You are on page 1of 13

A model for underwater sound levels generated by marine impact pile driving

Alexander MacGillivray

Citation: Proc. Mtgs. Acoust. 20, 045008 (2013); doi: 10.1121/2.0000030


View online: https://doi.org/10.1121/2.0000030
View Table of Contents: https://asa.scitation.org/toc/pma/20/1
Published by the Acoustical Society of America

ARTICLES YOU MAY BE INTERESTED IN

A split-step Padé solution for the parabolic equation method


The Journal of the Acoustical Society of America 93, 1736 (1993); https://doi.org/10.1121/1.406739

Improved equivalent fluid approximations for a low shear speed ocean bottom
The Journal of the Acoustical Society of America 98, 3391 (1995); https://doi.org/10.1121/1.413789

Compressional and shear wave properties of marine sediments: Comparisons between theory and data
The Journal of the Acoustical Society of America 117, 137 (2005); https://doi.org/10.1121/1.1810231

Comparison of algorithms for solving parabolic wave equations


The Journal of the Acoustical Society of America 100, 178 (1996); https://doi.org/10.1121/1.415921

Marine biological choruses observed in tropical waters near Australia


The Journal of the Acoustical Society of America 64, 736 (1978); https://doi.org/10.1121/1.382038

Pile driving zone of responsiveness extends beyond 20 km for harbor porpoises (Phocoena phocoena (L.))
The Journal of the Acoustical Society of America 126, 11 (2009); https://doi.org/10.1121/1.3132523
Volume 20 http://acousticalsociety.org/

166th Meeting of the Acoustical Society of America


San Francisco, California
2 - 6 December 2013
Physical Acoustics: Paper 2aPAa5

A model for underwater sound levels generated


by marine impact pile driving
Alexander O. MacGillivray
JASCO Applied Sciences, Victoria, BC, Canada; alex@jasco.com

Marine impact pile driving generates very high underwater sound pressures, which can harm aquatic life.
Environmental assessments for pile driving projects typically require acoustic impact zones for marine
mammals and fish to be estimated in advance. A computer model that predicts the radiated acoustic field
from impact driving of cylindrical piles has been developed. A lumped-mass model of the hammer, which
predicts the force generated at the top of the pile, is coupled to a 1-D finite-difference model of radial and
axial stress waves in a cylindrical pile. The radiated pressure is computed by matching the velocity
boundary condition at the pile wall using a superposition of monopole sources distributed over the length
of the pile in a layered 2-D fluid medium. The transfer function for the monopoles is computed using the
near-field Hankel transform for radial particle velocity at the pile wall. Standard ocean acoustic modeling
techniques are used to compute the Mach wave propagating away from the pile. The model considers the
physical characteristics of the pile-hammer system, such as the hammer energy, pile dimensions, sediment
properties, and pile-hammer impedance ratio. The model’s predictions are compared to field measurements
obtained in a riverine environment.

Published by the Acoustical Society of America

© 2015 Acoustical Society of America [DOI: 10.1121/2.0000030]


Received 07 January 2015; Published 06 April 2015
Proceedings of Meetings on Acoustics, Vol. 20 045008 (2015) Page 1
A. MacGillivray A marine pile driving sound model

1. Introduction

Marine impact pile driving generates very high underwater sound pressures, which can harm
aquatic life. Environmental assessments for pile driving projects typically require advance
estimates of acoustic impact zones for marine mammals and fish. Models of peak and rms sound
pressure level (SPL) and sound exposure level (SEL) from planned pile driving activities are
needed to predict these zones.

Although practical spreading loss models are widely applied in environmental


assessments (ICF Jones & Stokes and Illingworth and Rodkin, 2009), their ability to predict
noise is limited because they only extrapolate from existing data. Recently, more sophisticated
acoustic models of pile driving have been developed using finite element methods (Reinhall and
Dahl, 2011) and semi-analytical approaches (Hall, 2013). This article describes a computational
model that combines a finite difference model of hammer and pile vibration with a near-field
wave-number integration model for coupling the pile wall vibration to the surrounding acoustic
medium. The model generates a vertical array of monopole sources, which can then be input into
standard ocean acoustic propagation models to predict the radiated sound field of the pile.

