produced that exhibits both strong whipping and local response and damage
effects. This would most likely be achieved by reducing the strength of the hull plate material in
the central section of the segmented model.

The simulations have also revealed some significant limitations of small-scale experiments of this
kind. Notably absent from the predicted responses is lifting of the targets due to the buoyancy
action of the expanded gas bubble. The main effect of the gas bubble collapse is local
deformations and damage in the two closed girder models. The segmented model possibly shows
some slight enhancement to the hull girder whipping action after the first collapse, but overall
there is little evidence that the gas bubble loads magnify the whipping action as was seen in
Chertock’s experiments [94]. The small size of the charge used seems to be an issue. The Froude
numbers associated with the modelled cases are in the range of 28.5–28.8, while in a comparable
full scale scenario the Froude number is less than five. As a further comparison, consider that
ܼҧߜ falls in the range of 0.15–0.22 in the modelled cases, whereas in a typical full scale case, values
close to 0.5 are achievable. Thus at full scale ܼҧߜ exceeds the critical value of 0.442 indicating that
buoyancy forces will exceed free surface repulsion, whereas at model scale the buoyancy is much
weaker. Given that gravity and atmospheric pressure cannot in any practical way be scaled, the
only way to lower the Froude number is to use a larger charge (and therefore a larger target
specimen), entailing much greater time and expense for an UNDEX test. Similarly, ܼҧߜ parameter
can be increased by either lowering the Froude number or increasing the charge depth, which
implies a larger target and a larger charge size.

Whereas carefully performed small scale UNDEX experiments have been successful in validating
hi-fidelity simulation tools for simple floating box configurations [116], validation has not yet
been demonstrated for slender ship-like targets. In the work presented in the present study,
hi-fidelity simulation tools have been used in the reciprocal role of helping to design UNDEX
experiments. This helps to better identify the most important configurations for testing and will
therefore allow limited resources to be better allocated. As a way forward from the present work,
it is recommended that ship-like targets configurations appropriate for charge sizes in the 5–40 g
range be designed and evaluated to determine if more realistic target response behaviour can be
achieved.

DRDC-RDDC-2017-R096 63
This page intentionally left blank.

64 DRDC-RDDC-2017-R096
References

[1] Van Aanhold, J. and van Everdinck, C. (2008), Close-Proximity UNDEX tests on a thick
plate at FOI Grindsjön, TNO Report, 2007-D-R1333, Netherlands Organisation for Applied
Scientific Research (TNO), Delft, The Netherlands.

[2] Alin, N., Sundel, T., and Helte, A. (2008), Comparison of Simulation and Measurement,
Technical report FOI-R-2545-SE, Swedish Defence Research Agency (FOI), Grindsjön,
Sweden.

[3] IABG (2014), Burk III UNDEX bubble simulation, Report No. B-TA-4064, IABG,
Ottobrunn, Germany.

[4] Gannon, L. (2014), Simulation of elastic plate response to close-proximity underwater


explosions, DRDC Scientific Report, DRDC-RDDC-2014-R119, Defence Research and
Development Canada.

[5] Slater, J. and Rude, G. (2000), Loading and Response of Flat Plate Targets From
Close-Proximity Underwater Explosions, In Proceedings of the 71st Shock and Vibration
Symposium, Arlington, VA.

[6] Slater, J. and Rude, G. (2004), Loading and Damage to Small-Scale Cylinder Targets from
Close-Proximity Underwater Explosions, DRDC Suffield Technical Report TR 2004-053,
Defence R&D Canada – Suffield.

[7] Slater, J., Rude, G., and Paulgaard, G. (2005), Experimental Study of Air-Backed and
Water-Backed Targets During Near-Contact Explosions, DRDC Suffield TR 2005-152,
Defence R&D Canada – Suffield.

[8] Lee, J., Smith, M., Huang, J., and Paulgaard, G. (2008), Deformation and Rupture of Thin
Steel Plates Due to Cumulative Loading from Underwater Shock and Bubble Collapse, In
Proceedings of the 79th Shock and Vibration Symposium, Orlando, FL.

[9] Lee, J., Smith, M., Rude, G., and Paulgaard, G. (2012), Deformation and rupture of stiffened
square plates subjected to close- proximity underwater explosions. In Proceedings of the
82nd Shock and Vibration Symposium, Baltimore, MD.

[10] Lee, J., Smith, M., Roseveare, D., and Paulgaard, G. (2014), 3D Image Correlation
Measurements on a Floating Panel Loaded by an Underwater Explosion, Presented at the
85th Shock & Vibration Symposium, Reston, VA.

[11] Office of Naval Research (1950), Underwater Explosion Research: A Compendium of


British and American Reports (3 volumes), Office of Naval Research, Department of the
Navy.

[12] Cole, R. (1948), Underwater Explosions, Princeton, NJ: Princeton University Press.

DRDC-RDDC-2017-R096 65
[13] Keil, A. (1961), The Response of Ships to Underwater Explosions, Report 1576, S-F013 04 04,
Society of Naval Architects and Marine Engineers, November 1961.

[14] Holt, M. (1977), Underwater Explosions, Annual Review of Fluid Mechanics, 9, 187–214.

