Sensor Array
Sensor Array
ARRAY SIGNAL
PROCESSING
Prabhakar S. Naidu
CRC Press
Boca Raton London New York Washington, D.C.
1195/Disclaimer Page 1 Monday, June 5, 2000 3:20 PM
Naidu, Prabhakar S.
Sensor array signal processing / Prabhakar S. Naidu.
p. cm.
Includes bibliographical references and index.
ISBN 0-8493-1195-0 (alk. paper)
1. Singal processing–Digital techniques. 2. Multisensor data fusion.
I. Title.
TK5102.9.N35 2000
621.382'2—dc21 00-030409
CIP
This book contains information obtained from authentic and highly regarded sources. Reprinted material
is quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable
efforts have been made to publish reliable data and information, but the author and the publisher cannot
assume responsibility for the validity of all materials or for the consequences of their use.
Neither this book nor any part may be reproduced or transmitted in any form or by any means, electronic
or mechanical, including photocopying, microfilming, and recording, or by any information storage or
retrieval system, without prior permission in writing from the publisher.
The consent of CRC Press LLC does not extend to copying for general distribution, for promotion, for
creating new works, or for resale. Specific permission must be obtained in writing from CRC Press LLC
for such copying.
Direct all inquiries to CRC Press LLC, 2000 N.W. Corporate Blvd., Boca Raton, Florida 33431.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation, without intent to infringe.
The tools of array signal processing remain the same, cutting across the
boundaries of different disciplines. For example, the basic tool of
beamformation is used in many areas mentioned above. The present book aims
at unraveling the underlying basic principles of array signal processing without
a reference to any particular application. However, an attempt is made to
include as many tools as possible from different disciplines in an order which
reflects the underlying principle.
In the real world, different types of wavefields are used in different
applications, for example, acoustic waves in sonar, mechanical waves in
seismic exploration, electromagnetic waves in radar and radio astronomy.
Fortunately, all wavefields can be characterized under identical mathematical
framework. This common mathematical framework is briefly summarized in
chapter 1. Here we have described the basic equations underlying different
wavefields and the structure of array signals and the background noise when the
noise sources follow some simple geometrical distribution. The topics covered
are wavefield in open space, bounded space including multipath propagation and
layered medium. Also covered is the weak scattering phenomenon which is the
basis for tomographic imaging. In chapter 2 we study different types of sensor
configurations. The emphasis is however on commonly used uniform linear
array (ULA), uniform circular array (UCA). Many practical sensor array
systems can be studied in terms of the basic ULA and UCA systems
(cylindrical array in radar and sonar, cross array in astronomy and seismology).
Like sensors, the sources can also be configured in the form of an array. The
Contents
Chapter One
An Overview of Wavefields
Chapter Two
Chapter Three
Chapter Four
Chapter Six
Source Estimation
Chapter Seven
Tomographic Imaging
Chapter Eight
8.1 Migration
8.2 Exploding reflector model
8.3 Extrapolation in ω-k plane
8.4 Focused beam
8.5 Estimation of wave speed
8.6 Exercises
An Overview of Wavefields
A sensor array is used to measure wavefields and extract information about the
sources and the medium through which the wavefield propagates. It is therefore
imperative that some background in different types of wavefields and the basic
equations governing the wavefield must be acquired for complete understanding
of the principles of Array Signal Processing (ASP). In an idealistic
environment of open space, homogeneous medium and high frequency (where
ray approach is valid), a thorough understanding of the wave phenomenon may
not be necessary (those who are lucky enough to work in such an idealistic
environment may skip this chapter). But in a bounded inhomogeneous medium
and at low frequencies where diffraction phenomenon is dominating, the physics
of the waves plays a significant role in ASP algorithms. In this chapter our
aim is essentially to provide the basics of the physics of the waves which will
enable us to understand the complexities of the ASP problems in a more
realistic situation. The subject of wave physics is vast and naturally no attempt
is made to cover all its complexities.
ρ d 2φ
∇2φ = (1.1a)
γφ 0 dt 2
γφ 0 κ
c= = (1.1b)
ρ ρ
1 d 2φ 1
= ρ(r)∇ ⋅ ( ∇φ)
2
c (r) dt 2
ρ(r)
(1.2a)
1
= ρ(r)∇( ) ⋅ ∇φ + ∇ 2 φ
ρ(r)
1 d 2 φ ∇ρ(r)
∇2φ − = ⋅ ∇φ (1.2b)
c 2 (r) dt 2 ρ(r)
where r stands for position vector. The acoustic impedance is equal to the
product of density and propagation speed ρc and the admittance is given by the
inverse of the impedance or it is also defined in terms of the fluid speed and the
pressure,
fluid speed ∇φ
Acoustic admittance = =
pressure jωφ
Note that the acoustic impedance in air is 42 but in water it is 1.53 x105.
1.1.2 Mechanical Waves in Solids: The physical quantity which propagates is
the displacement vector, that is, particle displacement with respect to its
stationary position. Let d stand for the displacement vector. The wave equation
in a homogeneous medium is given by [1, p142]
∂2d
ρ = (2µ + λ)grad divd − µcurlcurld
∂t 2
d = ∇φ + ∇ × ψ (1.3a)
1 ∂2φ
∇2φ =
α 2 ∂t 2
(1.3b)
1 ∂2 ψ
∇×∇×ψ = − 2 2 ∇.ψ = 0
β ∂t
e x e y ez
∂ ∂ ∂
∇ × F = det
∂x ∂y ∂z
f x f y f z
∂f z ∂f y ∂f ∂f ∂f y ∂f x
=( − )e x + ( x − z )e y + ( − )e z
∂y ∂z ∂z ∂x ∂x ∂y
d = (d x , d y , dz )
∂φ ∂ψ z ∂ψ y
dx = + −
∂x ∂y ∂z
∂φ ∂ψ x ∂ψ z
dy = + − (1.4)
∂y ∂z ∂x
∂φ ∂ψ y ∂ψ x
dz = + −
∂z ∂x ∂y
The first subscript refers to the plane perpendicular to the axis denoted by the
subscript and the second subscript denotes the direction in which the vector is
pointing. For example, sxx is a stress in a plane perpendicular to the x-axis
(i.e., y-z plane) and pointing along the x-axis.
sxz
sxy
sxx
Figure 1.1: An element of volume (cuboid) and stresses are shown on a face
perpendicular to the x-axis.
The stress components on different faces of a cuboid are shown in fig. 1.1. The
torque on the cuboid should not cause any rotation. For this, we must have
sxy = syx and similarly all other nondiagonal elements in the stress matrix.
Thus, s must be a symmetric matrix. The components of a strain matrix are
related to the displacement vector
∂d x ∂d ∂d
ε xx = , ε yy = y , ε zz = z
∂x ∂y ∂z
∂d ∂d
ε xy = ε yx = x + y ;
∂y ∂x
∂d y ∂dz
ε yz = ε zy = + ; (1.6)
∂z ∂y
∂dz ∂d x
ε zx = ε xz = + .
∂x ∂z
Finally, the stress and strain components are related through Hooke’s Law:
Using equations (1.4), (1.6) and (1.7) we can express all nine stress
components in terms of the scalar and vector potential functions, φ and ψ .
For example, it is possible to show that sxx is given by
∞
1
∫ ∫ Φ(u,v,ω)e
− u 2 + v 2 − k α2 z
φ(x, y, z,ω) = 2 e j (ux + vy) dudv (1.8a)
4π −∞
Recall that the pressure waves (p-waves) travel at speed α and the shear waves
travel at speed β , where α is generally greater than β . The displacement
vector is in the direction of the gradient of the scalar potential but it is in the
direction of curl of the vector potential (1.4), that is, normal to the vector
potential. Thus there is a fundamental difference in the nature of propagation of
the shear and the pressure waves. The shear waves are polarized; the
displacement vector is always perpendicular to the direction of wave
propagation. The displacement vector executes a motion depending upon the
phase difference between the components of the displacement vector; a line
∂B
∇×E= −
∂t Faraday’s law (1.9a)
∇⋅D = ρ
∂D
∇×H= J+
∂t Ampere’s law (1.9b)
∇⋅B = 0
where
D = εE ε: dielectric constant
B = µH µ: magnetic susceptibility (1.9c)
J = σE σ: conductivity
H=∇×ψ
∂ψ (1.10a)
Ε = −∇φ −
∂t
∂φ
∇ ⋅ ψ + εµ =0 (1.10b)
∂t
Using (1.10) in (1.9) we obtain two decoupled equations,
∂2φ 1
∇ φ − εµ 2 = − ρ
2
(1.11a)
∂t ε
H=∇×ψ
∂ψ (1.12)
Ε=−
∂t
and the vector potential satisfies
∂ 2ψ
∇ × ∇ × ψ − εµ = −µJ
∂t 2 (1.13)
∇⋅ψ = 0
Both electric and magnetic fields travel with the same speed unlike p and s
waves in solids. The electric and magnetic vectors lie in a plane ⊥ to the
direction of propagation. The tip of a field vector will execute a smooth curve
known as a polarization ellipse. For example, in a vertically polarized electric
field the tip of the electric vector lies on a vertical line while the magnetic field
lies on a horizontal line. The minor and major axes of the polarization ellipse
a
are a and b respectively. Define an angle χ = tan −1 ( ) . The major axis of the
b
polarization ellipse is inclined at an angle ζ (see fig. 1.2). Consider a plane
wave with its wave vector in y-z plane making an angle θ with z axis. The
electric vector will lie in a plane ⊥ to direction of propagation. This plane is
known as the x-θ plane. The electric vector can be split into two components,
π
cos(2γ ) = cos(2χ)cos(2ζ) 0≤γ ≤
2
-E x
θ Eθ
(a)
x
Eθ
b
a ζ
Ex
Polarization
ellipse
(b)
Figure 1.2: (a) EM wave propagation vector is in y-z plane. Electric and magnetic
vectors lie in a plane ⊥ to the propagation vector and it contains the x-axis. (b)
The tip of the electric vector executes a smooth curve, for example, an ellipse. The
electric vector can be split into two components, Ex and Eθ.
ω
where k= a k and A is a complex constant. a p and a k are mutually ⊥
c
unit vectors. The electric and magnetic field vectors are given by
∂φ 1
r→∞
→ O( )
∂r r
∂φ 1
+ jk φ r→∞
→ o( )
∂r r
(b) Time Reversal: All wave equations without time varying coefficients or
time derivatives of odd order share an interesting property of time reversal. If
φ(x, y, z,t) is a solution of the wave equation it is easy to verify that
φ(x, y, z,t0 − t) is also a solution of the same wave equation for any constant
t 0 [3]. This fact has been used for reverse time propagation of seismic
wavefields. The seismic wavefield observed on a surface is time reversed and fed
into a loudspeaker broadcasting the recorded wavefield.
1.1.5 Sensing of Wavefields: A device is required to convert a physical
wavefield, say, pressure wave, into an electrical signal which is then sampled
and digitized. The resulting numbers are stored in computer memory for further
processing. Such a device is known as a sensor. While the mechanism of
conversion of a physical wavefield into an electrical signal is not important
from the point of array signal processing, the speed of conversion and the
dynamic range are very relevant; accordingly the sampling rate and the number
of bits per sample are fixed. The speed of conversion controls the bandwidth of
a sensor, faster is the conversion larger is its bandwidth. As we have two types
of wavefields, scalar wavefield and vector wavefield, there are also two types of
where f (r,t) stands for any one of the wave types, for example, pressure or
one of the components of a vector field. We introduce the Fourier integral
representation of a wavefield,
∞
1
f (r,t) = 3 ∫ ∫ ∫ F(u,v,ω)H(u,v, z)e j (ωt −ux − vy) dudvdω (1.16)
8π − ∞
d 2 H(u,v, z) ω2
= (u 2
+ v 2
− )H(u,v, z) (1.17)
dz 2 c2
whose solution is
ω
where k= is known as a wavenumber. When (u 2 + v 2 > k , we choose
c
(-) sign for z>0 and (+) sign for z<0 so that the wavefield does not diverge. In
both cases the field will rapidly decay as z → ∞ . These are known as
evanescent waves. When (u
2
+ v 2 < k we get propagating waves whose
integral representation reduces to
f (x, y,t) =
∞
1 (1.19)
8π 3 ∫ −∫∞∫
k 2 −u 2 − v 2 z ) j ( ωt −ux − vy)
F(u,v, ω )e ± j ( e dudvdω
± j( k 2 −u 2 −v 2 z)
where the sign in e is selected depending on whether the
waves are diverging or converging. The convention is (-) sign for diverging
waves and (+) sign for converging waves (see fig. 1.3). Note that in a bounded
space both diverging and converging waves can coexist and hence it would be
necessary to use both signs in describing wavefields in a bounded space.
u = k sin θ cosϕ
v = k sin θsin ϕ
where ϕ and θ are respectively azimuth and elevation angles of a plane wave.
The elevation angle is an angle between the z-axis and the wave vector and
azimuth angle is an angle between the x-axis and the projection of the wave
vector on the x-y plane. The representation given in (1.19) is also known as a
plane wave decomposition of a wavefield. In (u, v, ω) space a single frequency
plane wave is represented by a point and a wideband wave by a line passing
through the center of the coordinate system (see fig. 1.4). The slope of the line
is inversely proportional to the direction cosines (defined later on page 17) of
the vector ⊥ to the plane wavefront and directly proportional to the speed of
propagation.
c
b
a
0
ω
(u 2 + v 2 ) ≤ ( )2 (1.20)
c
po c
= cz = (1.21a)
dt cos(θ)
or = oq cosϕ
os2 = or sin θ
= oqsin θ cosϕ
os2 oq
The travel time from s 2 to o is dt = = sin θ cosϕ . Hence the
c c
apparent speed along the x-axis is given by
oq c
cx = = (1.21b)
dt sin θ cosϕ
c
cy = (1.21c)
sin θsin ϕ
ω ω
cx = =
k sin θ cosϕ u
ω ω
cy = =
k sin θsin ϕ v
q
1 r ϕ
o q x
s2
s
1
1 1 1 1
2
+ 2+ 2 = 2 (1.23)
c x c y cz c
and that the apparent speeds are always greater than or equal to the wave speed.
Closely related to the apparent speed are the so-called ray parameters,
1 1 1
px = , py = , pz = (1.25a)
cx cy cz
From (1.24) and (1.25) the ray parameters can be expressed in terms of direction
cosines
α = px c, β = py c, γ = pz c
f (x, y, z1 ,t) =
rapidly vanishing and hence the phase shift is set to zero in this range.
1.2.5 Point Source: A point source is often used as a source of illumination,
but it will only generate spherical waves which produce a much more complex
(mathematically speaking) response from a target than what a plane wave does.
Fortunately, a point source wavefield may be written as a sum of infinitely
many plane waves; a plane wave decomposition was described in subsection
(1.2.1). Additionally, a response of a target may be obtained as a sum of plane
wave responses. To obtain the Fourier integral representation of a point source
wavefield we go back to (1.19) where we shall assume that the Fourier
transform of the wavefield has radial symmetry,
f (x, y, z,t)
∞
1
= 3 ∫ ∫ ∫ F0 (u,v, ω )e ± j ( k 2 −u 2 − v 2 z ) j ( ωt −ux − vy)
e dudvdω (1.27a)
8π − ∞
∞ 2π
1 1
2 ∫ 0
= F (s, ω )e jωt dω ∫ se ± j ( k 2 −s2 z)
ds ∫ e − j (sr cos( θ − ϕ )) dθ
4π 0
2π 0
∞ ∞
1
∫ F (s,ω)e dω ∫ sJ0 (sr)e ± j ( k 2 −s2 z)
f (r, z,t) = 2 0
jωt
ds (1.27b)
4π −∞ 0
2.4
1.6
0.8
0.0
-3.2 -1.6 0.0 1.6 3.2
u
Figure 1.7: Phase shift as a function of u (keeping v=0) for vertical separation equal
to wavelength (assumed to one).
π
where s = u 2 + v 2 = k sin γ where 0 ≤ γ ≤ . In (1.27b) we replace s by
2
k sin γ and rewrite (1.27b) as
π
∞ 2 z
1 jω(t ± cos γ )
f (r, z,t) = 2 ∫ dω ∫ F (k sin γ,ω)e k sin(2γ )J0 (kr sin γ )dγ
c 2
0
8π −∞ 0
(1.28)
Figure 1.8: In a homogeneous medium the wavefield due to a point source may be
expressed as a sum of plane waves traveling with an angle of incidence γ , where
π
0≤γ ≤ .
2
F0 (s,ω)e ± j( k 2 −s2 z)
= ∫ F(r, z,ω)rJ0 (sr)dr (1.29)
0
where F(r, z,ω) is a Fourier transform (temporal) of f (r, z,t) . On the z=0
surface (1.29) reduces to
∞
2
f (t,sin γ ) = ∫ f (r, z = 0,t) * rdr (1.31)
0
r
( sin γ )2 − t 2
c
2
h(t) =
r
( sin γ )2 − t 2
c
and then summed over all sensors. The impulse response function may be
expressed as an integral involving a delta function [5],
π
2
h(t) = = ∫ δ(t − t0 cosβ)dβ
r
( sin γ )2 − t 2 −π
c
r
where t0 = ( sin γ ) . Using the above integral representation of the impulse
c
response function in (1.31) we obtain
t 0
∞
2
f (t,sin γ ) = ∫ ∫ f (r, z = 0,t − t ′ ) 2 d t ′ rdr
0 0
−t t 0 − ( t ′ ) 2
∞ t 0 π
= ∫ ∫ f (r, z = 0,t − t ′ )dt ′ ∫ δ(t ′ − t0 cosβ)dβ rdr (1.32)
0 −t 0 −π
∞ π
= ∫ ∫ f (r, z = 0,t − t0 cosβ)dβ rdr
0 − π
π ∞
∆r
where ∆t0 = sin γ . The inner sum in (1.33) is the sum-after-delay
c
operation, commonly known as slant stacking in seismic exploration. Later in
chapter 2 we shall show how this operation is related to the radon transform.
x ∞
f (x,t) = f (t − ) ∑ δ(x − i∆x)
cx i = − ∞
∞ j (t − )ω
x ∞
1
=
2π −∫∞
F(ω)e cx
dω ∑ δ(x − i∆x)
i=−∞
Note that cx stands for wave speed in the direction of x-axis or apparent speed.
Taking the 2D Fourier transform of f(x,t)
+∞
∞ ∞ ∞ ω′
j (u − )x
= ∫ F(ω ′)δ(ω ′ − ω)dω ′ ∑ ∫ e δ(x − i∆x)dx
cx
−∞ i=−∞ −∞
∞ ∞ ω′
j (u − )i∆x (1.34)
= ∫ F(ω ′)δ(ω ′ − ω)dω ′ ∑ e
cx
−∞ i=−∞
∞ ∞
ω ′ 2π
= ∫ F(ω ′)δ(ω ′ − ω)dω ′ ∑ δ(u −
−∞ k =−∞
−
cx ∆x
k)
The spatial Fourier transform of the array output is sketched in fig. 1.9. The
signal spectrum is concentrated on sloping lines in the u-w plane as seen in fig.
1.9.
Depending upon the signal bandwidth and the apparent speed, an alias
will show up within the principal band. In fig. (1.9) there is no aliasing when
aliased
k=0
k=1
aliased
4π
−∆ - 2∆πx 0 2π u 4π
x ∆x ∆x
Principal
band
Figure 1.9: The Fourier transform of spatially sampled wavefield (plane wave) lies
on a set of sloping lines (with slope= 1/ cx ) corresponding to different values of k
in (1.34).
the signal bandwidth is as shown by the dark square, which in this case
λ min π
corresponds to a sensor spacing ∆x = or ∆x = cx . Note that,
2 ω max
π
since the sampling interval ∆t = , ∆x = ∆t cx . There is no aliasing
ω max
whenever the sensor spacing and the time sampling interval are such that
∆x
≤ cx . For the vertical angle of incidence, since cx = ∞ , there is no
∆t
aliasing effect for any ∆x .
1.2.7 Dispersion: A medium is said to be dispersive when a wavefront travels
at a speed which is a function of frequency. Consider a wavefront at a fixed
λ0 ω0
c ph = = =c
T0 k0
∆ω
ω0 +
2
1
f (r,t) =
2π ∫ F(ω)exp( j(ωt − k ⋅ r)dω
∆ω
(1.35a)
ω0 −
2
ω
where k = (u,v, ( )2 − u 2 − v 2 ) is the wave vector. We assume that the
c
propagation speed is a function of frequency, and hence the wave vector is also
a function of frequency. Using a Taylor’s series expansion of the wave vector
dk(ω)
k(ω) = k(ω 0 ) + (ω − ω 0 )+...
dω ω =ω 0
in (1.35a) we obtain
∆ω
2
1 dk(ω)
f (r,t) = F(ω 0 )e j (ω 0 t − k(ω 0 )⋅r)
2π ∫
∆ω
exp[ j(t −
dω ω =ω 0
⋅ r)ω̃]dω̃
−
2
∆ω dk(ω) ∆ω
= F(ω 0 )e j (ω 0 t − k(ω 0 )⋅r) sin c[(t − ⋅ r) ]
2π dω ω =ω 0 2
(1.35b)
0.5
Field
0
0
-0.5
-1
1 1
0.5 0.5
Field
0 0
Field
-0.5
-0.5
-1
-1
Time
Figure 1.10: A narrowband signal with a center frequency at 0.23 Hz and a bandwidth
equal to 0.02 Hz. The envelope travels at the group speed which in this is equal to
the phase speed (assumed to be one). The signal in the lower panel arrives 32 time
units later.
dk(ω)
cgp = 1 (1.36a)
dω ω =ω 0
or
1
cgp = (1.36b)
dk(ω)
dω ω =ω 0
The group speed differs from the phase speed only when the medium is
dispersive. In a nondispervive medium both speeds are equal. The modulating
waveform (second term in (1.35b) travels at the same speed as the carrier wave
(first term). An example of propagation of a narrowband signal is shown in fig.
∞
1
∫
j[ωt −u 0 ( x − x 0 )− v 0 ( y − y 0 )− k 2 − s 02 ( z − z 0 )]
f (r,t) = F(ω)e dω
2π − ∞
θi θr
Fluid Medium #1
Fluid Medium #2
θt
refracted
Figure 1.11: Laws of reflection and refraction: (1) Incident wave, reflected wave,
refracted wave and the normal to the interface at the point of incidence lie in the
sin θ i sin θ t
same plane. (2). θi=θr (3) = = p (ray parameter).
c1 c2
ρ2 c
cosθi − ( 2 )2 − sin 2 θi
ρ c1
r̂ = 1 (1.37a)
ρ2 c
cosθi + ( 2 )2 − sin 2 θi
ρ1 c1
θ' r
θ i θr
Elastic Medium #1
Elastic Medium #2
θt
θ't refracted(p)
refracted(s)
2 cosθi
tˆ = (1.37b)
ρ2 c
cosθi + ( 2 )2 − sin 2 θi
ρ1 c1
Reflection and refraction at an interface separating two elastic media are more
complex. An incident longitudinal wave will give rise to two reflected and two
refracted rays. The first ray is a p-wave and the second ray is a s-wave. The laws
of reflection and refraction for a p-wave incident at the interface are summarized
in fig. 1.12. The coefficients of reflection and refraction, ( r̂ p , r̂s , tˆp , tˆs ) , are
obtained by solving the following system of four linear equations [6, p3-101]:
r̂ p cosθi − r̂s sin θ′r + tˆp cosθt − tˆs sin θ′t = cosθi
(1.38a)
− r̂ p sin θi − r̂s cos θ′r + tˆp sin θt + tˆs cos θ′t = sin θi
where (λ1 ,µ1 ) are the elastic constants and (c pI ,csI ) are respectively the p-
wave and the s-wave speed in the upper medium. Other parameters refer to the
lower medium. For vertical incidence, that is, θi = 0 from (1.38), it follows
that r̂s = tˆs = 0 . This is a consequence of the fact that there is no s-wave
generation for vertical incidence.
Polarization in the plane of incidence: Finally, an EM wave at an interface
between two different media undergoes reflection and refraction. However, there
is only one reflected and one transmitted wave which travel at the same speed.
The process of reflection and refraction is strongly influenced by the
polarization angle. We shall consider one simple case of polarization in the
plane of incidence (fig. 1.13). The coefficient of reflection and transmission is
given by (Fresnel’s equations)
χ−ζ ˆ 2
r̂ E = , tE =
χ+ζ χ+ζ (1.39)
cosθt µc
where χ= and ζ = 1 1 . Note, when χ = ζ , r̂ E = 0 and tˆE = 1 ;
cosθi µ 2 c2
that is, there is no reflected energy. This phenomenon takes place at a specific
angle of incidence known as Brewster angle given by
1 − ζ2
sin 2 θ B = (1.40)
c
( 2 )2 − ζ 2
c1
Figure 1.13: The electric vector is in the plane of incidence, that is, in the plane
containing the incident ray, normal to interface and reflected and refracted rays.
free surface is (-1) and that at the bottom is r̂b (see (1.37)). Because of multiple
reflections at the free surface and at the bottom (see fig. 1.14) many waves will
reach a sensor array at different times. It is convenient to model all these waves
as emanating from a series of images whose position can be determined by
following simple geometrical optics rules. An infinite set of images is formed
between two parallel reflecting surfaces. We index these images with two
integers (i,k); the first integer represents a group and the second integer
represents an image within a group. There are four images in each group. For
example, in the i=0 group the images are: s01 and s02 which are equal but
opposite in strength, that is, a surface dipole, and s03 and s04 , which are
caused by the reflection of the surface dipole onto the bottom. The surface
dipole and its image are separated by a distance of 2H where H is the depth of
the fluid channel. The next group of images, that is, for i=1, is obtained by
sliding the surface dipole two depth units above and the image dipole two depth
units below the bottom. The vertical distance with respect to the top sensor in
the array to different images in the ith group is given by
Hi1 = 2iH + zr − zs
Hi2 = 2iH + zr + zs
(1.41)
Hi3 = 2(i + 1)H − zr − zs
Hi 4 = 2(i + 1)H − zr + zs i = 0,1,2,...
Air
zs
Source zr
s o1
H sensor
Fluid medium 1
Fluid medium 2
so 3
so
4
Figure 1.14: A shallow water channel with a sound source (•) at depth zs and a
sensor at depth zr. Rays starting from the source are reflected at the surface and
bottom and finally reach the sensor after one or more reflections. Some of the
images (•) are also shown.
