ArticlePDF Available

Physics-Based Time-Frequency Representations for Underwater Acoustics: Power Class Utilization with Waveguide-Invariant Approximation

Authors:

Abstract

Time-frequency (T-F) analysis of signals propagated in dispersive environments or systems is a challenging problem. When considering dispersive waveguides, propagation can be described by modal theory. Propagated signals are usually multicomponent, and the group delay of each mode (i.e., each component) is nonlinear and varies with the mode number. Consequently, existing T-F representations (TFRs) covariant to group delay shifts (GDSs) are not naturally adapted to this context. To overcome this issue, one solution is to approximate the propagation using simple models for which the dispersion properties do not vary with the mode number. If the chosen model is both simple and robust to uncertainties about the waveguide, it can be used to define adapted TFRs, such as the power-class with a suitable power coefficient. This article focuses on a context where this methodology can be applied: low-frequency acoustic propagation in shallow water. In this case, the global oceanic dispersion can be summarized using a single scalar called the waveguide invariant. This parameter can be used to approximate the group delay of each mode with a power law. Consequently, it is possible to use power-class TFRs with a -based power coefficient. Their practical use is demonstrated on two experimental data sets: a man-made implosion used for underwater geoacoustic inversion, and a right-whale impulsive vocalization that can be used to localize the animal.
IEEE SIGNAL PROCESSING MAGAZINE [120] NOVEMBER 2013 1053-5888/13/$31.00©2013IEEE
Digital Objec t Identifier 10.110 9/MSP .2013 .226 7651
Date of publica tion: 15 October 2013
T ime-frequency (T-F) analysis of signals propagated
in dispersive environments or systems is a challeng-
ing problem. When considering dispersive wave-
guides, propagation can be described by modal
theory. Propagated signals are usually multicompo-
nent, and the group delay of each mode (i.e., each component)
is nonlinear and varies with the mode number. Consequently,
existing T-F representations (TFRs) covariant to group delay
shifts (GDSs) are not naturally adapted to this context. To over-
come this issue, one solution is to approximate the propagation
using simple models for which the dispersion properties do not
vary with the mode number. If the chosen model is both simple
and robust to uncertainties about the waveguide, it can be used
to define adapted TFRs, such as the power-class with a suitable
power coefficient. This article focuses on a context where this
[
Julien Bonnel, Grégoire Le Touzé, Barbara Nicolas, and Jérôme I. Mars
]
[
Power class utilization with
waveguide-invariant approximation
]
Physics-Based
Time-Frequency
Representations for
Underwater Acoustics
T
h
e
o
r
y
a
n
d
A
p
p
l
i
c
a
t
i
o
n
s
o
f
T
i
m
e
-
F
r
e
q
u
e
n
c
y
A
n
a
l
y
s
i
s
© istockphoto.com/–m–i–s–h–a–
IEEE SIGNAL PROCESSING MAGAZINE [121] NOVEMBER 2013
methodology can be applied: low-frequency acoustic propaga-
tion in shallow water. In this case, the global oceanic dispersion
can be summarized using a single scalar b called the waveguide
invariant. This parameter can be used to approximate the group
delay of each mode with a power law. Consequently, it is possi-
ble to use power-class TFRs with a b-based power coefficient.
Their practical use is demonstrated on two experimental data
sets: a man-made implosion used for underwater geoacoustic
inversion, and a right-whale impulsive vocalization that can be
used to localize the animal.
INTRODUCTION
Many physical or technical situations involve the propagation of a
wave through a dispersive medium, e.g., ultrasonic waves in
bones [1], electromagnetic whistlers in the magnetosphere [2],
and ground-penetrating radar [3]. When considering broadband
propagation, each frequency travels with its own speed, and a
propagating nonstationary signal is distorted during its propaga-
tion. Dispersion usually tends to complicate the received signal,
particularly in complex propagation environments that are domi-
nated by multiple modal components, and this often limits the
ability to directly recover information from the signal. On
the other hand, the dispersion effects convey information about
the propagation medium and the source/receiver configuration. If
properly characterized using suitable signal processing methods,
dispersion can be used as the basis of inversion algorithms
(source localization and/or environmental estimation) [4]–[6].
When considering a transient source signal in a single receiver
configuration, the received signal is nonstationary, and T-F analy-
sis is the most common tool for dispersion analysis.
The application field considered in this article is underwater
acoustics, although the proposed methodology can be extended
to other dispersive situations. In particular, the article examines
the practical application of power class TFRs for underwater
acoustics, especially the low-frequency impulsive sources in
shallow water. In this context, the propagation is described by
modal theory. The oceanic environment acts as a dispersive
waveguide, and the received signal contains several nonlinear
frequency-modulated components called propagating modes.
T-F analysis of the received signal and modal filtering are
important issues for the underwater acoustic community. Using
a single receiver, this allows for source localization and environ-
mental estimation [4]–[6]. Dispersion analysis in the T-F
domain has been studied by the signal processing community.
Several approaches have been proposed, mainly for underwater
acoustics [6]–[11], but also for other fields of application [1],
[12]. The majority of these methods are physics-based and inject
a priori environmental knowledge into the T-F processing. For
such methods, the key is to make the best use of the available a
priori information, while at the same time being insensitive to
misknowledge in this a priori information.
Many TFRs have been proposed by the signal processing
community [13], but none of them can be used successfully in
every situation. All the TFRs have advantages and drawbacks,
and are adapted to analyze signals with particular properties and
particular T-F content. The TFRs are thus often classified based
on the properties they satisfy. Critical properties for T-F analysis
are the covariance properties. A TFR is said to satisfy a covari-
ance property if this TFR preserves certain T-F changes in the
signal. It is also said that the TFR is covariant to that property.
As an example, the Cohen’s class is covariant to constant time-
shifts and frequency-shifts, while the affine class is covariant to
constant time shifts and to T-F scaling [13]. However, because of
dispersion, none of these covariance properties are adapted to
our underwater acoustics context. There is the requirement for
covariance properties that take into account the non linear
group delay of the signal considered. Such a covariance prop-
erty is called the GDS covariance, and the corresponding TFR
class was introduced by Papandreou-Suppappola et al. [14], [15].
This article is centered on a subclass of the GDS covariant
class: the power class with an associated power time-shift covari-
ance property [16]–[18]. The power class is thus adapted to disper-
sive signals with power-law group delays. For physical and
underwater acoustic applications, the power class covariance
should allow the inclusion of any available a priori information
about the environment. To do that, the article establishes a link
between the power class TFRs and the waveguide invariant b [19],
[20]. The waveguide invariant is a scalar that summarizes the dis-
persion in the waveguide and allows the approximation of the
phase of each mode as a frequency power function. This connec-
tion allows power class TFRs to be used to analyze the propagation
of impulsive sounds in dispersive multimodal shallow water wave-
guides. This is demonstrated on both simulated and experimental
marine data sets that correspond to both active and passive under-
water acoustics (geoacoustic inversion and bioacoustics).
TIME-FREQUENCY REPRESENTATIONS COVARIANT
TO GROUP DELAY SHIFTS
The group delay shifT covariance
As stated in the introduction, the constant time shift covariance
property of many TFRs is not adapted to the context of disper-
sive propagation. One requires a TFR covariant to dispersive
time shifts corresponding to the nonlinear group delay function
of the medium.
Considering a signal
()
Xf in the frequency domain, the out-
put
()
Yf of an allpass dispersive system is given by ()Yf =
() .Xfe(/ )jff20
rp- The group delay information is given by
() (/ ).fdf
dff
0
xp
= (1)
It is proportional to the derivative of the one-to-one phase func-
tion (/ ),ff
0
p where f0
02 is a normalization frequency. To sat-
isfy the GDS covariance, a TFR T should preserve the
frequency-dependent time shift ().fx In a more general frame-
work, the GDS operator of a signal
()
Xf is ()DXf
() =
a
p
() ,Xfe(/ )jff20
rap- with a a real number. For a given phase func-
tion ,p the parameter a quantifies the amount of dispersion on
the signal ().DXf
()
a
p A TFR T is GDS covariant if the TFR of
the output
()
DXf
()
a
p corresponds to the TFR of the input
()
Xf
IEEE SIGNAL PROCESSING MAGAZINE [122] NOVEMBER 2013
shifted in time by the change in group delay
()
fax introduced
by the GDS operator
(, )( ()
,)
.TtfT
tf
f
DXX
() ax=-
a
p (2)
Successful use of the GDS covariant TFRs requires that the
group delay of the TFR in (2) matches the group delay of the
signal studied. Indeed, significant distortion can occur if this is
not the case [21]. Moreover, the GDS covariance follows directly
from the constant time-shift covariance according to the uni-
tary equivalence principle [22], and thus requires knowledge
about the time origin. These are two strong constraints for
practical/experimental T-F analysis. Both of these will be taken
into account in the context of underwater acoustics for
unknown low-frequency impulsive sources in shallow water.