In this work, impact hammer forces at the top of the pile are computed using a lumped-
mass force generator model (Parola, 1970). Equations of motion for the hammer are coupled to
those for the axial and radial vibrations of a thin cylindrical shell (Junger and Feit, 1972).
Damping of the pile vibration due to radiation loading is computed using the cylindrical-wave
impedance relation for Mach waves emanating from the pile wall. The resulting system of
equations for the pile hammer system is solved in the time-domain using the finite difference
method. A vertical array of discrete monopoles, centered on the pile, simulate the radiated
pressure waves. The amplitudes of the monopoles are chosen in such a way that the
superposition of their particle velocity fields—calculated using a near-field wave-number
integration model—matches the horizontal displacement boundary condition at the pile wall. All
of the model parameters are based on readily obtainable physical properties of the impact
hammer and of the pile itself.

2. Theory
2.1. Impact Hammer Dynamics
A model of the force imparted by the hammer at the top of the pile is needed to compute the
stress wave in a pile. In the civil engineering literature, wave-equation analysis models are used
to estimate drivability and load-bearing capacity of support piles. Wave-equation analysis
models use lumped-mass systems to simulate impact hammer dynamics. The model discussed in
this paper uses the basic force generator model, as described by Parola (1970), where the
hammer is modeled as a two-mass system consisting of a falling ram (𝑚1 ) coupled to a drive
head (𝑚2 ), by a linear elastic cushion with stiffness coefficient 𝜅 (Figure 1). The motion of the
ram and drive head is thus described by a system of ordinary differential equations (ODEs):

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 2


A. MacGillivray A marine pile driving sound model

𝑚1 𝑋̈1 = 𝑔 − 𝜅(𝑋1 − 𝑋2 ) (1a)

𝑚2 𝑋̈2 = 𝜅(𝑋1 − 𝑋2 ) − 𝜎𝑧𝑧 𝐴 (1b)

In the above equations, 𝑋1 is the vertical displacement of the ram, 𝑋2 is the vertical displacement
of the drive head, g is the acceleration due to gravity, 𝜎𝑧𝑧 is the vertical stress at the top of the
pile and A is the cross-sectional area of the pile. In solving these equations, the initial velocity of
the ram, 𝑋̇1 (𝑡 = 0), is determined by assuming that the hammer energy, E, is transformed into
kinetic energy as it falls. The vertical displacement of the drive head is assumed to be equal to
that at the top of the pile. Note that there is no tension in the pile cushion, so that 𝜅 = 0 when
𝑋1 < 𝑋2 .

Figure 1. Diagram of the lumped-mass force generator model, which is used for simulating impact
hammer dynamics.

In the force generator model, the shape of the force impulse at the top of the pile is
governed by the stress wave impedance ratio between the pile and the hammer:

𝑍𝑝 𝜌𝑝 𝑐𝑝 𝐴
= (2)
𝑍ℎ √𝑚1 𝜅

where 𝜌𝑝 is the density of the pile, 𝑐𝑝 is the stress-wave speed in the pile; the other parameters
are described in Eqn. 1. This ratio is a dimensionless quantity that is typically in the range 0.4 <
𝑍𝑝 ⁄𝑍ℎ < 1.5 for efficient pile driving. When the impedance ratio is small, the amplitude of the
force pulse is larger. Note that the period of the initial pulse in the force generator model is
proportional to a characteristic time:

𝑚1
𝑇 = 𝜋√ (3)
𝜅

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 3


A. MacGillivray A marine pile driving sound model

This quantity, which is closely related to the shape of the force pulse spectrum, is independent of
the hammer energy and the characteristic impedance of the pile.