[15] Swisdak, M. (1978), Explosion Effects and Properties Part II – Explosion Effects in Water,
NSWC/WOL TR 76-116, Naval Surface Weapons Center, Dahlgren VA, ADA056694.

[16] Scavuzzo, R. and Pusey, H. (2007), Naval Shock Analysis and Design, Richmond VA: The
Shock and Vibration Information Analysis Center/HI-TEST Laboratories, Inc.

[17] Reid, W. (1996), The Response of Surface Ships to Underwater Explosions, DSTO-GD-0109,
Defence Science and Technology Organisation Aeronautical and Maritime Research
Laboratory, Melbourne, Australia.

[18] Wardlaw, A. and Luton, J. (2000), Fluid-structure interaction mechanisms for close-in
explosions, Shock and Vibration, 7, 265–275.

[19] Young, Y., Liu, Z., and Xie, W. (2009), Fluid-Structure and Shock-Bubble Interaction
Effects During Underwater Explosions Near Composite Structures, Journal of Applied
Mechanics, 76 (5), 1–10.

[20] Church, P., Reynolds, M., Huntington-Thresher, W., Townsley, R., and Sharpe, K. (2004),
Underwater Plate Holing Studies, In Proceedings of the 21st International Symposium on
Ballistics, Adelaide, Australia, 9–16.

[21] Gregson, J., Dunbar, T., and Lee, J. (2007), Simulation of Structural Failure from Contact
Underwater Explosions, In Proceedings of the 78th Shock and Vibration Symposium,
Philadelphia, PA.

[22] Riley, M., Paulgaard, G., Lee, J., and Smith, M. (2010), Failure mode transition in
air-backed plates from near-contact underwater explosions, Shock and Vibration, 17(6)
723–740.

[23] Taylor, G. (1941), The Pressure and Impulse of Submarine Explosion Waves on Plates, In
Underwater Explosion Research: A Compendium of British and American Reports: Vol I – The
Shock Wave, Office of Naval Research, US Department of the Navy, 1155–1174.

[24] Kennard, A. (1944), The Effect of a Pressure Wave on a Plate or Diaphragm, In


Underwater Explosion Research: A Compendium of British and American Reports:
Vol III – The Damage Process, Office of Naval Research, US Department of the Navy, 11–64.

[25] Rajendran, R., Paik, J., and Kim, B. (2006), Design of warship plates against underwater
explosions, Ship and Offshore Structures, 1(4) 347–356.

[26] Jiang, J. and Olson, M. (1996), Non-linear transient analysis of submerged circular plates
subjected to underwater explosions, Computer Methods in Applied Mechanics and
Engineering, 134, 163–179.

66 DRDC-RDDC-2017-R096
[27] Shin, Y. and Hooker, D. (1996), Damage Response of Submerged Imperfect Cylindrical
Structures to Underwater Explosion, Computers and Structures, 60(5), 683–693.

[28] Hammond, L. and Grzebieta, R. (2000), Structural response of submerged air-backed plates
by experimental and numerical analyses, Shock and Vibration, 7, 333–341.

[29] Ramajeyathilagam, K., Vendhan, C., and Bhujanga Rao, V. (2001), Experimental and
numerical investigations on deformation of cylindrical shell panels to underwater explosion,
Shock and Vibration, 8, 253–270.

[30] Ramajeyathilagam, K. and Vendhan, C. (2004), Deformation and rupture of thin rectangular
plates subjected to underwater shock, International Journal of Impact Engineering, 30,
699–719.

[31] Hung, C., Hsu, P., and Hwang-Fuu, J. (2005), Elastic shock response of an air-backed plate
to underwater explosion, International Journal of Impact Engineering, 31, 151–168.

[32] Rajendran, R. and Narasimhan, K. (2006), Deformation and fracture behaviour of plate
specimens subjected to underwater explosion – a review, International Journal of Impact
Engineering, 32, 1945–1963.

[33] Rajendran, R. (2009), Effective shock factors for the inelastic damage prediction of
clamped plane plates subjected to non-contact underwater explosion, Journal of Strain
Analysis, 44, 211–220.

[34] Hung, C., Lin, B., Hwang-Fuu, J., and Hsu, P. (2009), Dynamic response of cylindrical
shell structures subjected to underwater explosion, Ocean Engineering, 36, 564–577.

[35] Jen C. and Tai, Y. (2010), Deformation behavior of a stiffened panel subjected to
underwater shock loading using the non-linear finite element method, Materials and
Design, 31, 325–335.

[36] Gupta, N., Kumar, P., and Hegde, S. (2010), On deformation and tearing of stiffened and
un-stiffened square plates subjected to underwater explosion – a numerical study,
International Journal of Mechanical Sciences, 52, 733–744.

[37] Geers, T. (1978), Doubly asymptotic approximations for transient motions of submerged
structures, Journal of the Acoustical Society of America, 64(5) 1500–1508.

[38] Fellipa, C. and DeRuntz, J. (1984), Finite element analysis of shock-induced hull cavitation,
Computer Methods in Applied Mechanics and Engineering, 44, 297–337.

[39] Sprague, M. and Geers, T. (2004), A spectral-element method for modelling cavitation in
transient fluid-structure interaction, International Journal for Numerical Methods in
Engineering, 60, 2467–2499.