The power class
A particular part of the GDS covariant class is the power class
[16]–[18]. It is adapted to the case of the power-law covariance
property, for which the phase and group delay functions are
power functions defined as
() () () ,
() () ,
ffsgnf f
ff
ff
f
00
1
pp
xx l
==
==
ll
l
l-
(3)
(4)
where l is the power parameter, f0 is a fixed reference fre-
quency, and
()
sgnf provides the sign of the frequency .f The
GDS covariance in (2) thus becomes the power time-shift cova-
riance. The lth power class TFRs satisfy two specific covariance
properties. As an extension of the affine class, the first covari-
ance property is covariance to scale change
(, )(
,TtfTat a
f
() ()
CXX
a
=
ll (5)
where Ca is the scaling operator that is defined as
() (/ )/
||
.CXf Xf
aa
a= The second covariance property is cova-
riance to power time shifts that matches the signal group delay
(, )( (),)
,,
TtfTtff
Tt ff
ff
() ()
()
DXX
X00
1
ax
al
=-
=-
lll
ll-
a
l
cm
(6)
where Da
l is the power time shift operator given by
() () .DXf Xfe(/ )jff20
=
a
lrap
-l Once again, the parameter a is a real
number that expresses the amount of dispersion in the signal
under study. It will be of particular interest in the section
“Waveguide Invariant-Based Time-Frequency Representations.”
Hlawatsch et al. have shown [18] that any member of the
th
l power class can be expressed as
(, )(/, /)
() () ,
Ttffffffe
Xf Xfdf df
1
()
*
Xjtf
f
f
f
f
12 2
1212
12
#
C=
3
3
3
3
lrl
pp
--
-
ll
eeoo
>
G
##
(7)
where C is a two-dimensional kernel that uniquely character-
izes the TFR. Note that the affine class [13] is a particular case
of the power class for .1l=
The power class can also be obtained using unitary warping
[18], [22]. The power-warping operator is
() (/ )(/)
|||/ |()|/
|,
WXf df
dff Xff
ff
Xf sgnf ff
1
1010
02
1
00
1
pp
l
=
=
lll
l
ll
--
-
^
`
h
j
(8)
where
()
f
1
pl
- is the inverse function of ().fpl Any member of
the power class T
()
l can be computed as an affine TFR T
()
A
(i.e., a TFR of the affine class) of the warped signal (),WXf
l fol-
lowed by a nonlinear T-F coordinate transform
(, )()
,(/),
|/ |,()| /| .
TtfT ff
tfff
Tff
tfsgn fff
() ()
()
W
W
XX
A
X
A
000
01
00
xp
l
=
=
l
ll
l
l
-
l
l
c
c
m
m
(9)
As an example, the power Wigner distribution corresponds to
the case where T
()
A is the Wigner distribution [13] and the pow-
ergram corresponds to the case where T
()
A is the scalogram
[13]. Note that as T
()
A is a member of the affine class, it is a qua-
dratic TFR. Both T
()
A and T
()
l are thus influenced by cross-
terms in the case of multicomponent signals. The cross-term
geometry is detailed in [17].
Power class definition is based on a single power group delay
().fxl Consequently, the power class cannot handle multicom-
ponent signals with different dispersion power laws. However,
power class definition is independent of the parameter .a Con-
sequently, power class is a suitable method to analyze multi-
component signals with different amounts of dispersion a (as
long as the phase function pl is the same for every component).
In the modal propagation context, the received signal contains
several components called modes. Successful use of the power
class requires that 1) each mode has a power group delay; 2) for
a given signal, the power parameter l is the same for every
mode; and 3) ideally, the value of l can be determined using
physical considerations. In the following, we will demonstrate
that it is possible to fulfill these three expectations.
ACOUSTIC PROPAGATION IN SHALLOW WATER
Modal propagaTion
When considering low-frequency acoustic waves in shallow
water, propagation can be described by normal mode theory
[20]. Considering an impulsive source, the received signal
()
Yf
after propagation in the acoustic waveguide over a range r con-
sists in N dispersive waves called modes
() () .Yf Afe()
m
m
N
jf
1
2m
=rz
=
-
/ (10)
The phase function of each mode ()
()
fr
kf
mm
z= depends on
the source/receiver range r and on the waveguide properties
through the horizontal wavenumber ().
kf
m The modal ampli-
tude
()Af
m depends both on the waveguide properties and on
the source/receiver configuration (range and depth). Because of
dispersion, each mode can be associated to a nonlinear group
IEEE SIGNAL PROCESSING MAGAZINE [123] NOVEMBER 2013
delay () (/
)(
).ddf
ff
mm
xz
= If the range and the waveguide
properties are known, the modal phase
()
f
m
z can be computed,
and it is possible to define GDS covariant TFRs that are perfectly
adapted to a given mode .m
When considering an ideal waveguide that consists of an iso-
velocity water column between a pressure-release upper bound-
ary and a rigid bottom, the modal phase can be expressed as [20]
() , ,fc
rff
ff
m
id
wmm
222z=- (11)
where cw is the water sound speed, and fm is the cutoff fre-
quency of mode m that depends linearly on the mode number
.m Equivalently, a time-domain received signal
() ()yt Bte
()
m
N
m
jt
1
2m
=r{
=
-
/
obtained as the inverse Fourier
transform of (10) can be considered. In an ideal waveguide, the
time-domain modal phase is [11], [20]
() (/ ),
/.
tftrctrc
m
id mw w
22
2{=- (12)
Note that from (12), the modal phase depends linearly on .fm It is
thus possible to decompose
()
t
m
id
{ into a mode-dependent
amount of dispersion fm and a mode-independent dispersive
phase (/
).
trcw
22
- If the waveguide properties and range are
known, it is thus possible to define covariant TFRs that are
adapted to the whole signal (i.e., to every mode at once) in an ideal
waveguide. Various studies have used the ideal waveguide (11) and
(12) to perform dispersion-based warping [9], [11]. Le Touzé et al.
also proposed modal-based TFRs using a similar model [10].
Except for this ideal simplistic waveguide, however, the dis-
persion characteristics evolve nonlinearly with mode number,
and it is impossible to define TFRs adapted to the whole
received signal. Although it would be possible to define GDS
covariant TFRs adapted to each mode, this solution does not
appear practical for experimental data, especially if the environ-
ment is not known. The next section presents the waveguide
invariant approximation that allows this issue to be overcome.
The waveguide invarianT
In the guided modal propagation context, interference structures
naturally appear in the
rf
- domain (where r is the source/
receiver range and f is the frequency). These interferences follow
a striation pattern that can be summarized by a single scalar b
called the waveguide invariant [19]. The waveguide invariant has
been widely used in underwater acoustics for various applications,
including passive source localization [23], [24], geoacoustic inver-
sion [25], [26], and active sonar [27], [28].
Classical uses of the waveguide invariant require range diver-
sity (i.e., several source/receiver ranges), which can be obtained
using a horizontal line array or the combination of a single
receiver and a moving source. By its nature, the waveguide invari-
ant b describes physical interferences between modes and sum-
marizes the dispersive behavior of the whole waveguide. However,
in 1999, D’Spain and Kuperman [29] provided an approximation
for single-mode dispersion using the waveguide invariant (see [29,
Appendix A]). In particular, they derived the frequency dependence
of the phase slowness ()
/,
Sf
kf
2
p
mmr= which was extended to
wavenumbers by Gao et al. [30]. Recalling that () (),fr
kf
mm
z=
one obtains the modal phase
()
[]
,fc
r
ff
mm
1
-
zc
-
bb
- (13)
where c is the average sound speed in the water column, and m
c
is a scalar that depends only on the mode number .m When con-
sidering a single receiver and an impulsive source, Gao et al. [30]
showed that it is possible to use (13) as the basis of a “dedispersion
transform.” This transform computes
() ,Ys YA
dedisp s
12
= by
projecting the received signal
()
Yf on the atoms
()
[]
.expAf jc
rfsf2
s
1
r=- -b
-
`
j
(14)
In the following, we will generalize the work of Gao and show
that (13) can be used as the basis of b-based power class TFRs and
warpings. Before this, note that (13) is based on several approxi-
mations that are inherent to the waveguide-invariant approxima-
tion. In particular, the waveguide invariant is actually invariant
only within a group of closely spaced modes and within a fre-
quency interval in which it is possible to define a functional rela-
tionship between group and phase speeds, which does not depend
on mode number [19], [20]. This relationship usually holds true
for low-frequency and long-range propagation in range-
independent shallow water, where the acoustic field consists of a
few modes that are weakly affected by refraction in the water col-
umn [20]. If a range-dependent environment is considered, it is
also required that the spatial variations of the waveguide are suffi-
ciently weak, so that an adiabatic modal approximation is valid
[20], [29]. Note that in classical shallow-water environments, it
can be demonstrated that .1-b In particular, 1b= in the ideal
waveguide presented in the section “Modal Propagation” [31]. The
link between the waveguide invariant and the ideal waveguide, i.e.,
the link between (11) and (13), will be presented in the section
“Comparison with Physics-Based Time-Frequency Representa-
tions Based on the Ideal Waveguide Model.”