An image whose second index is 1 or 2 lies above the bottom and an image
whose second image is 3 or 4 lies below the bottom. The strength of an image
depends upon the number of bounces the ray has undergone before reaching the
e −βli0 e −βli1
i +1 i
α i 0 = (−1) r̂
i i
b α i1 = (−1) r̂ b
li 0 li1
(1.43)
e −βli2 e −βli3
α i2 = (−1)i r̂bi+1 α i3 = (−1)i +1 r̂bi+1
li2 li3
where β is the attenuation coefficient in the top liquid layer. The signal
reaching the sensor (fig 1.14) may be expressed as
∞
1 ∞ 3
p(t) = ∑
2π i = 0
∑ ∫α
m=0 −∞
im P(ω)e j (t − τ im )ω dω (1.44)
where p(t) is the pressure field received by a sensor and P(ω) is the Fourier
transform of the waveform emitted by the source. τ im is propagation delay
lim
τ im = . The pressure field given by (1.44) is a complex field. The
c
magnitude and phase are found to vary rapidly with the depth of the sensor. For
example, for a channel with the following parameters:
H=100 m (meters)
d=2000 m
zs= 50 m
ρ1 = 1, c1 = 1500
ρ2 = 2,c2 = 1600
the magnitude and phase variations as a function of the depth of a sensor are
shown in fig. 1.15. Eighty images were taken into account in computing the
pressure field.
As the range increases, while the position of image sources remain
unchanged, the strength of the images may increase on account of the
phenomenon of total internal reflection. Recalling equation (1.37) it may be
j ∞ ∞
p(r, z,t) = ∫ ∑ φ n (z)φ n (zs )H01 (kn r)P(ω )e jωt dω (1.45)
2 −∞ n=0
d 2φ
2
+ γ 2n φ = 0
dz
φ(0) = 0 (1.46)
dφ
= 0 (rigid bottom)
dz z = H
ω2
where kn + γ n = 2 . kn is a horizontal wavenumber and γ n is a vertical
2 2
c (z)
wavenumber. The source is assumed to be at a depth zs . Note that both
wavenumbers are functions of frequency and also of depth when the sound speed
is a function of depth. Consider a special case of isospeed channel and large r.
Firstly, the Hankel function can be approximated as
π
2 j ( kn r + 4 )
H (kn r) ≈
1
e ; secondly both wavenumbers are independent of
kn πr
0
e jγ n z − e − jγ n z
φ n (z) = sin(γ n z) =
2j
1
(n − )π
where γ n = 2 . For this special case the pressure field in the fluid
H
channel with the rigid bottom is given by
p(r, z,t) =
π
j
4 ∞ ∞
e
2 ∫∑
1
2kn πr
{ }
sin(γ n zs ) e j ( γ n z + k n r +ωt ) − e j ( − γ n z + k n r +ωt ) P(ω)dω
−∞ n=0
(1.47)
From (1.47) it is possible to infer that each mode is a sum of two plane
γ
wavefronts traveling in the vertical plane at angle ±θ n , where tan θ n = n
kn
with respect to the horizontal plane. In three dimensions the wavefront is a
π
conical wavefront (see fig.1.16); the angle of the cone is − θ n . Note that the
2
direction of propagation of the wavefronts is solely dependent on the channel
characteristics.
The comparison of ray and normal mode approaches to propagation in
a shallow water channel is instructive. The ray approach is essentially a high
frequency approximation to the exact solution of the wave equation. The
normal mode approach on the other hand is a series approximation to the exact
solution of the wave equation. The conical wavefronts obtained by
decomposing each mode may be looked upon as a result of constructive
interference of spherical wavefronts from the source and its images [8]. At low
frequency the accuracy of the ray approach is enhanced if one were to use the
concept of beam displacement. At a point of reflection the beam is found to be
laterally displaced, a phenomenon first observed in optics [10]. Both numerical
and experimental evidence in support of the fact that beam displacement does
help to increase the accuracy of the ray approach in relation to the exact
solution has been widely reported [11, 12].
1.3.4 Propagation Through Layered Medium: In seismic exploration a
horizontally layered medium is often used as a model. Vertically
incidentlongitudinal waves (p-waves) are preferred as there is no loss of energy
through conversion into s-waves at each interface. Also, as s-waves arrive
slightly later,
Wavefronts
Shallow
water Source
Rigid Bottom
Figure 1.16: A mode can be decomposed into two conical wavefronts. In the vertical
γn
plane the wavefronts travel at the angle ±θ n , where tan θ n = .
kn
ρ1 c1
ρ2c2
ρn cn
Figure 1.17: A stack of uniform layers. All layers are of the same thickness but with
different impedances.
they would interfere with late arriving p-wave signals. A layered medium is
modeled as uniform horizontal layers stacked one above the other (see fig.
1.17). A vertically propagating plane wavefront is repeatedly reflected and
transmitted at each interface, thus producing a complex reverberation pattern
which may be conveniently described within the frame work of filter theory
[13].
Figure 1.18: Reflected and transmitted waves at an interface. (a) Incident from above
and (b) incident from below.
∞
1 r0 + r1e − jω jωt
2π −∫∞
f refl (t) = F (ω) e (1.48a)
1 + r0 r1e − jω
0
Layer 0
^t 0 ^t 0 ^r 1 - ^t 0 ^r 1 r 0
^ ^
- ^t 0 ^r 1 r 0 ^r 1
travel time
unit return
Layer 1
Layer 2
^t 0 ^t ^ ^
1 - ^t 0 r 1 r 0 ^t 1
Transmission
Figure 1.19: Repeated reflections at the two faces of a layer produce a sequence of
reflected and transmitted waves. Some amount of energy is also trapped inside the
layer. All waves travel vertically though for the sake of clarity the waves are shown
as inclined.
ω
∞ −j
1 t0 t1e 2 jωt
2π −∫∞
f trans (t) = F (ω) e (1.48b)
1 + r0 r1e − jω
0
r0 + r1z −1
1
t0 t1 −
R(z) = , T(z) = z 2 (1.49)
1 + r0 r1z −1 1 + r0 r1z −1
We can now express the z-transforms of the reflected and the transmitted
waveforms as
It is interesting to note that T(z) has the form of a first order AR process
filter and R(z) has the form of a ARMA process filter of order (1,1) [14].
Since r0 r1 ≤ 1, the only pole of the transfer function (1.49) lies within the
unit circle, making R(z) and T(z) stable. The signal flow diagrams of
transmission and reflection filters are shown in fig. 1.20.
∞
1
3 ∫ ∫∫
f (x, y,t) = dF(u,v,ω)e j (ωt −ux − vy)
8π − ∞
∞
1
3 ∫ ∫∫ f
C f (∆x, ∆y, τ) = S (u,v,ω)e j (ωτ −u∆x − v∆y) dudvdω (1.50)
8π − ∞
where C f (∆x, ∆y, τ) = E{f (x, y,t) f (x + ∆x, y + ∆y,t + τ)} is the
covariance function and S f (u,v,ω) is the frequency wavenumber spectrum. In
the spectrum analysis of time series (1.50) is known as the Wiener-Khinchin
relation of great significance [14]. S f (u,v,ω) represents power received at a
given temporal frequency ω and spatial frequencies u and v. Since u and v are
related to the direction cosines, S f (u,v,ω) may be looked upon at a fixed
temporal frequency as a function of the direction cosines.
1.4.2 Open Space: Consider noise sources in the far field region, distributed
over a sphere or a circle (in two dimensions). Each point source emits a
stationary stochastic waveform uncorrelated with all other sources. Let f i (t)
be the stochastic waveform emitted by the ith source at angular distance
ϕ 0 + θi where θi is a random variable uniformly distributed over an interval
±θ 0 (see fig. 1.21). The signal received at the upper sensor is given
d
by f i (t − sin(ϕ 0 + θi )) and that at the lower sensor is
2c
d
f i (t + sin(ϕ 0 + θi )) where d is the sensor separation. The total signal
2c
obtained by summing over all sources is given by
d
g1 (t) = ∑ f i (t − sin(ϕ 0 + θi ))
i 2c
(1.51a)
d
g2 (t) = ∑ f i (t + sin(ϕ 0 + θi ))
i 2c
We shall now replace the random function in (1.51a) with its spectral
representation [14] and rewrite it as
d
1 ∞ j (ω(t − sin(ϕ 0 +θ i ))
g1 (t) = ∑
2π ∫− ∞ i
dFi (ω)e 2c
θi
2θ 0
ϕ0
Figure 1.21: Point sources are uniformly distributed over an arc of a large circle in
the x, y plane. An elementary array of two sensors is on the y-axis.
d
1 ∞ j (ω(t + sin(ϕ 0 +θ i )))
g2 (t) = ∑
2π ∫− ∞ i
dFi (ω)e 2c
(1.51b)
We will now compute the cross-covariance function between the two outputs,
where we have assumed that the sources are uncorrelated with an identical
spectrum. Since θi is a uniformly distributed random variable the expected
value in (1.52b) may be replaced by an integral
θ0 d
1 ∞ 1 j (ω( τ + sin(ϕ 0 +θ)))
c12 (τ) = c12 (τ) = ∫
2π − ∞
S0 (ω)dω
2θ 0 ∫e
−θ 0
c
dθ (1.53a)
θ0 d
j (ω( τ + sin(ϕ 0 +θ)))
∫e
−θ 0
c
dθ
ωd sin(nθ 0 )cos(nϕ 0 )
cos(ωτ) ∑ δ n Jn ( c ) n
+
n = 0,2,4...
ωd sin(nθ 0 )sin(nϕ 0 )
sin(ωτ) ∑ 2J n (
c
)
n
n =1,3,5...
=2
ωd sin(nθ 0 )cos(nϕ 0 ) (1.53b)
sin(ωτ) ∑ δ n Jn ( ) −
n = 0,2,4... c n
j
ωd sin(nθ 0 )sin(nϕ 0 )
cos(ωτ) ∑ 2Jn ( )
n =1,3,5... c n
negligible even for a separation of the order of one wavelength. When the
sources are uniformly distributed on a sphere of large radius, the spatial
sin(kd)
covariance function is given by c12 (0) = [17].
kd
1.4.3 Channel: In this model we shall consider a situation where the noise
sources are uniformly distributed on one of the faces of a channel. This
situation is close to an ocean channel where all noise sources are on or close to
the ocean surface (fig. 1.23) [18].
Array of
Sensors
channel
Figure 1.23: The noise sources on the surface of a ocean channel. Correlation
structure of wavefield in the vertical direction is investigated.
∞ 2π
1
Sψ (ω) = 2
4π ∫ ∫ sS
0 0
ψ (s,α,ω)dαds
For wind generated noise on the ocean surface the frequency wavenumber
spectrum is approximately modeled as
2p s 2 m −1
Sψ (s,α,ω) = 2 Γ(m)(1 − 2 ) S0 (ω) s<k
k k
=0 s≥k
∞
1
ψ(r,ϕ,t) = ∫
2π − ∞
dΨ(r,ϕ,ω)e jωt (1.54a)
+∞
1
3 ∫∫∫
ψ(r,ϕ,t) = dΨ(u,v,ω)e j (ωt −ux − vy) (1.54b)
(2π) − ∞
The pressure field due to a small element of noise sources at a sensor at a depth
z1 is given by (1.45),
∞
1 ω
p1 (t) = ∫ dΨ(r,ϕ,ω)e jωt ∑ φ m (z1 )φ*m (zs )H01 (km r)rdrdϕ
2π − ∞ 2H m
(1.55)
∞
2 1
∫ H (k r)J0 (sr)rdr =
1
π km2 − s 2
0 m
0
and obtain
∞ +∞
1 ω jωt φ (z )φ* (z )
p1 (t) = 3 ∫ e ∫ ∫ dΨ(u,v,ω)∑ m 21 m2 s (1.57a)
2π − ∞ 2H −∞ m km − s
∞ +∞
1 ω jωt φ (z )φ* (z )
p2 (t) = 3 ∫ e ∫ ∫ dΨ(u,v,ω)∑ m 22 m2 s (1.57b)
2π − ∞ 2H −∞ m km − s
∞ ∞
1 1
where p1 (t) = ∫
2π − ∞
dP1 (ω)e jωt and p2 (t) = ∫
2π − ∞
dP2 (ω)e jωt .
For spatially white noise sources, Sψ (s,ω) = S0 (ω) , and the integral in
(1.59) reduces to
∞
sSψ (s,ω) 4 k S (ω)
∫ (k
0
2
m
2 *2
− s )(k − s )
n
2
ds =
π 2
ln( m* ) 2 0 *2
kn (km − kn )
For details on the derivation and numerical results the reader is urged to see [20]
and [21] where a more general model is dealt with. As an illustration of the
general variation of the spectrum and the coherence, the numerical results for a
channel where H= 4λ are shown in fig. 1.24. The noise sources are located on
an annular ring with inner radius = 100λ and outer radius =1000λ. The noise is
presumed to have been generated by wind (m=1). For coherence calculation one
sensor is kept fixed at 2λ and the other sensor is moved along the depth axis.
We considered two types of bottoms, rigid and soft bottoms (ρ 1 =1gms/cc,
c1=1500 m/sec, ρ2=2.0 gms/cc, c2=1600 m/sec). The results are shown in fig.
Sz (ω)
1.24. The y-axis in fig. 1.24a represents a ratio, where Sz (ω) is the
Sψ (ω)
spectrum of the pressure field at z. It is interesting to note that in the hard
bottom channel, because of the trapping of energy, the spectrum is always
greater than that in the soft bottom channel and it increases with depth. The
coherence as a function of the sensor separation (fig.1.24b) is highly
oscillatory, roughly following the interference pattern of vertically travelling
modes.
1
0.01
0.00
0 1 2 3 4
Depth/lambda
(a)
Coherence Magnitude Phase
0 0.5 1.0 -180 -90 0 90 180
-2
-1
(2)
Separation in wavelengths
(t)
(t)
(2)
(b)
Figure 1.24: Spectrum and Coherence as a function of depth. (a) Spectrum. Empty
circles represent soft bottom and the filled circles, hard bottom. (b) Coherence as a
function of sensor separation. (1) Hard bottom. (2) Soft bottom. The magnitude is
shown in the left panel and phase in the right panel.
where
g0(t): waveform transmitted by a source.
t : arrival time of unperturbed ray at nth sensor.
n
τ k : relative arrival time of kth micro-ray.
δτ k (t) : relative delay due to time varying scattering effect.
ak : coefficient of attenuation for kth ray.
Nrays : Number of rays.
1
∞
1 jωt
f n (t) =
2π −∫∞
dG0 (ω) ∑
Nrays k
ak e − jω(t n + τ k +δτ k (t ))
e
(1.61)
∞
1
= ∫
2π − ∞
dG0 (ω)H(ω,t)e jωt
where
1
H(ω,t) = ∑ ak e − jω(tn + τ k +δτ k (t ))
Nrays k
Tr Rc
prism
(a)
A B
Tr Rc
(b)
1
E{Hn (ω,t)} = E ∑ ak e − jω(t n + τ k +δτ k (t ))
Nrays k
where An =
1
∑ E{ak },
Nrays k
{
Φ 0 (ω) = E e − jωτ k } and Φ1 (ω,t) =
{ }
E e − jωδτ k (t ) . For a uniformly distributed random variable in the interval
∆t
± ,
2
∆t
sin(ω )
Φ 0 (ω) = 2
∆t
ω
2
1.5.2 Point Scatterers: In radio communication, the electromagnetic waves
travelling along straight line ray paths may encounter obstacles which would
reflect or scatter the incident wavefield. Since frequency used is very high (800-
1000 MHz, λ ≈ 1meter ) most of the obstacles are likely to be much larger
than the wavelength. There will be significant reflections and corner
diffractions. The air medium is however assumed to be homogeneous and free
from any scattering. Furthermore, a transceiver used in a modern wireless
communication system is likely to be in motion causing a significant Doppler
shift. One consequence of the reflection or the scattering of waves is the
possibility of more than one ray path connecting a transmitter and a receiver.
Such multipath propagation in wireless communication is quite common. As a
result of this, the different components of the wavefield reach a receiver at
slightly different time instants, at different angles and with different Doppler
shifts, but coherently. Thus, multipaths are characterized by the following
attributes: (i) delay diversity (0-10 micro seconds), (ii) angular diversity (5-6
degrees) and (iii) Doppler shift (0 ± 50 Hz).
Delay diversity: The signals arrive at a receiver (usually a single sensor) at
different time intervals, the delay being due to different path lengths. A long
delay implies a weak signal due to multiple reflections, attenuation in the air
and also by geometrical spreading. The quantity of great interest is the power
received at a sensor as a function of delay, known as power delay profile. A
typical power delay profile is sketched in fig. 1.26. Referring to this figure we
define an excess delay spread τ e as the delay within which 90 percent of the
total power reaches the sensor. There are two other parameters commonly used
to characterize a power delay profile,
.1
τe
.01
0 1 2
Delay (in microsecs)
Figure 1.26: A sketch of the power delay profile. Excess delay parameter τe is
shown.
∑p k
where pk is the power in the kth path arriving with a delay τ k . A cumulative
m m
plot of power, a plot of ∑pk =1
k versus ∑τ
k =1
k , is useful in deciphering different
groups of multipaths.
∑p k
to ∆ω max = 200π .
It is interesting to compute a power delay profile for a simple model of
a street lined with buildings which act as obstacles. For simplicity we have
assumed regularly spaced building (spacing=15m and gap=5m) on either sides of
a street, twenty meters wide (see fig. 1.28a). A stationary source in the middle
of the street emits a spike waveform. A suite of rays starting from the source
and ending at the base station antenna, which is 200 meters away from the
street, were traced taking into effect all possible reflections at different
buildings. A sample of five rays thus traced are shown in fig. 1.28a. Perfect
reflection was assumed (reflection coefficient =1) but geometrical spreading was
taken into account. The power delay profile was computed by noting the power
received at computed delay. The computed power delay profile is shown in fig.
1.28b. Other parameters are mean delay=0.042 microsec, rms delay=0.149
microsec and excess delay spread=0.05 microsec.
1.5.3 Local Scattering: A signal emitted by a transceiver is often scattered by
point scatterers in the immediate neighborhood of the transceiver. What reaches
a distant array is a collection of plane wavefronts, differing in phase, angle of
arrival and amplitude but all wavefronts remain correlated. Let (θ 0 + δθ k ) be
the direction-of-arrival (DOA) of the scattered wavefront from the kth scatterer
and θ 0 is the nominal DOA of the direct wavefront from the source (see fig.
1.29). We
v
c ω0
Doppler shift
Base station
0
-2 -1 0 1 2
angle
Figure 1.27: A moving car and stationary scatterer will produce a Doppler shift at
the base station antenna. Inset shows the Doppler shift as a function of angle.
assume that there are L scatterers in the vicinity of the source. The array output
may be expressed as
L −1
f(t) = ∑ α k e − jω c δt k a(θ 0 + δθ k ) f 0 (t) (1.65)
k = 0
a(θ 0 + δθ k ) =
ωcd ωc 2d ω c ( M −1)d
−j sin(θ 0 +δθ k ) −j sin(θ 0 +δθ k ) −j sin(θ 0 +δθ k )
c c c
[1,e ,e ,...,e ]
for a equispaced (spacing = d) linear array (see chapter 2 for more details on the
array response), α k is the coefficient of scattering (complex) and δ tk is the
delay (with respect to direct wavefront) of a wavefront from the kth scatterer.
Note that k=0 refers to the direct wavefront, for which α 0 = 1 and δ t0 = 0 . It
Using (1.66) with the first two terms only in (1.65) we obtain the following
approximate result [24]
2 πd
φ1 2πd j sin θ 0
f(t) ≈ φ 0 1, (1 + j cosθ 0 )e λ , + ... f 0 (t) (1.67b)
φ0 λ
L −1 L −1
where φ 0 = ∑ α k e − jω c δt k and φ1 = ∑ α k δθ k e − jω c δt k . The covariance
k =0 k =0
matrix of the array output is of interest in the DOA estimation. Using the first
order approximation in (1.67a), we obtain
θ0
2∆
×
×
Source
× ×
× Scatterers
{c } f mn
= Lσ 2f 0 σ α2 E{a(θ 0 + δθ)a H (θ 0 + δθ)}mn
2 πd
j ( m − n) sin θ 0 j 2 πd ( m − n)δθ cos θ 0 (1.69a)
= Lσ 2f 0 σ α2 e λ
E e λ
where σ 2f 0 is the variance of the source signal and σ α2 is the variance of the
coefficient of scattering. Assuming δθ is uniformly distributed over a range
±∆ the expected value in (1.69a) may be shown to be
d
j 2 λπd ( m − n)δθ cos θ 0 sin 2π λ ∆(m − n)cosθ 0
E e = d
(1.69b)
2π ∆(m − n)cosθ 0
λ
r ≈ 2∆ cosθ 0 where [x] stands for the largest integer greater than x.
Md
λ
The eigenvectors corresponding to the significant eigenvalues are known as
discrete prolate spheroidal sequences (DSSP) [25]. We have computed the rank
of the matrix Q as shown in table 1.1
1 ∂2 f
∇ f−
2
=0 (1.71)
c(x, y, z)2 ∂t 2
Table 1.1: Rank of Q matrix for different values of ∆cosθ 0 and M=64 and
λ=d/2.
where c is the acoustic wave speed which is a function of the space coordinates.
We shall assume that the density remains constant. (Equation (1.2b) deals with
both speed and density variations.) Let c = c0 + δc(x, y, z) where c0 is the
mean wave speed in the medium and δc(x, y, z) is the fluctuation around the
mean value. A medium is said to be weakly inhomogeneous if
δc(x, y, z) << c0 . The wave equation for inhomogeneous medium (1.71)
reduces to
1 ∂2 f 2δc ∂ 2 f
∇ f − 2 2 =− 3
2
(1.72)
c0 ∂t c0 ∂t 2
where the term on the right hand side of (1.72) represents the contribution due
δc
to speed fluctuations in the medium. Let us represent = εδc̃ where δc̃ is
c0
a normalized function with unit root mean square (rms) magnitude and ε (<<1)
is a constant. Eq(1.72) may be expressed as
1 ∂2 f 2ε ∂ 2 f
∇2 f − = − δc̃ 2 (1.73)
c02 ∂t 2 c02 ∂t
We shall now try to find a series solution of (1.73). Let the series solution be
given by
f = f 0 + εf 1 + ε 2 f 2 +...+ε P f P (1.74)
1 ∂2 f 0
∇2 f 0 − =0
c02 ∂t 2
1 ∂ fp ∂ 2 f p−1
2
2
∇2 f p − = − δc̃
c02 ∂t 2 c02 ∂t 2
2 ∂ 2 f i −1
+∞ − c̃( x ′ , y ′ , z ′ )
1 c02 ∂t 2 jk⋅(r − r ′ )
4π ∫− ∫∞ ∫
f i (r,t) = e dx ′dy ′dz ′ (1.76)
r − r′
+∞
1 2k02δc̃e j (k 0 r − r ′ )
4π ∫− ∫∞ ∫
f 1 (r, t) = f 0 (r' , t)dx ′dy ′dz ′ (1.77)
r − r′
e j ( k0 r − r ′ ) e j ( k0 ρ + z )
2 2
=
r − r′ ρ2 + z 2
(1.78)
∞ 2π
1 λdλ
∫ ∫
− λ2
− k 02 z−z′
= e j (λ( x − x ′ ) cos θ+ λ( y − y ′ ) sin θe dθ
2π 0 0 λ2 − k02
to simplify (1.77)
where Γ represents the space occupied by the scattering medium (see fig. 1.30).
Define u ′ = λ cos θ and v ′ = λ sin θ . It follows that du ′dv ′ = λ dλ dθ .
Equation (1.79) can be expressed as
f 1 (x, y,ω) =
+∞
− jk02
∫ ∫ ∫ δc̃( x ′, y′, z ′)e
− j (( u ′ −u 0 ) x ′ +( v ′ − v 0 ) y ′ +( k 02 − u ′ 2 − v ′ 2 − w 0 ) z ′
dx ′dy ′dz ′
4π 2 −∞Γ
− j ( k 02 −( u ′ 2 + v ′ 2 ) − w 0 )l
e
e j ( u ′x + v ′y) du′dv ′
k − (u ′ + v ′ )
2
0
2 2
(1.80)
Notice that the inner integral over Γ represents a three dimensional Fourier
transform of δ c̃( x ′, y ′, z ′ ) . Hence, (1.80) can be written in the frequency
domain as follows:
f 1 (x, y, ω)
∆c̃(u′ − u0 , v ′ − v0 , k02 − u′ 2 − v ′ 2 − w0 )
2 +∞
− jk (1.81)
=
4π
0
2
−∞
∫∫e − j ( k 02 −( u ′ 2 + v ′ 2 ) − w 0 )l
e j ( u ′x + v ′y) du′dv ′
k − (u ′ + v ′ )
2
0
2 2
j k 02 −( u ′ 2 + v ′ 2 )l
Note that the factor e rapidly decays for (u′ 2 + v ′ 2 ) > k02 and
l > 0 . Such waves correspond to evanescent waves, which will be significant
only in the immediate neighborhood of the scattering object. The presence of
density fluctuation merely introduces an extra term in (1.81). We will not go
into the details but cite a reference where the density fluctuation is accounted for
[27]. For a two dimensional object the scattered field has been obtained by Kak
[28].
z
l
Illumination
Figure 1.30: Γ represents the space occupied by the scattering medium. δ c=0 outside
Γ. The scattered field is evaluated on observation plane l units above Γ .
+∞ j ( k 02 − u ′ 2 − v 0 )l
jk 2 e
f 1 (x, ω) = 0 ∫ ∆c̃(u′ − u , k − u′ − v0 ) e ju ′x du′
2 2
0 0
4π −∞ k − u′
2
0
2
(1.82)
1.6.2 Far Field Approximation: When the scattering object is finite and the
array is placed quite far from the object we shall apply in (1.77a) the far field
e jk⋅(r − r ′ ) e jk (r − r ′ ⋅r̂)
approximation, namely, ≈ where r̂ is the unit vector in
r − r′ r
the direction of r . The error due to this approximation, specially in the
numerator, is illustrated in fig. 1.31. To assess the quantitative effect, consider
the binomial expansion of
r′ r′
( )2 − 2 cos θ
r − r′ = r 2 + r ′ 2 − 2rr ′ cos θ ≈ r(1 + r r +...)
2
r ′2
= (r + − r ′ cos θ +...) ≈ (r − r ′ cos θ )
2r
y r − r′
r
r' o'
x
o
r′2
The error is of the order of in the first term of binomial expansion. This
2r
πr ′ 2
error will introduce a phase error, . For the phase error to be small we
λr
must have r ′ << λr or the largest dimension of the object must be much
2
f 1 (r, t) =
e − j (ωt − k 0 r ) (1.83)
∫ 2k δc̃( x ′, y′, z ′)e
− jk 0 ( r ′ ⋅ r̂) − j (u 0 x ′ + v 0 y ′ + w 0 z ′ )
2
0 e dx ′dy ′dz ′
4πr
Γ
f 1 (r, t) =
e − j (ωt − k 0 r )
4πr ∫Γ
2k02δc̃( x ′, y ′, z ′ )e − j[(u −u0 ) x ′ +(v − v0 ) y ′ +(w − w 0 ) z ′ ]dx ′dy ′dz ′
e − j (ωt − k 0 r ) 2
= 2k0 ∆c̃(u − u0 ,v − v0 , w − w0 ) (1.84)
4πr
where ∆c̃(.) is the Fourier transform of δc̃ . The result derived in (1.84) has
some significance. In the far field region the first order scattered field has the
form of a point scatterer (term outside the square brackets), that is, spherical
waves. The term inside the square brackets depends on the Fourier transform of
the speed fluctuations, evaluated at spatial frequencies determined by the
direction of illumination and the direction of sensor. We shall later in chapter 6
exploit this result for reconstruction of speed fluctuations.