[40] DeRuntz, J. (1980), The underwater shock analysis code and its application, In Proceedings
of the 60th Shock and Vibration Symposium, Virginia Beach VA, 89–107.

DRDC-RDDC-2017-R096 67
[41] LSTC (2011), USA LS-Dyna User’s Manual, USA Release 7.4, Livermore Software
Technology Corp., Livermore, CA.

[42] Trouwborst, W. and Bosman, T. (2002), A simplified fluid-structure interaction model for
underwater shock loading of naval vessels, In Proceedings of the 73rd Shock and Vibration
Symposium, Newport, RI.

[43] Deiterding, R., Cirak, F., and Mauch, S. (2008), Efficient Fluid-Structure Interaction
Simulation of Viscoplastic and Fracturing Thin-Shells Subjected to Underwater Shock
Loading, In Fluid-Structure Interaction: Theory, Numerics and Applications, Kassel:
Kassel University Press, 65–80.

[44] Liu, Z., Xie, W., and Young, Y. (2009), Numerical modeling of complex interactions
between shocks and composite structures, Computational Mechanics, 43, 239–251.

[45] Shin, Y. and Santiago, L. (1998), Surface ship shock modelling and simulation:
two-dimensional analysis, Shock and Vibration, 5, 129–137.

[46] Van Aanhold, J., Meijer, G., and Lemmen, P. (1998), Underwater shock response analysis
of a floating vessel, Shock and Vibration, 5, 53–59.

[47] Shin, Y. and Park, S. (1999), Ship Shock Trial Simulation of USS John Paul Jones
(DDG 53) Using LS-DYNA/USA: Three Dimensional Analysis, In Proceedings of the
70th Shock and Vibration Symposium, Albuquerque, NM.

[48] Shin, Y. and Schneider, N. (2003), Ship Shock Trial Simulation of USS Winston S.
Churchill (DDG 81): Modeling and Simulation Strategy and Surrounding Fluid Volume
Effects, In Proceedings of the 74th Shock and Vibration Symposium, San Diego, CA.

[49] Liang, C. and Tai, Y. (2006), Shock response of a surface ship subjected to noncontact
underwater explosions, Ocean Engineering, 33, 748–772.

[50] Sprague, M. and Geers, T. (2006), A spectral-element/finite element analysis of a ship-like


structure subjected to an underwater explosion, Computer Methods in Applied Mechanics
and Engineering, 195, 2149–2167.

[51] Gong, S. and Lam, K. (2006), On attenuation of floating structure response to underwater
shock, International Journal for Numerical Methods in Engineering, 32 1857–1877.

[52] Zhang, A., Zhou W., Wang, S., and Feng, L. (2011), Dynamic response of non-contact
underwater explosions on naval equipment, Marine Structures, 24, 396–411.

[53] Herring, C. (1949), Theory of the Pulsations of the Gas Bubble Produced by an Underwater
Explosion, In Underwater Explosion Research: A Compendium of British and American
Reports: Vol II – The Gas Globe, Office of Naval Research, US Department of the Navy,
35–130.

68 DRDC-RDDC-2017-R096
[54] Taylor, G. (1942), Vertical Motion of a Spherical Bubble and the Pressure Surrounding it,
In Underwater Explosion Research: A Compendium of British and American Reports:
Vol II – The Gas Globe, Office of Naval Research, US Department of the Navy, 131–144.

[55] Willis, H. (1941), Underwater Explosions: Time Interval between Successive Explosions,
In Underwater Explosion Research: A Compendium of British and American Reports:
Vol II – The Gas Globe, Office of Naval Research, US Department of the Navy, 13–33.

[56] Snay, H. (1962), Underwater Explosion Phenomena: The Parameters of Migrating Bubbles,
NAVORD Report 4185. US Naval Ordnance Laboratory, White Oak MD.

[57] Hicks, A. (1972), The Theory of Explosion Induced Ship Whipping Motions,
Report No. NCRE/R579, Naval Construction Research Establishment, Dunfermline, UK.

[58] Vernon, T. (1986), Whipping response of Ship Hulls from Underwater Explosion Bubble
Loading, Technical Memorandum 86/225, Defence Research Establishment Atlantic.

[59] Heaton, K. (1984), Migration of the Gas Globe from Underwater Explosions: the Effects of
Drag and Radiative Energy Loss, Transactions of the 2nd Army Conference on Applied
Mathematics and Computing, ARO Report 85-1, US Army Research Office, 535–583.

[60] Penny, W. and Price, A. (1942), On the Changing Form of a Nearly Spherical Submarine
Bubble, in Underwater Explosion Research: A Compendium of British and American
Reports: Vol II – The Gas Globe, Office of Naval Research, US Department of the Navy,
145–181.

[61] Heaton, K. (1986), The Effects of Non-Sphericity and Radiative Energy Loss on the
Migration of the Gas Bubble from Underwater Explosions, Transactions of the 3rd Army
Conference on Applied Mathematics and Computing, ARO Report 86-1, US Army
Research Office, 353–401.

[62] Keller, J. and Kolodner, I. (1956), Damping of Underwater Explosion Bubble Oscillations,
Journal of Applied Physics, 27(10), 1152–1161.