WAVEGUIDE INVARIANT-BASED TIME-FREQUENCY
REPRESENTATIONS
power class and waveguide invarianT
According to (13), the modal phase can be decomposed into
a nondispersive component (/)rc f that corresponds to a
plane wave with constant group delay /
rc
t0=
a dispersive component (/ )rc f(/ )
m1
c-b- with group delay
(/)( /) .rc f1(/)
m
11
bc
b
--
As ,1-b the dispersive component corresponds to a disper-
sive wave with high frequencies arriving before low frequencies,
with the limit that infinite frequency arrives first at time .
tt
0
=
One can consider only the dispersive component by making t0 the
new time axis origin. This operation is not a practical problem, as
t0 corresponds to the time when energy first arrives at the receiver.
The modal phase thus becomes
() .fc
rf
mm
1
zc=-
bb
-
u (15)
IEEE SIGNAL PROCESSING MAGAZINE [124] NOVEMBER 2013
Equation (15) shows that the power class with power parameter
/1
lb
=- is adapted to every mode. One can thus define the
b-power phase and group delay
(16)
() ()|| ,
() .
fsgn ff
fff
f1
/1
00
1
p
xb
=
=-
bb
b
b
b
-
-
+
(17)
(In the following, the chosen notations are not consistent with
classical power class notations; however, no ambiguity arises and
these concise notations will be used throughout.) The correspond-
ing b-warping operates as
()
|/ |
|| () .WXf
ff
Xfsgnf f
f
02
100
b
=
bb
b
+
-
cm
(18)
The b-power warping linearizes the modal phase of each mode,
while the b-power TFRs are adapted to the whole received signal.
Note that inverse b-warping is given by .
WW
/
11
=
bb
-
To normalize dispersion to mode 1, i.e., to have ,1
1
a= one
must define
,
fr
c
m
m
0
1
1
c
ac
c
=-
=
b-
cm
(19)
(20)
where f0 is the reference frequency, and m
a is the amount of
dispersion of mode .m As a reminder, the concept of the
amount of dispersion was introduced in (2). Physically, it is a
good thing that the amount of dispersion depends on the
mode number .m In terms of signal processing, the modal
expansion of the acoustic field corresponds to the power signal
expansion over a finite number of power impulses [16], [18]
(each power impulse being a physical mode). As a result, the
inverse Fourier transform of the frequency-warped signal
()
WXf
b shows the modes as a discrete set of impulsions [18]
[see the section “Waveguide Invariant-Based Time-Frequency
Representations” and Figure 1(b)]. As a consequence, we will
call the inverse Fourier transform of the warped frequency
domain the modal domain. In the modal domain, mode m lies
at the position
/.
f
m0
a
coMpuTaTion and inTerpreTaTion
According to (9), any member of the b-power class can be
computed as an affine TFR of the b-power warping signal fol-
lowed by a nonlinear T-F coordinate transform. Until the end
of the article, we focus on the case of the b-powergram, for
which the original affine TFR is the scalogram. This choice
minimizes the cross-terms between the different components
of the signals, i.e., between the modes. This is an important
feature for practical underwater acoustics applications with
more than two propagating modes, as it prevents overlap
between modes and cross terms. In the following, we present
the equations for the b-powergram with power class formal-
ism and show that an equivalent formulation can be obtained
using a physics-based approach.
POWER CLASS FORMALISM
The powergram is obtained when the considered affine TFR T
()
A
is the scalogram. The scalogram (, )Stf
X of the signal
()
Xf can
be computed as the squared magnitude of the wavelet transform
(, )|
,|
,StfXH,
Xt
f2
12
= (21)
where H,
tf
is the time-shifted and scaled version of the wavelet
:H .
))
(|(/)| (( /)
ff
Hf fHffe
,/
tf jtf
012 02
=r
ll
lNote that /
ff
0 is the
scale coefficient. Next, the scalogram of the b-warped signal
(, )
St
f
WX
b can be obtained using unitary equivalence [22]:
(, )| ,||,
|.WW
StfXHXH
,/
,WXt
ft
f
212
12
12
==
bb
b (22)
Finally, the b-powergram (, )Stf
()
X
b is obtained from (9) by the
nonlinear T-F coordinate transform:
/)f/)f
)
(, )()
,(/)
(||
(/ )
(.
StfS ff
tfff
f
fXf f
f
Hf ff
f
ed
f
()
/
*() (
W
XX
jf
tf
000
012
0
1
00
02
2
0
xp
b
p
p
=
=
3
3
b
bb
b
b
b
brxp
-
-
-
+
-
b
bb
ll
ll
l
c
c
m
m
#
(23)
PHYSICAL APPROACH
It is also possible to obtained a b-based scalogram (, )
St
f
()phy
X
with a physical approach that does not use power class for-
malism or unitary equivalence. The change in formalism
does not provide a better TFR, but it helps to understand the
equations and to physically interpret the results. To compute
(, ),
St
f
()phy
X the wavelet must be adapted to the studied signal.
This can be done using a wavelet family for which the wave-
let phase corresponds to the modal phase. As the modes have
the phase () (/ ),
ff
f
mm0
zap=b one can choose the wavelet
family
{}
G,mf with
)/)f
)(( ,Gf Gf e
,(
mf f
jf
2m0
=ra p-b
ll l (24)
where )(Gf
fl is a window function that is concentrated
around frequency .f As an example, )(Gf
fl can be a scaled
version of the wavelet G: .
))
(|(/)| (( /)
ff
Gf fG
ff
/
f012 0
=
ll
This
choice allows the easy computing of (, )
St
f
()phy
X by modifying
an existing scalogram code. Whatever the choice of )
(,
Gf
fl
the family
{}
G,mf leads to a mode-frequency wavelet
transform
(,), ,SfXG
() ,
mod
X
mm
f
12
a= (25)
where the square magnitude is similar to ,SWX
b apart from a
compressional factor /f10 in the modal dimension. Conse-
quently, mode m lies at position m
a in the modal dimension
associated to ,S
()
mod
X while it lies at position /f
m0
a in the modal
dimension associated to .SWX
b
The final step to obtain a b-based scalogram (, )
St
f
()phy
X
using the physical approach is to go from the mode-frequency
domain to the T-F domain. This is achieved by substituting m
a
with the time t that corresponds to the mode m group delay
() /().tf
tf
mm
+ax
ax==
bb
(26)
IEEE SIGNAL PROCESSING MAGAZINE [125] NOVEMBER 2013
The following b-based T-F scalogram is thus obtained:
/)f
))
(, )| (/ (),)|
(( .
StfStff
Xf Gf
ed
f
() ()
*() (
phymod
XX
fjf
tf
2
20
x=
=
3
3
b
rxp
-
-
bb
2
ll l
l
# (27)
Note that in (27), the phase in the integral is equal to the
phase in the integral for powergram computation [see
(23)]. Moreover, Gf can be chosen so that the wavelet
transform
S()
mod is invertible (it must thus respect the clo-
sure condition). Note also that Gf can be chosen so that
.
SS
()
()
phy
=
b The corresponding Gf is obtained by compar-
ing (27) and (23).
siMulaTion exaMple
The proposed methodology is applied to realistic simulated data
that mimics shallow water conditions during the summer. The
configuration considered consists of a 100-m-deep water col-
umn over a layered bottom. The sound speed profile in the
water column is as follows:
constant sound speed 1,525 m/s from the surface z0=
m to the depth z45= m
linear variation 1,525 m/s "1,495 m/s from z45= m to
z55= m
constant sound speed 1,495 m/s from z55= m to the
bottom z100= m.
The seabed is modeled as a fluid (a classical assumption in
underwater acoustics). It consists of a homogeneous sediment
layer on top of a homogeneous semi-infinite basement. The
seabed parameters are
sediment: width 10 m, sound speed 1,800 m/s, density 1.8
basement: sound speed 2,000 m/s, density 2.
The source/receiver range is r10= km. Both the source and
the receiver lie on the bottom, so that the modes are (nearly)
equally excited. The frequency-domain impulse response of the
waveguide was computed using the ORCA modal code [32]. The
source is assumed to be perfectly impulsive. The received signal is
thus the simulated impulse response of the waveguide. No noise
has been added to the simulation, so that the results are more
easily understood. The robustness of the method against noise
will be demonstrated on real data in the next section.