1.6.3 Multisource Illumination: The basic fact used in tomography is that
when an object is illuminated from different directions the scattered field
contains useful information for three dimensional reconstruction. This property
of the wavefield is elaborated in this simple example. Consider a source and
sensor array on opposite sides of an object to be imaged (see fig. 1.32). Let the
a
mth source be fired and the scattered field be sensed by the sensor array. Let r m
be a vector to the mth source and r be a vector to the nth sensor. r ′ is vector
b
n
to a scattering element. The scattered field due to the scattering element at nth
sensor is given by
jk r b − r ′ jk r ′ −r a
1 2k02δc̃e 0 n e 0 m
∆f 1 (m, n) = dx ′dy ′dz ′ (1.85)
(4π)2 r′ − r bn r ′ − r am
Note that the source and sensor arrays are equispaced linear arrays in the y=0
a b
plane. Therefore, the tips of the vectors r m and r n will lie at
[md,0, −La , m = 0, ±1, ±2,⋅⋅⋅] and [nd, 0, Lb , n = 0, ±1, ±2,⋅⋅⋅],
respectively. Using Sommerfeld formula (p. 61) to express
r′
object
La Lb
Sensor array
Source array
x
y
z
Figure 1.32: An array of sources is used to illuminate an object and the scattered
wavefield is received by another array of sensors. The object Fourier transform can
be directly related to the 2D Fourier transform of the received wavefield.
jk r a − r ′ +∞
e − j k 0 −u − v ( z ′ + La ) j[(md − x ′ )u − y ′v ]
2 2 2
e 0 m −j
2 π ∫− ∫∞ k02 − u 2 − v 2
= e dudv (1.86a)
r am − r ′
z ′ < La
jk r b − r ′ +∞
e − j k 0 −u − v (− z ′ + Lb ) j[(nd − x ′ )u − y ′v ]
2 2 2
e 0 n −j
2 π ∫− ∫∞
= e dudv (1.86b)
r′ − r bn k02 − u 2 − v 2
z ′ < Lb
DFT a
r m − r′ m
(1.87a)
+∞ − j k 02 −u 2 − v 2 z ′ + La
−j e
= ∫ ∫
2π − ∞ k0 − u − v
2 2 2
δ(ud − u1 )e − j[ x ′u + y ′v]dudv
e jk 0 r n − r ′
a
DFT
r′ − r n
a
n
(1.87b)
+∞ − j k 02 −u 2 − v 2 − z ′ + La
−j e
= ∫ ∫
2π − ∞ k02 − u 2 − v 2
δ(ud − u2 )e − j[ x ′u + y ′v]dudv
The subscripts m and n on the left hand side refer to discrete Fourier transforms
with respect to index m and index n, respectively. Using (1.87) in (1.85) we
obtain the 2D Fourier transform of the response of a scattering element
∆F1 (u1 , u2 ) =
− j k 02 −( u1 ) 2 − v ′ 2 ( z ′ + La ) − j k 02 −( u2 ) 2 − v 2 ( − z ′ + Lb )
e d
e d
+∞
−k02 u u (1.88)
2 ∫ ∫
k0
2
− ( 1 2
) − v ′ 2
k0
2
− ( 2 2
) − v 2
dx ′dy ′dz ′
2π − ∞ d d
− j[( u1 + u2 ) x ′ +( v ′ + v) y ′ ]
δc̃e d d dv ′dv
Equation (1.88) is now summed over all scattering elements covering the entire
object. We obtain
F1 (u1 , u2 ) =
+∞
−k02 u1 u2 u u
2π 2 ∫ ∫ ∆c̃( d +
−∞
d
, v + v ′, k02 − ( 1 )2 − v 2 − k02 − ( 2 )2 − v ′ 2 )
d d
F1 (u1 , u2 ) =
∞
k02 u1 u2 2 u1 2 u
π ∫ ∆c̃( d +
−∞
, k0 − ( ) − v 2 − k02 − ( 2 )2 − v 2 )
d d d
u1 2 2 u2 2 2
− j k 02 −( ) − v La − j k 02 −( ) − v Lb
d d
e e
× dv (1. 90)
u u
k − ( 1 )2 − v 2
2
0 k − ( 2 )2 − v 2
2
0
d d
If we were to use a line source in place of a point source (1.90) reduces to, by
letting v=0,
F1 (u1 , u2 ) =
u1 u2 u u
2k02 ∆c̃( + , k02 − ( 1 )2 − k02 − ( 2 )2 ) ×
d d d d (1.91)
u u
− j k 02 −( 1 ) 2 La − j k 02 −( 2 ) 2 Lb
d d
e e
u u2 2
k02 − ( 1 )2 k02 − ( )
d d
exp( jk0 r − r ′ )
E1 (x, y, z, ω ) = jω ∫ J1 ( x ′, y ′, z ′, ω ) dx ′dy ′dz ′ (1.93a)
Γ
4 π r − r′
exp( jk0 r − r ′ )
H1 (x, y, z, ω ) = ∫ ∇ × J1 ( x ′, y ′, z ′, ω ) dx ′dy ′dz ′ (1.93b)
Γ
4 π r − r′
and
(1.95) it may be inferred that a strong similarity between the scattered acoustic
field and the scattered EM field exists.
§1.7 Exercises:
ω
−j R ∞ d
c jω(t − cos θ)
1 e
f 1 (t) ≈
2π R ∫ dF(ω,θ,ϕ)e
−∞
2c
where θ,ϕ are respectively elevation and azimuth of the point source. Place
d
another sensor, also on the z-axis, at a distance - . Show that the cross
2
correlation between the outputs is given by
∞
1 ωd
c12 (τ) ≈ ∫
2π − ∞
S f (ω)sin c( )e jωτ dω
c
References
1) To localize a source.
2) To receive a message from a distant source.
3) To image a medium through which the wavefield is
propagating.
In this chapter we shall study the basic structure of a sensor array system and in
the sequel learn how the above objectives are achieved. The most commonly
used array geometries are uniform linear array (ULA) and uniform circular array
(UCA). A uniform planar array (UPA) where sensors are placed on an
equispaced rectangular grid is more common in large military phased array
systems. A wavefront which propagates across the array of sensors is picked up
by all sensors. Thus, we have not one but many outputs which constitute an
array signal. In the simplest case, all components of the array signals are
simply delayed replicas of a basic signal waveform. In the worst case, individual
sensor outputs are strongly corrupted with noise and other interference, leaving
a very little resemblance among them. Array processing now involves
combining all sensor outputs in some optimal manner so that the coherent
signal emitted by the source is received and all other inputs are maximally
discarded. The aperture of an array, that is, the spatial extent of the sensor
distribution, is a limiting factor on resolution. However, the aperture can be
synthetically increased by moving a source or sensor. The synthetic aperture
concepts are extensively used in mapping radars and sonars. In this chapter we
concentrate on sensor array systems which will form the basic material for the
subsequent chapters.
f(t)
θ
wave vector
ULA
7 2 1 0
Reference sensor
Figure 2.1: Uniform linear array of sensors. Note the convention of sensor
indexing. The left most sensor is the reference sensor with respect to which all time
delays are measured.
first sensor is f 0 (t) = f (t) , the output of the second sensor is f 1 (t) = f (t − ∆t)
and so on. Thus, the output of the mth sensor is f m (t) = f (t − m∆t). Some
times it is convenient to represent the sensor output in the frequency domain
∞ md
1 jω(t − sin θ)
f m (t) = ∫
2π − ∞
dF(ω)e c
(2.1)
∞
1 M −1 M −1 − jω md
1 sin θ
jωt 1
g(t) = ∑ m 2π ∫
M m=0
f (t) = dF(ω)e
M
∑= 0 e c
−∞ m
∞
1
= ∫
2π − ∞
dF(ω)H(ωτ)e jωt (2.2)
M
M −1 jω md sin θ sin( ωτ) j M −1 ωτ
1
H(ωτ) = ∑e
M m=0
c
= 2
ωτ
e 2 (2.3a)
M sin
2
1 M −1
md
sin θ
∑
jω
H(ωτ) = am e c
(2.3b)
M m=0
A few samples of the frequency response function (magnitude only) are shown
in fig. 2.2 for different values of M, that is, array size. The response function is
periodic with a period 2 π . The maximum occurs at ωτ = 2nπ . The peak at
n=0 is known as the main lobe and other peaks at n = ±1, ±2,... are known as
grating lobes. Since the magnitude of the array response is plotted, the period
becomes π as seen in fig. 2.2. The grating lobes can be avoided if we restrict
the range of ωτ to ±π , that is, at a fixed frequency the direction of arrival
d 1 π
must satisfy the relation sin θ ≤ . For θ in the interval ± this
λ 2 2
d 1
requirement is satisfied if ≤ . If the range of θ is reduced it is possible to
λ 2
π π
increase the sensor spacing, for example, for − ≤θ+ the sensor spacing
4 4
d 1
need satisfies the constraint ≤ . The phase of the frequency response is a
λ 2
linear function of ωτ . This useful property of a ULA is lost when the sensors
are nonuniformly spaced (see p. 94).
The array response is a function of the product of frequency ω and
d
delay τ or, more explicitly, ω sin θ . The implication of this dependence is
λ
that two wavefronts whose waveform is a simple sinusoid but with different
frequencies (ω1 , ω 2 )arriving at different angles (θ1 , θ 2 ) will produce identical
array response if ω1 sin θ1 = ω 2 sin θ 2 . We shall discuss later such ambiguity
issues when we look into the broadband beamformation. The response function
has a main lobe which is surrounded by many sidelobes of decreasing magnitude
just as we find in spectral windows. The first zero is at
M=8
0.8
0.6
0.4
0.2
0
1
0.8 M=64
Response (magnitude)
0.6
0.4
0.2
0.8 M=128
0.6
0.4
0.2
0
-6 -4 -2 0 2 4 6
Radians
Figure 2.2: Array response function (magnitude) for different values of M. Notice
that the main lobe becomes sharper as the array size is increased.
1 M −1
g(t) = ∑ f m (t + mτ)
M m=0
∞ M −1 j ( τ − sin θ )ωm
d
1 jωt 1
= ∫ ∑
0
dF(ω)e e c (2.5)
2π − ∞ M m=0
∞
1 d
= ∫
2π − ∞
dF(ω)H((τ − sin θ 0 )ω)e jωt
c
d
where we have assumed that the DOA is θ 0 . Let τ = sin θ. Then the array
c
response is maximum whenever θ = θ 0 . We say that the array is steered in the
direction θ 0 , that is, in the direction of arrival of the incident wavefront. The
array response is now a function of DOA. This is demonstrated in fig. 2.3. It is
interesting to note that the width of the main lobe increases with increasing
DOA. To further understand this broadening effect we shall study the array
d
response function around its maximum, that is, at τ = sin θ 0 . The first zero
c
will occur at
M d
sin(θ 0 − ∆θ) = π
d
ω sin θ 0 − ω (2.6a)
2 c c
λ
∆θ = θ 0 − sin −1 sin θ 0 − (2.6c)
Md
The dependence of ∆θ on the DOA for different array sizes is illustrated in fig.
2.4. The broadening of the main lobe is due to reduction in the array aperture
for a wavefront which is incident away from the broadside. The response is
maximum whenever
d
ω (sin θ 0 − sin θ) = 2πn
c
or
d
(sin θ 0 − sin θ) = n
λ
d 1 d 1
For ≤ , the acceptable solution is θ = θ 0 for which n=0. For > there
λ 2 λ 2
d
is more than one solution, one for each grating lobe. For example, let = 1; a
λ
solution of θ = sin −1 (sin θ 0 − n) exists only for n=0 and ±1.
d k
Now, let τ = . The array response function can be written as a
c M
2πd
discrete Fourier transform of a complex sinusoid, exp(− j sin θ 0 m) ,
λ
2 πd 2 πkm d
1 M −1 − j sin θ 0 m
∑e
j
H(k) = λ
e M λ
(2.7)
M m=0
Now H(k) is the kth discrete Fourier transform coefficient which should
k
correspond to the array response at a steering angle, sin −1 ( ) . The array
M
response is thus computed only at a set of discrete angles. Since M is finite,
π
usually a few tens, the angular range of ± is coarsely sampled. To overcome
2
this limitation it may be necessary to pad zeros to the complex sinusoid before
computing the discrete Fourier transform. However, it must be remembered that
0.6
0.4
0.2
0
1
0
60
0.8
Response (magnitude)
0.6
0.4
0.2
1
0
72
0.8
0.6
0.4
0.2
0
-1.5 -1 -0.5 0 0.5 1 1.5
DOA in radians
Figure 2.3: The effect of angle of arrival of a wavefront on the array response. The
mainlobe broadens and the sidelobes become asymmetric.
0.8
Width in radians
16
0.6
0.4
128
0.2
1024
0
0 0.5 1 1.5 2
DOA in radians
Figure 2.4 Width of the main lobe as a function of DOA for three different sizes of
the sensor array (M=16, 128,1024). The sensor spacing is assumed to be λ/2.
this step does not enhance the resolution but only improves the sampling of the
otherwise windowed Fourier transform. Use of discrete Fourier transform for
beamformation was first suggested in [2].
2.1.3 Broadband Source: Often a remote source such as broadband radar, engine
noise, or earthquake, etc. emits a broadband stochastic waveform. The simplest
approach to DOA estimation in such a situation is to compute the spectrum of
the sum of the sensor outputs. From (2.2) we have
∞
1 M −1 1
g(t) = ∑
M m=0
f m (t) = ∫
2π − ∞
dF(ω)H(ωτ)e jωt (2.8)
2
Now consider a plot of H(ωτ) as a function of ω . There is always one peak
at ω =0 and a stream of peaks caused by the incident wavefront [3], at positions
given by the equation below
c
ω peak = 2πn (2.10a)
d sin θ
where n=0, ± 1, ± 2,... We introduce a quantity called minimum array
c
sampling frequency ω min = 2π . An array of sensors may be considered as a
d
waveform sampler which samples the waveform as it propagates across the
d
array. The sampling interval is sin θ and the maximum interval or minimum
c
π
sampling frequency occurs when θ = . In terms of the minimum array
2
sampling frequency the peak may be written as
ω min n
ω peak = (2.10b)
sin θ 0
Evidently, ω peak must be in the range ω min n ≤ ω peak ≤ ∞ . For the sake of
illustration, let the signal spectrum be of infinite width. Now, a plot of
2 ω
H(ωτ) will show an infinite set of peaks spaced at an interval min . For
sin θ
2
example, for θ = 450 an idealised plot of H(ωτ) is shown in fig. 2.5a. A
numerical example is shown in fig. 2.5b where we have assumed a 16 sensor
ULA with spacing d=15 meters. A broadband signal with bandwidth = ( ±200
Hz) is incident at DOA angle equal to 45o. The average spacing of peaks is
42.0 Hz against the theoretical value of 42.43 Hz.
Angular Spectrum: Recall that the frequency wavenumber spectrum of a plane
wave is a line passing through the origin with a slope inversely proportional to
the direction cosines of the wave vector, in particular, on pages 15-17, chapter 1
we have
c c
ω= u, ω = v (2.11)
α β
0 ω min
ω
400
350 (b)
300
250
Sg (ω)
200
S f (ω)
150
100
50
0
0 0.1 0.2 0.3 0.4 0.5
frequency(Hz) x512
1.5
Power
0.5
0
10 20 30 40 50 60 70 80
Angle in deg
Figure 2.6: Angular spectrum obtained by averaging along the radial lines. Two
broadband sources (solid line: 80-120 Hz) and two narrowband sources (dashed line:
98-102 Hz). 16 sensor array, 128 samples with sampling interval 0.005 sec.
snr=0dB.
S̃ f (u,v) = ∫ S (u,v,ω)dω
f
freq band
(2.12)
Such a method was in fact proposed in [3] where the integrated wavenumber
spectrum S̃ f (u, v) was called zero-delay wavenumber spectrum.
Slant Stacking: Closely related to the idea of array steering (or beamforming) is
slant stacking used extensively in seismic exploration. Stacking is also related
to the radon transform [5]. Consider a wavefield, f(t,x), where we shall replace t
1 u
by τ + px x where px = = . The stacking operation is defined as
cx ω
f̂ (τ, px ) = ∫ f (τ + p x, x)dx
−∞
x (2.13)
+∞
+∞
1 j (ωt − p x ωx − p y ωy)
3 ∫∫
f (t, x, y) = ω 2 F(ω, px ω, pyω)e dωdpx dpy
8π − ∞
f̂ ( τ , px , py ) =
+∞ +∞
1 j ( ωτ +( p x′ − p x ) ωx +( p y′ − p y ) ωy)
∫∫ ∫∫ω F(ω , px ω , py ω )e dωdpx′ dpy′
2
dxdy 3
8π
−∞ −∞
∞
1
∫ F(ω , p′ ω , p′ ω )e
jωτ
= x y dω
2π
−∞
Hence,
F̂(ω, px , py ) = F(ω, px ω, pyω)
(2.16a)
2.1.4 Matrix Formulation: When the incident signal is a narrowband signal the
output of an array, in particular a ULA, may be conveniently represented in a
matrix format which reveals some interesting properties. This is also true of a
broadband signal but the processing has to be in the frequency domain.
Representation of Narrowband Signals: A narrowband signal f nb (t) may be
represented as
where
τ0 τ τ
f nb ((t − )) = f i (t)cos(ω c (t − 0 )) − f q (t)sin(ω c (t − 0 ))
4 4 4
≈ f i (t)sin(ω c t) + f q (t)cos(ω c t)
= f nbHilb (t)
τ0
f nb (t) + jf nb (t − ) = f i (t)e jω c t + jf q (t)e jω c t
4 (2.17b)
= f c (t)e jω c t
Figure 2.7. Quadrature filter structure. Since the phase change due to propagation
appears in the complex sinusoid at the output it is easy to introduce phase
adjustments for beamformation.
d d τ
f m (t) = f nb (t − m ) + jf nb (t − m − 0 )
cx cx 4
(2.17c)
d
jω c t − jmω c
= f c (t)e cx
The term representing propagation delay now appears in the complex sinusoid.
Naturally, in order to form a beam, it is convenient to introduce phase
adjustments. Let w0 , w1 ,..., wM −1 be a set of complex weight coefficients for
beamformation. The beam output will be given by
M−1 d
∑w e
jω ct+ jmω c
cx
output= f c (t) m .
m=0
= f c (t)φ(θ 0 )
c
where, it may be recalled that, the apparent speed c x = . Further let the
sin θ 0
sensor response matrix be α(θ 0 ) = diag{α 0 (θ 0 ), α1 (θ 0 ),...α M−1 (θ 0 )}, in which
each element represents the response of a sensor as a function of the angle of
incidence of the wavefront. φ(θ 0 ) represents the propagation effect of the
medium on a wavefront propagating across the array. φ(θ 0 ) and α(θ 0 ) together
form a direction vector a(θ 0 ) = α(θ 0 )φ(θ 0 ) representing the response of an
array to a wavefront incident at angle θ 0 (DOA). Finally, the array output may
be expressed as follows:
When there are P narrowband sources radiating simultaneously the array output
may be expressed as a linear combination of P terms of the type shown in
(2.17e)
f c0 (t)
f c1 (t)
.
α(θ 0 )φ(θ 0 ),α(θ1 )φ(θ1 ),...,α(θ P −1 )φ(θ P −1 )
f(t) = . + η(t)
(M × P)
.
f c P−1 (t)
(P × 1)
(2.18a)
where η(t) is the noise vector assumed to be uncorrelated with the signal
terms. Equationq (2.18a) may be written in a more compact form where P
columns
{
a H (θ)C f a(θ) = σ 2s0 E a H (θ)a(θ 0 )
2
}
2 (2.19b)
d
= σ M H(ω (sin θ − sin θ 0 ))
2
s0
2
{ }
where C f = E f (t)f H (t) is the spatial covariance matrix (SCM). Whenever
θ = θ 0 , that is, when the steering angle is equal to the DOA, the left hand side
of (2.19b) equals σ 2s0 M 2 giving the power of the source.
The M dimensional steering vector will span an M-dimensional space
known as an array manifold. The tip of the steering vector traces a closed curve
in the array manifold or a closed surface when the steering vector is a function
of two variables, for example, azimuth and elevation. Consider the case of
identical sensors, that is,
α 0 (θ 0 ) = α1 (θ 0 ) =...= α M −1 (θ 0 ) = α(θ 0 )
d 2d (M-1)d
- jω sin θ 0 - jω sin θ 0 - jω sin θ 0
a(θ 0 ) = α(θ 0 )col{1,e c
,e c
...e c
}
In the event of sensors being omnidirectional, that is, α(θ 0 ) = constant, the
array manifold becomes a closed curve on a sphere (in M-dimensional space).
For uniqueness the array manifold must not intersect; otherwise, at the point of
intersection, the steering vector will point to two different directions, θ1 and
d
θ 2 , such that a(θ1 ) = a(θ 2 ). Such a possibility exists only when > 0.5. To
λ
show this, consider the steering vector for omnidirectional sensors. Let θ1 and
θ 2 be two such directions for which a(θ1 ) = a(θ 2 ), that is, for all m
d
[sin θ1 − sin θ 2 ] = 1
λ
or
λ
sin θ1 = + sin θ 2 (2.20)
d
λ λ
A solution of (2.20) exists only when < 2 ; for example, when = 1. 2 the
d d
following pairs of directions are the possible solutions: (36.87o , -36.87o ),
(23.58o,-53.13o), and (11.54o,-90o).
The steering vector satisfies the following properties:
Property (a) implies a wavefront coming from the north and another
symmetrically opposite from the south (a and b in fig 2.8) cannot be
distinguished (north-south ambiguity). Property (b) implies a wavefront coming
from the east and another symmetrically opposite from the west (a and c in fig.
2.8) can be distinguished only if the signal is complex (east-west ambiguity).
To show this recall (2.17a) and compare the outputs of a ULA for a real input
signal incident at angle θ and −θ. Let f θ (t) be output of a ULA for an
incident angle, θ , and f −θ (t) be the output for an incident angle, −θ. For a real
signal f θ (t) = f ∗−θ (t) but for a complex signal f θ (t) ≠ f ∗−θ (t). Property (c)
π
implies that there is no grating lobe in the range ± when the sensor spacing
2
λ
is d ≤ .
2
The steering vector is closely related to the array response function. To
show this we define a unit vector, 1=col{1, 1, 1,...,1}, and consider a dot
product
c a
π −θ
−θ θ
x
E
b
.
M −1 jωm d sin(θ)
a(θ) 1 =
H
∑e
m=0
c
(2.21a)
which follows from (2.3). We have assumed for the sake of simplicity that all
sensors are identical and omnidirectional. In real array, the individual sensor
response is likely to be directional and varying. Then, the array response is
given by
M −1 d
∑ α m (θ)e
jωm sin(θ)
a(θ) H 1 = c
(2.21b)
m=0
d 1
H((τ − sin ϕ 0 )ω) = b H φ(r) (2.23)
c M
d
where b = col{α 0 (θ 0 )α ∗0 (θ), α1 (θ 0 )α1∗ (θ),...α P−1 (θ 0 )α ∗P−1 (θ)} and r=ω
c
× (sin θ − sin θ 0 ) . The power output of the array is simply proportional to
the square of the transfer function
1 H
power ∝ H = b φ(r)φ H (r)b
2
2
(2.24)
M
1 e jr e j 2r e j 3r ... e j ( p−1)r
− jr
e 1 e jr e j 2r ... e j ( p−2)r
H = φ(r)φ H (r) = e − j 2r e − jr 1 e jr ... e j ( p−3)r (2.25)
...
− j ( p−1)r − j ( p−2)r
e e ... 1
Sensor
No sensor
Figure 2.9: A four sensor array spread over a 3 λ aperture array will produce all
entries of the matrix on (2.25).
15 0, 1, 2, 3, 7, 11, 15
0, 1, 3, 6, 10, 14, 15
0, 1, 4, 8, 13, 14, 15
0, 2, 4, 5, 8, 14, 15
Table 2.1: Sensor locations for minimum redundancy. For array length of six there is
only one arrangement but for array lengths 15 and 31 there are 77 and 888
arrangements, respectively.
It may be observed that H is a toeplitz matrix, that is, along any diagonal the
entries are repeated even though they refer to different sensors in the array. For
example, consider the second upper diagonal where the entries refer to a pair of
sensors whose indices are m, and m-2, where m=2,3,...,M-2; explicitly, the
pairs of sensors involved in creating the terms on this diagonal are (2,0), (3,1),
(4,2), etc. Thus, insofar as the second diagonal is concerned there are several
redundant pairs of sensors. This redundancy can be removed by selectively
removing sensors [8]. For example, consider the seven sensors uniform array
shown in fig. 2.9. All entries in the H matrix (2.25) can be obtained from just
four sensors shown by filled circles; for example, the first diagonal may be
obtained from sensors at position 5 and 6, the second diagonal from sensors at
prob{∆xl = k ⋅ d} = (1 − p) p k −1 (2.26)
(a)
1
Transfer function (magnitude)
.1
.01
.001
0 32 64 96 128 160 192 224 256
Frequency in Hz
10
(b)
1
Transfer function (magnitude)
.1
.01
.001
0 32 64 96 128 160 192 224 256
Frequency in Hz
Figure 2.10: (a) Transfer function of a 16 sensor ULA and (b) Transfer function of a 7
sensor minimum redundancy array of same length as ULA in (a). Sensor spacing=15
meters, wave speed=1500 m/s and θ = 450 .
n
ω
+∞ j (ωt − sin θ 0 ∑ ∆x l )
1
∫
c
f n (t) = dF(ω)e l=0
2π − ∞
(2.27)
+∞ n ω
1 −j sin θ 0 ∆x l
= ∫ dF(ω)e j (ωt ) ∏ e c
2π − ∞ l=0
1 N −1
g(t) = ∑ f n (t − τ n )
N n=0
+∞ ω
(2.28)
N −1 n − j (sin θ 0 −sin θ)∆x l
1 j (ωt ) 1
= ∫
2π − ∞
dF(ω)e ∑ ∏
N n=0 l=0
e c
We shall now evaluate the expected value of the delay and sum processor given
by (2.28)
+∞
1 j (ωt ) 1
N −1 n
− j ωc (sin θ 0 −sin θ)∆x l
E{g(t)} =
2π −∫∞
dF(ω)e ∑ ∏ E e
N n=0 l=0
(2.29)
ω
−j d (sin θ 0 −sin θ)
(1 − p)e c
= ω (2.30)
−j d (sin θ 0 −sin θ)
1 − pe c
ω n
+∞
N −1
− j d (sin θ 0 −sin θ)
(1 − p)e
c
1 1
E{g(t)} = ∫
2π − ∞
dF(ω)e j (ωt ) ∑
N n=0 − j d (sin θ 0 −sin θ)
ω
1 − pe c
+∞
j (ωt ) 1 1 − Q
N
1
2π −∫∞
= dF(ω)e (2.31)
N 1− Q
where Q stands for the quantity inside the square brackets of (2.31). Note that
ω
−j d(sin θ 0 −sin θ)
for p=0, Q = e c and (2.31) reduces to a known expression for the
response function of a ULA. The response of an array with missing sensors,
given by
1 1 − QN
H(ωτ) =
N 1− Q
d
where τ = (sin θ 0 − sin θ) , is shown in fig. 2.11 for two different values of
c
probability of malfunctioning. It is assumed that the total number of live
sensors is the same in both cases, namely, 16. Notice that the magnitude
response has fewer sidelobes but the phase characteristics appear to be grossly
different from those of a ULA, which is a linear function of ωτ (see (2.3)). A
nonlinear phase response results in a distortion of the received waveform
particularly when it is broadband.