[63] Prosperetti, A. and Lezzi, A. (1986), Bubble dynamics in a compressible liquid. Part 1.
First-order theory, Journal of Fluid Mechanics, 168, 457–478.

[64] Geers, T. and Hunter, K. (2002), An integrated wave-effects model for an underwater
explosion bubble, Journal of the Acoustical Society of America, 111(4), 1584–1601.

[65] Van Aanhold, J. (2012), Private communication.

[66] Anderson, C. (1987), An Overview of the Theory of Hydrocodes, International Journal of


Impact Engineering, 5, 33–59.

[67] Benson, D. (1992), Computational methods in Lagrangian and Eulerian hydrocodes,


Computer Methods in Applied Mechanics and Engineering, 99, 235–394.

DRDC-RDDC-2017-R096 69
[68] Mair, H. (1999), Review: Hydrocodes for structural response to underwater explosions,
Shock and Vibration, 6, 81–96.

[69] Landsberg, A., Kee, A., Dare, A., and Ruben, K. (2003), Parallel, coupled hydrocode
simulations of weapon/target interactions for naval applications, In Proceedings of the 2nd
International Conference on Fluid Structure Interaction, 241–250.

[70] Wardlaw, A., Luton, J., Renzi, J., and Kiddy, K. (2003), Fluid-structure coupling
methodology for undersea weapons, Proceedings of the 2nd International Conference on
Fluid Structure Interaction, 251–262.

[71] Hsu, C., Liang, C., Nguyen, A., and Teng, T. (2014), A numerical study on the underwater
explosion bubble pulsation and the collapse process. Ocean Engineering 81, 29–38.

[72] Smith, M. (2016), Benchmark Simulations of Underwater Explosion Gas Bubble Dynamics
and Interactions, DRDC Scientific Report, DRDC-RDDC-2016-R171, Defence Research
and Development Canada.

[73] Menon, S. (1996), Experimental and Numerical Studies of Underwater Explosions, Georgia
Institute of Technology report. ADA317378.

[74] Brett, J. (1998), Numerical Modelling of Shock Wave and Pressure Pulse Generation by
Underwater Explosions, DSTO-TR-0677, DSTO Aeronautical and Maritime Research
Laboratory, Melbourne, Australia.

[75] Barras, G., Souli, M., Aquelet, N., and Couty, N. (2012), Numerical simulation of
underwater explosions using an ALE method. The pulsating bubble phenomena. Ocean
Engineering, 41, 53–66.

[76] Webster, K. (2007), Investigation of Close Proximity Underwater Explosion Effects on a


Ship-Like Structure Using the Multi-Material Arbitrary Lagrangian Eulerian Finite Element
Method. Master of Science thesis. Virginia Polytechnic Institute and State University,
Blacksburg, VA.

[77] Mader, C. (1972), Detonations near the Water Surface. Report LA-4958, Los Alamos
Scientific Laboratory, Los Alamos, NM.

[78] Li, J. and Rong, J. (2011), Bubble and free surface dynamics in shallow underwater
explosion, Ocean Engineering, 38, 1861–1868.

[79] Petrov, N. and Schmidt, A. (2015), Multiphase phenomena in underwater explosion,


Experimental Thermal and Fluid Science, 60, 367–373.

[80] Blake, J., Taib, B., and Doherty, G. (1986), Transient cavities near boundaries. Part 1. Rigid
boundary, Journal of Fluid Mechanics, 170, 479–497.

[81] Blake, J., Taib, B., and Doherty, G. (1986), Transient cavities near boundaries. Part 2. Free
surface, Journal of Fluid Mechanics, 181, 197–212.

70 DRDC-RDDC-2017-R096
[82] Best, J. and Kucera, A. (1992), A numerical investigation of non-spherical rebounding
bubbles, Journal of Fluid Mechanics, 245, 137–154.

[83] Best, J. (1993), The formation of toroidal bubbles upon the collapse of transient cavities,
Journal of Fluid Mechanics, 251, 79–107.

[84] Wang, Q., Yeo, K., Khoo, B., and Lam, K. (1996), Nonlinear interaction between gas
bubble and free surface, Computers and Fluids, 25, 607.

[85] Zhang, Y., Yeo, K., Khoo, B., and Wang, C. (2001), 3D Jet Impact and Toroidal Bubbles,
Journal of Computational Physics, 166, 336–360.

[86] Wang, Q. (2004), Numerical simulation of violent bubble motion, Physics of Fluids, 16(5),
1610–1619.

[87] Best, J. (2002), The Effect of Non-spherical Collapse on Determination of Explosion


Bubble Parameters, DSTO-RR-0238, DSTO Systems Sciences Laboratory, Edinburgh,
Australia.

[88] Zhang, A., Wang, S., Chao, H., and Wang, B. (2011), Influences of initial and boundary
conditions on underwater explosion bubble dynamics, European Journal of Mechanics
B/Fluids, 42, 69–91.

[89] Chapman, N. (1985), Measurement of the waveform parameters of shallow explosive


charges. Journal of the Acoustical Society of America, 78, 672–681.

[90] Chahine, G. and Bovis, A. (1983), Pressure field generated by non-spherical bubble
collapse, ASME Journal of Fluids Engineering, 105, 356–363.