The scalogram of the simulated signals is presented in Fig-
ure 1(a) with mode numbers labeled in white. Figure 1(a)
shows seven modes with relatively good T-F resolution. High-
order modes are closer to each other than low-order modes,
and they suffer from T-F interference. Mode 1, which has the
lowest frequency content, suffers from poor time resolution.
Figure 1(b) shows the scalogram of the inverse Fourier trans-
form of the b-warped signal. As explained in the section
“Waveguide Invariant-Based Time-Frequency Representa-
tions,” warped modes are impulsive signals. Note that warped
modes are not perfect impulses, which shows that the wave-
guide invariant, and particularly (13), is an approximation.
However, waveguide invariant approximation is accurate
enough such that the T-F representation of the warped modes
shows them well separated. The b-warping is thus a useful
tool for modal filtering, which is an important issue for the
underwater-acoustics community. Indeed, filtered modes serve
as input for matched mode processing [33] and other inver-
sion schemes [5], [34]. Figure 1(c) shows the b-powergram.
The seven modes are resolved with good resolution. Note that
both power warping and the powergram were computed using
..
12b= This value was chosen empirically to match the T-F
curvature of the modes.
As stated in the section “The Waveguide Invariant,” the wave-
guide-invariant value is usually 1-b in shallow water. However,
b is known to change according to the environmental conditions
[35]. For example, b can be linked to seabed properties in a
100
Frequency (Hz)
90
80
70
60
50
40
30
20
10
0 0.5 1
Time (s)
1.5
100
Frequency (Hz)
90
80
70
60
50
40
30
20
10
0 0.5 1
Time (s)
1.5
60
Warped Frequency (Arbitrary Unit)
55
50
45
40
35
30
25
20
15 02 4
Modal Dimension (Arbitrary Unit)
(a) (c)(b)
6
1
2
2
2
3
3
3
4
4
4
5
5
5
6
1
1
7
6
6
7
7
[FIG1] The simulated results. (a) A scalogram of the simulated signal. (b) A scalogram of the inverse Fourier transform of the b-warped
signal. (c) b-powergram. The power warping and powergram were computed using
..
12b= The scalograms and powergram were
computed using Morlet wavelets.
IEEE SIGNAL PROCESSING MAGAZINE [126] NOVEMBER 2013
Pekeris waveguide [36]. Depending on the available a priori
knowledge of the environmental context, the b value can be cho-
sen empirically or defined accurately using the knowledge of the
environment. In either case, it is important that the correspond-
ing TFR is robust to a relative mismatch in the b value. The
robustness of the b-powergram is considered in Figure 2. It pres-
ents powergrams that were computed for b= 0.5, 1, 1.5, and 3.
All of these powergrams show a good T-F resolution for each
mode, which demonstrates that the method is robust to mis-
match in the b value. However, Figure 2(a) and (d), which corre-
sponds to b=0.5 and 3, respectively, shows relatively poor T-F
resolution for high-order modes. Indeed, these modes are the
modes that interact most with the seabed, which explains why
they are more sensitive to an incorrect b value (and thus to an
incorrect environmental model).
coMparison wiTh physics-based
TiMe-frequency represenTaTions
based on The ideal waveguide Model
The b-power class provides a general and theoretical framework
to analyze the signals received in a dispersive waveguide. In this
section, this is compared with other physics-based solutions that
allow for dispersion T-F analysis [7], [8], [10]. Physics-based TFRs
inject a priori knowledge about the waveguide properties to mod-
ify the T-F tiling. This knowledge is usually based on a simple
waveguide model [7], [10], although it can be adaptively refined
using optimization [8]. The most common model for dispersion-
based signal processing is the ideal waveguide presented in the
section “Modal Propagation” [7], [9], [11]. It is thus interesting to
compare it with our b-based model.
The modal phase in the ideal waveguide is given by (11), while it
is given by (13) under the waveguide-invariant model. If the fre-
quency f is well above the cutoff frequency ,fm the ideal model and
the waveguide-invariant model with 1b= are both equivalent to a
nondispersive wave that would travel at the water sound speed
() ~()~ .
ff
c
rf
m
id
fmfw
1
zz
""33
b= (28)
This is consistent with modal propagation. In shallow water, high
frequencies correspond to low grazing angles (the energy propa-
gates nearly horizontally), while low frequencies correspond to
high grazing angles (the energy propagates nearly vertically).
Consequently, high frequencies are not strongly influenced by the
waveguide properties. In the limit of infinite frequency, the waves
travel horizontally and the waveguide does not influence propaga-
tion anymore (there is no dispersion). On the other hand, low fre-
quencies interact strongly with the bottom. It is thus necessary to
model the seabed correctly to study low-frequency dispersion,
which is the superiority of the waveguide-invariant model.
Indeed, the ideal waveguide model has a perfectly rigid bottom
100
Frequency (Hz)
80
60
40
20
00.5 1
Time (s)
1.5
100
Frequency (Hz)
80
60
40
20
00.5 1
Time (s)
1.5
(a) (b)
(c) (d)
100
Frequency (Hz)
80
60
40
20
00.5 1
Time (s)
1.5
100
Frequency (Hz)
80
60
40
20
00.5 1
Time (s)
1.5
[FIG2] The simulated b-powergram dependence on the value of .b (a)
..
05b= (b) .1b= (c)
..
15b= (d) .3b= The powergrams
were computed using Morlet wavelets.
IEEE SIGNAL PROCESSING MAGAZINE [127] NOVEMBER 2013
that does not model seabed propagation (i.e., no energy pene-
trates the seabed) while the waveguide invariant conveys informa-
tion about a more realistic seabed [25], [ 36]. Consequently, the
b-based TFRs can be used to perform accurate dispersion analy-
sis in complex waveguides. This is particularly suitable when sea-
bed interactions are important and can be modeled using a
suitable b value.
Waveguide invariant approximation, however, does not cor-
rectly model frequency dispersion between the cutoff frequency
and the Airy phase. For a given mode, the Airy phase is the fre-
quency at which the group speed is minimum, and thus it corre-
sponds to the last energy arrival in the T-F domain. The energy
that corresponds to frequencies between the cutoff frequency and
the Airy phase was called ground wave by Pekeris [37]. It propa-
gates in the seabed with high group speed and high attenuation.
Its T-F content is a slow upsweep going from cutoff frequency to
Airy phase, creating a T-F turnaround at the Airy phase. As an
example, mode 3 cutoff frequency and Airy phase can be seen in
Figure 1(a) at 28 Hz and 31 Hz, respectively. As ground waves are
highly attenuated, they are usually masked by noise in experi-
mental marine data (see Figures 3 and 4). However, ground waves
can be encountered when considering powerful explosions [4].
Analysis of the T-F turnaround associated with ground waves can-
not be achieved for all modes at once. This would require the
GDS covariant TFR based on accurate single mode dispersion.
EXPERIMENTAL DATA APPLICATION
This section demonstrates the practical use of the b power class
on experimental data. Two different underwater acoustic con-
texts are considered: active geoacoustic inversion and passive
study of marine mammal vocalization. In both cases, the con-
sidered source is impulsive and the environment is a complex
waveguide (nonisovelocity sound speed profile in the water col-
umn, layered seabed, and weak range dependence).
lighTbulb daTa: shallow waTer 2006 experiMenT
First, we consider an experimental signal that was recorded on
the continental shelf of New Jersey (United States) during the
Shallow Water 2006 Experiment [38]. The considered source is
a lightbulb implosion, and the corresponding acoustic field was
recorded by the Marine Physical Laboratory vertical line array (a
single hydrophone is used for this study). In this active context,
the T-F dispersion of the signal can be used to infer information
about the seabed properties (geoacoustic inversion). More
details about the environment and this specific signal can be
found in [5].
The source/receiver range, r7- km, is relatively short for
such a low-frequency inversion configuration. The scalogram of
the received signal is presented in Figure 3(a). The modes are
labeled in white for easier understanding. It can be seen that they
are close to each other and interfere together. Classical scalo-
grams cannot be used to infer information about the modal dis-
persion. The b-powergram
(.
)125b= is presented in Figure
3(b). It shows better resolution than the classical scalogram, and
it can be used to infer dispersion information. The value
.125b= has been chosen empirically to match the T-F curva-
ture of the modes.
righT-whale vocalizaTion: fundy bay daTa
Next, we consider an experimental signal that was recorded in
Fundy Bay (Canada) as part of the 2003 Workshop on Detec-
tion, Localization, and Classification of Marine Mammals
Using Passive Acoustics [39]. The source considered is a right-
whale impulsive vocalization that was recorded on a single
bottom-moored hydrophone (labeled H during the experi-
ment). In this passive context, the T-F dispersion of the signal
can be used to estimate the range between the whale and the
receiver. More information about the environment can be
found in [39], and about this specific signal in [40].