2.2. Equations of Motion of the Pile


The axial and radial stress waves in the pile are modeled using the equations of motion for a thin
cylindrical shell (Junger and Feit, 1972), which are a set of coupled partial differential equations
that describe the radial acceleration, 𝑤̈ , and axial acceleration, 𝑢̈ , of the pile versus length, z, and
time, t:

𝑤̈ 1 − 𝜈 2 𝜈 𝜕𝑢 𝑤 ℎ4 𝜕 4 𝑤
= 𝑝𝑎 (𝑧, 𝑡) − − 2− (4a)
𝑌/𝜌 𝑌ℎ 𝑎 𝜕𝑧 𝑎 144 𝑎 𝑑𝑧 4

𝑢̈ 𝜕 2 𝑢 𝜈 𝜕𝑤
= + (4b)
𝑌/𝜌 𝜕𝑧 2 𝑎 𝜕𝑧

The above equations are expressed in terms of the material properties of the pile: Young’s
modulus, Y, Poisson’s ratio, ν, radius, a, wall thickness, h, and bulk density, 𝜌𝑝 . The term
𝑝𝑎 (𝑧, 𝑡) is the pressure loading at the pile wall, which accounts for energy loss due to radiated
sound (see Section 2.3). The thin-shell equations of motion are derived under the assumption that
the thickness is small compared to the radius (h << a) and that the displacement is small
compared to the thickness (w << h). These assumptions are generally satisfied for real-world
driving of hollow pipe piles, where typically the radius is on the order of 0.1–1 m, the thickness
is on the order of 10−2 m, and the radial displacement is on the order of 10−4 m.

The equations of motion are solved subject to three boundary conditions. The first
boundary condition assumes continuity of vertical stress between the drive head and the top of
the pile:

𝑌 𝜕𝑢 ℎ2 𝜕 2 𝑤
𝜎𝑧𝑧 |𝑧=0 = ( − ) (5)
1 − 𝜈 2 𝜕𝑧 12𝑎 𝜕𝑧 2

The second boundary condition assumes that the radial velocity gradient is equal to zero at the
pile tips:

𝜕 2𝑤
| =0 (6)
𝜕𝑧 2 𝑧=0,𝐿

The third is an impedance boundary condition that assumes a reflection coefficient, −1 < R < 1,
for axial stress waves incident at the bottom of the pile:

𝜕𝑢 1 1 − 𝑅 𝜕𝑢
| =− ( ) (7)
𝜕𝑧 𝑧=𝐿 √𝑌/𝜌 1 + 𝑅 𝜕𝑡

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 4


A. MacGillivray A marine pile driving sound model

The impedance boundary condition accounts for energy dissipation due to soil resistance
at the bottom of the pile. Note that the value of the reflection coefficient, R, determines the total
penetration of the pile into the soil. Values of 𝑅 → −1 correspond to hard soils and small
displacements (i.e., refusal), whereas values of 𝑅 → 1 correspond to soft soils and large
displacements. A suitable value for R can be selected by matching the total pile displacement to
the expected driving conditions.

2.3. Radiation Loading


Energy loss due to radiation loading at the pile wall is computed based on the assumption that
the pile is immersed in a fluid with depth-dependent sound speed, 𝑐𝑓 (𝑧), and density, 𝜌𝑓 (𝑧). In
the frequency domain (𝜔), the local acoustic pressure at the pile wall (𝑟 = 𝑎) is computed from
the impedance relationship for cylindrical waves:

(1)
𝑃𝑎 (𝜔) 𝑖𝜌𝑓 𝜔 𝐻0 (𝑘𝑟 𝑎)
𝑍𝑎 (𝜔) = =− (8)
𝑊̇ (𝜔) 𝑘𝑟 𝐻 (1) (𝑘𝑟 𝑎)
1

where 𝑃𝑎 (𝜔) is acoustic pressure, 𝑊̇ (𝜔) is horizontal particle velocity, 𝑘𝑟 is the horizontal
(1)
wave-number in the fluid, and 𝐻𝑛 is the nth order Hankel function of the first kind. This
expression assumes that only the horizontal component of particle velocity couples acoustically
to the surrounding medium, and that radiation loading occurs only on the outer wall of the pile.