Random Array: A third type of nonuniform array is one where the sensors are
spaced at random intervals, in particular, an exponential distribution for which a
closed form solution can be derived. Let x n , n = 0,1,... N − 1 be the locations of
the sensors; all of which, however, lie on a straight line. Let us assume an
exponential distribution for the sensor spacing
where pdf stands for probability density function and β is a parameter in the
exponential distribution. The output of the nth sensor may be written as in
(2.27). The delay and sum type of processing would result in the following
array output (from 2.29)
0.8 (a)
p=0.2
0.6
Magnitude
0.4
p=0.4
0.2
0.0
0.0
-0.5
p=0.2
-1.5
-2.0
-2.5
0.0 0.8 1.6 2.4 3.2
ωτ
Figure 2.11: (a) Magnitude response and (b) Phase response of an array with
malfunctioning sensors. The total number of live sensors is assumed to be the same
in both cases, namely, 16.
N
β
1−
ω
+∞ j (sin ϕ 0 − sin ϕ) + β
E{g(t)} =
1 1 c
2π −∫∞
dF(ω)e j (ωt )
N 1− β
ω
j (sin ϕ 0 − sin ϕ) + β
c
(2.33b)
The array response may be expressed in terms of the product of the wavelength
and parameter β,
N
βλ
1−
1 j2π(sin θ 0 − sin θ) + βλ
H(ν,βλ) = (2.33c)
N 1− βλ
j2π(sin θ 0 − sin θ) + βλ
where ν = 2π(sin θ 0 − sin θ). We have plotted the array response function for
different values of βλ in fig. 2.12. While the magnitude response is free from
sidelobes, the phase response is highly nonlinear in the range, ν=0.0 to 1.6,
where the magnitude response is significant. This behavior was also noticed in
the case of an array with missing sensors.
2.1.6 Flexible Array: We have so far considered a sensor array which is rigidly
fixed to the ground or to a platform. We now consider an array where the
sensors are held in position by means of a flexible rope which allows a sensor
to move over a circular arc of fixed radius. The sensor spacing, however,
remains unchanged. An important consequence of this freedom allowed to the
array is to alter the shape of the array when it is being towed or it is under the
influence of ocean currents. A commonly used array model is that M sensors are
separated by straight line segments of fixed length d [11] (see fig 2.13a). Let
(x m , ym , zm ) represent coordinates of the mth sensor with respect to m+1st
sensor.
0.8 (a)
βλ=2
0.6
Magnitude
βλ=1
0.4
0.2
0.0
0.0 1.6 3.2 4.8 6.4
v
0.0
βλ=1
-0.5 (b)
βλ=2
-1.0
Phase
-1.5
-2.0
0.0 1.6 3.2 4.8 6.4
v
Figure 2.12: Frequency response of a random linear array. (a) magnitude response.
(b) phase response. 16 sensors spaced at random intervals having an exponential
distribution.
Since the distance between the sensors remains fixed (x m , ym , zm ) must satsify a
β2 d
y
α2
x
(a)
0.5
0.4
0.3
y-axis
0.2
0.1
0.0
0.0 2.0 4.0 6.0
x-axis
(b)
Figure 2.13: (a) Model of a flexible array where the adjacent sensors are held at a
fixed distance but the azimuth and elevation of the line segment joining the sensors
are random variables. (b) A bow shaped 16 sensor array in the x-y plane. The
distance between the adjacent sensors is 0.5 λ . All dimensions are in units of
wavelength .
xm = d sinβ m sin α m
ym = d sinβ m cosα m (2.34a)
zm = d cosβ m
where (α m ,β m ) are azimuth and elevation of a line segment joining mth sensor
with m+1st sensor (see fig 2.13a). It is assumed that x 0 = y0 = z0 = 0 .
m ω
ω c t − ∑ xi c sin θ 0 sin ϕ 0
i=0 c
f m (t) = s0 (t) cos + ηm (t) (2.34b)
− ∑ yi ω c sin θ 0 cos ϕ 0 − ∑ zi ω c sin θ 0
m m
i=0 c i=0 c
Transforming into a complex analytical signal (2.17d) the array output may be
expressed in a matrix form
γ m = γ m −1 + sinβ m cosα m
ε m = ε m −1 + sinβ m sin α m
ξ m = ξ m −1 + cosβ m
0.8
0.6
Response
0.4
0.2
0
0 20 40 60 80 100
Angle in deg.
Figure 2.14: Response of a bow shaped array (solid curve) shown in fig. 2.13b and
the dashed curve is for an undeformed array.
Notice the extremely broad main lobe which is fortunately located at the right
position. If we further increase the deformation, for example, the bow height is
increased to 2.3λ the array response is found to be totally distorted. Even the
main lobe is found to be wrongly placed.
Circular Array
Triangular Array
Figure 2.15: Planar array geometries are shown above. The square or rectangular
array is a natural extension of ULA and other geometries are sparse versions of
square or rectangular array.
+∞
1
f (t, x, y) = ∫
2π − ∞
F(ω)e j (ωt −ux − vy) dω (2.35)
ω ω
where u = sin θ cos ϕ and v = sin θ sin ϕ , is incident on a UPA. The
c c
output of the (m,n)th sensor is
θ
y
ϕ
x
d2
d1
+∞
1
f m1 m2 (t) = ∫
2π − ∞
F(ω)e j (ωt −um1 d1 − vm2 d 2 ) dω (2.36)
M1 M2
∑ ∑a
1
g(t) = m1 m 2 f m1 m2 (t)
M1 M2 m1 =1m 2 =1
+∞ M1 M 2
∑ ∑a
1 1
∫ F(ω )e
jωt − j (um1 d1 + vm 2 d 2 )
= m1 m 2 e dω
2π M1 M2 m1 =1m 2 =1
−∞
(2.37)
+∞
1
∫ F(ω )e
jωt
= H(ud1 , vd2 )dω
2π
−∞
where
M1 M2
1
H(ud1 ,vd2 ) =
M1 M2
∑ ∑a
m1 =1m 2 =1
m1 m 2 e − j (um1 d1 + vm2 d 2 ) (2.38a)
m1d1 md
τ m1 ,m2 = sin θ cosϕ + 2 2 sin θsin ϕ
c c
where ϕ and θ are respectively azimuth and elevation to which the array is
required to be steered. In place of (2.37) we have
M1 M2
g(t) = ∑∑f
m1 =1m 2 =1
m1 m 2 (t + τ m1 ,m2 )
+∞ M M
1 1 2
= ∫
2π − ∞
F(ω)e jωt
∑ ∑
m1 =1m 2 =1
e − j[(u0 −u)m1 d1 +(v0 − v)m2 d 2 ]dω (2.39)
+∞
1
2π −∫∞
= F(ω)e jωt H((u0 − u)d1 ,(v0 − v)d2 )dω
where
+∞ M M 2 πkm1 2 πlm 2
j( + )
1 1 2
g(t) =
2π −∫∞
F(ω)e jωt
∑∑
m1 =1m 2 =1
e − j (u 0 m1 d1 + v 0 m 2 d 2 )
e M1 M2
dω
(2.40a)
+∞
1
2π −∫∞
= F(ω)e jωt DFT{e − j (u0 m1 d1 + v0 m2 d 2 )}dω
and
This result is an extension of (2.7) for a ULA to a UPA. The spatial frequencies
u and v, which appear in (2.40b), are linearly related to the frequency numbers,
k and l, in DFT
2πk 2πl
u= , v=
M1d1 M1d1
Given the frequency numbers, we can get the azimuth and elevation angles. For
λ l
example, assuming d1 = d2 = and M1 = M2 = M , ϕ = sin −1 and
2 k + l2
2
2 k2 + l2 M M
θ = sin −1 . Note that 0 ≤ k, l ≤ and 0 ≤ k 2 + l 2 ≤ . Thus, the
M 2 2
acceptable domain for k and l is as shown in fig. 2.17. For those values of k
and l lying outside this domain we get nonpropagating waves (see chapter 1,
page 11).
Random Array: An array with its sensors at random locations, modelled as an
independent, identically distributed random variable, is known as a totally
random array [12]. The root mean square (rms) error in the estimation of the
M/2
Propagating
domain
k
M/2-1 0 M/2
M/2-1
Figure 2.17: Acceptable domain for frequency numbers, k and l, is shown above. For
those values of k and l lying outside this domain we get non propagating waves.
azimuth angle (measured with respect to the x-axis) is given in [12] and
reproduced here without derivation
ϕ x
a
Figure 2.18: Sensors are uniformly spaced on the circumference of a circle of radius
a. A plane wave is incident at an azimuth angle ϕ and an elevation angle θ .
∞ ωa 2 πm 2 πm
1 j[ωt − (cos ϕ sin θ cos +sin ϕ sin θ sin )]
f m (t) = ∫
2π − ∞
F(ω)e c M M
dω
(2.41)
∞ ωa 2 πm
1 j[ωt − (sin θ cos( −ϕ))]
= ∫
2π − ∞
F(ω)e c M
dω
Note that time is measured with respect to the time of arrival of the wavefront
at the center of the array. First, we evaluate the frequency response function.
The sum of all outputs in the frequency domain is given by
∞ M −1 − j[ ωa 2 πm
1 (sin θ cos( −ϕ))]
jωt 1
g(t) =
2π −∫∞
F(ω)e ∑e
M m=0
c M
dω
(2.42)
∞
1
= ∫
2π − ∞
F(ω)H(ω,ϕ,θ)e jωt dω
For large M (for example, M> 48 when a=6λ and M>32 when a=4λ) the
summation in (2.43a) may be replaced by an integral and the result is
ωa
H(ω , ϕ , θ ) ≈ J0 ( sin θ ) (2.43b)
c
We shall call such a UCA a fully populated array. The most interesting
property of a circular array is that the frequency response function is independent
of ϕ . The property arises from (2.43b). Taking the distance to the first zero as
the effective half width of the main lobe, the angular width will be equal to
c
∆θ = sin −1 (2. 45 ). The height of the first (largest) sidelobe is 0.4025 at
ωa
c
θ = sin −1 (0.8 ).
ωa
A circular array may be steered to any desired direction just like a ULA
or a UPA. A delay τ m is introduced at each sensor output before summation,
where
a 2πm 2πm
τ m = [ (cosϕ sin θ cos + sin ϕ sin θsin )]
c M M
and ϕ and θ respectively are the desired azimuth and elevation angles. The
delayed outputs of all sensors are summed, for the time being without any
weighting.
1 M −1
g(t) = ∑ f m (t + τ m )
M m=0
∞ M −1 − j[ ωa 2 πm 2 πm
1 (sin θ 0 cos( −ϕ 0 )−sin θ cos( −ϕ))]
jωt 1
=
2π −∫∞
F(ω)e ∑e
M m=0
c M M
dω
(2.44)
where ϕ 0 and θ 0 respectively are the unknown azimuth and elevation angles of
the incident wavefront. Let
ωa 2 πm 2 πm
ωa 1 M −1 − j[ (sin θ 0 cos( −ϕ 0 )−sin θ cos( −ϕ))]
H( ,θ 0 ,ϕ 0 ,θ,ϕ) =
c
∑e
M m=0
c M M
ωa
2
output power= F(ω)H( ,θ 0 ,ϕ 0 ,θ,ϕ)
c
Earlier we had noted that steering of an array is equivalent to the spatial Fourier
transform of the array output. This result holds in a slightly different form for a
circular array. We will demonstrate how the spatial Fourier transform can be
used for estimation of azimuth [13] and elevation angles. Consider the spatial
discrete Fourier transform of the circular array output.
2 πm
1 M −1 −j
∑
k
gk (t) = f m (t)e M
M m=0
∞ M −1 − j[ ωa (sin θ cos( 2 πm − ϕ ))]
1 jωt 1
=
2 π −∫∞
F(ω )e ∑e c
M m=0
M
dω
(2.45)
∞
1
= ∫
2π −∞
F(ω )H(ω , ϕ , θ )e jωt dω
π
ωa jk
Gk (ω ) ≈ F(ω )Jk ( sin θ 0 )e 2 e − jkϕ 0 (2.46)
c
ωa
which is valid for k < kmax ≈ [15] and for sensor spacing approximately
c
λ
equal to [14]. Consider the following quantity:
2
ωa
Jk +1 ( sin θ 0 )
Gk +1 (ω) c
= je − jϕ 0 (2.47)
Gk (ω) ωa
Jk ( sin θ 0 )
c
2k
Jk +1 (x) = Jk (x) − Jk −1 (x) ,
x
Jk +1 (x) 2k Jk −1 (x)
= −
Jk (x) x Jk (x)
which we use in (2.47) and derive a basic result for the estimation of ϕ 0 and
θ0
Gk +1 (ω) 2k G (ω)
− je jϕ 0 = − je − jϕ 0 k −1 (2.48)
Gk (ω) ωa Gk (ω)
sin θ 0
c
Equation (2.48) may be solved for ϕ 0 and θ 0 As an example, we consider a 16
sensor circular array of 3λ radius and a source in a far field emitting a
bandlimited random signal. The center frequency is 100Hz and the bandwidth is
10Hz. The azimuth and the elevation angle of the source are respectively 10o
(0.1745 rad) and 45o (0.7854 rad). The sampling rate was 500 samples/sec. The
estimates were averaged over all frequency bins lying within the bandwidth. The
results are shown in fig. 2.19. Notice that the standard deviation of the
estimates decreases considerably when a reference sensor is used at the center.
The decrease is more pronounced at a very low snr, e. g., at 0 dB the decrease is
by a factor of three or more. An analysis of errors has shown that the standard
deviation is dominated by a few outliers which are caused by random noise in
the array output. Unless these outliers are eliminated the mean and the standard
deviation of the estimate gets severely affected. To overcome this problem
median in place of mean may be considered. It was observed through computer
simulation that the median is a better estimate of the azimuth than the mean.
When there is more than one source (say, P sources), equation (2.46)
takes a form
π
p−1
ωa
Gk (ω) = ∑ Fi (ω)Jk (
jk
sin θi )e 2 e − jkϕ i + ηk (ω) (2.49a)
i=0 c
We shall rewrite (2.49a) in a matrix form. For this purpose we define the
following vectors amd matrices:
ωa ωa ωa
Di (ω ) = diag J0 ( sin θ i ), J1 ( sin θ i ), ... Jr −1 ( sin θ i ),
c c c
0.12
0.10
Standard deviation
0.08
0.06
0.04
0.02
0.00
512 1024 1536 2048
Data length
− j (ϕ i − π2 ) − j 2(ϕ i − π2 ) π
− j (r −1)(ϕ i − )
Zi = col 1,e ,e ,...e 2
ηr (ω) = col{η0 (ω), η1 (ω),...ηr −1 (ω)}
where r is an integer (P ≤ r ≤ M) .
P −1
G r (ω) = ∑ Di (ω)Zi Fi (ω) + ηr (ω) (2.49b)
i=0
Exploiting the above signal structure a subspace approach has been developed in
[14, 16] for the estimation of azimuth and elevation.
Steering Vector: For circular array we define a steering vector as
a(ϕ , θ )
− j[ ωa (sin θ cos( ϕ ))] − j[ ωca (sin θ cos( 2Mπ − ϕ ))] − j[
ωa
(sin θ cos(
2 π ( M −1)
− ϕ ))]
= col e c ,e ,...e c M
(2.51)
Each term in the steering vector can be expanded in a series form [15]
ωa 2 πm ∞ 2 πm
− j[ (sin θ cos( −ϕ))] ωa −ϕ)
∑
jk (
e c M
= (− j)k Jk ( sin θ)e M
k =−∞ c
Define a matrix
m = 0,1,... M − 1
2 πkm
j
{W}km = e M
k = 0, ±1,...∞
and a vector
π
ωa − j ( +ϕ)k
{V(ϕ,θ)}k = Jk ( sin θ)e 2 , k = 0, ±1,...∞
c
In terms of matrix W and vector V we can express (2.51) as
a(ϕ , θ ) = W V(ϕ , θ )
M ×1 M × ∞ ∞ ×1
ωa
e sin θ
ωa 1 c
Jk ( sin θ) ≈ ( )k
c 2πk 2k
ωa
Hence, the size of vector V(ϕ, θ) will be of the order of e sin θ . For
c
π
example, for a = 8λ and θ = the size of the vector is about 140. The
2
steering vector for a circular array possesses some interesting properties,
namely,
Property (a) implies a wavefront coming from the north can be distinguished
from one coming from the south (north-south ambiguity). Property (b) implies
that a complex signal coming from the east can be distinguished from the one
coming from the west (east-west ambiguity; see fig. 2.20 for an illustration)
and property (c) implies that for any sensor spacing there is no grating lobe in
the range of ±π . A circular array differs from a linear array in respect of
properties (a- c).
Boundary Array: An array of sensors for localization of an object in the near
field region may take a form of a boundary array where the sensors are placed all
around the object as its approximate location is known before hand. A circular
array enclosing a source is an example of boundary array. With three coplanar
sensors and accurate time delay measurements it is possible to localize a point
target in the plane of array. There is a vast literature on time delay estimation
but, on account of space limitation, the topic of time delay estimation will not
be covered in this book. It is, on the other hand, possible for a boundary array
to determine the time delays from the phase measurements. The basic idea is to
consider a pair of sensors and estimate the phase difference at a fixed frequency
ω . The source will fall on one of the phase trajectories drawn with a phase
difference ω∆τ = 2πs + φ where ∆τ is the unknown time delay, s is an integer
and φ is the actual observed phase difference. To estimate the time delay ∆τ ,
c a
π −ϕ
ϕ
x E
−ϕ
Figure 2.20: Circular array does not suffer from north-south ambiguity, that is,
wavefronts a and b can be distinguished. There is no east-west ambiguity for
complex signals, that is, wavefronts a and c can be distinguished.
ω∆τ
we need to know the integer constant s in the range ± s0 where s0 = Int
2π
and Int[ x ] stands for the largest integer less than x. For a given φ , ∆τ will
assume a set of 2s0 + 1 values and for each value of s there corresponds a locus
of points called the phase trajectory [17]. For every pair of adjacent sensors we
can draw a suite of trajectories as shown in fig. 2.21a. The unknown source
must lie on any one of the trajectories. Next, consider another sensor pair and
draw another suite of trajectories. Any one of the points of intersection is a
possible location of the unknown source (see fig. 2.21b). Since there are M
pairs of sensors there will be M suites of trajectories. The true position of the
unknown source is then given by the intersection of all M trajectories, one
from each suite. At true source location all phase estimates obtained from
different pairs of sensors must match with the theoretically evaluated phases.
Let φ̂1k , k = 0,1,..., M − 1 be the estimated phases from M pairs of adjacent
sensors and φ1k , k = 0,1,..., M − 1 be the theoretically computed phases. Define
an error vector
(b)
F igure 2.21: (a) For every pair of adjacent sensors we draw a suite of equiphase
trajectories. (b) Intersection of two sets of equiphase trajectories. The unknown
source must lie at one of the intersections. For every adjacent pair of sensors the
search is carried out within the dotted quadrilateral.
{ 0 0 1 1 M −1
ε = col (e jφ1 − e jφ̂1 ),(e jφ1 − e jφ̂1 ),...,(e jφ1
M −1
− e jφ̂1 ) } (2.52)
H
and error power= ε ε . The error power will be zero at the true location of the
source. This property may be utilized to spot the true location from all
available intersections of any two suites of trajectories. Evaluate the error power
at each and every intersection. That intersection which yields zero (or
minimum) error power is the most likely location of the unknown source.
Finally, to reconstruct the true phase we need to know the integers which may
be obtained from the order of the trajectory for every pair of sensors passing
through the source location. However, in any practical problem, this step may
not be required as the interest is usually in localization and not in the true phase
retrieval. The results of a numerical experiment are shown in fig. 2.22. The
question of minimum size (number of sensors) of the array, required for unique
localization, has not been answered. However, numerical experiments suggest a
minimum array size of five sensors placed the circumference of a large circle. It
is not necessary that the array be perfectly circular. Any closed curve will do,
provided the phase trajectories are drawn for each pair.
Figure 2.23: Randomly distributed planar dipole sensors. Each dipole consists of
two identical sensors displaced by ∆ (vector).
F1
F̃ = (2.53)
F 2
where
F1 = col{F0+ (ω), F1+ (ω)... FM+ −1 (ω)}
F 2 = col{F0− (ω), F1− (ω)... FM− −1 (ω)}
where superscript + refers to the upper sensor and - refers to the lower sensor.
Let P plane wavefronts from P sources be incident on the array.
[ ]
F1 = a 0 ,a1 ,...a p−1 F(ω) + dη1 (ω) (2.54)
where
− j ∆ ⋅δ i − j ω (r1 + ∆ )⋅δ i ω ∆
− j (r M −1 + )⋅δ i
a i = col e 2 ,e c 2 ,...,e c 2
∆⋅δ i
− j ωc r1 ⋅δ i ω
− j r M −1 ⋅δ i − j
= col 1,e ,...,e c
e 2
∆⋅δ p−1
− j ∆⋅δ 0 − j ∆⋅δ1 −j
Γ + = diag e 2
,e 2
,...,e 2
and
1 1 ... 1
− j ω r ⋅δ − j ω r ⋅δ ω
− j r 1 ⋅δ p−1
e c 1 0 e c 1 1 e c
.
A=
.
.
ω ω ω
e − j c r M −1 ⋅δ 0 e − j c r M −1 ⋅δ1 ... e − j c r M −1 ⋅δ p−1
Similarly,
where
∆⋅δ p−1
j ∆⋅δ 0 j ∆⋅δ1 j
Γ − = diag e 2
,e 2
,...,e 2
We shall now compute the spectral matrix of dipole array output. For this the
incident signal will be treated as a stationary stochastic process. Then, in place
of the ordinary Fourier transform we need to invoke the generalized Fourier
transform of a stationary stochastic process (see p.40).
where
{
Γ = Γ − Γ H+ = diag e jω∆⋅δ 0 ,e jω∆⋅δ1 ,...,e
jω∆⋅δ p−1
}
{
S f = diag S f 0 , S f 1 ,..., S f p−1 }
Equation (2.56) may be expressed as
I 0 σ η1
H 2
A A
S f̃ = S f + 2 (2.57)
AΓ AΓ 0 I σ η2
T
j ωdc sin θ j 2 ωdc sin θ j ( M −1)
ωd
sin θ
j (ωt +ϕ)
f x = − cos γ 1,e ,e ,...,e c
Ee
= − cos γ aEe j (ωt +ϕ)
(a)
z
x
(b)
Figure 2.24: A ULA of electric dipoles oriented in the x direction (a) and in the y
direction (b).
T
j ωd sin θ j 2 ωdc sin θ j ( M −1)
ωd
sin θ
j (ωt +ϕ)
f y = sin γ cosθe 1,e cjκ
,e ,...,e c
Ee
= sin γ cosθe jκ a Ee j (ωt +ϕ)
where
Γ1 = −diag{cos γ 0 ,cos γ 1 ,...,cos γ P −1}
A = {a 0 ,a1 ,...,a P −1}
{
S0 = col E0 e jϕ 0 , E1e jϕ1 ,..., EP −1e
jϕ p−1
}
and η1 (t) = col {η0′ (t), η1′ (t),..., η M−1
′ (t)} is the background noise vector. A
similar expression for an array of dipoles oriented along the y-axis is given by
{
Γ 2 = diag sin γ 0 cosθ 0 e jκ 0 ,sin γ 1 cosθ1e jκ 1 ,...,sin γ P −1 cosθ P −1e jκ P−1 }
η2 (t) = col{η′′0 (t), η1′′(t),..., η′′M −1 (t)}
2π c 2π 2π
∆ω = = = (2.60)
M d sin θ Md τ
cx
where τ is the time required to sweep across the array aperture. Note that the
bandwidth is infinite when the wavefront is incident on the broadside, that is,
2 πc
the array is steered in the direction of source; but it is equal to when the
Md
wavefront is incident from the endfire. For a circular array, however, the
wavefront has to sweep a constant aperture equal to the diameter of the circle,
independent of the azimuth angle. This essential difference between a ULA and a
UCA is illustrated in fig. 2.25. The bandwidth of a long ULA or a large
diameter UCA is very small and hence much of the energy of a broadband
source will be lost, unless the array is steered in the direction of the desired
source. Further, as pointed in §2.1, when there is more than one source
radiating at different frequencies there is a possibility for ambiguity. Let P
narrowband sources with DOAs θ i , i = 0,1,..., P − 1 and center frequencies ω i ,
i = 0,1,..., P − 1 be incident on a ULA. Further, we assume that the center
frequency and DOA pair satisfies the following relation:
d
ωi sin θi = τ 0 (cons tan t), i = 0,1,..., P − 1 (2.61)
c
where τ 0 is the delay per sensor introduced by all sources. Now, through a
process of sum and delay, the array is steered simultaneously, to all sources; in
other words, the array will “see” all P sources at the same time. When the
sources are broadband with overlapping spectrum we can always find a set of
frequencies which satisfies (2.61). As a result, when the array is steered to one
of the sources the output may be contaminated with the power derived from
other sources. Such an interference is unacceptable particularly when waveform
(a)
2a
(b)
Figure 2.25: The essential difference between a ULA and UCA is that the effective
aperture (dashed line) is azimuth dependent for a ULA but independent of the azimuth
for a UCA. (a) Linear array, the effective aperture= (M − 1)d sin θ . (b) Circular
array, the effective aperture = diameter of the circle.
∞
1
f(t) = ∫
2π − ∞
a(ω,θ 0 )dF0 (ω)e jωt (2.62)
where
∞
1
f −θ 0 (t) = ∫
2π − ∞
a∗ (ω,θ 0 )dF0 (ω)e jωt ≠ f θ 0 (t)
except in the unlikely event of dF0 (ω ) being real. Thus, the east-west
ambiguity shown for the narrowband complex signal does not apply to the
broadband signals.