(a)
0.05 0.1 0.15
Time (s)
1.2
(b)
0.05 0.1 0.15
Time (s)
0
240
220
200
180
160
Frequency (Hz)
140
120
100
80
60
40
240
220
200
180
160
Frequency (Hz)
140
120
100
80
60
40
11
2
2
3
3
4
4
[FIG3] SW06 experimental results. (a) Scalogram of the received signal. (b) b-powergram of the received signal, computed using
.125b=. The scalogram and the powergram were computed using Morlet wavelets.
IEEE SIGNAL PROCESSING MAGAZINE [128] NOVEMBER 2013
In this passive context, the source/receiver range is not known.
It has been estimated in various studies, and it should be r8-
km. The scalogram of the received signal is shown in Figure 4(a).
It is nearly impossible to distinguish the modal content of the sig-
nal, and the modes are labeled in white to facilitate understanding.
Note that because of a low-speed sediment layer, mode 1 is proba-
bly trapped in the seabed and so does not appear in the T-F
domain. The b-powergram
(.)08
b= is shown in Figure 3(b),
and it has relatively good modal resolution. The b-scalogram is
thus a valuable tool to analyze experimental low-frequency data,
even when they are recorded in a passive context and in a compli-
cated environment. Note that the value .
08
b= was chosen
empirically to match the T-F curvature of the modes.
CONCLUSIONS
The power-class TFRs, or, more generally, the GDS covariant
TFRs, are suitable TFRs to analyze signals with dispersive group
delay. However, their use is limited to monocomponent signals,
or to multicomponent signals for which the dispersive proper-
ties do not vary from one component to the other. Successful
use of the power-class TFRs to analyze signals propagated in a
dispersive medium requires that the dispersion can be modeled
using a power law. Another desirable property is that the power
coefficient l can be defined using robust physical a priori infor-
mation about the medium.
This article focuses on dispersive propagation in an oceanic
waveguide; especially the low-frequency impulsive sounds propa-
gated in shallow water. This context is particularly challenging as
propagation is described by modal theory. Indeed, the signals con-
sidered are multicomponent, and each component (called a
mode) has a nonlinear group delay that varies from mode to
mode. However, it is possible to approximate the waveguide dis-
persion with a power law ,f/1 b- where b is the waveguide
invariant. The waveguide invariant is a scalar widely used in
underwater acoustics that characterizes propagation in shallow
water. It is possible to define a class of TFRs that are adapted to the
modal propagation context. This class benefits from the formalism
of the power class (with power coefficient
/)
,1
lb
=- while at the
same time allowing the inclusion of robust physical a priori infor-
mation through the waveguide invariant .b Note that single-
receiver modal filtering using b-warping can be obtained as a
by-product of the power class methodology.
This method is illustrated both on realistic simulation data
and experimental data. Two marine data sets that correspond to
different underwater acoustic contexts are considered: active
geoacoustic inversion and passive bioacoustics. In these difficult
practical situations, the proposed method allows us to resolve
several modes in the T-F domain, which was impossible with
classical TFRs. The b-power class is thus a promising method
for analysis of underwater acoustic single-receiver data.
AUTHORS
Julien Bonnel (julien.bonnel@ensta-bretagne.fr) received the
Ph.D. degree in signal processing from Grenoble-Institut
National Polytechnique (INP), France, in 2010. Since 2010, he
has been an assistant professor at Lab-STICC (UMR 6285), Ecole
Nationale Supérieure de Techniques Avancées Bretagne in
Brest, France. His research in signal processing and underwater
acoustics include time-frequency analysis, source detection/
localization, geoacoustic inversion, acoustical tomography, pas-
sive acoustic monitoring, and bioacoustics. He is a Member of
the IEEE and the Acoustical Society of America.
Grégoire Le Touzé (gregoire.le-touze@gipsa-lab.grenoble-
inp.fr) received the master’s degree in signal processing from
Grenoble-INP, France, in 2004 and the Ph.D. degree in 2007.
From 2008 to 2010, he was a postdoctoral researcher at the
[FIG4] The Fundy Bay experimental results. (a) Scalogram of the received signal. (b) b-powergram of the received signal, computed
using
..
08b= The scalograms were computed using Morlet wavelets.
(a)
0.2 0.4 0.6
Time (s)
0.8
(b)
0.200 0.4 0.6
Time (s)
0.8
100
90
80
70
60
Frequency (Hz)
50
40
30
20
10
100
90
80
70
60
Frequency (Hz)
50
40
30
20
10
2
3
4
2
3
4
IEEE SIGNAL PROCESSING MAGAZINE [129] NOVEMBER 2013
Laboratoire de Mécanique et d’Acoustique, Marseille, France,
and from 2010 to 2012 at the GIPSA-Lab, Grenoble, France. His
research interests include seismic and acoustic signal process-
ing, wavefield separation methods, time-frequency and chirplet
anaysis, and array processing.
Barbara Nicolas (barbara.nicolas@gipsa-lab.grenoble-inp.fr)
received the M.S. and Ph.D. degrees in signal processing from
Grenoble-INP, France, respectively, in 2001 and 2004. In 2005,
she held a postdoctoral position at the French Atomic Energy
Commission (C.E.A.) in medical imaging. Since 2006, she has
been a CNRS researcher at the GIPSA-Lab in the Image Signal
Department, Grenoble, France. In 2008, she was a visiting sci-
entist at SipLab in Faro, Portugal. Her research in signal pro-
cessing and underwater acoustics include TFRs, array
processing, acoustic tomography, underwater source detection
and localization, and geoacoustic inversion.
Jérôme I. Mars (jerome.mars@gipsa-lab.grenoble-inp.fr)
received the Ph.D. degree in signal processing (1988) from the
Institut National Polytechnique of Grenoble. From 1989 to 1992,
he was a postdoctoral researcher at the Centre des Phénomènes
Aléatoires et Geophysiques, Grenoble. From 1992 to 1995, he was
a visiting lecturer and scientist at the Materials Sciences and Min-
eral Engineering Department at the University of California,
Berkeley. He is currently a professor of signal processing in the
Image-Signal Department at GIPSA-Lab (UMR 5216), Grenoble
Institute of Technology. He is the head of the Image-Signal
Department. His research interests include seismic and acoustic
signal processing, source separation methods, time-frequency,
and time-scale characterization. He is a Member of the IEEE and
the European Association of Geoscientists and Engineers.
REFERENCES
[1] K. Xu, D. Ta, a nd W. Wang , “Multiridge-based analysis for separating individual
modes from multimodal guided wave signals in long bones,” IEEE Trans. Ultrason.
Ferroelectr., Freq. Control, vol. 57, no. 11 , pp. 24802490, 2010 .
[2] D. Mihovilovic´ and R. Bracewell, “Whistler analysis in the time-frequency plane
using chirplets,” J. Geophys . Res., vol. 97, no. A 11, pp. 1719 9 17204, 1992.
[3] J. der Kr uk, N. Diamanti, A. Giannopoulos, and H. Vereecken, “Inversion of
dispersive GPR pulse propagation in waveguides with heterogeneities and rough
and dipping interfaces,” J. Appl. Geophys., vol. 81, pp. 88–96 , June 2011.
[4] G. Potty, J. Miller, J. Lynch, and K. Smith, “ Tomographic inversion for sediment
parameters in shallow water,” J. Acoust. Soc. Amer., vol. 10 8, no. 3, pp. 973986, 2000.
[5] J. Bonnel, C. Gervaise, B. Nicolas, a nd J. Mars, “Single receiver geoacoustic inver-
sion using modal reversal,” J. Acoust. Soc. Amer., vol. 131, no. 1, pp. 119128, 201 2.
[6] C. Ioa na, A. Jarrot, C. Gervaise, Y. Stéphan, and A. Quinquis, “Localization
in underwater dispersive channels using the time-frequency-phase continuity of
signals,” IEEE Tran s. Signal Processing, vol. 58 , no. 8, pp. 40934107, 2010.
[7] C. Chen , J. M iller, G. Boudreaux-Bartels, G. Potty, and C. Lazauski, “Time-
frequency representations for wideband acoustic signals in shallow water,” in Proc.
OCEANS 2003, vol. 5, pp. SP2903SP2907.
[8] J. Hong, K. Sun, a nd Y. Kim, “Dispersion-based short-time Fourier transform
applied to dispersive wave analysis,” J. Acoust. Soc. A mer., vol . 117, no. 5, pp. 2949
2960, 2005.
[9] Y. Jiang and A. Papandreou-Suppappola, “Discrete time-frequency character-
izations of dispersive linear time-varying systems,” IEEE Trans . Signal Processin g,
vol. 55, no. 5, pp. 2066 2076, 2007.