To calculate the radiation loading, the impedance relation is evaluated for Mach waves
propagating away from the pile wall. The Snell’s law approximation is used to compute the
horizontal wave-number for an acoustic wave in the fluid with vertical wave-number equal to
that of the axial stress wave in the pile:

𝜔 2
𝑘𝑟 = 𝑐 √1 − (𝑐𝑓 ⁄𝑐𝑝 ) , 𝑐𝑓 < 𝑐𝑝 (9)
𝑓

Because the equations of motion are solved in the time-domain, the frequency-domain
impedance relationship of Eqn. 8 is transformed to a time-domain impulse response function via
the inverse Fourier transform:

∞ (1)
𝜌𝑓 𝑐𝑓 −𝑖𝐻0 (𝑘𝑟 𝑎)
𝜁𝑎 (𝑧, 𝑡) = ∫ (1)
𝑒 −𝑖𝜔𝑡 𝑑𝜔 (10)
2 −∞ 𝐻 (𝑘𝑟 𝑎)
2𝜋√1 − (𝑐𝑓 ⁄𝑐𝑝 ) 1

By applying the convolution theorem to Eqn. 8, the radiation loading at the pile wall is computed
by convolving the radial velocity with the impulse response function of Eqn. 10:

𝑝𝑎 (𝑧, 𝑡) = 𝜁𝑎 (𝑧, 𝑡) ∗ 𝑤̇ (𝑧, 𝑡) (11)

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 5


A. MacGillivray A marine pile driving sound model

2.4. Equivalent Monopole Array

Sound radiation from the pile is simulated with a vertical array of N monopole sources (Figure
2). The amplitudes of the monopoles are chosen so that the horizontal particle velocity equals the
radial pile velocity, 𝑤̇𝑎 (𝑧, 𝑡), at N boundary points along the pile wall. A near-field wave-number
integration model is used to compute the monopole amplitudes so that their superposition
satisfies the horizontal particle velocity boundary condition at r = a. The N monopole
amplitudes, 𝐶𝑖 (𝜔), are calculated in the frequency-domain by inverting a matrix expression for
the superposition of N monopole fields:

𝐶1 (𝜔) 𝐺11 (𝜔) ⋯ 𝐺1𝑁 (𝜔) −1 𝑊̇1 (𝜔)


( ⋮ )= ( ⋮ ⋱ ⋮ ) ( ⋮ ) (12)
𝐶𝑁 (𝜔) 𝐺𝑁1 (𝜔) ⋯ 𝐺𝑁𝑁 (𝜔) 𝑊̇𝑁 (𝜔)

where 𝑊̇𝑛 (𝜔) is the Fourier transform of the radial pile velocity at boundary point n, and
𝐺𝑛𝑚 (𝜔) is the transfer function for particle velocity between monopole n and boundary point m.

Figure 2. Geometry (r-z plane) of the equivalent monopole array used to model sound radiated by a
cylindrical pile. Red dots indicate monopole sources at the pile center. Yellow dots indicate receiver
points along the pile wall boundary.

The wave-number integration model computes the near-field transfer function using the
Hankel transform for horizontal particle velocity in a stratified medium:

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 6


A. MacGillivray A marine pile driving sound model


𝑖
𝐺𝑗𝑘 (𝜔) = ∫ 𝑔(𝑧𝑗 , 𝑧𝑘 , 𝜔, 𝑘𝑟 )𝐽1 (𝑘𝑟 𝑎)𝑘𝑟2 𝑑𝑘𝑟 (13)
𝜔𝜌𝑓
0

where 𝑔(𝑧𝑗 , 𝑧𝑘 , 𝜔, 𝑘𝑟 ) is the depth-dependent Green’s function between a source at depth 𝑧𝑗 and
a receiver at depth 𝑧𝑘 .

2.5. Numerical Implementation


The equations of motion for the pile hammer system (Eqns. 1–7) are discretized using the finite
difference method (Jensen et al., 2011) and implemented in a computer program. Solutions to the
partial differential equations are computed on a finite depth and time mesh. The equations of
motion are discretized to second order, and Lagrange extrapolation is used to compute
differentials at the pile tips. The program outputs the radial velocity of the pile wall as a function
of depth and time, 𝑤̇ (𝑧, 𝑡).

The equivalent monopole array is calculated in a separate computer program, which takes
the radial velocity of the pile wall as input. This program performs direct numerical integration
of the near-field Hankel transform of Eqn. 13, and solves for the frequency-domain monopole
amplitudes using Eqn. 12. The current version of the monopole program computes the Green’s
function for a two-layer fluid environment (water over sediment) but could be extended to handle
more layers using standard methods. This program outputs a set of N time-domain pressure
signatures (referenced to 1 m) that can be input into standard ocean propagation models for
computing far-field sound pressure.