2.3.2 Broadband Signals: In the case of thebroadband signal, since a snapshot
may vary rapidly, it is necessary that many snapshots must be collected at
different time instants in the past; for example, f (t), f (t − ∆t), f (t − 2∆t),...,
f (t − ( N − 1)∆t) are N past or delayed snapshots (see fig. 2.26). An alternate
approach, in the case of broadband signal, is to go over to the frequency domain
(temporal frequency). The output of each sensor, consisting of N samples, is
subjected to Fourier analysis (DFT). A collection of the Fourier coefficients,
one from each sensor at a fixed frequency, constitutes a frequency snapshot. The
array signal processing of broadband signals using the frequency snapshots
closely follows the time domain approach for narrowband signals (after
removing the carrier), widely used in radar signal processing. In place of a
covariance matrix we use a spectral matrix which indeed is a spatial covariance
matrix (SCM). In this book we shall emphasize the frequency domain approach
as the time domain approach for wideband signals turns out to be conceptually a
bit more involved. We begin with time domain approach. First, let us introduce
(2.63)
As an example, consider a two sensor array (M=2) and two delayed snapshots
(N=2). The STCM is given by
C STCM =
d sin θ 0
where τ 0 = is a propagation delay per sensor. We express the stacked
c
vector by using (2.62) in (2.63),
∞
1
f stacked (t) = ∫
2π − ∞
ζ(ω) ⊗ a(ω,θ 0 )dF0 (ω)e jωt (2.65)
where
and ⊗ stands for the Kronecker product. Define a direction vector h(ω , θ 0 ) =
ζ(ω ) ⊗ a(ω , θ 0 ) and rewrite (2.65) as
Figure 2.26: The output of an eight sensor ULA is sampled at different time instants
in the past. All outputs taken at the same time instant are grouped into a vector
called a temporal snapshot. We have as many temporal snapshots as the number of
samples or taps taken from each sensor output.
∞
1
f stacked (t) = ∫
2π − ∞
h(ω,θ 0 )dF0 (ω)e jωt (2.66)
which may be considered as an extended output of the array (stacked vector) due
to a source at azimuth angle θ 0 . Using (2.66) we can express STCM as
(2.67)
Premultiply on both sides of (2.68) with the steering vector in some desired
direction and evaluate the expected value of the magnitude square. Dividing by
(M ⋅ N ) 2 we get the output power
1
h H (ω, θ)S f (ω)h(ω, θ) =
(M ⋅ N )2 stacked
2 2 (2.69)
h H (ω, θ)h(ω, θ 0 ) a H (ω, θ)a(ω, θ 0
S0 (ω) = S0 (ω)
M⋅N M
d
ωi sin θi = τ 0 (cons tan t), i = 0,1,..., P − 1,
c
the steering vector a(ω i ,θi ) remains unchanged. This, however, is not true
for a ULA with a tapped delay line, as the steering vectors, h(ω i , θ i ) =
ζ(ω i ) ⊗ a(ω i , θ i ), i = 0,1,..., P − 1, will be different because of ζ(ω i ). It may
be noted that for a circular array there is no frequency-direction ambiguity of the
type described above even when no tapped delay line is used [18].
and the response function of the sensor array in the direction of the scatterer is
given by
+∞
2π
where u = k sin θ cos ϕ , v = k sin θ sin ϕ and k = is the wavenumber. Let
λ
ρ (u, v) be the reflection coefficient as a function of the azimuth angle ϕ and
the elevation angle θ . The total response of the source/sensor array is given by
O(u,v) = T(u,v)ρ(u,v)R(u,v)
= t( p)r(q)
Thus, we note that the combined PSF of a source array along the x-axis and a
sensor array along the y-axis is equal to the product of two weighting functions;
+∞
The PSF at point (p,q) is obtained by summing the product of two weight
functions over two lines p=x+x´ and q=y+y´ where (x,y) refer to a point in the
source array and (x´,y´) refer to a point in the sensor array. Such a coarray is
also known as a sum coarray. Since we have a discrete set of point sources and
sensors, the discrete version of equation (2.71) is given by R T T where
T = col {t0 , t1 ,...t M−1 } and R = col {r 0 , r1 ,... r M−1 } .
An arbitrary PSF may be synthesized through eigen-decomposition.
Let P be the desired PSF given in the form of a matrix which we shall assume
to have a hermitian symmetry. We can express its eigen-decomposition as
M−1
P= ∑λ u
l=0
l
H
l u l where λ l is lth eigenvalue (real) and u l is the corresponding
eigenvector. We let T = R = u l , that is, both source and sensor arrays are
weighed by the same eigenvector of the desired PSF matrix, and thus an image
is formed. This step is repeated over all significant eigenvectors. All such
images are linearly combined after weighting each with the corresponding
eigenvalue [20].
Let us consider an example of circular array of eight transceivers
uniformly distributed on the circumference. A transceiver consists of a source
and a sensor physically placed at the same place. The transceivers are located at
where a stands for radius of the circle. The sum coarray consists of 64 locations
whose coordinates may be found by summing the 1st column and the 1st row
entries from table 2.2. The actual coordinates, thus obtained, are also shown in
N2
the table. There are in all + 1 distinct nodes. The coarray nodes lie on a set
2
of concentric circles of radii (0, 0.76a, 1.141a, 1.85a, 2a) as shown in fig. 2.28.
Source Array
(b)
PSF
Figure 2.27: (a) Source and sensor array. (b) Point spread function (PSF) on a grid,
known as coarray.
2π 2π
t (a cos(m ), asin(m ))
M M
(−
a
, a−
a
, (0, 2a) −
a
, − 2a, −
a
, − 2a, 0 −a−
a
, (0,0)
2 2 2 2 2
a a a 2a a a
) a+ −a +
2 2 2 2 2
(0,− a) (a,-a) a (0,0) a (0.2a) a (-a,-a) a
, − , − , ,
2 2 2 2
a a a a
2
−a −a + −a − −a −
2 2 2
−(
a
, a−
a
, (0,0) −
a
, − 2a, 0 −
a
, − 2a, −a−
a
, (0,
2 2 2 2 2
a a a a − 2a a − 2a)
) − a− −a − − ,
2 2 2 2 2
( − a,0) (0,0) a
− a, (-a,a) −a−
a
, (-a,-a) −a−
a
, (-2a,0) −a+
a
,
2 2 2 2
a a a a
− , −
2 2 2 2
(
a
, a+
a
, ( 2a, 0)
a
, (0,0) a
, 0, a
− a, 2a,
2 2 2 2 2
a a a a − 2a a
− 2a
− ) − a− −a − −
2 2 2 2 2
Row entry
Table 2.2: The sum coarray consists of 64 locations whose coordinates are obtained
by summing the 1st column (left shaded column) and the respective entries from the
1st row (top shaded row). The coordinates are shown in clear cells.
PSF( p, q)
2π 2π 2π 2π
= t(a cos(m ), a sin(m ))r(a cos(m ), a sin(m ))
M M M M
m=n
t(a cos(i 2 π ), a sin(i 2 π ))r(a cos((i + M ) 2 π ),
M M 2 M
−1 a sin((i + M ) 2 π )) +
M
2
= ∑
i = 0 r(a cos(i
2π
2 M
2π M 2π
), a sin(i ))t(a cos((i + ) ),
M M 2 M
M 2π
a sin((i + ) ))
2 M
M
m−n =
2
2π 2π 2π 2π
= t(a cos(m ), a sin(m ))r(a cos(n ), a sin(n ))
M M M M
(2.72)
2π 2π 2π 2π
+r(a cos(m ), a sin(m ))t(a cos(n ), a sin(n ))
M M M M
m≠n
M
m−n ≠
2
PSF(0,0) = 8t0 r0
(2.73)
PSF( p 2 + q 2 ) = 2t0 r0
The PSF has a tall peak at the center, four times the background level.
(b)
Figure 2.28 (a) A circular array (radius=1.5cm) of eight transceivers. (b) Coarray:
Nodes are shown by empty circles. There are 33 nodes. The diameter of the outermost
circle is 3.0cm. The PSF is defined at the nodes. For uniform array, with constant
source strength and sensor sensitivity, the PSF at the center is equal to 8 t0 r 0 and
elsewhere it is equal to 2 t0 r 0 .
Or
+ ∞
I(u,v) = Ρ(u,v) ∫ ∫ ∫ ∫ r1 (x, y)r2 (x + p, y + q)dxdy e j (up+ vq) dpdq
2
p q − ∞
(2.75b)
The quantity inside the square brackets in (2.75b) represents the cross
correlation of two weighting functions.
+∞
λ
p1 p2 = 2l0 (2.77)
L − λ2
2
where l0 is the height of the moving transceiver above ground. The smallest
size of an object that can be seen on the ground with the help of an array with
aperture given by (2.77) will be approximately equal to L2 − λ 2 . The
underlying assumption is that p1 p2 >> λ which would allow us to simplify
the Rayleigh resolution criterion (see p. 222) and then the result follows. The
requirement that p1 p2 >> λ is easily met by selecting a small antenna as a
transmitter. The two way (one way in a passive system) travel time from the
scatterer to the transceiver is given by
Figure 2.29 Difference coarray obtained from L-shaped receive array shown in fig.
2.24(a).
2ωcs cs t
ω(t) = ω −
c l + (cs t)2
2
0
(a)
Beam Beam
Ground
p2 p1
x
l
2θ 0 (b)
l0 θ
Beam Beam
x'
Ground
Figure 2.30: (a) A simple illustration of the principle of synthetic aperture. The
energy scattered by a particle on the ground is received by a transceiver at p. (b) The
two way travel time from the scatterer at x´ to the transceiver at x.
2ω cs cs t 2ω cs
∆ω = − =− sin θ (2.80)
c l02 + (cs t)2 c
In the case of passive synthetic aperture sonar, the Doppler shift can be used for
estimation of the direction of arrival of an unknown source [21].
At a fixed place and time the signal received consists of a sum of the
scattered wavefields from a patch on the ground which is coherently illuminated
(see fig. 2.31). The size of the patch is equal to the synthesized aperture. The
wavefield from different scatterers reaches the sensor with delays as given by
(2.78). The receiver output is given by a convolutional integral [22],
(2.81)
L
where w(x − x ′ ) =1 for x − x ′ ≤ and L=p1 p 2 , the length of the synthetic
2
aperture. For x − x' << l0 we can approximate (2.81) as
p2
(x − x' )2
f (x,t) = ∫ r0 (x' )w(x − x ′ )cos(ω(t − τ 0 − ) + ϕ)dx ′
p1
cl0
(2.82a)
which simplifies to
∞
(x − x' )2 ϕ
f (x,t) = cos(ω(t − τ 0 )) ∫ r0 (x' )w(x − x ′ )cos(ω( ) + )dx ′
−∞
cl0 2
∞
(x − x' )2 ϕ
+ sin(ω(t − τ 0 )) ∫ r0 (x' )w(x − x ′ )sin(ω( ) + )dx ′
−∞
cl0 2
(2.82b)
We can recover r 0 (x) from f (x, t) , using the Fourier transform method, that
is,
−1
r̃0 (x) = FT
FT f̂ (x) { }
(2.83)
FT w(x)cos(ω( x ))
2
cl0
where
f̂ (x) = ∫ f (x,t)cos(ω(t − τ
one period
0 ))dt
Figure 2.31: At a fixed place and time the signal received consists of a sum of
scattered wavefields from a patch of scatterers which is coherently illuminated. The
width of the patch is equal to the synthesized aperture.
result of geometry of the data collection as shown in fig. 2.30. Indeed, the
transceiver need not move at all during pulse transmission and reception.
§2.5 Exercises:
1. The angle of arrival of a broadband signal at ULA may be estimated from the
spectrum of the sum of the output of all sensors (without delays) and the
spectrum of the output of anyone sensor. Find the bandwidth required to
estimate the angle of arrival equal to 45o given that c=1500 m/s and d=15m.
What is the role of the sensor spacing?
2. It is desired to measure the speed of wave propagation in a medium. A ULA
with sensor spacing d meters is employed for this purpose. A broadband signal
from a known direction is sweeping across the array. How do you go about
estimating the speed of propagation?
3. Consider a circular transmit and receive array as shown in fig. 2.32. (a) Show
the sum coarray along with weight coefficients assuming that the physical array
has unit coefficient. (b) Do the same for the difference coarray.
4. Compare the outputs of crossed electric dipole arrays given by (2.58 & 2.59)
with that of displaced identical sensors pairs (also known as dipoles) given by
equations (2.55 a & b). Obtain an expression for the spectral matrix, similar to
the one given in (2.57), of the output of the crossed dipole arrays.
Figure 2.32: A circular sensor and source array. Compute sum and difference coarray
a b
ω < ω max
H(s,ω) = 1 ω
s < c (s = u 2 + v 2 ) (3.1)
Hor
=0 otherwise
which is illustrated in fig. 3.1. The filter will pass all propagating waves whose
c
angle of elevation lies between 0o and ± sin −1 ( ) (the elevation angle is
cHor
measured with respect to the vertical axis). Fail and Grau [3] were the first to
introduce the fan filter in 1963. Independently, Embree [4] came up with a
similar filter which he called a “Pie Slice Filter”. Since the filter has a circular
symmetry in a spatial frequency plane we need a series of concentric uniform
circular arrays (in effect, a planar array) to measure the direction of arrival in the
horizontal plane and horizontal apparent speed, when the wave speed is known.
For the purpose of filter design it is enough if we consider any one radial line.
Slope= c hor
0 u
Figure 3.1: Fan filter. The passband is a cone with an axis along the ω-axis and the
apex is at the center. All waves whose horizontal apparent speed lies inside the cone
are allowed and others are attenuated.
The filter may be rotated about the vertical axis to yield a circularly symmetric
filter. For simplicity we consider a single radial line of sensors coinciding with
the x-axis and assume that all waves are propagating along the x-axis with
different horizontal apparent speeds.
In order to compute the fan filter coefficients we take the inverse
Fourier transform of the transfer function given in (3.1),
+∞
1
∫ ∫ H(ω,u)e
j (ωn∆t +um∆x )
hm,n = 2 dωdu
4π −∞
hm,n =
+ω max
1 m∆x m∆x
2 ∫ sin( c − n∆t)ω + sin( + n∆t)ω dω (3.3)
2π m∆x 0 Hor cHor
We are free to choose the spatial and temporal sampling intervals but within the
constraints imposed by the sampling theorem. These are discussed in §3.2. Let
∆x
the temporal sampling interval ∆t be so chosen that ∆t = . In practice
cHor
∆x is held fixed; therefore, to alter cHor it is necessary to resample the whole
signal with a different sampling interval, consistent with the above choice.
Equation (3.3) is easily evaluated to yield
Equation (3.4) takes particularly a simple form if we shift the origin to the half-
way point between two sensors
1 1
hm,n = (3.5)
π ∆x∆t m − n 2
2 2
1 1 1
where m = ± , ±1 , ±2 ,... (see fig. 3.2). Note that since m ≠ n , hm,n
2 2 2
∆x
1 1 1 1
-1 - m= 1
2 2 2 2
Figure 3.2: Centered array (ULA) with origin lying between two sensors. This is also
the physical center of the array.
will always be finite. The frequency response (magnitude) of a fan filter given
by (3.5) is shown in fig. 3.3
3.1.3 Fast Algorithm: In order to implement the filter the following
convolution sum will have to be evaluated
The computational load for evaluating the convolutional sum can be greatly
reduced by exploiting the filter structure [5]. We define a new index
µ = (1 + 2 m ′ ) where m ′ = 0, ±1, ±2,... . We note that the new index takes
values, µ = ±1, ± 3, ± 5, ... , for different values of m ′ . Equation (3.5) may
be reduced to
1 1
hm,n = 2 2
π µ
− n2
4
0.7
2
-1
-2
-3
-3 -2 -1 0 1 2 3
u
Figure 3.3: Frequency response (magnitude) of a fan filter (64 sensors and 64 time
samples).
1 1 1
hm,n = 2 +
π µ µ − n µ + n
2 2
1
[
= rµ ,n + qµ ,n
µ
]
The components rµ ,n and qµ ,n possess many symmetry properties as listed
below:
µ
• rµ ,n and qµ ,n are antisymmetric in time index (n) about a point n = and
2
µ
n=− , respectively.
2
1
hm,n = r µ −1 − r µ +1 (3.7)
π µ 1,n − 2
2
1,n +
2
1
f̂ (n∆t) = ∑ ∑ 2
r µ −1 − r µ +1 f (µ∆x,(n − n′ )∆t)
n′ µ π µ 1, n ′ − 2 1, n ′ +
2
where l=M-1 and M (an even number) stands for the number of sensors. In (3.8)
there is only one convolution to be carried out in place of M/2 convolutions in
(3.6). The inner summation in (3.8) stands for spatial filtering and the outer
summation stands for temporal filtering. Let the output of the spatial filter be
f 1 (n∆t) , where
f (µ∆x, (n − n′ − µ − 1 )∆t)
1
l
2
f 1 (n∆t) = ∑
µ = −l µ µ −1
− f (µ∆x, (n − n′ + )∆t)
2
N1
1
R1 (z) = ∑ 1
zn
n = − N1 +1 ( − n)
2
N1 −1 N1 −1
1 1
= −z ∑ zn + ∑ z−n (3.10)
1 1
n=0 ( + n) n=0 ( + n)
2 2
= −zRN1 (z) + RN1 (z −1 )
where
N1 −1
1
RN1 (z) = ∑ 1
zn
n=0 ( + n)
2
and the number of time samples are assumed to be 2N1 . Note that RN1 (z)
−1
acts on future time and RN1 (z ) acts on past time. We assume that the data are
pre-recorded; hence the question of causality is irrelevant. It is shown in [5] that
as N1 → ∞ , RN1 (z) may be approximated by a stable pole-zero filter, in
particular,
2(1 − 0.65465z)
RN1 (z) →
1 − 0.98612z + 0.13091z 2 (3.11)
N1 → ∞
5
Magnitude
1
0 0.5 1 1.5 2 2.5 3
radians
Figure 3.4: A comparison of the filter response N1 =64 (dashed line) and N1 →∞
(solid line).
z +
Temporal
Processor
+
1/3 1 -1 -1/3
Spatial
Processor
-
++ -
++ -
++ -
++
-2 +1 -1 0 0 -1 +1 -2
z z z z z z z z
-3 -1 µ= 1 3
Figure 3.5: A block diagram of fan filter implementation. The filter structure has two
parts, namely, spatial and temporal parts, which are independent of each other.
ω =au
- ωmax
The passband is bounded with two radial lines with slopes a and b respectively.
Further the passband is terminated by two horizontal lines ω = ±ω max where
ω max is cut-off temporal frequency. When b > 1 and a < 1 the passband will
be terminated by two cut-off lines as shown in fig. 3.6. Draw a diagonal line
and divide the passband into two triangles. Straight forward integration yields
m ∆x
the filter impulse response function for m ≠ 0, ( + n) ≠ 0 and
a ∆t
m ∆x
( + n) ≠ 0 [6].
b ∆t
ω max ωa
1
hm,n =
4π 2 ∫ ∫ 2 cos(m∆xu + n∆tω)du dω
0
ωb
(b − a)ω 2max
h0,0 = m=n=0
4π 2 ab
m ∆x
(1 − cos( + n)ω max ∆t)
−1 b ∆t m ∆x
hm,n = ( + n) = 0, n ≠ 0,
2π 2 m ∆x a ∆t
n∆x∆t( + n)
b ∆t
m ∆x
( + n) ≠ 0
b ∆t
m ∆x
(1 − cos( + n)ω max ∆t)
1 a ∆t m ∆x
hm,n = ( + n) = 0, n ≠ 0,
2π 2 m ∆x b ∆t
n∆x∆t( + n)
a ∆t
m ∆x
( + n) ≠ 0
a ∆t
(3.12b)
∆x
We now set ∆x =1 and ∆t =1, hence =1. Using these settings (3.12a)
∆t
1
reduces to equation (6) in [6]. The highest temporal frequency is f hi = Hz (or
2
ω hi = π in angular frequency) and the corresponding lowest wavelength,
2 0.9
0.1
1
Temporal Frequency
0.5
0
0.5
-1
0.1
-2
0.9
-3
-3 -2 -1 0 1 2 3
Spatial frequency
Figure 3.7: Frequency response of a quadrant filter (a=1, b=2). 16 sensors and 64
time samples are assumed. The maximum temporal frequency is 0.8π. The contour
values are as shown in the figure.
λ lowest = 2 . The sensor spacing will be one. In real terms, consider a signal
with f hi = 1500Hz propagating underwater where the wave speed is 1500
m/s. Since the wavelength is 1 meter, the sensor spacing is ∆x = 0.5 meter.
1
The sampling interval is ∆t = ms. Let us redefine units of time and distance.
3
1
Let one unit of time be ms and one unit of distance be 0.5 meter. Here
3
onwards we shall measure the time and the distance in terms of these newly
defined units. The response function of a quadrant filter for parameters listed
below is shown in fig. 3.7 Parameters: 64 sensors, 64 time samples,
ω max =0.8π, a=1 and b=4. In (3.12a) if we let a=-b=1, cHor = 1 , ∆t = 1, and
ω max =π the resulting filter becomes a fan filter with its axis turned 90o and the
filter coefficients are
∑∑h
m′ n′
r
mn m − m ′ ,n − n ′ = gmn (3.14a)
where
π π
∆x ∆t
1
∫ ∫ W(ω ,u)e
+ j ( ωn∆t −um∆x )
rm,n = dω du
4π 2 π π
− −
∆x ∆t
and
π π
∆x ∆t
1
∫ ∫ W(ω ,u)H(ω ,u)e
+ j ( ωn∆t −um∆x )
gm,n = dω du
4π 2 π π
− −
∆x ∆t
Note that, when W(ω,u)=1 over the entire plane, (3.14a) reduces to the expected
result, namely,
π π
∆x ∆t
1
∫ ∫ H(ω ,u)e
+ j ( ωn∆t −um∆x )
hm,n = dω du
4π 2 π π
− −
∆x ∆t
In order to understand the role of the weighting function, consider (3.14a) with
finite limits on the summation signs
M −1 N −1
∑ ∑h
m′ =0 n′ =0
r
mn m − m ′ ,n − n ′ = gmn (3.14b)
M −1 N −1
where [ F ] finite = ∑ ∑ f mne − j ( ωm +un) stands for the finite discrete Fourier
m=0 n=0
2π 2π
transform (DFT) for ω= k and u = l . Note that [ F ] finite → F as
M N
M → ∞ and N → ∞ . From (3.15) we obtain
[WH ] finite
[ H ] finite = (3.16)
[W ] finite
(b)
Figure 3.8: The role of a weighting function is shown here. (a) unweighted case. (b)
weighted case (equal to six inside a fan shaped region bounded by two radial lines).
Notice how a side lobe falling within the region of high weight function has been
attenuated.
ω max ∆x
cLo ≥ = (3.17)
B0 ∆t
N
H(ω ′ ) = h0 + 2 ∑ hk cos(ω ′k) (3.18)
k =1
which may be expressed as a polynomial in cos(ω ′ ) ,
c Hi
c' cLo
- B0 B0 2B
0 u
- ω maxa b
Figure 3.9: The aliasing is caused by intrusion of a radial line with slope cLo from
the neighboring rectangles into the principal rectangle. Triangles abc and a´b ´c´ are
the aliased pass regions. Aliasing can be prevented if we were to sample an analog
quadrant filter according to (3.17).
N
H(ω ′ ) = ∑ bk cos k (ω ′ ) (3.19)
k =0
where the coefficients bk , k=0, 1, ... N are expressed in terms of the FIR filter
coefficients. A point on the frequency axis is mapped onto a closed contour in
the frequency wavenumber space using a linear transformation
P Q
cos(ω ′ ) = ∑ ∑ t pq cos( pω)cos(qu) (3.20a)
p= 0 q = 0
-1
-2
-3
-3 -2 -1 0 1 2 3
Figure 3.10: Mapping produced by (3.21). The inner contours corresponding to low
frequency are close to circular shape but the outer contours corresponding to high
frequency are only approximately circular.
or
ω′ P Q
ω u
sin 2 ( ) = ∑ ∑ t pq ′ sin 2 p ( )sin 2q ( ) (3.20b)
2 p= 0 q = 0 2 2
1
t00 = −
2
1
t01 = t10 = t11 =
2
1
cos(ω ′ ) = {−1 + cos(u) + cos(ω) + cos(u)cos(ω)} (3.21)
2
k
P Q
N
H(ω,u) = ∑ bk ∑ ∑ t pq cos( pω)cos(qu) (3.22)
k =0 p= 0 q = 0
P Q
a) When ω ′ =0, ω = u = 0 . This requires ∑∑t
p= 0 q = 0
pq = 1. The
condition ensures that a lowpass filter remains a lowpass filter even after the
transformation, that is, H(0,0)=H(0).
P Q
b) When ω ′ =π, ω = u = π . This requires ∑∑t
p= 0 q = 0
pq (−1) p+ q = −1.
The condition ensures that a highpass filter remains a highpass filter after
transformation, that is, H(π,π)=H(π).
P
c) When ω =0, H(0,u) = H(u) . This requires ∑t
p= 0
p1 = 1 and
P
∑t
p= 0
pq = 0 for q ≠ 1 .
Q
d) When u =0, H(ω,0) = H(ω) . This requires ∑t
q=0
1q = 1 and
Q
∑t
q=0
pq = 0 for p ≠ 1 .
For circularly symmetric mapping, conditions (b) and (c) must hold good. For
P=Q=1, a general solution for the transformation coefficients satisfying the
1 1
above conditions is given by t00 = − , t01 = t10 = t11 = . By relaxing (b),
2 2
the solution is given by
having one free constant which may be optimally chosen for the best fit.
The next important question is to determine the coefficients bk ,
k=0,1,2,..., N in (3.19) given the coefficients of a digital filter. Let
h0 ,h1 ,h2 ,...hN be the given 1D filter coefficients.