[10] G. Le Tou , B. Nicolas, J. Mars, and J. Lacoume, “Matched representa-
tions and filters for guided waves,” IEEE Trans. Signal Processing, vol. 57, no. 5,
pp. 178 317 95 , 2009.
[11] J. Bonnel, B. Nicolas, J. Mars, and S. Wal ker, “Estimation of modal group
velocities with a single receiver for geoacoustic inversion in shallow water,” J.
Acoust. Soc. Amer., vol. 128, no. 2, pp. 719 727, 2010.
[12] L. De Marchi, M. Ruzzene, B. Xu, E. Baravelli, and N. Speciale, “Warped basis
pursuit for damage detection using lamb waves,” IEEE Trans. Ultrason. Ferro-
electr., Freq. Control, vol. 57, no. 12, pp. 27342741, 2 010.
[13] F. Hlawatsch an d G. Boudreaux-Bartels, “Linear and quadratic time-frequency
signal representations,” IEEE Signal Process. Mag., vol. 9, no. 2, pp. 2167, 1992.
[14] A. Papandreou-Suppappola, “Generalized time-shift covariant quadratic time-
frequency representations with arbitrary group delays,” in Proc. 29th A silomar
Conf. Signals, Systems a nd Computers, 1995, pp. 5535 57.
[15] A. Papandreou-Suppappola, R. Murray, B. Iem, a nd G. Boudreaux-Bartels,
Group delay shift covariant quadratic time-frequency representations,” IEEE
Trans. Sig nal Proces sing, vol. 49, no. 11, p p. 25492564, 2 001.
[16] F. Hlawatsch, A. Papandreou, and G. Boudreaux-Bartels, “The power classes
of quadratic time-frequency representations: A generalization of the affine and hy-
perbolic classes,” in Proc. 27th Asilomar Conf. Signals, Systems a nd Computers,
1993, pp. 1265–1270.
[17] A. Papandreou-Suppappola, F. Hlawatsch, a nd G. Boudreaux-Bartels, “Power
class time-frequency representations: Interference geometry, smoothing, and
implementation,” in Proc. IEEE -SP Int. Sym p. Time-Fre quency and Time-Scale
Analysis, 1996, pp. 193196 .
[18] F. Hlawat sch, A. Papandreou-Suppappola, a nd G. Boudreaux-Bartels, “The
power classes-quadratic time-frequency representations with scale covariance and
dispersive time-shift covariance,” IEEE Trans. Signal Processing, vol . 47, no. 11,
pp. 30673083, 1999.
[19] S. Chuprov, “Interference structure of a sound field in a layered ocean,” in
Ocean Acoustics, Modern State, 1982 , pp. 7191.
[20] F. Jensen, W. Kuper man, M. Por ter, and H. Schmidt, Computational Ocean
Acoustics, 2nd ed. Ne w Yor k: Amer ican Institute of Physics, 2011.
[21] A. Papandreou-Suppappola and G. Boudreaux-Bartels, “The effect of mis-
matching analysis signals and time-frequency representations,” in Proc. IEEE-SP
Int. Symp. Time-Freque ncy and Time- Scale Analysis, 1996 , pp. 149 152.
[22] R. Baraniuk a nd D. Jones, “Unitary equivalence: A new twist on signal process-
ing,” IEEE Tran s. Signal Processing, vol. 43, no. 10, p p. 22692282, 1995.
[23] K. Cockrell and H. Schmidt, “Robust passive range estimation using the wave-
guide invariant,” J. Acoust. So c. Amer., vol. 127, no. 5, pp. 27802789, 2010.
[24] S. Rakotonarivo and W. Kuper man, “Model-independent range localization
of a moving source in shallow water,” J. Acoust. Soc. Amer., vol. 132, no. 4, pp.
22182223, 2012.
[25] K. Heaney, “Rapid geoacoustic characterization using a surface ship of oppor-
tunity,” IEEE J. Oceanic E ng., vol. 29, no. 1, pp. 8899, 2004.
[26] Q. Ren and J. Her mand, “Acoustic striation processing for ocean bottom char-
acterization,” in Proc. OCE ANS, 2011 IEEE-Spain, 2 011, pp. 1–6.
[27] R. Goldhahn, G. Hickman, and J. Krolik, “A waveguide invariant adaptive
matched filter for active sonar target depth classification,” J. A coust. Soc. A mer.,
vol. 129, no. 4, pp. 1813 1824, 2 011.
[28] D. Rouseff and L. Zurk, “Striation-based beamforming for estimating the
waveguide invariant with passive sonar,” J. Acoust. Soc. Amer., vol. 130, no. 2,
pp. EL76 EL81, 2 011.
[29] G. D’Spain and W. Kuperma n, “Application of waveguide invariants to analysis
of spectrograms from shallow water environments that vary in range and azimuth,”
J. Acoust. Soc . Amer., vol. 10 6, no. 5, pp. 24542468, 1999.
[30] D. Gao, N. Wang, and H. Wang, “A dedispersion transform for sound propagation
in shallow water waveguide,” J. Comput. Acoust., vol. 18, no. 3, pp. 245257, 2010.
[31] L. Brekhovskikh, Y. Lysanov, I. Lysanov, and I. Lysanov, Fundamentals of
Ocean Acoustics. Ne w York: Spr inger Verlag, 2003.
[32] E. Westwood, C. Tindle, and N. Chapman, “A normal mode model for acousto-
elastic ocean environments,” J. Acoust. So c. Amer., vol. 100, no. 6, p. 3 631, 199 6.
[33] T. Ya ng , “Effectiveness of mode filtering: A comparison of matched-field and
matched-mode processing,” J. Acoust. So c. Amer., vol. 87, no. 5, p. 2072, 1990.
[34] A. Newhall, Y. Lin, J. Lynch, M. Baumgartner, and G. Gawarkiewicz, “Long
distance passive localization of vocalizing Sei whales using an acoustic normal
mode approach,” J. Acoust. So c. Amer., vol. 1 31, no. 2, p. 1814 , 2012 .
[35] D. Rouseff and R. Spindel, “Modeling the waveguide invariant as a distribu-
tion,” AIP Conf. Proc., vol. 621, pp. 137–15 0, May 2001.
[36] E. Shang, J. Wu, and Z. Zhao, “Relating waveguide invariant and bottom
reflection phase-shift parameter P in a Pekeris waveguide,” J. Aco ust. Soc. Ame r.,
vol. 131, no. 5, pp. 36913697, 2012.
[37] C. Pekeris , “Theory of propagation of explosive sound in shallow water,” Mem-
oir Geol. Soc. Amer., vol. 27, 1948.
[38] D. Ta ng, J. Moum, J. Lynch, P. Abbot, R. Chapm an, P. Dahl, T. Duda, G.
Gawarkiewicz, S. Glenn, J. Goff, H. Graber, J. Kemp, A. Maffei, J. Nash, and A.
Newhall, “Shallow water 06: A joint acoustic propagation/nonlinear internal wave
physics experiment,” Oceanography, vol. 20, no. 4, pp. 156–167, 20 07.
[39] F. Desharnais, M. Laurinolli, D. Schillinger, and A. Hay, “A description of the
workshop data sets,” Can. A coust., vol. 32, no. 2 , pp. 3338, 2004.
[40] J. Bon nel, G. Le Tou , B. Nicolas, J. Mars, and C. Gervaise, “Automatic
and passive whale localization in shallow water using gunshots,” in Conf. IEEE
OCEANS 2008, 2009, pp. 1–6.
[SP]
... It can be concluded that the propagation of broadband pulse signals in shallow water shows a frequency dispersion phenomenon (for a sufficiently large propagation distance). As a result, the signal waveform is elongated in the time domain [35]. ...
Article
Full-text available
In marine environments with strong pulse interference, distinguishing water entry sounds from pulse interference with high accuracy presents a challenge. Conventional algorithms that only utilize the singular values of intrinsic mode functions (IMFs) as the classification feature can have many misclassified signals in noisy environments. To identify the water entry sound more precisely, this study introduces a classifier based on selected singular value and correlation coefficient features. Specifically, the correlation coefficient between each IMF and the original signal are used to select crucial singular values and eliminate uninformative components. Furthermore, the correlation coefficient also combines with singular values as classification features for water entry sounds. The classification experiment results of 2460 groups of water entry sound and pulse interference indicate that the proposed classifier improves classification accuracy by approximately 11.5% compared to using singular values alone and by approximately 2.4% compared to classification accuracy without eliminating uninformative IMFs.