[91] Krieger, J. and Chahine, G. (2005), Acoustic signals of underwater explosions near
surfaces, Journal of the Acoustical Society of America, 118, 2961–2974.

[92] Chahine, G., Frederick, G., Lambrecht, C., Harris, G., and Mair H. (1995), Spark generated
bubbles as laboratory-scale models of underwater explosions and their use for validation of
simulation tools, In Proceedings of the 66th Shock and Vibration Symposium, Biloxi, MS.

[93] Chertock, G. (1953), The Flexural Response of a Submerged Solid to a Pulsating Gas
Sphere, Journal of Applied Physics, 24(2), 192–197.

[94] Chertock, G. (1970), Transient Flexural Vibrations of Ship-Like Structures exposed to


Underwater Explosions, Journal of the Acoustical Society of America, 48(1) Pt 2, 170–180.

[95] Ogilvy, I. and Birkhead, G. (2001), A Numerical Study of the Survivability Implications for
a Trimaran Hull Form to UNDEX Loading, In Proceedings of the 72nd Shock and Vibration
Symposium, Destin, FL.

[96] Zhang, A., Zeng, L., Cheng, X., Wang, S., and Chen, Y. (2011), The evaluation method of
total damage to ship in underwater explosion, Applied Ocean Research, 33, 240–251.

DRDC-RDDC-2017-R096 71
[97] Zhang, N. and Zong, Z. (2012), Hydro-elastic-plastic response of a ship hull girder
subjected to an underwater bubble, Marine Structures, 29, 177–197.

[98] Wang, H., Zhi, X., Cheng, Y.S., and Liu, J. (2014), Experimental and numerical
investigation of ship structure subjected to close-in underwater shock wave and following
gas bubble pulse, Marine Structures, 39, 90–117.

[99] Zhang, Z., Wang, Y., Zhao, H., Qian, H., and Mou, J. (2015), An experimental study on the
dynamic response of a hull girder subjected to near field underwater explosion, Marine
Structures, 44, 43–60.

[100] Tipton, R., Steinberg, D., and Tomita, Y. (1992), Bubble Expansion and Collapse Near a
Rigid Wall, JSME International Journal, Series II, 35(1) 67–75.

[101] Klaseboer, E., Hung, K., Wang, C., Wang, C.W., Khoo, B., Boyce, P., Debono, S., and
Charlier, H. (2005), Experimental and numerical investigation of the dynamics of an
underwater explosion bubble near a resilient/rigid structure, Journal of Fluid Mechanics,
537, 387–413.

[102] Van Aanhold, J. and Bosman, T. (2005), Nearby UNDEX Tests on a Thick Plate, In
Proceedings of the 76th Shock and Vibration Symposium, Destin, FL.

[103] Brett, J.M. and Yiannakopolous, G. (2008), A study of explosive effects in close proximity to
a submerged cylinder, International Journal of Impact Engineering, 35, 206–225.

[104] Li, J. and Rong, J. (2012), Experimental and numerical investigation of the dynamic
response of structures subjected to underwater explosion, European Journal of Mechanics
B/Fluids, 32, 59–69.

[105] Imakita, A. and Yasuda, A. (2009), Response of Cylindrical Floating Structure to an


Underwater Explosion, In ClassNK Technical Bulletin, 27, 1–9, Nippon Kaiji Kyokai
(ClassNK), Tokyo, Japan.

[106] Chen, Y., Tone, Z., Hua, H., Wang, Y., and Gou, H. (2009), Experimental investigation on
the dynamic response of scaled ship model with rubber sandwich coatings subjected to
underwater explosion, International Journal of Impact Engineering, 36, 318–328.

[107] Chisum, J. and Shin, Y. (1997), Explosion Gas Bubbles Near Simple Boundaries, Shock
and Vibration, 4(1), 11–25.

[108] Link, R., Lin, F., Whitehouse, D., and Slater, J. (2002), CFD Modeling of Close Proximity
Underwater Explosions, In Proceedings of the 73rd Shock and Vibration Symposium,
Newport, USA.

[109] Dunbar, T. (2009), Modelling of Close-Proximity Underwater Explosion Loads and


Structural Response – Phase 1, DRDC Atlantic CR 2008-272, Martec Limited, Halifax, NS.

72 DRDC-RDDC-2017-R096
[110] Riley, M. (2010), Modeling Gas Bubble Behaviour and Loading on a Rigid Target due to
Close-Proximity Underwater Explosions, DRDC Atlantic TM 2010-238, Defence R&D
Canada – Atlantic.

[111] Riley, M., Smith, M., van Aanhold, J., and Alin, N. (2012), Loading on a Rigid Target
from Close Proximity Underwater Explosions, Shock and Vibration, 19 (4), 555–571.

[112] Link, R., Ripley, R., Norwood, M., Josey, T., Donahue, L., and Slater, J. (2003), Analysis
of the Loading and Response of Flat Plate Targets Subjected to Close-Proximity
Underwater Explosions, In Proceedings of the 74th Shock and Vibration Symposium, San
Diego, CA.