N
H(ω ′ ) = ∑ h̃k cos(kω ′ )
k =0
(3.24)
N
= ∑ h̃k T k (cos(ω ′ ))
k =0
k
T k (cos(ω ′ )) = ∑c
m=0
k
m cos m (ω ′ ) (3.25)
Using (3.25) in (3.24) and comparing with (3.19) we obtain the following
system of equations:
N N N
b0 = ∑ c0n h̃n , b1 = ∑ c1n h̃n , ... bk = ∑ ckn h̃n , ... bN = cNN h̃N
n=0 n =1 n=k
n
The coefficients ck are listed in [10, p. 795]. Finally, the 2D filter transfer
function for the case P=Q=1 can be written as
G(ω,u) = ∑ bk F(ω,u)
k =0 +t11 cos(u)cos(ω) (3.26b)
N
= ∑ bk [ H0 (ω,u)] F(ω,u)
k
k =0
whereH0 (ω ,u) = t00 + t01 cos(u) + t10 cos(ω ) + t11 cos(ω )cos(u) . Note
that F0 (ω,u), F1 (ω,u), ..., FN (ω,u) do not depend upon the filter
coefficients but only on mapping coefficients and the input. Using (3.26c), the
filter output may be written as
N
G(ω,u) = ∑ bk Fk (ω,u) (3.27)
k =0
1
t00 = t11 = 0 and t01 = −t10 = (3.28a)
2
and the mapping function is given by [11]
1
cos(ω ′ ) = [cos(ω) − cos(u)] (3.28b)
2
The contours generated by (3.28b) are shown in fig. 3.12. The zero valued
π
diagonal contour corresponds to . When ω ′ = 0 , ω = 0 and
ω′ = ±
2
u = ±π ; and when ω ′ = ±π , ω = ±π and u = 0 . The positive contours
π π
correspond to 0 ≤ ω′ ≤ and the negative contours to < ω′ ≤ π .
2 2
π
Ideally, a lowpass filter having a unit response in the range ± and zero outside
2
150 .0 -.8
- -
100
50
+
0
0 50 100 150 200 250
Figure 3.12: Contours generated by a mapping function given by (3.28a). The two
diagonals make up the zero contour which divides the display area into four
π
triangles. The positive contours correspond to 0 ≤ ω′ ≤ and the negative
2
π
contours to < ω′ ≤ π .
2
when mapped into two dimensions using the mapping function (3.28b) will
result in a fan filter with a unit response in the top and bottom triangles and a
zero response in the left and right triangles. The value of the 2D filter transfer
function may be found by first computing the frequency corresponding to the
−1
index of each contour, that is, ω ′ = cos (contour index) and then evaluating
the 1D filter response at the desired frequency.
It is interesting to note that the mapping coefficients given by (3.28a)
do not satisfy any of the conditions listed on page 167. This is expected as the
1 1 1 1
cos(ω ′ ) = cos(ω) − cos(u) − cos(v) + (3.29)
2 2 2 2
§3.3 Multichannel Wiener Filters:
In Wiener filtering the goal is to make the filter output as close as possible, in
the least square sense, to a desired signal. The sensor output and the desired
signal are assumed to be stationary stochastic signals which are characterized
through their covariance functions. The Wiener filters are known after Norbert
Wiener who did pioneering work on the prediction of a trajectory of a moving
object from its past observations [13]. A recursive algorithm for the solution of
a discrete version of the Wiener-Hopf equation was developed by Levinson [14]
and by Durbin [15] in the context of time series model fitting. Multichannel
extension was made by Wiggins and Robinson [16]. In this section we briefly
describe the Wiener filter as applied to array signals where we like to extract a
signal traveling in some known direction and to optimally suppress all other
propagating waves and noise. Here, a straightforward solution of the Wiener-
Hopf equation requires invertion of a large block toeplitz covariance matrix,
leading to a dramatic increase in the computational load over the single time
series version. Therefore, it is worthwhile to spend some effort to understand
the principles of the Levinson-Durbin recursive algorithm for the multichannel
Wiener filter [16].
3.3.1 Planar Array: We consider a planar array of sensors, not necessarily
uniformly distributed (see fig. 3.13). Let f p (t) be the output of the pth sensor
located at (x p , y p ) . We shall model the output as a sum of two random
processes, namely, a desired signal, ξ p (t) , and unwanted noise, η p (t) ,
P −1 ∞
f̂ (t) = ∑ ∫ hp (τ) f p (t − τ)dτ (3.31)
p= 0 0
x 1y 1
x 2y 2
x 0y 0
Figure 3.13: A distribution of sensors in a planar array. All sensors are identical but
for their position.
{
E f̂ (t) − ξ 0 (t)
2
} = min
that is, the mean square error (mse) is minimum. This requires minimization of
an expression for mse,
P −1 ∞
mse = cξ 0 (0) − 2 ∑ ∫ hp (τ)cξ 0 f p (τ)dτ
p= 0 0
(3.32)
P −1 P −1 ∞ ∞
+∑ ∑ ∫ ∫ h (τ)h (τ′)[c
p= 0 q = 0 0 0
p q ξ pξq ]
(τ − τ ′ ) + cη p η q (τ − τ ′ ) dτdτ ′
Now, differentiate the sum on the right hand side in (3.33) with respect to
hp (n∆τ) for fixed n. The result is ∆τcξ ξ (t − n∆τ) . A similar approach is
0 p
P −1 ∞
∑ ∫ h (τ′)c
q=0 0
q f p fq (τ − τ ′ )dτ ′ = cξ 0 f p (τ), p = 0,1,... P − 1 (3.34a)
The minimum mean square error (mmse) may be derived by using (3.34a) in
(3.32). We obtain
P −1 ∞
mse min = cξ 0 (0) − ∑ ∫ hp (τ)cξ 0 f p (τ)dτ (3.34b)
p= 0 0
[ ]
T
h p = hp (0), hp (∆τ), hp (2∆τ), ... hp ((N − 1)∆τ)
and
P −1
∑c
q=0
f p fq hq = cξ 0 f p , p = 0,1,2,... P − 1 (3.35a)
P −1
mse min = cξ 0 (0) − ∑ c Tξ 0 f p h p (3.35b)
p= 0
where
where
[ ]
T
c 0 (n∆τ) = cξ 0 f 0 (n∆τ), cξ 0 f 1 (n∆τ), cξ 0 f 2 (n∆τ), ... cξ 0 f P−1 (n∆τ)
As an example consider a three sensor array with two time samples (N=2). The
above quantities, C , h , and C0 , become
and
[ ]
T
C0 = cξ 0 f 0 (0),cξ 0 f 1 (0),cξ 0 f 2 (0),cξ 0 f 0 (1),cξ 0 f 1 (1),cξ 0 f 2 (1)
C h = C0 (3.36a)
∑ H (ω)S
q =1
q f p fq (ω) = Sξ 0 f p (ω), p = 0,1,... P − 1 (3.37a)
∞
Hq (ω) = ∫ hq (τ)e jωτ dτ
0
∞
P −1
mse min = ∫− ∞ ξ 0
S (ω) − ∑ H p (ω)Sξ 0 f p (−ω)dω (3.37b)
p= 0
To write (3.37) in a matrix form define the following vectors and matrix:
and
for all ω in the range ±∞ . Formally, the solution of (3.38) may be expressed
as
H(ω) = S−1
f (ω)S 0 (ω) (3.38b)
Now consider a plane wave sweeping across an array of sensors. Let the
background noise be spatially white. The spectral matrix for this model is given
by
where
1 + T(ω) T(ω)e j (u0 x1,0 + v0 y1,0 ) ... T(ω)e j (u0 x P−1, 0 + v0 y P−1, 0 )
− j (u x + v y ) j (u x +v y )
T(ω)e 0 1,0 0 1,0 1 + T(ω) ... T(ω)e 0 P−1, 1 0 P−1, 1
...
A(ω) =
...
...
T(ω)e − j (u0 x P−1, 0 + v0 y P−1, 0 ) T(ω)e − j (u0 x P−1, 1 + v0 y P−1, 1 ) ... 1 + T(ω)
S0 (ω)
T(ω) = , x p q = x p − xq , y p q = y p − yq , S0 (ω) is the signal
Sη (ω)
spectrum, and Sη (ω ) is the noise spectrum common to all sensors. Similarly,
the vector on the right hand side of (3.38) may be expressed as
[
S 0 (ω) = 1, e
− j (u 0 x1,0 + v 0 y1,0 )
,e
− j (u 0 x 2,0 + v 0 y 2,0 )
, ..., e
− j (u 0 x P−1, 0 + v 0 y P−1, 0 )
]S (ω)
0
= B(ω)S0 (ω)
(3.40)
S0 (ω)
A(ω)H(ω) = B(ω)
Sη (ω) (3.41a)
= B(ω)T(ω)
[ ]
T
a(ω) = e j (u0 x 0 + v0 y0 ) , e j (u0 x1 + v0 y1 ) , e j (u0 x 2 + v0 y2 ) , ... e j (u0 x P−1 + v0 y P−1 )
(3.42b)
P −1 ∞
f̂ (t) = ∑ ∫ hp (τ)ξ p (t − τ)dτ (3.43)
p= 0 0
Assume a signal model where a plane wave is sweeping across the array
maintaining its waveform unchanged. The outputs of any two sensors differ
only in propagation delays. The signal at the pth sensor may be given by
ξ p (t) = ξ 0 (t − τ p ) where τ p is the propagation delay at the p t h
sensor ( τ0 = 0 when delays are measured with respect to the 0th sensor). Since
the geometry of the array, speed of propagation and direction of arrival are
known or can be estimated independently the propagation delays are presumed to
be known. The output of each sensor is advanced to make it in phase with the
output of the reference sensor. Equation (3.43) reduces to
P −1 ∞
f̂ (t) = ∑ ∫ hp (τ)ξ 0 (t − τ)dτ
p= 0 0
(3.44)
∞
P −1
= ∫ ∑ hp (τ) ξ 0 (t − τ)dτ
0 p= 0
The filters are to be chosen to satisfy the condition that f̂ (t) = ξ 0 (t) . From
(3.44) it is clear that this constraint can be satisfied if
P −1
∑ h (t) = δ(t)
p= 0
p (3.45a)
P −1
∑ H (ω) = 1
p= 0
p (3.45b)
Thus for distortion free extraction of waveforms the filters must satisfy (3.45)
[17]. While an individual filter transfer function is allowed to be of any form,
the sum must be a constant.
Another type of constraint arises when it is required to minimize the
background noise power. The noise in the output of an array processor is given
by
P −1 ∞
η̂(t) = ∑ ∫ hp (τ)η p (t − τ)dτ
p= 0 0
∞∞
2
η̂
p q
{ }
σ = ∑ ∑ ∫ ∫ hp (τ)hq ( τ ′ )E η p (t − τ)ηq (t − τ ′ ) dτdτ ′
0 0
∞∞
= ∑ ∑ ∫ ∫ hp (τ)hq ( τ ′ )c pq ( τ ′ − τ )dτdτ ′
p q 0 0
(3.46)
∞
1
= ∑∑ ∫ H p (ω)Hq* (ω)Sη pq (ω)dω
p q 2π 0
P −1 ∞
1
σ =∑
2
∫ dω σ 2η
2
η̂
H p (ω)
p= 0 2π 0
The noise power in the array output shall be minimum whenever the filter
transfer functions satisfy the condition
P −1 ∞
1
∑
2
∫
p= 0 2π 0
H p (ω) dω = min (3.47)
P −1 ∞
1
∑
2
∫
p= 0 2π 0
H p (ω) dω = NRF < 1 (3.48)
AULA (ω) can be expressed as (I + T(ω)a(ω)a H (ω)) (see (3.42)) where the
vector a(ω) for a ULA is given by
Remember that in a ULA the right most sensor is conventionally taken as the
reference sensor. Using the Woodbury’s identity [18] we obtain its inverse in a
closed form,
T(ω)
H0 (ω) =
1 + PT(ω)
T(ω)
H1 (ω) = e − ju0 d
1 + PT(ω)
T(ω)
H2 (ω) = e − j 2u0 d
1 + PT(ω)
.
(3.50)
.
.
T(ω)
HP −1 (ω) = e − ju0 ( P −1)d
1 + PT(ω)
The frequency wavenumber response of the Wiener filter for T=4 and sixteen
sensor array (ULA) is shown in fig. 3.14. Notice that whenever PT(ω) >>1,
S0 (ω)
that is, either T(ω) = >>1 or P>>1 or both or the spectra of the
Sη (ω)
signal and noise are nonoverlapping, the Wiener filter reduces to a simple delay
filter
1 − jk u0 d
Hk (ω) = e (3.51)
P
− ju R 2π 2π
− j[u 0 R cos + v 0 R sin ]
e 0 e P P ...
a(ω) = col (3.52b)
2π 2π
− j[u0 R cos( P −1) P + v 0 R sin( P −1) P ]
e
and
BUCA (ω) = a(ω) (3.53)
The reference point is at the center of the circle; however, there is no sensor
physically present there. The Wiener filter is designed to predict the waveform
as seen by an hypothetical sensor kept at the center of the circle. To solve for
the Wiener filter (3.41b) we need to invert the A(ω) matrix. We shall once
again use Woodbury’s identity and obtain
−1
AUCA (ω) = I −
T(ω)
1 + PT(ω)
[
a(ω)a H (ω) ] (3.54)
T(ω)
H(ω) = a(ω) (3.55)
1 + PT(ω)
The Wiener filter for a circular array is similar to that for a linear array
except for the difference arising out of the definition of a(ω) . The frequency
wavenumber response of 16 sensor circular array of radius 10λ is shown in fig.
3.15. Although the mainlobe width is narrow, the sidelobe level is quite high.
This is clearly brought out in a cross-sectional plot passing through the
maximum (see fig. 3.16). When we increase the number of sensors to 64 the
sidelobe level is brought down considerably, but the main lobe width remains
practically unchanged. It may be emphasized that the sensors need not be spaced
at ≤ 0.5λ as in a ULA [19]. What is gained by increasing the number of
sensors (keeping the radius fixed) is the reduction of the sidelobe level. In
contrast, in case of a ULA, by increasing the number of sensors the array
aperture is increased which in turn sharpens the mainlobe but does not reduce
the sidelobe level.
0.0
2
0.0
1 0.9
Temporal frequency
0.0
0.0
.
0 0.5
0.5
0.0
-1
0.9 0.0
0.0
-2
0.9
-3
-3 -2 -1 0 1 2
Spatial frequency
Figure 3.14: Frequency wavenumber response of the Wiener filter for ULA. The
direction of arrival (DOA) is assumed to be known. In the above example it is equal
to 30o .
3.4.3 Robustification: The Wiener filters given by (3.51) and (3.55) require a
knowledge of a(ω) for which we need to know the apparent speed of the
wavefront sweeping across the array. Prior to waveform estimation it is a
common practice to estimate the direction of arrival (DOA) of a wavefront.
(This topic is covered in some detail in chapter 5). The DOA estimation is not
without error. Hence, it would be nice if the Wiener filters are made robust so
that the degradation in its performance is minimum. We shall confine to a
linear array (ULA). An error in DOA estimation will introduce an error in the
wavenumber. For a ULA, the erroneous wavenumber may be expressed as
(u0 + ε) where ε is auniformly distributed random variable. Naturally, there
AULA (ω) and BULA (ω) . We shall use the
will be an error in each element of
stochastically averaged AULA (ω) and BULA (ω) matrices in (3.41b). This
approach was suggested in [20] in the context of optimal velocity filters in
1
v
-1
-2
-3
-2 0 2
u
Figure 3.15: Frequency wavenumber response of Wiener filter for UCA. The DOAs
are known (azimuth=elevation=45o and ω=π /2). 16 sensors and constant snr=4 are
assumed.
AULA (ω ) =
(3.56a)
ju 0 ( P −1)d
1+T (ω ) ju 0 d
T (ω ) sin c(ε 0 d )e ... T (ω ) sin c(ε 0 ( P −1)d )e
T (ω ) sin c(ε 0 d )e − ju0 d 1+T (ω ) ... T (ω ) sin c(ε 0 ( P −2)d )e 0
ju ( p−2)d
...
...
...
− ju 0 ( P −1)d − ju 0 ( P −2)d
T (ω ) sin c(ε 0 ( P −1)d )e T (ω ) sin c(ε 0 ( P −2)d )e ... 1+T (ω )
and
1, sin c(ε 0 d)e − ju0 d , sin c(2ε 0 d)e − ju0 2 d , ...,
BULA = col (3.56b)
sin c((P − 1)ε 0 d)e − ju0 ( P −1)d
0.8
0.6
Response
16
0.4
0.2
0 64
-4 -2 0 2 4
Spatial frequency
Figure 3.16: A cross-section of the Wiener filter response taken through the peak.
Number of sensors are 16 and 64. The sidelobe level relative to the peak has been
reduced when the number of sensors is increased from 16 to 64. Radius of circular
aperture is 10 units.
sensors and constant snr (=4). Notice the splitting of the peak particularly in the
higher temporal frequency range. Next, we compute the frequency wavenumber
response of a UCA designed to tolerate an error of 2o . The frequency
wavenumber response is shown in fig. 3.17b. The main lobe shape remains
practically unchanged but there is an increase in the sidelobe level. Cross-
sectional plots passing through the maximum of the response function of the
UCA with and without DOA error are shown in fig. 3.18. While the shape of
the main lobe remains practically unchanged the side lobe level seems to have
slightly increased. This is the price one has to pay for the lack of exact
knowledge of the DOA.
3.4.4 Levinson-Durbin Algorithm: In (3.36), in order to solve for h, we have
to invert a large covariance matrix (e.g., with the array size, M=24, and filter
length, N=16, the size of the covariance matrix will be 384x384). The
0.2
Temporal frequency
2
1
0.4
0.8
1
0
-2 -1 0 1 2
Spatial frequency
(a)
ω 0
-1
-2
-3
-2 0 2
u
(b)
Figure 3.17: Frequency wavenumber response of Wiener filter when the error in DOA
estimate is ±2.0 o . (a) ULA and (b) UCA. 16 sensors and 64 time samples are
assumed.
1.5
Response
0.5
0
-4 -2 0 2 4
Spatial frequency
Figure 3.18: A cross-section through the maximum (v0=0.7854) for UCA. Solid line
shows response when there is no error and the dashed line shows one with DOA error
( ±2.0 ). 16 sensors and 64 time samples are assumed.
o
computational load for inverting such a large matrix will be very high. We
shall outline a recursive method applicable to a ULA. For any other geometry
of an array the covariance matrix becomes a block symmetric matrix, but only
for a ULA does the covariance matrix become toeplitz. This important property
enables us to devise a recursive algorithm, known as Levinson-Durbin
algorithm, which requires inversion of a matrix of size M × M in place of a
matrix of size MN × MN. Briefly, the algorithm is as follows [16]: Let h N
[ ]
( h N = col h(0), h(∆τ), ... h((N − 1)∆τ) ) be the solution of the Nt h
order normal equations, (3.36). Let us now increase the size of the covariance
matrix by padding one row and one column of covariance matrices as below,
C N c(N)
h N C0
=
.
. (3.57)
.
0 γ N
c(N) ... c(0)
Note that the square matrix on the left hand side of (3.57) is a block covariance
matrix of size M(N + 1) × M(N + 1) ; therefore, (3.57) is similar to (3.36)
but of order N+1. Indeed, if we subtract (3.57) from (3.36) of the same order,
we shall obtain
N
h 0
C N +1 h N +1 − =
(3.58)
0 c (N) − γ
0 N
b N = col{b NN ,b NN −1 ,...b N1 , I}
( N +1) M × M
b NN h N +1 (N)
b 0
NN −1h N +1 (N)
C N +1 =
(3.60a)
M β N h N +1 (N)
h N +1 (N)
0
C N +1b N = (3.60b)
β
N
We shall once again increase the order of C N +1 in (3.60b) by padding one more
row and one more column. We get
where
α N
C N +1a N = (3.61a)
0
where
[
α N = c(0)I + c(1)a N 1 + ... + c(N)a NN ]
M × M
Let us now increase the size of C N +1 in (3.61a) by padding one more row and
one more column of covariance matrices. We obtain
α N
a N
C N + 2 = 0 (3.61b)
0
α ′N
where
[
α ′N = c(N + 1)I + c(N)a N 1 + ... + c(1)a NN ]
M × M
We linearly combine (3.60c) and (3.61b) such that the resulting equation is the
(N+2)th order equivalent of (3.60b). Let the linear combination be given by
β′N + α N δ N = 0
or
δ N = −α −1
N β ′N (3.62b)
The resulting equation is equivalent to (3.60) but of order N+2. Then, we have
0 a N
b N +1 = + δ N (3.63a)
b N 0
β N +1 = β N + α ′N δ N (3.63b)
Further we take a linear combination of (3.60c) and (3.61b) such that the
resulting equation resembles (3.61a) but it is of order N+2. This may be
achieved by requiring
∆ N β N + α ′N = 0
or
∆ N = −β −1
N α ′N (3.64a)
a N 0
a N +1 = + ∆ N (3.64b)
0 b N
α N +1 = α N + β′N ∆ N (3.64c)
Eqs. (3.63a) and (3.64b) form a set of recursive relations to compute a N +1 and
b N +1 given a N and b N . Similarly, equations (3.63b) and (3.64c) form another
h N +1 (N) = [β N ] (c 0 (N) − γ N )
−1
∞ 2
E f 1 (t) − ∫ f 2 (t ′ − t)hpred (t ′ )dt ′ = min
0
∞ S (ω) + S (ω) H 2
1 pred (ω) −
∫ 1 dω
2
2π S (ω)H ∗ (ω) − S ∗ (ω)H
− ∞ 12 pred 12 pred (ω)
We obtain
and
h1
M2
⊃ + h
h2 f2(t)
Figure 3.19: Source is in near field and noise is in far field. Microphone M2
(reference microphone) receives very little of the signal.
[ ]
∞
1 2
Error min = ∫
2π − ∞
S1 (ω) − S2 (ω) H pred (ω) dω (3.66b)
∞
1
2π −∫∞
Error min = S0 (ω)dω
1
H pred (ω) =
H1 (ω)
∞
1
∫ [S (ω) − S (ω) H
2
Output power = 1 2 pred (ω) ]dω
2π
−∞
∫ {S (ω)(1 − H (ω) }
1 2 2
= 0 1 H pred (ω) ) dω Signal power
2π
−∞
∞
∫ {S }
1 2 2
+ η1 (ω)(1 − H2 (ω) H pred (ω) ) dω Noise power
2π
−∞
=
{S (ω)(1 − H (ω) H
0 1
2
pred (ω) )
2
}
{S (ω)(1 − H (ω) H (ω) )}
SNRoutput
2 2
η1 2 pred
2 2
[ SNR ]output 1 − H1 (ω) H pred (ω)
G= [ SNR ]input
= 2 2 (3.67)
1 − H2 (ω) H pred (ω)
2 2
For G>1 we must have H1 (ω) < H2 (ω) .
3.5.2 Source in Far Field: Both signal and noise sources are in the far field
region but the DOAs of their wavefronts are different. Let τ 0 and τ1 be the
incremental delays produced by the signal wavefront and noise wavefront
respectively. The array can be steered to receive the signal or noise at the same
time (see fig. 3.20)[21]. Let f 1 (t) be the output of an array when it is steered
to the signal wavefront and f 2 (t) be the output of an array when steered to the
noise wavefront. Since the array response function has finite side lobes, some
∞
1
2π −∫∞
f 1 (t) = ξ 0 (t) + Ν1 (ω)H(ω(τ 0 − τ1 ))e jωt dω (3.68a)
and
∞
1
2π −∫∞
f 2 (t) = η1 (t) + Ξ0 (ω)H(ω(τ1 − τ 0 ))e jωt dω (3.68b)
where Ξ0 (ω) is the Fourier transform of the signal and Ν1 (ω) is that of the
noise. By comparing (3.65) with (3.68) it is possible to write
1
H2 (ω) = (3.69b)
H (ω(τ 0 − τ1 ))
Sη1 (ω)
Sξ 0 (ω)H ∗ (ω(τ1 − τ 0 )) + ∗
H (ω(τ 0 − τ1 ))
H pred (ω) =
2 Sη1 (ω)
Sξ 0 (ω) H(ω(τ1 − τ 0 )) + 2
H(ω(τ 0 − τ1 ))
1 + SNRinput
= 2 H(ω(τ 0 − τ1 ))
1 + SNRinput H(ω(τ1 − τ 0 ))
2
(a) When SNRinput >>1 and SNRinput H(ω(τ1 − τ 0 )) >>1
1
H pred (ω) ≈
H(ω(τ1 − τ 0 ))
If this filter is used on f 2 (t) (see eq.(3.68b)) for predicting the noise in f 1 (t) ,
the signal component will be restored causing the cancellation of the signal.
1
1
1
1
4τ
3τ
2τ
τ
3τ0
2τ0
4τ0
0
τ
Noise
Signal
Figure 3.20: An array of sensors can be steered simultaneously in the direction of the
signal and in the direction of noise. When the array is steered in the direction of
signal the output f 1 (t) is mostly signal and when it is steered in the direction of
noise the output f 2 (t) is mostly noise.
2
(b) SNRinput H(ω(τ1 − τ 0 )) <<1
where s1 (t) and s2 (t) are the first and the second sinusoid, respectively and
ω 0 is the frequency of the sinusoids. Solving (3.70) we obtain, for
2
H(ω 0 (τ 0 − τ1 )) < 1,
f 1 (t) − f 2 (t)H(ω 0 (τ 0 − τ1 ))
s1 (t) =
1 − H 2 (ω 0 (τ 0 − τ1 ))
(3.70b)
f 2 (t) − f 1 (t)H(ω 0 (τ 0 − τ1 ))
s2 (t) =
1 − H 2 (ω 0 (τ 0 − τ1 ))
∞
1
f 2 (t) = ∫
H
H null (ω)Ν1 (ω)e jωt dω (3.71a)
2π − ∞
where
a (ω)a 0H (ω)
H null (ω) = I − 0 a1 (ω)
M
1.5
0.5
Amplitude
-0.5
-1
-1.5
-2
0 50 100 150 200 250 300
1
(b)
0.5
Amplitude
-0.5
-1
0 50 100 150 200 250 300
1
(c)
0.5
Amplitude
-0.5
-1
0 50 100 150 200 250 300
Time
Figure 3.21: (a) sum of two sinusoids, (b) first sinusoid after subtraction and (c)
second sinusoid after subtraction.
1
∞ a1H (ω)a 0 (ω)
2
2π −∫∞
f 2 (t) = M 1 − Ν1 (ω)e jωt dω (3.71b)
M2
Define a filter
a 0H (ω)a1 (ω)
w= M (3.71c)
a1 (ω)a 0 (ω)
H 2
M 1 −
M2
If we now pass f 2 (t) through above filter (3.71c) the output will be exactly
the same as the noise term in (3.68a); therefore it may be removed by simple
subtraction.