... Transforming y obs ðtÞ with the warping function vðtÞ ¼ ðt 2 þ t 2 r Þ 1=2 as shown in Eq. (1) has been successfully used to isolate normal modes in shallow-water fields. [13][14][15][16] The signal time origin is t r ¼ r=c w , with c w being the sound speed in the waveguide. Bonnel et al. 2 have shown that the warping is a robust transformation, which can be applied to most low-frequency shallow-water scenarios, even in non-ideal waveguides. ...
Article
Full-text available
This Letter proposes a joint geoacoustic inversion method for modal group velocity dispersion and amplitudes of waveform by incorporating a Pearson correlation constraint. Numerical simulations show that this joint inversion leads to improved geoacoustic inversion performance with smaller uncertainties compared to separate inversion methods when applied to data from a single receiver. Additionally, the effective use of the Wasserstein metric from optimal transport theory is explored and compared to the more-common L2 norm misfit measure. The Letter also presents a qualitative representation of joint inversion convergence obtained through multiple independent runs of genetic algorithms. The algorithm is applied to simulated data.
... The main objective of warping is to linearize the mode phase φ m and facilitate modal separation of waveguide signals. Warping transform maps y(t) into the warped signal W h y(t) W h y obs (t) = |χ (t)|y obs [χ(t)], (8) where the warping function χ(t) = (t 2 + t 2 r ) 1/2 has a derivative χ (t) relative to t [14,20,28,29]. t r = r/c w is the signal time origin, with c w the sound speed in the waveguide. ...
Article
Full-text available
The inversion of acoustic field data to estimate geoacoustic parameters has been a prominent research focus in the field of underwater acoustics for several decades. Modal-dispersion curves have been used to inverse seabed sound speed and density profiles, but such techniques do not account for attenuation inversion. In this study, a new approach where modal-dispersion and waveform envelope data are simultaneously inversed under a multi-objective framework is proposed. The inversion is performed using the Multi-Objective Bayesian Optimization (MOBO) method. The posterior probability densities (PPD) of the estimation results are obtained by resampling from the exploited state space using the Gibbs Sampler. In this study, the implemented MOBO approach is compared with individual inversions both from modal-dispersion curves and the waveform data. In addition, the effective use of the Wasserstein metric from optimal transport theory is explored. Then the MOBO performance is tested against two different cost functions based on the L2 norm and the Wasserstein metric, respectively. Numerical experiments are employed to evaluate the effect of different cost functions on inversion performance. It is found that the MOBO approach may have more profound advantages when applied to Wasserstein metrics. Results obtained from our study reveal that the MOBO approach exhibits reduced uncertainty in the inverse results when compared to individual inversion methods, such as modal-dispersion inversion or waveform inversion. However, it is important to note that this enhanced uncertainty reduction comes at the cost of sacrificing accuracy in certain parameters other than the sediment sound speed and attenuation.
... Though the received signal is complicated after the propagation in dispersive underwater channel, the information of source spectrum and environment is combined in received signals. 3,5 Once this information is obtained accurately, both for passive and active application scenarios, many underwater acoustical problems could be solved such as environmental inversion, 6-8 communication, 9,10 source localization [11][12][13][14] and identification. 15 Consequently, the key is to separate the Green's function and the source spectrum from the field and reconstruct their characteristic information. ...
Article
Full-text available
In shallow water, there exist regular interference striations on the structured sound intensity pattern versus frequency and element location of a horizontal array. Beamforming along the striation with frequency-shifted pressure can improve the correlation among array signals, but the beam focusing performance is limited by the source spectrum (e.g., random phase spectrum generated from ship-radiated noise) and signal truncation. In this paper, the cross-correlation phase relationship among the array signals is deduced, and striation-based beamforming with such cross-correlated data can produce a similar directivity to the conventional beamforming, while the effect from signal truncation and source spectrum can be eliminated. This striation-correlation-based beamforming also outputs the signal corresponding to the Green’s function between the source and the reference array element, moreover, recovers the interference structure that characters the channel. The striation-based array signal processing also provides the proposed Green’s function estimation method with the ability to work effectively under the condition of low signal-to-noise ratio. Simulation results are presented and performance variations related to signal frequency, signal-to-noise ratio, striation slope estimation error, element number, and array aperture are also analyzed numerically.
... Ainsi la méthode proposée se démarque par exemple de méthodes ayant besoin de connaître la nature de la source ou ne fonctionnant que pour certains types de sources comme les méthodes de warping (méthodes utilisées pour la représentation en temps -82 4.2. Etude de performances fréquence) [Bonnel et al., 2013]. Cependant, la prise en compte du bruit se fait sur le plan fréquence -distance, notre algorithme suppose donc que la source dispose d'un spectre relativement blanc. ...
Thesis
L’objectif de ces travaux de thèse est de proposer une méthode permettant de reconstruire les nombres d’onde associés à des modes propagatifs au fil des fréquences dans le cas de sources large-bande émettant dans un milieu petit fond. Nous présentons dans un premier temps la propagation acoustique régissant notre cas d’étude, ce faisant nous pourrons décrire le signal reçu par un capteur comme la somme d’un petit nombre de composantes modales. Tant que l’acquisition du signal se fait de façon régulière dans le domaine spatial, une analyse de Fourier classique permet d’estimer ces nombres d’onde.Toutefois l’approche par transformée de Fourier est sujet à la disposition d’un nombre important de capteurs ce qui résulte au déploiement d’antenne de plusieurs kilomètres de long.Nous verrons ensuite une nouvelle méthode de traitement de signal permettant de reconstruire un signal en disposant d’un plus petit nombre de mesures que celles nécessaires pour l’analyse de Fourier. Après un état de l’art sur les méthodes de reconstruction parcimonieuse utilisées dans le cadre de la reconstruction de nombre d’onde nous aborderons le cas des approches continues permettant d’affiner la détection des nombres d’onde. La prise en compte d’une relation physique liant les nombres d’onde d’un même mode à deux indices fréquentiels successifs dans une approche continue de type gloutonne nous permet de proposer l’algorithme CPOMP.Nous soumettrons notre approche à une étude quantitative mené sur deux jeux de données synthétiques issues de deux modèles : un guide de Pekeris et un guide plus générique car stratifié. Nous proposons de plus une analyse qualitative de notre approche menée sur des jeux de données réelles issus de la campagne sismique menée en Mer du Nord ainsi que de la campagne SBCEX.
... The first one is the study of the oceanic environment parameters variations influence on the interference pattern. [10][11][12][13] The second one is related to the development of interferometric methods for hydroacoustic signals processing to solve a set of important applied problems (passive localization, [14][15][16][17][18][19][20][21] active sonar, 22,23 dispersion compensation, 24,25 source selection, 26 etc.). In both of these areas, significant advances have been made in recent years. ...
Article
An interferometric signal processing method for localizing a broadband moving sound source in an oceanic waveguide is proposed and studied theoretically and experimentally. The field of a moving sound source in waveguide creates a stable interference pattern of the intensity distribution (interferogram) I(ω,t) in the frequency-time domain. Sound intensity is accumulated along interference fringes over the observation time. The two-dimensional Fourier transform (2D-FT) is applied to analyze the interferogram I(ω,t). The result of the 2D-FT F(τ,ν) is called the Fourier-hologram (hologram). The mathematical theory of hologram structure F(τ,ν) is developed in the present paper. It is shown that the hologram F(τ,ν) allows the coherent accumulation of sound intensity of the interferogram in a relatively small area focal spots. The presence of these focal spots is the result of interference of acoustic modes with different wave numbers. The main result of this paper is a simple relationship between the focal spots coordinates on the hologram and the source range, velocity, and motion direction. The proposed interferometric signals processing method for source localization is validated using experimental observations and numerical modeling in the band 80-120 Hz. The estimations of source range, velocity, and motion direction are performed for different cases of source motion.
Article
Dispersion and multipath effects contribute to the complexity of the shallow water acoustic field. However, this complexity contains valuable information regarding both the waveguide and the acoustic source. The horizontal wavenumber and relative amplitude of the modes comprising the acoustic field are crucial pieces of information for addressing acoustic inversion problems in shallow water. However, when employing a horizontal array to extract this information, limitations arise due to array aperture and signal-to-noise ratio constraints. To attempt to solve these challenges, the approach of spatial domain dedispersion transform and frequency domain accumulation is proposed. The objective can be attained by leveraging broadband source with slowly varying phase spectrum or known phase spectrum under the constraints of small aperture arrays and low signal-to-noise ratio. Additionally, the approach is validated on dual-hydrophone horizontal array by relaxing the signal-to-noise ratio requirement. In this paper, theoretical proof of the algorithms' performance is provided, accompanied by analysis of the impact of parameters such as acoustic source bandwidth, the number of elements and array aperture. The effectiveness of the algorithms are validated through simulations and experimental data.