3. Model Data Comparison


Calibrated pressure waveforms from impact driving of cylindrical steel piles in the Hudson River
were measured during the 2012 Tappan Zee Bridge Pile Installation Demonstration Project
(PIDP) (Martin et al., 2012). Predictions of the pile driving model were compared with data for a
244 cm diameter, 68 m long steel pile driven with a Menck MHU-800S hydraulic hammer.

Model input parameters were based on observations of environmental and hammering


conditions logged during the measurements (Table 1). Sound speed in water was calculated
based on temperature and salinity profiles taken during the acoustic measurements. Geoacoustic
properties were based on borehole samples taken near the pile driving site, which indicated the
bottom type was clay (mean grain size φ ≈ 8). The construction contractor provided hammer
specifications and driving energy logs. Hydrophone data were obtained at 10 m range from the
center of the pile in 5.5 m deep water. A cushion stiffness value (𝜅) for the force generator model
could not be determined from the hammer specifications; therefore, the stress wave impedance
ratio between the pile and hammer was varied within a physically reasonable range (0.4 <
𝑍𝑝 /𝑍ℎ < 0.8).

Because the construction contractor did not provide the penetration logs for this pile, the
axial reflection coefficient at the pile toe was assumed to be R = −0.5, which yielded a
penetration rate of 0.8 cm per blow. This is a reasonable value, since logs for a 122 cm diameter

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 7


A. MacGillivray A marine pile driving sound model

pile, driven at the same location using the same hammer, recorded penetration rates in the range
0.5–1.5 cm per blow. Furthermore, the modelled SPL and SEL values were not very sensitive to
the assumed value of R, since most of the sound energy is contained in the initial down-going
wave generated by the hammer impact at the top of the pile.

Table 1. Model input parameters for the 2012 Tappan Zee Bridge PIDP test case.
Parameter Value Units
Hammer energy (E) 800 kJ
Pile length (L) 68 m
Pile radius (a) 122 cm
Pile wall thickness (h) 5.08 cm
Ram mass (m1) 45.3 tons
Drive head mass (m2) 18.2 tons
Impedance ratio (Zp/Zh) 0.4–0.8 non-dimensional
Pile stress wave speed (cp) 5 km/s
Pile bulk density (ρp) 8 g/cm3
Pile toe reflection coefficient (R) -0.5 non-dimensional
Water sound speed (cw) 1.463 km/s
Sediment sound speed (cb) 1.562 km/s
Sediment density (ρb) 1.458 g/cm3
Water depth (d) 5.5 m
Monopole separation (Δz) 1 m

The computer programs described in Section 2.5 were used to solve the equations of
motion for the hammer motion and pile vibration (Figures 3 and 4) and to calculate pressure
signatures for an array of equivalent monopoles (Figure 5). A near-field wavenumber integration
model was then used to calculate time-dependent acoustic pressure on a 75 × 75 m radially-
symmetric grid that surrounded the pile (Figure 6).

Model predictions for a receiver at r = 10 m and z = 5 m were compared with hydrophone


data for three different values of the stress wave impedance ratio (𝑍𝑝 /𝑍ℎ = 0.4, 0.6, and 0.8). The
peak SPL, rms SPL, and SEL from the model were a good match to the data for 𝑍𝑝 /𝑍ℎ =0.6
(Figure 7). The modeled spectrum was consistent with hydrophone data at frequencies below
600 Hz, but underestimated the measured spectral energy at higher frequencies (Figure 8).

Figure 3. Force at the top of the pile versus time, as predicted by the model for the Tappan Zee Bridge
PIDP test case (𝑍𝑝 /𝑍ℎ =0.6). The characteristic period of the force pulse from Eqn. 3 is T=5.48 ms.

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 8


A. MacGillivray A marine pile driving sound model

Figure 4. Axial displacement of the pile top, pile toe, and ram (top) and radial velocity of the pile wall
(bottom) versus time, as predicted by the model for the Tappan Zee Bridge PIDP test case (𝑍𝑝 /𝑍ℎ =0.6).