[113] Gregson, J., Link, R., and Lee, J. (2006), Coupled Simulation of the Response of Targets
to Close Proximity Underwater Explosions in Two and Three Dimensions, In Proceedings
of the 77th Shock and Vibration Symposium, Monterey, CA.

[114] Dunbar, T. (2009), Modelling of Close-Proximity Underwater Explosion Loads and


Structural Response – Phase 2, DRDC Atlantic CR 2009-180, Martec Limited, Halifax, NS.

[115] Kan, K., Stuhmiller, J., and Chan, P. (2005), Simulation of the collapse of an underwater
explosion bubble under a circular plate, Shock and Vibration, 12, 217–225.

[116] Smith, M. and Lee, J. (2016), Coupled Eulerian-Lagrangian Simulations of an Underwater


Explosion near a Floating Target, In Proceedings of the 24th International Symposium on
Military Aspects of Blast and Shock Conference (MABS24), Halifax, NS.

[117] Yoon, D., Park, J., Eccles, T., Powell, A., Manley D., and Widholm, A. (2010), Joint
Investigation Report On the Attack Against ROK Ship Cheonan, Ministry of National
Defence, Republic of Korea.

[118] Duncan, J., Milligan, C., and Zhang, S. (1996), On the interaction between a bubble and a
submerged compliant structure, Journal of Sound and Vibration, 197(1), 17–44.

[119] Kalumuck, K., Duraiswami, R., and Chahine, G. (1995), Bubble Dynamics Fluid-Structure
Interaction Simulation by Coupling Fluid BEM and Structural FEM Codes, Journal of
Fluids and Structures, 9, 861–883.

[120] Chahine, G. and Kalumuck, K. (2002), Simulation of Surface Piercing Body Coupled
Response to Underwater Bubble Dynamics Utilizing 2DynaFS, a Three-Dimensional BEM
Code, Presented at IABEM 2002 (International Association for Boundary Element
Methods), Austin, TX.

[121] Zhang, A., Yao X., and Li, J. (2008), The interaction of an underwater explosion bubble
with an elastic-plastic structure, Applied Ocean Research, 30, 159–171.

[122] Costanzo, F.A. and Gordon, J.D. (1983), A Solution to the Axisymmetric Bulk Cavitation
Problem, The Shock and Vibration Bulletin, Part 2: Fluid Structure Dynamics and
Dynamic Analysis, 53, 33–51.

DRDC-RDDC-2017-R096 73
[123] Arons, A., Yennie, D., and Cotter, T. (1950), Long Range Shock Propagation in
Underwater Explosion Phenomena II, In Underwater Explosion Research, Vol. 1. The
Shock Wave, Office of Naval Research – Department of the Navy, USA, pp. 1473–1583.

[124] Wedberg, R. (2015), Testing a physically based model for cloud cavitating water in
underwater explosion simulations. FOI Technical Report: FOI-D-0698-SE, Swedish
Defence Research Agency (FOI), Grindsjön, Sweden.

[125] Kedrinskii, V. (2005), Hydrodynamics of explosions: experiment and models, Berlin:


Springer-Verlag.

[126] Youtube; Australian Collins class submarine torpedo (online),


https://www.youtube.com/watch?v=SLrKOOXcOhM, (Access date: 3 Mar. 2017).

[127] Tse, F., Morse, I., and Hinkle, R. (1978), Mechanical Vibrations: Theory and Applications.
Boston, MA: Allyn and Bacon.

[128] Lewis, F. (1929), The Inertia of Water Surrounding a Vibrating Ship, Transactions of
SNAME, 37, 1–20.

[129] Martec Limited (2012), Vibration and Strength Analysis, Version 9.1, User’s Manual,
Halifax, NS: Martec Limited.

[130] Vernon, T., Bara, B., and Hally, D. (1988), A Surface Panel Method for the Calculation of
Added Mass Matrices for Finite Element Models, Technical Memorandum 88/203,
Defence Research Establishment Atlantic.

[131] Martec Limited (2015), ChinookEXP Input Manual, Martec Software Manual #SM-13-04
Rev 11, Halifax, NS: Lloyd’s Register Applied Technology Group.

[132] LSTC (2012), LS-DYNA Keyword User’s Manual. Version 971 R6.1.0, Livermore, CA:
Livermore Software Technology Corporation.

[133] Pointwise (2015), Pointwise User Manual, Fort Worth TX: Pointwise, Inc.

[134] Li, Y. (1967), Equation of State of Water and Sea Water, Journal of Geophysical
Research, 72(10), 2665–2678.

[135] Lee, E., Horning, H., and Kury, J. (1968), Adiabatic expansion of high explosives
detonation products, Report TID 4500-UCRL 50422, Lawrence Livermore National
Laboratory, Livermore, CA.

[136] Jones, H. and Miller, A. (1948), The detonation of solid explosives, Proceedings of the
Royal Society of London, A-194, 480–507.

[137] Wilkins, M. (1964), The equation of state of PBX 9404 and LX04-01, Report UCRL-7797,
Lawrence Radiation Laboratory, Livermore CA.

74 DRDC-RDDC-2017-R096
[138] MacAdam, T. (2015), SubSAS User’s Guide, Technical Report # TR-15-63, Lloyd’s
Register Applied Technology Group, Halifax, NS.