Article
The interference pattern of the sound field of a broadband source in a shallow water waveguide is studied theoretically and experimentally in this paper. In the ocean waveguide, the sound source generates the interference pattern of the intensity distribution (interferogram) in the frequency-time domain. The mathematical theory of the interferogram structure is developed. The source interferogram consists of a set of quasi-parallel interference fringes in the frequency-time domain. It is shown that the slope of the interference fringes depends on the distance, velocity, and direction of motion of the sound source. The relationship between the slope angle of the interference fringes in the interferogram and the source parameters is derived in the paper. The two-dimensional Fourier transform (2D-FT) is used to analyze the interferogram. The result of the 2D-FT is called the Fourier hologram (hologram). It is shown that the hologram consists of a few focal spots in a relatively small area. The presence of these focal spots is the result of the interference of acoustic modes with different wavenumbers. The mathematical theory of the hologram structure is developed in this paper. The relationship between the coordinates of the focal spots on the hologram and the source parameters is considered. Consequently, the position of the focal spots can be used to estimate the source parameters (range, velocity, and direction of motion). The theoretical conclusions are verified in the context of computer modeling and the results of the acoustic experiment on the Pacific shelf (Yellow Sea, April, 2004) in the band 180–220 Hz.
Preprint
Full-text available
Citation: Ehrhardt, M.; Pereselkov, S.; Kuz'kin, V.; Kaznacheev, I.; Rybyanets, P. Title. Sensors 2022, 1, 0. https://doi.org/
Article
Full-text available
We investigate the influence of interface roughness, heterogeneous media, and dipping layers on the inversion of dispersive GPR pulse propagation in a surface waveguide, using 3D FDTD modelling. For both broadside and endfire source–receiver configurations, we calculated responses for different interface roughnesses, heterogeneities in dielectric properties, and dipping interfaces. For increasing roughness and heterogeneity, increased backscatter energy is visible in the data. The use of multiple source–receiver offsets to calculate the phase-velocity spectrum produced a relatively good signal-to-noise ratio. For low interface roughness the medium properties could be reasonably well reconstructed. For the largest interface roughness, significant diffracted energy was present and the medium properties could not be reliably reconstructed. For models having stochastic relative permittivity variations with a Std up to 15% and correlation lengths between 0.1 and 0.5m, the model parameters could still be relatively well reconstructed. For models with a dipping interface, the shot and countershot configuration clearly indicate lateral changes. Single-mode inversions return a wide range of medium property values, whereas combined TE-TM inversion return permittivity values that are remarkable close to the ε1 and ε2 values for both the shot and countershot configuration. In general, the interface roughness and heterogeneous media have a relative small influence on the inversion results. This is probably due to the use of multi-offset data for calculating the phase-velocity spectrum which is reducing the random noise.
Article
Since the mid-1970s, the computer has played an increasingly pivotal role in the field of ocean acoustics. Faster and less expensive than actual ocean experiments, and capable of accommodating the full complexity of the acoustic problem, numerical models are now standard research tools in ocean laboratories. The progress made in computational ocean acoustics over the last thirty years is summed up in this authoritative and innovatively illustrated new text. Written by some of the field's pioneers, all Fellows of the Acoustical Society of America, Computational Ocean Acoustics presents the latest numerical techniques for solving the wave equation in heterogeneous fluid–solid media. The authors discuss various computational schemes in detail, emphasizing the importance of theoretical foundations that lead directly to numerical implementations for real ocean environments. To further clarify the presentation, the fundamental propagation features of the techniques are illustrated in color. Computational Ocean Acoustics conveys the state-of-the-art of numerical modeling techniques for graduate and undergraduate students of acoustics, geology and geophysics, applied mathematics, and ocean engineering. It is also an essential addition to the libraries of ocean research institutions that use propagation models.
Article
For sound propagation in a shallow water waveguide, a dedispersion transform which can remove the dispersive effect of signal is described in this paper. The transform is a modified Fourier transform with two variables: translation and dispersion. Dispersive effect of signal for normal modes can be removed when these two variables match to real parameters. The validity of the present approach is verified in simulation and real data processing.
Article
This paper shows the benefits of using chirplet analysis of multipath magnetospheric whistlers. Higher resolution can be achieved than with the customary dynamic spectra (Gabor diagram), in which the cells characterizing resolution in time and frequency are the same over the whole diagram. Here we not only allow the aspect ratio of a resolution cell to vary within the one diagram but also introduce oblique cells corresponding to a new kind of elementary signal not discovered by Gabor. Chirplet analysis is also applicable to arbitrary signals and is most striking where certain kinds of finely detailed structure are present. We also describe a user interactive system that allows regions of interest to be selected in the time-frequency plane and enhanced by selecting chirplets appropriate to the local structure.
Article
This book attempts to present a theory of sound propagation in the ocean in its most fundamental form. However, the main theoretical chapters are prefaced with a first chapter where the most interesting experimental facts are presented. The book is intended for students, post graduates, researchers and practical workers dealing with ocean acoustics.
Article
The waveguide invariant summarizes in a single scalar parameter the dispersive propagation characteristics for a given environment. In this paper, an expression is derived for the invariant in shallow water environments where the bottom bathymetry varies with range and azimuth. It then is applied to the interpretation of broadband, single-element spectrograms collected during SWellEx-3, a shallow water acoustics experiment off the coast of San Diego. The unusual behavior of the interference patterns seen in the spectrograms from this experiment during the transmission of a 75-150-Hz pseudo-random noise sequence by a source towed over range-varying bottom bathymetry is well matched by the model predictions. (C) 1999 Acoustical Society of America. [S0001-4966(99)00911-X].
Article
A normal mode method for propagation modeling in acousto-elastic ocean waveguides is described. The compressional (p-) and shear (s-) wave propagation speeds in the multilayer environment may be constant or have a gradient (1/c(2) Linear) in each layer. Mode eigenvalues are found by analytically computing the downward- and upward-looking plane wave reflection coefficients R(1) and R(2) at a reference depth in the fluid and searching the complex k plane for points where the product R(1)R(2)=1. The complex k-plane search is greatly simplified by following the path along which \R(1)R(2)\=1. Modes are found as points on the path where the phase of R(1)R(2) is a multiple of 2 pi. The direction of the path is found by computing the derivatives d(R(1)R(2))/dk analytically. Leaky modes are found, allowing the mode solution to be accurate at short ranges. Seismic interface modes such as the Scholte and Stonely modes are also found. Multiple ducts in the sound speed profile are handled by employing multiple reference depths. Use of Airy function solutions to the wave equation in each layer when computing R(1) and R(2) results in computation times that increase only linearly with frequency. (C) 1996 Acoustical Society of America.
Article
The "invariant parameter" called "beta" is often useful for describing the acoustic interference pattern in a waveguide. For some shallow water waveguides, the measured acoustic intensity might contain contributions from several propagating acoustic modes. For each pair of these modes, a different value for the waveguide invariant might apply. If the acoustic intensity is measured over some distributed aperture and finite bandwidth, it may become difficult to assign a single value to beta. In the present work, the waveguide invariant is treated as a distribution. An algorithm for estimating this distribution for a general measurement geometry is developed. The algorithm is exercised for different classes of shallow water waveguides. When the propagation is dominated by modes interacting with the sea surface, the distribution can be sharply peaked. For cases where the sound speed profile creates a duct, the distribution is more diffuse. The effects of source/receiver depth, range, bandwidth and bottom attenuation are quantified.
Article
In a previous paper [T. C. Yang, J. Acoust. Soc. Am. 8 2, 1736–1745 (1987)] an eigenvector (EIG) mode decomposition method was used as the basis for modal beamforming (matched‐mode processing). In this article, it is pointed out that matched‐field processing is equivalent to a modal beamforming method with a different mode decomposition algorithm (denoted as the MFP‐mode decomposition method). The difference in performance between the two processing schemes can be understood in terms of the effectiveness of mode filtering of the mode decomposition method used. It is found that the difference between the MFP‐ and EIG‐mode decomposition is small if the modes associated with the signal and noise are adequately sampled by a (filled) vertical array, and potentially large if they are not. The difference is analyzed analytically and illustrated numerically using vertical arrays in an Arctic and Pacific environment. Array gain and peak‐to‐sidelobe ratio are evaluated for the MFP and EIG processors under identical environmental conditions, and compared with the matched‐mode processing assuming perfect mode decomposition. Examples are given to illustrate the use of mode filtering to suppress the sidelobes caused by sound‐speed mismatch in the upper (<1000 m) water column in the Pacific. It is found that higher‐order modes are essential in keeping the sidelobe level down, despite the fact that they are not as well resolved as the lower‐order modes. The (lower) modes that are responsible for the sidelobes (due to the sound‐speed mismatch) are identified. It is concluded that source localization is possible in a mismatch environment, if proper mode filtering is applied to the data.