Figure 5. Monopole pressure signatures, versus time and depth, for the Tappan Zee Bridge PIDP test
case (𝑍𝑝 /𝑍ℎ =0.6). The total number of monopoles is N = 64. Note that Gibbs oscillations near the
interfaces result from discontinuities in the radial velocity boundary condition.

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 9


A. MacGillivray A marine pile driving sound model

Figure 6. Contours of instantaneous sound pressure (kPa), versus range and depth, as predicted by the
model for the Tappan Zee Bridge PIDP test case (𝑍𝑝 /𝑍ℎ =0.6).

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 10


A. MacGillivray A marine pile driving sound model

Figure 7. Comparison of measured sound level data with model predictions for the Tappan Zee Bridge
PIDP test case. Model predictions are shown for three different values of the stress wave impedance ratio
(Zp/Zh). Numbers on the right are mean data and numbers of the left are modeled values for 𝑍𝑝 /𝑍ℎ =0.6.

Figure 8. Comparison of measured 1/3-octave band level data with the modeled pressure spectral
density. Model predictions are shown for three different values of the stress wave impedance ratio (Zp/Zh).

4. Discussion and Conclusions


The computational model described in this paper can predict underwater sound levels from
impact hammer driving of cylindrical piles. The model uses readily obtained physical parameters
for the hammer, pile, water, and seabed, and is thus suited to predict acoustic injury zones for
marine life exposed to noise from in-water pile driving operations.

When the model predictions were compared to data, they showed that the model does a
good job of predicting SPL, SEL, and the spectrum at frequencies below 600 Hz. The mismatch
in the spectrum above 600 Hz suggests that the hammer generates more high-frequency vibration
in the pile than the model predicts. This may be related to the spectrum of the forcing function,

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 11


A. MacGillivray A marine pile driving sound model

or it may be related to the excitation of high-frequency vibrational modes in the pile. The high-
frequency mismatch is a subject of ongoing investigation.

In general, model predictions are most sensitive to the value of the stress wave
impedance ratio between the pile and the hammer (𝑍𝑝 /𝑍ℎ ). Model predictions are not as
sensitive to the reflection coefficient at the pile toe (R)—and therefore to the penetration rate—
since the majority of sound energy is produced by the primary stress wave from the hammer,
rather than by secondary stress waves reflected from the pile tips.

The current version of the model neglects several physical effects present when real piles
are driven: multiple seabed layers, skin friction at the pile wall, shear-wave coupling to elastic
seabed layers, and non-linear soil dynamics. Future versions of the model could accommodate
these effects, and, furthermore, could allow for different pile geometries (e.g., sheet piles, solid
piles), more complex hammer models, and different kinds of hammers (e.g., vibratory hammers,
diesel hammers).

5. References
Hall, M. (2013). "A semi-analytical model for non-mach peak pressure of underwater acoustic
pulses from offshore pile driving," Acoustics Australia 41, 42-51.

ICF Jones & Stokes, and Illingworth and Rodkin, Inc. (2009). "Technical Guidance for
Assessment and Mitigation of the Hydroacoustic Effects of Pile Driving on Fish,"
(Prepared for California Department of Transportation (CALTRANS), Sacramento, CA),
p. 298.

Jensen, F. B., Kuperman, W. A., Porter, M. B., and Schmidt, H. (2011). Computational Ocean
Acoustics (AIP Press - Springer, New York).

Junger, M., and Feit, D. (1972). Sound, Structures and their Interaction (MIT Press, Cambridge
Massachusets).

Martin, B., MacGillivray, A., MacDonnell, J., Vallarta, J., Deveau, T., Warner, G., Zeddies, D.,
Mouy, X., and Krebs, J. (2012). "Underwater Acoustic Monitoring of the Tappan Zee
Bridge Pile Installation Demonstration Project: Comprehensive Report," (Technical
report for AECOM by JASCO Applied Sciences), p. 139.

Parola, J. F. (1970). "Mechanics of Impact Pile Driving," in Department of Civil Engineering


(University of Illinois), p. 250.

Reinhall, P. G., and Dahl, P. H. (2011). "Underwater Mach wave radiation from impact pile
driving: Theory and observation," Journal of the Acoustical Society of America 130,
1209-1216.

Proceedings of Meetings on Acoustics, Vol. 20, 045008 (2015) Page 12

You might also like