[139] Zhu, D., Mobasher, B., Rajan, S., and Peralta, P. (2011), Characterization of Dynamic
Tensile Testing Using Aluminum Alloy 6061-T6 at Intermediate Strain Rates, ASCE
Journal of Engineering Mechanics, 137 (10), 669–679.

DRDC-RDDC-2017-R096 75
This page intentionally left blank.

76 DRDC-RDDC-2017-R096
List of symbols/abbreviations/acronyms/initialisms

a Instantaneous bubble radius


A Johnson-Cook parameter
A1, A2 Empirical constants
‫ܣ‬ெ௡ Maximum radius of explosion gas bubble in the nth cycle
ALE Arbitrary Lagrangian-Eulerian
atm Atmospheres
B Overall breadth, also a Johnson-Cook parameter
bg Breadth of box girder (in segmented model)
BEM Boundary element method
c Speed of sound in water
C Johnson-Cook parameter, also a Cowper-Symonds parameter, also shape
coefficient
C1, … C4 Design criteria
C-4 Composition 4 (explosive)
CEL Coupled Eulerian-Lagrangian
CFD Computational fluid dynamics
cm centimetres
cos Cosine
cosh Hyperbolic cosine
d Charge depth, or water depth at target plate
D Overall depth of model
D1,…, D5 Design objectives
dg Depth of box girder (in segmented model)
dn Depth of gas bubble during nth period or pulse
ds Depth of stiffener
dw Depth of vertical side shell
DAA Doubly asymptotic approximation
DIC Digital image correlation
DRDC Defence Research and Development Canada
E Elastic modulus, or total energy

DRDC-RDDC-2017-R096 77
Et Tangent modulus
EOS Equation of state
݂௡ Frequency in the nth natural mode
FEM Finite element method
Fr Froude number
FSI Fluid-structure interaction
g gram
g Acceleration of gravity
‫ܩ‬ሺܽሻ Internal energy of the gas bubble
GPa Gigapascals
h Hull thickness
hb Bulkhead thickness
hg Thickness of box girder (in segmented model)
Hn Hydrostatic water head in the nth bubble period
hs Stiffener thickness
Hz Hertz
I Second moment of area
J Joules
J Correction factor in the Lewis fluid mass formula
JWL Jones-Wilkins-Lee
k Isentropic gas law constant
K Kelvin
K1, K2, K5, K6 Empirical constants
kg kilograms
kn Wavenumber
L Overall length
m Metres
m Mass per unit area
mf Fluid added mass per unit length
ml Mass per unit length
mm Millimetres
MOU Memorandum of Understanding

78 DRDC-RDDC-2017-R096
ms Milliseconds
n Johnson-Cook strain exponent
nb Number of bulkheads
p Cowper-Symonds parameter
P Pressure
ܲ௔ Atmospheric pressure
Pa Pascal
ܲ௖ Cavitation closure pressure
ܲ௚ Gas pressure
ܲ௩ Vapour pressure
ܲஶ Undisturbed pressure at depth z
ܲ଴ Peak pressure
Q Explosive energy per unit mass
R Standoff distance
rn Fraction of explosive energy in the nth bubble pulse
RDX Research Department explosive (cyclotrimethylenetrinitramine)
ROKS Republic of Korea Ship
R&D Research & Development
s Seconds
SI Système International
sin Sine
sinh Hyperbolic sine
t time
Tn Bubble period of the nth period or pulse
TNT Trinitrotoluene
UNDEX Underwater explosion
‫ݒ‬ Fluid particle velocity
V volume
‫ݒ‬௝௘௧ Water jet velocity
W Charge mass
wn(x) Eigenfunction
x Spatial coordinate

DRDC-RDDC-2017-R096 79
X,Y,Z Cartesian coordinates
z Depth
ܼҧ Normalized depth
1D One dimensional
2D Two dimensional
2NV Two-node vertical
3D Three dimensional
α Surface correction factor
ߙ௖ Cavitation fraction
ߛ Ratio of specific heats
ߜ Buoyancy parameter
ߝ௣ Effective plastic strain
ߝሶ Averaged strain rate
‫כ‬
ߝሶ Normalized effective strain rate
ߠ Decay constant
ߤ Non-structural mass per unit length
ߥ Poisson ratio
ߦ Plate mass ratio
ߩ Density
ߪ௬ Yield stress
߱௡ Natural frequency

80 DRDC-RDDC-2017-R096
DOCUMENT CONTROL DATA
(Security markings for the title, abstract and indexing annotation must be entered when the document is Classified or Designated)
1. ORIGINATOR (The name and address of the organization preparing the document. 2a. SECURITY MARKING
Organizations for whom the document was prepared, e.g., Centre sponsoring a (Overall security marking of the document including
contractor’s report, or tasking agency, are entered in Section 8.) special supplemental markings if applicable.)

DRDC – Atlantic Research Centre UNCLASSIFIED


Defence Research and Development Canada
9 Grove Street
2b. CONTROLLED GOODS
P.O. Box 1012
Dartmouth, Nova Scotia B2Y 3Z7
(NON-CONTROLLED GOODS)
Canada
DMC A
REVIEW: GCEC DECEMBER 2013
3. TITLE (The complete document title as indicated on the title page. Its classification should be indicated by the appropriate abbreviation (S, C or U) in
parentheses after the title.)