3.5.3 Adaptive Filter: We have derived the multichannel Wiener filter in §3.3.
A single channel version may be derived along the same lines (see [22] for
derivation). Here we shall state the final result. The Wiener filter which predicts
f 1 (t) from f 2 (t) is given by h = C −1 f 2 C f 1 f 2 where C f 2 is the covariance
matrix of f 2 (t) and C f 2 f 1 is the cross-covariance matrix between f 2 (t) and
f 1 (t) . For real time estimation of the filter and also to account for temporal
variations in the covariance functions it is appropriate to devise an adaptive
approach which in the limiting case reduces to the Wiener solution. Let
h = [h0 , h1 , h2 ... hN −1 ] be the prediction filter vector and
T
T
be the data vector. The filter output is given by h f 2 , which is required to be
as close as possible to f 1 (t) . This filter is known as a transversal filter acting
on the delayed outputs (see fig. 3.22). For this we need to minimize the mean
square error
{
E ε(t)
2
} = E{ f (t) − h f
1
T
2
2
}
h0 h1 hN-1
+
Output
Figure 3.22: Structure of transversal filter. The filter coefficients are made adaptable
to the changing input.
with respect to the prediction filter coefficients. To minimize the error power
we need to compute a gradient of {
E ε(t)
2
} with respect to
h0 , h1 , h2 ... hN −1 and go down the path of the steepest descent until a
minimum (possibly a local minimum) is encountered. The difficulty, however,
is in estimating {
E ε(t)
2
}, which requires averaging over a finite interval
(ideally infinite) of time. Instead, in the least mean squares (LMS) algorithm, it
is proposed to use ε(t) in place of E ε(t)
2
{ 2
}. The gradient of ε(t)
2
is
now easily computed,
In the steepest descent search method the current filter vector is adjusted by an
amount proportional to negative of the gradient of the error function, ∇ ε(t)
2
[24]. The idea of adaptation is illustrated in fig. 3.23. The current filter vector is
updated by an amount proportional to the product of prediction error and current
input,
where we have assumed that hi and f 2 (t) are independent. Let h̃ represent
filter coefficients obtained by solving the normal equation (3.36), that is,
h̃ = Cf−12 Cf 2 f 1 . Equation (3.74) reduces to
{ } { }
E hi +1 − h̃ = E hi − h̃ + 2µh̃Cf 2 − 2µCf 2 E{hi } (3.75a)
f2(t)
hpred
Figure 3.23: The idea of adaptive prediction filter is illustrated in the above figure.
The prediction error modifies the filter so as to reduce the prediction error.
§3.6 Exercises:
1. A tapered fan filter is defined as [3],
1 u u
H(u, ω ) = rect( ) • rect( )
ω ω 2ω
where • stands for the convolution in u. Sketch the filter frequency response
function. Compute the impulse response function.
2. In the weighted least squares filter design the minimum mean square error,
which is given by
WH − [WH ] finite
+ π 2
1
ε = ∫∫
2
dudv ,
4π 2
min
− π
W
becomes independent of the weight function as the filter size increases, ideally
at infinity.
3. The output of a ULA with its sensors spaced at one meter apart is sampled at
the rate of 5kHz (Nyquist rate). A quadrant filter is desired with upper and lower
cut off speeds 7 km/sec and 3 km/sec, respectively. Sketch the pass regions
including the aliased part, if any.
4. The following are the low pass 1D filter coefficients:
h(0)=0.52
h(1)=0.3133176
h(2)=-0.01808986
h(3)=-0.09138802
h(4)=0.01223454
h(5)=0.04000004
References
Source Localization:
Frequency Wavenumber Spectrum
In this chapter we consider the most important problem in sensor array signal
processing which is estimation of the coordinates of a source emitting a signal
(passive localization) or a point target illuminated by an external signal (active
localization). A point in three dimensional space is defined by three parameters,
namely, range (r), azimuth ( ϕ ) and elevation (θ) . The range is often measured
by means of return time of travel in active systems and by means of time delays
measured at a number of sensors in passive systems. The azimuth and elevation
angles are obtained from the measurements of direction of arrival (DOA) by an
array of sensors. A horizontal array of sensors is required for azimuth
measurement and a vertical array for elevation measurement. The basic quantity
used for estimation of location parameters is the frequency wavenumber (ω , k)
spectrum (see chapter 2). A source is assumed to be present where there is a
concentration of power. We shall describe three different methods; namely,
beamformation, Capon spectrum and maximum entropy spectrum. The last two
methods fall under the nonlinear category while the first method belongs to the
linear category. The important difference between the linear and nonlinear
methods lies in their response to an input which consists of a sum of two or
more uncorrelated signals. The output of a linear method will be a sum of the
spectra of input signals but the output of a nonlinear method may contain an
additional cross term. In spite of this drawback the nonlinear methods have
become quite popular [1].
(2 π ) − π
ω
where k= and s = u + v . Using the stochastic properties of the
2 2
c
spectral representation of a wavefield the frequency wavenumber (ω , k)
spectrum may be given by
1
S f (ω,u,v)dωdudv
(2π)3
1 1
= E 3
dF(ω,u,v) 3
dF * (ω,u,v) (4.1)
(2π) (2π)
∞
1
3 ∫ ∫∫ f
S f (ω,u,v) = c (x, y, τ)e j (ωτ −ux − vy ) dxdydτ (4.2)
(2π) − ∞
where
c f (x, y, τ) = E{ f (x, y,t) f * (x + x, y + y,t + τ)}
M −1 + ∞ −j
2π
∑∫
mk
− jωt
F(ω, k) = f (t, md)e dt e M
m=0 −∞
M −1 jm( ω d − 2 π k ) + ∞
= ∑e
m=0
cx M
∫f 0 (t)e − jωt dt (4.4)
−∞
ω 2π
= F0 (ω)H( d− k)
cx M
ω 2π
where H( d− k) is the response function of a ULA of length M. Note
cx M
that
ω2
v
ω1
ω 2π ω
H( d− k) → δ( d − u)
cx M cx
as M → ∞
2π
where k → u as M, k → ∞ . To avoid the aliasing error in the spectrum
M
the temporal sampling interval and sensor spacing must satisfy the relation
shown in (1.34b) which, when expressed in terms of ω max and d, reduces to
d c
ω max ≤ (4.5a)
π sin θ
For fixed ω max , d and c, in order to avoid aliasing error, the angle of incidence
will have to satisfy the following inequality,
π c
θ ≤ sin −1 ( ) (4.5b)
ω max d
d' ωmax d
c'
c
2
π π π 2π u
d - d d d
b b'
a - ωmaxa'
Figure 4.2: Aliasing error due to spatial sampling of a broadband plane wave.
λ π
From (4.5b) it may be seen that for d= and ∆t = there is no
2 ω max
aliasing for any angle of incidence. Aliasing error will occur whenever the
above requirements are not satisfied. As an example, consider a stochastic plane
wave, incident on a ULA with 16 sensors at an angle of 45o. The sensors are
15 meters apart. The bandwidth of the waveform is ±100Hz and it is sampled
with a sampling interval of 0.005 sec. The aliasing error is present in the top
left and bottom right corners (see fig. 4.3).
4.1.3 Spectral Matrix: The output of an array of sensors may be treated as a
collection of time series or vector time series. A spectral matrix whose
elements
80
60
40
Frequency (Hz)
20
0
-20
-40
-60
-80
Figure 4.3: Aliasing error due to spatial undersampling. A stochastic plane wave is
incident on a ULA at 45°. The ULA consists of 16 sensors spaced 15 meters apart.
The wave speed is 1500m/s. The bandwidth of the waveform is ±100Hz and it is
sampled with a sampling interval of 0.005sec.
∞
1
∫ ∫ dF(ω,u)e
j (ωt −umd )
f m (t) = f (t, x = md) = 2 (4.6)
4π −∞
Using (4.6), the cross-covariance function between two sensor outputs is given
by
cmn (τ) = E{ f m (t) f n (t + τ)}
∞
1 (4.7)
∫ ∫ s (ω,u)e
jωτ − j ( m − n)ud
= f e dωdu
4π 2 −∞
Further, the spectral representation of a cross-covariance function [2] is
Comparing (4.7) and (4.8) we obtain the following relation between elements
of the spectral matrix and (ω , k) spectrum:
∞
1
2π −∫∞
Smn (ω) = S f (ω,u)e jd ( m − n)u du
(4.9a)
∞
1
2π −∫∞
Smm (ω) = S f (ω,u)du
The reverse relation, that is, (ω , k) spectrum in terms the elements of spectral
matrix, is
∞ ∞
S f (ω,u) = ∑ ∑S
m=−∞ n=−∞
mn (ω)e − jd ( m − n)u (4.9b)
The spectral matrix has Hermitian symmetry. Additionally, for a ULA, it has
toeplitz symmetry. Consider an example of a stochastic plane wave incident on
a ULA. The output of the mth sensor is given by
∞ d
1 jω(t − m sin θ)
2π −∫∞
f m (t) = dF(ω)e c
∞ d
1 jω( τ −( m − n) sin θ)
2π −∫∞
cmn (τ) = S f (ω)e c
dω (4.10)
From (4.10) the cross-spectrum between the outputs of the mth and nth sensors
is given by
d
− jω( m − n) sin θ
Smn (ω) = S f (ω)e c
(4.11a)
ω
where u= sin θ . Comparing (4.11b) with (4.9) we obtain
c
S f (ω,u) = S f (ω)δ(ω − u) .
The spectral matrix for this model has a very useful representation, that
is, as an outer product of two vectors
where
− jω dc sin θ − j ( M −1)ω sin θ
d
a(ω) = col 1,e ,...,e c
is the direction vector of the incident plane wave. Eq. (4.12a) may be easily
generalized for P uncorrelated sources,
P −1
S f (ω) = ∑ S f i (ω)a i (ω)a i H (ω)
i=0
{ }
S0 (ω) = diag S f 0 (ω), S f 1 (ω),..., S f P−1 (ω)
A(ω) = [a 0 (ω),a1 (ω),...,a P −1 (ω)]
The spectral matrix for a case of P uncorrelated waves and uncorrelated white
background noise is given by
A1 ( P × P)
A(ω) =
A 2 ( M − P × P)
Since A(ω) has full column rank, there is a unique linear operator known as
propagation matrix, Γ (P × M − P) , such that Γ A1 = A 2 which may also
H
be written as
Γ
AH = AHQ = 0 (4.13a)
−I
It follows that Q spans the null space of A . Now let us use the partitioned
A(ω) matrix in (4.12b)
P M − P
Note that the spectral matrix has also been partitioned so that
where I P stands for a unit matrix of size PxP. It may be shown from (4.13b)
that
G 2 = Γ H A1S0 A1H = Γ H (G1 − σ 2ηI P )
and hence,
Thus, the propagation matrix may be derived from the partitioned spectral
matrix. The background noise variance is assumed to be known (see Exercises,
4.4). Q may be used to find the direction of arrival in place of eigenvectors
corresponding to noise eigenvalues as in the MUSIC algorithm to be described
later in chapter 5 [4, 5].
4.1.4 Eigenstructure: The spectral matrix possesses interesting eigenstructure.
S f (ω) is a hermitian symmetric toeplitz (only for a ULA) matrix; hence its
e(ω) be
eigenvalues are real. Further, they are positive as shown below: Let
H
some arbitrary vector and consider a quadratic form e (ω)S f (ω)e(ω) . It
follows from (4.13a) for any e(ω) ,
P −1
e H (ω)S f (ω)e(ω) = ∑ S f i (ω)e H (ω)a i (ω)a i H (ω)e(ω)
i=0
P −1
= ∑ S f i (ω) e H (ω)a i (ω) ≥ 0
2
i=0
ω max
1 ω
S(θ) =
2ω max ∫
−ω max
S f (ω,
c
sin θ)dω (4.15b)
In the (ω , k) domain the integration is carried out over a radial line sloping at
angle θ (see fig. 4.4). A peak in the angular spectrum is an indication of wave
energy arriving from a direction where the peak is found.
The angular spectrum defined in (4.15b) is also the beam power integrated over
the frequency bandwidth as a function of the look angle.
4.1.6 Parametric Spectrum: The frequency wavenumber spectrum defined in
(4.15a) assumes a plane wave model which is more appropriate in open space
(see chapter 1 for wavefield representation in open and bounded space). In
bounded space the wavefronts are far from planar. Such nonplanar wavefronts
may be represented in terms of source location parameters measured with
reference to the bounded space geometry. For example, in s hallow water the
ωmax
0 u
Figure 4.4: In the frequency wavenumber plane the spectrum is averaged over a series
of radial lines. The spectrum of a plane wave which is incident at angle θ 0 is shown
by the bold line.
source location parameters are range, depth (measured from the surface) and
azimuth. In place of frequency wavenumber spectrum, where the parameters of
interest are frequency and wavenumbers which depend on azimuth and elevation,
we introduce a similarly defined quantity called parametric spectrum
where θ now stands for generalized location parameters and a(θ) is the
wavefield which the array would sense if the source were to be located at θ .
S(ω ,θ) is computed over a range of values of θ spanning the entire
parameter space. The actual source position is indicated by the position of the
maximum in S(ω ,θ) . Evidently, a(θ) must be computed for the assumed
geometry and the boundary conditions there on. Since the central idea is to
match the computed field with the observed field, the processing is also known
as matched field processing. It was first introduced by Bucker [6] in 1976 and
since then a lot of research effort has been devoted toward its development as a
tool in underwater detection. An account of this effort is summarized in a
monograph [7]. The chief obstacle in the use of matched field processing lies in
the requirement of an exact knowledge of the propagation environment for the
purpose of computing the wavefield.
3.06, 0.0, -3.06, -5.65, -7.39, -8.00, -7.39, -5.66, -3.06, 0.0, 3.06, 5.66, 7.39
(rounded to second decimal place). All these delays are with respect to a
∆ω
ω0 +
2 d
1 jω(t − m sin θ 0 )
f m (t) ≈
2π ∫ ∆ω
Fnb (ω)e c dω
ω0 −
2
ω (t − m d sin θ )
∆ω 0 0
j c
2 d
1 +δω(t − m sin θ 0
)
=
2π ∫ F(ω
∆ω
0 + δω)e c dω (4.17a)
−
2
d
− jω 0 m sin θ 0 )
=e c f nb (t)
where the subscript nb stands for narrowband. The approximation is valid when
d
the bandwidth satisfies the condition, ∆ω (m sin θ 0 ) << 2 π for all m,
c
which implies that the time taken for a wave to sweep across the array must be
much less than the inverse of the bandwidth, expressed in Hertz. In vector
notation (4.17a) may be expressed as
{
S(ω 0 ,θ) = E a H a 0 f nb (t)
2
}= a aH
0
2
σ 2f (4.17c)
Window: The sensor outputs are often weighted before summation, the purpose
being to reduce the sidelobes of the response function just as in spectrum
estimation where a window is used to reduce the sidelobes and thereby reduce
the power leakage. As this topic is extensively covered under spectrum
estimation, for example see [2,9], we shall not pursue any further. Instead, we
like to explain the use of a weight vector to reduce the background noise
variance or to increase the snr. Let us select a weight vector, w , such that the
signal amplitude is preserved but the noise power is minimized.
c −1
η a0
w= H −1
(4.18b)
a0 cη a0
It may be observed that for spatially white noise c η = σ 2ηI and therefore
a0
w= . In other words, the weights are simply phase shifts or delays as in
M
beamformation. The variance of the noise in the output is equal to
σ 2η̂ = σ 2η /M.
A weight vector may be chosen to maximize the snr. The output
power of an array with a weight vector w , when there is no noise, is given by
w H c s w and when there is no signal, the output power is w H c η w . We select
that weight vector which will maximize the output power ratio
w H csw
= max (with respect to w )
w H c ηw
2 2
s(ω 0 , θ ) = a H a1 σ 2f 1 + a H a 2 σ 2f 2
2
In order that each signal gives rise to a distinct peak, a H a1 σ 2f 1 when plotted
2
as a function of θ should not overlap with a H a 2 σ 2f 2 . A condition for
nonoverlap is necessarily arbitrary as the array response to an incident wavefront
is strictly not limited to a fixed angular range The Rayleigh resolution criterion
states that two wavefronts are resolved when the peak of the array response due
to the first wavefront falls on the first zero of the array response due to the
λ
second wavefront. The first zero is located at an angle, sin −1 , away from
Md
the direction of arrival (broad side). Thus, two wavefronts are resolved,
according to the Rayleigh resolution criterion when their directions of arrival
λ
differ by sin −1 . An example of resolution is shown in fig. 4.5. For a UCA
Md
we can derive a simple expression when it is fully populated. In this case, its
response function is a Bessal function of 0th order (see eq. 2.43b). The first zero
of the Bessal function of 0th order is at 2.45. Two wavefronts are said to be
resolved according to the Rayleigh resolution criterion when the angular
separation is greater than sin −1 (
1. 225λ
). Let us compare the resolution
πa
properties of a ULA and a UCA having equal aperture, for example, a 16 sensor
ULA with 7.5λ aperture and the corresponding UCA with a radius equal to
3.75λ but fully populated with more than 32 sensors. The relative performance
0.8
0.6
Spectrum
0.4
0.2
0.0
-90 -70 -50 -30 -10 10 30 50 70 90
Angle in deg
Figure 4.5: Two uncorrelated wavefronts with DOA’s, 0o and 7.18o, are incident on a
16-sensor ULA. The waves are clearly resolved. The DOAs were chosen to satisfy the
Rayleigh resolution criterion.
Table 4.1: The Rayleigh resolution angle as a function of the number of sensors
λ
(ULA with sensor spacing).
2
is shown in table 4.1. The performance of the UCA is marginally better than
that of the ULA. Beamformation in the frequency domain requires the 2D
Fourier transform. For a fixed temporal frequency, the magnitude of the spatial
Fourier transform coefficients is related to the power of a wave coming from a
d d
H(ω) = a H (ω sin θ)a(ω sin θ 0 )
c c
where θ 0 is DOA of the incident wave and θ is the steering angle. We shall
model two types of phase errors, namely, those caused by position errors and
phase errors caused by all other sources lumped into one. The corrupted
direction vector has the following form:
where ∆d1 is the position error of the ith sensor and φ1 is the phase error. We
have assumed that the first sensor is a reference sensor and hence there is no
(a)
10
Magnitude
0
0 2 4 6 8 10 12 14
20
(b)
15
Magnitude
10
0
0 8 16 24 32 40 48 56
Frequency Numbers
Figure 4.6: The role of padding zeros or introducing dummy sensors is to correctly
position the peak (a) before padding zeros and (b) after padding zeros. The beam
width remains unchanged.
position error. We shall illustrate the effect of position and phase errors on the
array response function. We assume that the ULA has 16 sensors which are
λ
equispaced but with some position error. Let d= and ∆d be a uniformly
2
0.8
0.6
Response
0.4
0.2
0
-100 -50 0 50 100
Angle in deg
Figure 4.7: Response of a ULA with position errors which are uniformly distributed
λ
in the range ± (solid curve). Compare this with the response of the ULA without
4
any position errors (dashed curve).
λ
distributed random variable in the range ± . The resulting response is shown
16
in fig. 4.7. The array response due to phase errors, caused by other factors, is
shown in fig. 4.8. The phase errors seem to cause less harm compared to the
position errors. The sensor position and phase errors largely affect the sidelobe
structure of the response function while the main lobe position and the width
remain unchanged.
4.2.2 Broadband Beamformation: Beamformation with a broadband signal can
also be written in a form similar to that for a narrowband signal. We must first
Fourier transform (temporal) the broadband signal output from each sensor and
treat each Fourier coefficient as a Fourier transform of a narrowband signal
whose bandwidth is approximately equal to the inverse of the time duration of
the signal. The frequency wavenumber spectrum in this case is given by
0.8
0.6
Response
0.4
0.2
0
-100 -50 0 50 100
Angle in deg
Figure 4.8: Response of a ULA with phase errors which are uniformly distributed in
π
the range ±
4
{
S(ω,θ) = E a H a 0 F(ω)
2
}= a aH
0
2
S f (ω) (4.20)
1,e jud ,...,e ju( M −1)d ;e jω∆t ,e j (ud + ω∆t ) ,...,e j (ud ( M −1)+ ω∆t ) ;...;
A = col jω ( N −1)∆t j (ud + ω ( N −1)∆t )
e ,e ,...,e j (ud ( M −1)+ ω ( N −1)∆t )
(4.21b)
It is easy to show that the response function can be expressed as an inner
product of two vectors defined in (4.21)
M −1 N −1
H(u,ω) = ∑∑w
m=0 n=0
mn e − j (umd +ωn∆t )
(4.22)
=A w H
1 ω2ω b H
Output energy = w 2 ∫ ∫ AA dudω w
H
(4.23a)
4π ω 1 ω a
1 ω max π d H
Total energy = w 2 ∫ ∫ AA dudω w
H
(4.23b)
4π ω min − π d
where a and b are slopes of the radial lines defining the passband (see fig. 3.6),
ω 2 and ω1 are respectively the upper and the lower cut-off frequency for the
beam, and ω max and ω min refer to the maximum and the minimum frequency
present in the signal, respectively. The problem may be expressed as a problem
Figure 4.9: The structure of a 2D filter for broadband beamformation. M=8 and N=4.
w H Γ 1w
in maximization of a ratio, = λ , which is achieved by solving the
w HΓ2w
following generalized eigendecomposition problem:
Γ 1w = λ Γ 2 w (4.24)
ω2ω b ω max π d
1 1
where Γ1 = ∫ω ω∫aAA dudω and Γ 2 = 4 π 2 ∫ ∫ AA dudω . The
H H
4π 2 1 ω min − π d
solution is given by the eigenvector corresponding to the largest eigenvalue and
the maximum relative power is equal to the largest eigenvalue.
To evaluate Γ1 and Γ 2 , we must first simplify the elements of
AA H ,
The principle of maximizing the power in the passband was first suggested in
[11] in the context of optimum window for spectrum estimation of time series.
Later, this principle with additional constraints on the magnitude and derivative
has been applied in beamformation [12, 13]. In fig. 4.10 we show a numerical
example of spatio-temporal filter for broadband beamformation. The desired
response is unity in the region bounded by two radial lines and the upper and
the lower frequency cutoff lines. The actual response of the maximum energy
filter is contoured in the same figure. It is observed that the maximum side lobe
level is about 4dB less than that in the simple quadrant filter shown in fig. 3.7.
§4.3 Capon’s ω-k Spectrum:
We consider a stochastic plane wave incident on a ULA. Let us represent the
array output in a matrix form,
∞
1
f(t) = ∫
2π − ∞
dF(ω)e jωt (4.26a)
where
d
dF(ω) = dF 0 (ω)a(ω sin θ) + dη(ω) (4.26b)
c
2
0.5
0.9
1
Temporal frequency
0.1
0.1
-1 0.5
0.9
-2
-3
−π
-3 -2 -1 0 1 2 3 π
Spatial frequency
Figure 4.10: Response of filter for broadband beamformation. The slopes of the
radial lines are a=1 (45o ) and b=2 (63.4o ). ω1 = 0 and ω 2 = 0.6π . The maximum
energy in the passband is 94% ( λ max = 0.94). 16 sensors and 16 delayed samples.
∆x = 1 and ∆t = 1.
d d
S f (ω) = S0 (ω)a(ω sin θ)a H (ω sin θ) + S η (ω)
c c
where both S f (ω) and S η (ω) are MxM matrices but S0 (ω) is a scalar. We
like to find a weight vector w acting on the array output such that it
minimizes the power output of the array and is transparent to all waves
propagating through a narrow cone with a cone angle ∆θ and its axis pointing
in the direction of θ 0 . Thus, the beamwidth is made intentionally wider to
allow for possible variation in the direction of arrival. This model is useful
w H S f (ω)w = min
(4.27a)
w H Γw = 1
where
∆θ
θ0 +
2
1 d d
Γ=
∆θ ∫ ∆θ
a(ω sin θ)a H (ω sin θ)dθ
c c
(4.27b)
θ0 −
2
and S f (ω) is the array signal spectrum under the assumption that the source
bearing lies in the range θ 0 ± ∆θ .
The constrained minimization problem specified in (4.27) is solved by
the Lagrange method,
S f w = λΓw (4.29a)
or
λ−1w = S−1
f Γw (4.29b)
d d
Γ = a(ω sin θ 0 )a H (ω sin θ 0 )
c c
Equation (4.29b) now becomes
d
By premultiplying both sides by a H (ω sin θ 0 ) we find that
c
d d
λ−1 = a H (ω sin θ 0 )S−1
f a(ω sin θ 0 ) (4.31a)
c c
It turns out that
d
S−1
f a(ω sin θ 0 )
w= c (4.31b)
d d
a H (ω sin θ 0 )S−1
f a(ω sin θ 0 )
c c
satisfies (4.30). We can express the array output power, which we shall call as
Capon spectrum,
1
sCap (ω,θ 0 ) = (4.32)
d d
a (ω sin θ 0 )S−1
H
f a(ω sin θ 0 )
c c
Capon [14], who first suggested the above measure of spectrum, however,
called it maximum likelihood spectrum. It is also known as the minimum
variance distortionless response (MVDR) beamformer or a linearly constrained
minimum variance (LCMV) beamformer [15]. Since θ is related to the spatial
d
frequency, u = ω sin θ , sCap (ω,θ) is indeed a (ω , k) spectrum as a
c
function of θ or u.
4.3.1 Resolution: The Capon spectrum has a better resolution compared to the
BT ω-k spectrum. We shall demonstrate this by considering two uncorrelated
wavefronts in the presence of white noise. The spectral matrix is given by
where θ 0 and θ1 are directions of arrival and s0 and s1 are powers of two
plane wavefronts and σ η is noise variance. The inverse of the spectral matrix
2
where
s0
a 0 (ω,θ 0 )a 0 H (ω,θ 0 )
1 σ 2
V −1 = 2 I − η (4.34b)
ση s0
1+ 2 M
ση
Using (4.34a) in (4.32) we obtain the Capon spectrum for the two source
model,
σ 2η
sCap (ω, θ) = 2 (4.35)
s0 H H
2 a a 0 a 0 a1
s1 H σ η
a a1 −
s0 H 2 σ 2η s
1 + 02 M
a a0
σ 2η ση
M− −
s
1 + 02 M s0 H 2
ση 2 a 0 a1
s1 σ η
1+ 2 M −
ση 1 +
s0
M
σ 2η
where for the sake of compactness we have dropped the arguments of the vectors
a , a 0 , and a1 . When the steering vector points to one of the sources, for
example, when a = a0
σ 2η σ 2η
sCap (ω, θ 0 ) = 2 ≈ s0 + (4.36a)
s0 H M
2 Ma 0 a1
s1 H σ η
a 0 a1 −
σ 2η s
1 + 02 M
M ση
−
s
1 + 02 M s0 H 2
ση 2 a 0 a1
s1 σ η
1+ 2 M −
ση 1 +
s0
M
σ 2η
and when a = a1
M−
s
1 + 02 M
ση
The approximation shown in (4.36) is valid for a1H a 0 << M . From the
above it is clear that when the wavefronts are well resolved the peak amplitude
approximately equals the power of the source. The noise power is reduced by a
factor equal to the number of sensors.
We will examine the resolution power of the Capon spectrum.