Ship-like target design for underwater explosion experiments


4. AUTHORS (last name, followed by initials – ranks, titles, etc., not to be used)

Smith, M.J.; Lee, J.J.

5. DATE OF PUBLICATION 6a. NO. OF PAGES 6b. NO. OF REFS


(Month and year of publication of document.) (Total containing information, (Total cited in document.)
including Annexes, Appendices,
etc.)
July 2017
92 139
7. DESCRIPTIVE NOTES (The category of the document, e.g., technical report, technical note or memorandum. If appropriate, enter the type of report,
e.g., interim, progress, summary, annual or final. Give the inclusive dates when a specific reporting period is covered.)

Scientific Report
8. SPONSORING ACTIVITY (The name of the department project office or laboratory sponsoring the research and development – include address.)

DRDC – Atlantic Research Centre


Defence Research and Development Canada
9 Grove Street
P.O. Box 1012
Dartmouth, Nova Scotia B2Y 3Z7
Canada
9a. PROJECT OR GRANT NO. (If appropriate, the applicable research 9b. CONTRACT NO. (If appropriate, the applicable number under
and development project or grant number under which the document which the document was written.)
was written. Please specify whether project or grant.)

01EA
10a. ORIGINATOR’S DOCUMENT NUMBER (The official document 10b. OTHER DOCUMENT NO(s). (Any other numbers which may be
number by which the document is identified by the originating assigned this document either by the originator or by the sponsor.)
activity. This number must be unique to this document.)

DRDC-RDDC-2017-R096
11. DOCUMENT AVAILABILITY (Any limitations on further dissemination of the document, other than those imposed by security classification.)

Unlimited
12. DOCUMENT ANNOUNCEMENT (Any limitation to the bibliographic announcement of this document. This will normally correspond to the
Document Availability (11). However, where further distribution (beyond the audience specified in (11) is possible, a wider announcement
audience may be selected.))

Unlimited
13. ABSTRACT (A brief and factual summary of the document. It may also appear elsewhere in the body of the document itself. It is highly desirable that
the abstract of classified documents be unclassified. Each paragraph of the abstract shall begin with an indication of the security classification of the
information in the paragraph (unless the document itself is unclassified) represented as (S), (C), (R), or (U). It is not necessary to include here abstracts in
both official languages unless the text is bilingual.)

The design of floating ship-like targets is considered with the objective of reproducing at
reduced scale most of the loading and response effects that would occur in a full-scale vessel in
a close proximity underwater explosion event. Some candidate designs are produced based on
simple parameterized cross sections. Design parameters are then iteratively adjusted so as to
satisfy parameter constraints and the overall design objectives, which include alignment
between the fundamental natural period and the explosion gas bubble pulsation period.
Candidate designs are proposed, and the shock and gas bubble loading and target response in
each case are extensively studied using a 3D coupled Eulerian-Lagrangian simulation
methodology. The results of these simulations show that the details of the loading and the
response among the four models are significantly different despite the similarity of the test
configurations. Local deformation and damage are observed in some models, and a global
whipping response is observed in other models. Because of the small scale of the configurations
selected, buoyancy effects are very small and this may limit the applicability of the
configurations to full-scale.
---------------------------------------------------------------------------------------------------------------

On envisage la conception de cibles flottantes semblables à des navires dans le but de reproduire
à une échelle réduite la plupart des effets de charge et de réaction qui se produisent sur un navire
à échelle réelle lors d’une explosion sous-marine à proximité immédiate. Des configurations
possibles ont été produites à partir de simples coupes transversales paramétrées. Les paramètres
de conception ont alors été rajustés de façon répétitive afin de satisfaire aux contraintes et aux
objectifs généraux, notamment la concordance entre la fréquence propre fondamentale et la
fréquence d’oscillation de la bulle de gaz lors de l’explosion. Dans l’étude, on propose des
configurations possibles; la charge de choc et de la bulle de gaz, ainsi que la réaction de la cible
sont étudiées en détail à l’aide de simulations 3D selon la méthode Euler-Lagrange. Les
résultats de ces simulations révèlent que les détails de la charge et de la réaction des quatre
modèles diffèrent considérablement, malgré la similitude des configurations mises à l’essai. On
a observé une déformation et des dommages locaux chez certains modèles, et une vibration de
frottement généralisée chez d’autres. En raison de l’échelle réduite de la configuration choisie,
les effets de flottabilité sont très légers et pourraient en limiter l’application à échelle réelle.

14. KEYWORDS, DESCRIPTORS or IDENTIFIERS (Technically meaningful terms or short phrases that characterize a document and could be helpful
in cataloguing the document. They should be selected so that no security classification is required. Identifiers, such as equipment model designation,
trade name, military project code name, geographic location may also be included. If possible keywords should be selected from a published thesaurus,
e.g., Thesaurus of Engineering and Scientific Terms (TEST) and that thesaurus identified. If it is not possible to select indexing terms which are
Unclassified, the classification of each should be indicated as with the title.)

Underwater explosion; close proximity UNDEX; whipping; gas bubble; water jet; hydrocode;
coupled Eulerian-Lagrangian

You might also like