Consider again two equal power wavefronts incident at angles θ 0 and θ1 . The
peaks corresponding to two wavefronts are resolved when a valley is formed in
between them. Let sCap (ω , θ̃ ) be the spectrum at θ̃ midway between θ 0 and
θ1 . Define the ratio ρ as
sCap (ω,θ 0 )
ρ=
sCap (ω, θ̃)
a 0H a1
2
ã H a1
2
1 − α − (1 + α) 2
s1 M M M ã H
a
1 + 2 −α 1
ση ã a1a1 a 0 a 0 ã
H H H M (4.37a)
+2α Re
M3
= H 2
a a
1− α 0 1
M
sM
Assuming 1 >> 1 (4.37a) simplifies to
σ2
η
s1 M sM
where α = ≈ 1 for 1 2 >> 1. A valley is formed iff ρ > 1. Let
1 + s1 M ση
a 0 = a1 = ã , that is, when two wavefronts merge into a single wavefront we
notice that ρ = 1 for all snr, which means that these two wavefronts can never
2 2
aHa ã H a1
be resolved. Next, let 0 1 = ≈ 0 which means that the wavefronts
M M
sM
are well separated. Then, ρ = 1 + 1 2 >1 except when s1 =0. The wavefronts
ση
can then always be resolved. All the above conclusions follow from common
∆θ
sense. We now consider two wavefronts with DOAs ± respectively and
2
sM
compute ρ for different 1 2 and ∆θ . A plot of ∆θ for which ρ is just
ση
greater than one as a function of array snr is shown in fig. 4.11.
4.3.2 Robust Beamformation: The sensitivity of beamformation to errors in the
sensor position and other phase errors has been demonstrated in figs. 4.7 and
4.8. These drawbacks may be reduced through an appropriate choice of
weighting coefficients. In this section we shall show how such coefficients can
be obtained [17] following a constraint used in deriving Capon’s filter and the
{ }
associated spectrum. Let w = col w0 , w1 ,...wM −1 be a coefficient vector.
The array response may be expressed as
d
H(ω) = w H a(ω sin θ) (4.38)
c
We shall model two types of phase errors, namely, those caused by position
errors and those caused by all other sources of errors lumped into a single phase
6
Angle in deg.
0
.01 .1 1 10 100
Array SNR
Figure 4.11: Resolution properties of the Capon spectrum as a function of array snr.
A 16 sensor ULA is assumed. The angles of incidence are 30o and 30o + angular
separation as shown on the y-axis. Simple beamformation (BT ω-k Spectrum) yields
a resolution of 7.18o shown in the figure by a thick line for comparison. Resolution
gain by the Capon Spectrum is possible only for high array snr.
error. The steering vector given by (4.19). We have assumed that the first
sensor is a reference sensor and hence there is no position error. Let H0 be the
desired response, for example, equal to 1 and H̃ be corrupted response,
H̃ = w H ã (4.39)
2
mse = w H Qw − (H0 P H w + H0H w H P) + H0 (4.40b)
where
Qw 0 = P (4.41)
mse = ( w 0 − w ) Q( w 0 − w ) + H0 − w 0H Qw 0
H 2
(4.42)
H
Output power= w S f w =min (4.43a)
( w 0 − w ) H Q( w 0 − w ) ≤ ε 2 (4.43b)
2
where ε 2 = mse − H0 + w 0H Qw 0 is a prescribed number which represents
an excess error over the minimum that can be achieved by satisfying (4.41).
Note that S f is a spectral matrix of the array output. Using the standard
primal-dual method we can solve the constrained optimization problem [18].
The solution is given by
w = w 0 − (S f + λ Q)−1 S f w 0 (4.44)
a circular array consisting of two concentric rings with random interring spacing
the array gain remains practically unaltered if the actual spacing is well within
the bounds used in the design of the weight coefficients. However, the sidelobe
characteristics of the response function of the weight coefficients are not
known.
4.3.3 High Resolution Capon Spectrum: The resolution capability of Capon’s
frequency wavenumber spectrum may be improved by noise cancellation by
subtracting an estimated white noise power from the diagonal elements of the
spectral matrix [19] and stretching the eigenvalue spread of the spectral matrix
[20]. A predetermined quantity is subtracted from the diagonal elements of the
spectral matrix, thus increasing the ratio between the maximum and minimum
eigenvalues. This process is called stretching the eigenvalue spread. The spectral
matrix must, however, remain positive definite. Consider a model of P plane
waves and white background noise. The spectral matrix, given in (4.12b), is
subjected to stretching of the eigenvalue spread by subtracting a fixed number
σ 20 from the diagonal elements,
S̃ f (ω) = S f (ω) − σ 20 I
(4.46)
= vΓ 0 v H
where
{ }
Γ 0 = diag α m + σ 2η − σ 20 , m = 1,2,..., M − 1
= diag{γ m , m = 1,2,..., M − 1}
1
s̃Cap (ω , θ ) = P −1 (4.48)
1 H 2 M −1 1 H 2
∑ a v m + m∑= P γ a v m
m=0 γ m m
In the denominator of (4.48) the second term dominates whenever a does not
belong to the set of direction vectors of the incident wavefronts and vice versa.
Hence,
2
As an example, consider a single source, that is, P=1, then a H v 0 = M and
α 0 = Mp0 . On account of (4.14c) a H v m = 0 (but γ m ≠ 0 ) for m=1, 2,...,
M-1. Therefore, s̃Cap (ω,θ 0 ) = p0 (power of the incident wavefront). For all
other values of θ s̃Cap (ω,θ) ≈ 0 . To demonstrate the resolution power of the
high resolution Capon spectrum we consider a 16 sensor ULA and two
wavefronts incident at angles 30o and 35o and 0dB snr. The resolution of the
Capon spectrum as shown in fig. 4.12 has dramatically improved when the
eigenvalue spread is increased from 30 to 1000. The above result is for an error-
free spectral matrix, that is, with infinite data. To study the effect of finite data
1.4 a
1.2
Spectrum
1
0.8
0.6
0.4
0.2
0
10 20 30 40 50
1.2
1.0
1
0.8
Spectrum
0.6
0.4
0.2
0
10 20 30 40 50
DOA in deg
Figure 4.12: (a) Capon spectrum. (b) High resolution Capon spectrum. Eigenvalue
spread is 1000. 16 sensor ULA with sensor spacing of λ/2 and two uncorrelated
wavefronts incident at 30o and 33o were assumed. The high resolution Capon
spectrum yields correct amplitude and DOA information.
a numerical experiment was carried out [20] on an eight sensor ULA with two
wavefronts incident at angles 45o and 53o in presence of white background
∞
1
Smn (ω) = ∫
2π − ∞
S f (ω,u)e − jd ( m − n)u du (4.9a)
∞
1
4π −∫∞
∆H = log S f (ω,u)du (4.49)
13
11
Angle in deg.
5
-15 -10 -5 0
SNR in dB
1
S f (ω,u) = p= p 0 (4.50)
∑λ e
p= − p 0
p
jup
1
S f (ω,u) =
H(u)H H (u)
1 (4.51)
= p0 p0
∑h e ∑h e
m=0
m
jum
m=0
m
− jum
1
SfH = δ (4.52a)
h0
1 −1
H= Sf δ (4.52b)
h0
where { }
a = col 1,e ju ,e j 2u ,...,e jp0 u is the steering vector.
We shall look at an alternate approach which will lead to an equation
identical to (4.52b). At a fixed temporal frequency the output of an array (ULA)
may be expressed as a sum of complex sinusoids,
∞ p d
1 0 jω(t − m sin θ p )
f m (t) = ∫ ∑
2π − ∞ p= 0
dGp (ω)e c
(4.53a)
Let
P ωd
− jm sin θ p
dFm (ω) = ∑ dGp (ω)e c
, m = 0,1,..., M − 1 (4.53b)
p= 0
where we have assumed that P plane wavefronts are incident on a ULA. The
sources are assumed to radiate stationary stochastic but uncorrelated signals.
Clearly, (4.53b) is a sum of P random spatial complex sinusoids. A sum of
P random sinusoids (real) are known as a deterministic random process and it
can be predicted without error from 2 P past samples [2]. In the case of P
complex sinusoids we will require P past samples for prediction. Error-free
prediction is not possible when there is background noise. The prediction
equation is given by
P
dFm (ω) + ∑ hp dFm − p (ω) = ηm (4.54a)
p=1
H H dF m (ω) = ηm (4.54b)
where {
dF m (ω) = col dF m , dF m −1 ,..., dF m − p0 } and h =1. The prediction
0
error is given by
H H S f (ω)H = σ 2η (4.55)
S f (ω)H = σ 2ηδ
or
H = σ 2ηS−1
f (ω)δ (4.56)
which is identical to (4.52b), except for a scale factor. Using (4.56) in (4.55)
1
the minimum prediction error is equal to σ 2η = −1
. The
min δ S (ω)δ
H
f
prediction filter vector corresponding to minimum error is given by
S−1 (ω)δ
H = H f −1 (4.57)
δ S f (ω)δ
(δ )
2
H
S−1
f (ω)δ
SME (ω,u) = 2 (4.58)
a H S−1
f (ω)δ
Note that the numerator of (4.58) is equal to the first element of the first
−1
column of S f . The maximum entropy spectrum given by (4.52c) and the
spectrum obtained by minimizing the prediction error, (4.58), are identical.
Thus, the maximum entropy principle leads to a simple interpretation in the
form of linear prediction.
4.4.2 Resolution: We shall now examine some of the properties of the
maximum entropy spectrum and compare them with those of the Capon
spectrum. As earlier we shall consider two wavefronts incident on a ULA. The
spectral matrix is given by (4.33). Using the inversion formula (4.34) in
(4.52c) we obtain an expression for the maximum entropy spectrum for two
wavefronts in presence of white noise.
h02 σ 4η
sME (ω,θ) = (4.59)
den
where
The height of the spectral peak grows with the array signal-to-noise ratio,
increasing to infinity as sM → ∞ . This is demonstrated in fig. 4.15 for the
two wavefront model, whose spectrum is plotted in fig. 4.14.
From figs. 4.14a and 4.14b it may be conjectured that the maximum
entropy spectrum has a better resolution property than that of the Capon
spectrum. The depth of the valley for the maximum entropy spectrum is much
larger than the one for the Capon spectrum. We have carried out a series
computation to find out the minimum snr required to resolve two equal
amplitude wavefronts separated by a specified angular distance. The criterion for
resolution was formation of a nascent valley between the spectral peaks. While
this is not a quantitative criterion it serves the purpose of comparison. The
results are shown in fig. 4.16 which may be compared with fig. 4.11 for the
Capon spectrum. Clearly, the maximum entropy spectrum has a better
resolution capability, but its peak amplitude does not bear any simple relation
to the actual spectrum value. In the time series context it was shown in [21]
that the Capon spectrum and maximum entropy spectrum of different orders are
related,
80
60
Spectrum
40
20
0
0 20 40 60
DOA in deg
1.2
0.8
Spectrum
0.6
0.4
0.2
0
0 20 40 60
DOA in deg
Figure 4.14: (a) Maximum entropy spectrum and (b) Capon spectrum. Two unit
amplitude plane wavefronts are incident at angles 30o and 35o on a 16 sensor ULA.
The amplitude of the peaks of Capon spectrum is close to the actual amplitude but the
amplitude of ME spectrum is much higher. According to (4.53), valid for large
separation, the computed amplitude is equal to 200.
10 5
10 4
Peak spectrum
10 3
10 2
10 1
10 0
0 20 40 60 80 100
Array SNR
Figure 4.15: Peak spectrum as a function of array snr, that is, Msnr. Two equal
amplitude uncorrelated wavefronts incident at 30o and 35o on a 16 sensor ULA.
1 1 M −1 1
= ∑ m (ω)
SCap (ω) M m = 0 SME
(4.61)
where M stands for the size of the covariance matrix. In array processing, M
stands for array size.
4.4.3 Finite Data Effects: So far we have tacitly assumed that the spectral
matrix is known and the incident wavefield and noise confirm with the assumed
model; for example, the wavefronts are planar and uncorrelated and the noise is
white. In practice, however, the spectral matrix needs to be computed from the
available data. Since the spectral matrix is a statistically defined quantity
involving the operation of expectation, there is bound to be some error in its
estimation when only finite length data is available. The effect of the errors in
the spectral matrix on wavenumber spectrum has been a subject of investigation
by many researchers [22, 23]. Here we shall briefly outline the important
results. The mean and variance of the BT frequency wavenumber spectrum
(linear), the Capon frequency wavenumber spectrum (nonlinear), are tabulated in
6
Angle in deg.
0
.01 .1 1 10 100
Array SNR
table 4.3. Unfortunately, we do not have simple expressions for mean and
variance of ME spectrum estimates; only experimental results are available
[22]. The ME spectrum is more variable and hence it needs much larger data to
1 N
Ŝ f (ω) = ∑
N 1
Fi (ω)FiH (ω)
When a new time snapshot arrives a new segment is formed with the newly
received snapshot and the past snapshots to form a required length segment and
then a new frequency snapshot is formed. The spectral matrix is updated by
incorporating the outer product of the newly formed frequency snapshot,
1 N
Ŝ Nf +1 (ω) = ∑
N +1 1
Fi (ω)FiH (ω) + F N +1 (ω)F HN +1 (ω)
(4.62a)
N 1
= Ŝ Nf (ω) + F N +1 (ω)F HN +1 (ω)
N +1 N +1
The recursion may be commenced with an initial value
Ŝ (ω) = F1 (ω)F (ω) . We can obtain a recursive expression for the (ω , k)
1
f
H
1
spectrum by using (4.62a) in (4.16)
N S N (ω,θ) +
N +1
N +1
S (ω,θ) = 2
(4.62b)
1 a H (ω d sin θ)F (ω)
N + 1 c
N +1
Using the matrix inversion formula given in (4.34) we can get a recursive
relation between the inverse of spectral matrix Ŝ Nf (ω) and Ŝ Nf +1 (ω)
[ ] [ ] (4.62c)
1 N −1 −H
Ŝ f (ω) F N +1 (ω)F HN +1 (ω) Ŝ Nf (ω)
N +1 N
[ ]
−1
Ŝ f (ω) − N
[ ]
1 −1
N 1 + F HN +1 (ω) Ŝ Nf (ω) F N +1 (ω)
N
The above recursion can be commenced only after Ŝ Nf (ω) becomes full rank.
This will require a minimum of M frequency snapshots. There is yet another
recursive approach to spectral matrix inversion. It is based on diagonalization of
the spectral matrix through a transformation,
or equivalently,
[Ŝ (ω)]
−1
N
f = QΓ −1Q H (4.64)
§4.5 Exercises:
1. The spatial undersampling has resulted into aliasing of the frequency
wavenumber spectrum as illustrated in figs. 4.2 and 4.3. Now consider temporal
undersampling. Sketch the (ω , k) spectrum of a stochastic plane wave which
has been undersampled temporally.
2. Apply the principle of Rayleigh resolution to wideband signals. In chapter 2
we have shown how the DOA of an incident wideband signal can be estimated
from the position of the spectral peaks of the transfer functions. Show that for
resolution the wavefronts must be separated by an angle greater than ∆θ ,
where
tan θ
∆θ =
M
tr(H 2 π)
σ 2η =
tr(π)
It is proposed to findw 0 which shall minimize the noise variance and at the
same time preserve the signal power. Show that w 0 is equal to the eigenvector
corresponding to the least eigenvalue of Γ S ηΓ where S η is the spectral
H
(see 4.62a). Show that the eigenstructure of the spectral matrix of the reduced
output is identical to that of the spectral matrix of the normal array output, in
particular, B a ⊥ v η . This property forms the basis for the beamspace
H
subspace method where B acts as a spatial filter to restrict the incident energy
to a preselected angular sector [26, 27].
8. Let the columns of B be the eigenvectors corresponding to the significant
eigenvalues of Q in (1.70). Assume that noise sources are distributed over an
References
The location parameters are estimated directly without having to search for
peaks as in frequency-wavenumber spectrum (the approach described in chapter
4). In open space the direction of arrival (DOA), that is, azimuth or elevation or
both, is estimated using the subspace properties of the spatial covariance matrix
or spectral matrix. MUSIC is a well known algorithm where we define a
positive quantity which becomes infinity whenever the assumed parameter(s) is
equal to the true parameter. We shall call this quantity as a spectrum even
though it does not possess the units of power as in the true spectrum. The
MUSIC algorithm in its original form does involve scanning and searching,
often very fine scanning lest we may miss the peak. Later extensions of the
MUSIC, like root MUSIC, ESPRIT, etc. have overcome this limitation of the
original MUSIC algorithm. When a source is located in a bounded space, such
as a duct, the wavefront reaching an array of sensors is necessarily nonplanar
due to multipath propagation in the bounded space. In this case all three
position parameters can be measured by means of an array of sensors. But the
complexity of the problem of localization is such that a good prior knowledge
of the channel becomes mandatory for successful localization. In active
systems, since one has control over the source, it is possible to design
waveforms which possess the property that is best suited for localization; for
example, a binary phase shift key (BPSK) signal with its narrow
autocorrelation function is best suited for time delay estimation. Source
tracking of a moving source is another important extension of the source
localization problem.
λ m = α m + σ 2η (4.14a)
Equation (4.14b) implies that the space spanned by the columns of A , that is,
the direction vectors of incident wavefronts, is orthogonal to the space spanned
by the eigenvectors, v m , m = P, P + 1,..., M − 1, often known as noise
1
SMusic (ω,θ) = (5.3)
a (ω,θ)v ηv ηH a(ω,θ)
H
SMusic (ω,θ) , also known as the eigenvector spectrum, will show sharp peaks
whenever θ = θ 0 ,θ1 ,...,θ P −1 . The subscript Music stands for M ultiple
Signal Classification. This acronym was coined by Schmidt [1] who discovered
the subspace algorithm. At about the same time but independently Bienvenu
and Kopp [2] proposed a similar algorithm. Pisarenko [3] had previously
published a subspace based algorithm in the context of harmonic analysis of
time series. Note that, although we refer to SMusic (ω,θ) as spectrum, it does
not have the units of power; hence it is not a true spectrum. Let us express the
steering vector in terms of the product of frequency and time delay
d
τ= sin θ ,
c
a(ωτ) = col{1,e jωτ ,e j 2ωτ ,...,e j ( M −1)ωτ }
shall show that the signal subspace is the same as the space spanned by the
columns of the matrix A . Since A is a full rank matrix its polar
decomposition gives
A = TG (5.4)
where G is a full rank PxP matrix and T is a MxP matrix satisfying the
following property,
v mH AC0 A H v m = α m (5.5a)
or in matrix form
[ ]
v sH T GC0G H T H v s = diag{α m , m = 0,1,..., P − 1} (5.5b)
A = v s v sH A (5.6d)
1
SMusic (ω,θ) = (5.7)
a (ω,θ)(I − v s v sH )a(ω,θ)
H
Signal Eigenvalues and Source Power: The signal eigenvalues and source power
are related, although the relationship is rather involved except for the single
source case. Let us first consider a single source case. The covariance matrix is
given by C0 = a 0 s0 a 0 which we use in (5.5a) and obtain
H
α0
s0 = H 2 (5.8)
v a0
0
s0 0
C f = [a 0 ,a1 ] [ a 0 ,a1 ]
H
0 s1
[ ] [
diag{α 0 ,α1} A H v s ]
−1 −1
C0 = v sH A (5.9b)
Equation (5.9b) has been used to estimate the powers of two equal power
(snr=1.0) wavefronts incident on a 16 sensor ULA and the results are tabulated
in Table 5.1. The estimation is error-free right down to a half-a-degree
separation; however, this good performance deteriorates in the presence of
model and estimation errors.
Aliasing: We pointed out earlier (p.211) that the sensor spacing in a ULA must
be less than half the wavelength so that there is no aliasing in the frequency-
wavenumber spectrum. The same requirement exists in all high resolution
methods, namely the Capon spectrum, Maximum entropy spectrum and Music
spectrum. The basic reason for aliasing lies in the fact that the direction vector
is periodic. Consider anyone column of the matrix A ,
λp λp
Now let d= − δ p where 0 ≤ δ p ≤ , then a p (ωτ )
sin θ p 2sin θ p
becomes
d δ
We can find an angle θ̂ p such that sin θ̂ p = − p sin θ p and hence
λp λp
d d
a p (ω p sin θ p ) = a p (ω p sin θ̂ p )
c c
δ
θ̂ p = sin −1 p sin θ p (5.10)
d
d
As an example, consider an array with sensor spacing, = 1 and a wavefront
λp
incident at an angle θ p =60°. For this choice of array and wave parameters
δ p = 0.1547 and from (5.10) we get the angle where the aliased peak will be
located, θ̂ p = −7.6993°. The wave number spectrum computed by all four
methods is shown in fig. 5.1.
Aliasing is on account of periodicity of a direction vector which in
turn is caused by periodicity present in a ULA. Thus, to avoid aliasing, it
would be necessary to break this periodicity; for example, we may space the
sensors nonuniformly. In a circular array, though sensors are uniformly spaced
(e.g. UCA), the time delays are nonuniform; therefore a UCA will yield an
alias-free spectrum [4]. This is demonstrated in fig. 5.2 where we consider a
wavefront which is incident at 60o (with respect to x-axis) on a circular array
consisting of 16 sensors uniformly spread over a circle of radius 8λ . The
Capon spectrum is shown for this case. The aliasing phenomenon is not
encountered in random arrays where the sensors are spaced at random intervals.
But as shown in chapter 2 the random array possesses a highly nonlinear phase
response.
6
Aliased Actual
4
2
0
1.5
(b)
1
0.5
Spectrum
0
6000
5000 (c)
4000
3000
2000
1000
0
6
x 10
5
4 (d)
0
-100 -50 0 50 100
Angle in deg
Figure 5.1 The aliasing effect due to undersampling of the wavefield (d=λ). All four
methods of spectrum estimation have been used. (a) Bartlett spectrum, (b) Capon
spectrum, (c) Maximum entropy spectrum and (d) Music spectrum. While the actual
peak is at 60o the aliased peak appears at -7.69o.
S0 = s ρ s H (5.11)
where
s = diag { s0 , s1 ,..., sP −1 }
s0 , s1 ,..., sP −1 represent the power of P sources and ρ is the coherence matrix
whose (m,n)th element represents the normalized coherence between the mth
and nth sources. The signal eigenvalues of the spectral matrix for P=2 are given
by [5]
M
λ0 = (s0 + s1 ) + M s0 s1 Re{ρ12 }ψ +
2
{[ }
1
M
]
(s0 + s1 ) + 2 s0 s1 Re{ρ12 }ψ − 4s0 s1 (1 − ψ )(1 − ρ12 )
2 2 2 2
+ σ 2η
2
M
λ1 = (s0 + s1 ) + M s0 s1 Re{ρ12 }ψ −
2
{[ }
1
M
]
(s0 + s1 ) + 2 s0 s1 Re{ρ12 }ψ − 4s0 s1 (1 − ψ )(1 − ρ12 )
2 2 2 2
+ σ 2η
2
(5.12a)
where
d
M(sin θ 0 − sin θ1 ))
sin(π
ψ(M) = λ
d
M sin(π (sin θ 0 − sin θ1 ))
λ
(λ 0 − σ 2η ) + (λ1 − σ 2η ) =
(5.12b)
M(s0 + s1 ) + 2M s0 s1 Re{ρ12 }ψ(M)
0.5
0
-100 -50 0 50 100
Angle in deg.
Figure 5.2: No aliasing effect is seen with a circular array. Sixteen sensor UCA with
radius= 8λ (sensor spacing=3.12l) is used. A plane wavefront is incident on the
array at 60o .
1
M
λ 0 = (s0 + s1 ) +
2
M
2
{ 2
(s0 + s1 )2 − 4s0 s1 (1 − ψ (M) ) } 2
+ σ η2
(5.13)
1
M
λ 1 = (s0 + s1 ) −
2
M
2
{ 2
(s0 + s1 )2 − 4s0 s1 (1 − ψ (M) ) } 2
+σ 2
η
Also, note that when the sources are in the same direction λ 0 = M(s0 + s1 )
and λ1 = 0 .
The source spectral matrix may be modified by spatial smoothing of
the array outputs. This is achieved by averaging the spectral matrices of
subarray outputs over all possible subarrays (see fig. 5.3). The ith subarray
(size µ ) signal vector at a fixed temporal frequency is given by
{ }
Fi = col Fi (ω ), Fi +1 (ω ),..., Fi + µ −1 (ω ) 0 ≤ i ≤ M − µ +1
= I i, µ F
4th subarray
Figure 5.3: Overlapping subarrays are formed as shown above. Each subarray has
four sensors. It shares three sensors with its immediate neighbours.
M −µ +1 M −µ +1
1 1
S= ∑
M − µ + 1 i=0
S i,µ = ∑
M − µ + 1 i=0
H
I i,µ S I i,µ
(5.14a)
M −µ +1
1
= ∑
M − µ + 1 i=0
H
I i,µ A(ω)S0 (ω)A H I i,µ + σ 2ηI
where we have used the spectral matrix of the signal model plane waves in the
presence of white noise (4.12b). We can show that
M −µ +1
1
S =  ∑ φ i S0 (ω) φ Hi  H + σ 2ηI (5.15a)
M − µ + 1 i=0
where
The quantity inside the square brackets in (5.15a) may be computed by actual
multiplication followed by summation,
M −µ +1 d
1 − j 2 π i ( sin θ m −sin θ n )
[ ]mn = sm sn ρmn ∑
M − µ + 1 i=0
e c
(5.15b)
d
− jπ ( M −µ ) ( sin θ m −sin θ n )
= sm sn ρmn ψ(M − µ + 1)e c
From (5.15b) it follows that the coherence after spatial smoothing may be
written as
d
− jπ ( M −µ ) ( sin θ m −sin θ n )
ρmn = ρmn ψ(M − µ + 1)e c
(5.16)
[
µ (s0 + s1 ) + 2 s0 s1 Re{ρ̃01}ψ(µ)ψ(M − µ + 1)
]
2 2
+ ση
2
2 −4s s (1 − ψ(µ) )(1 − ρ ψ(M − µ + 1) )
2 2
01 01
µ
λ1 = (s0 + s1 ) + µ s0 s1 Re{ρ̃01}ψ(µ)ψ(M − µ + 1) −
2
1
[
µ (s0 + s1 ) + 2 s0 s1 Re{ρ̃01}ψ(µ)ψ(M − µ + 1)
] 2
2
+ ση
2
2 −4s s (1 − ψ(µ) 2 )(1 − ρ ψ(M − µ + 1) 2 )
01 01
(5.17)
d
jπ ( M − µ ) ( sin θ 0 −sin θ 1 )
where ρ̃01 = ρ01e
λ
. The sum of the signal eigenvalues is
given by
(λ 0 − σ 2η ) + (λ1 − σ 2η ) =
M(s0 + s1 ) + 2M s0 s1 Re{ρ̃01}ψ(µ)ψ(M − µ + 1)
d
π (M − µ + 1) (sin θ 0 − sin θ1 ) ≈ π
λ
then ψ(M − µ + 1) ≈ 0 , λ 0 = Ms0 and λ1 = Ms1 and the desired subarray size
would be
2
µ ≈ (M + 1) −
(sin θ 0 − sin θ1 )
1
Let (sin θ 0 − sin θ1 ) ≈ . Then the desired subarray size is
µ
1
µ ≈ (M + 1) (5.18)
3
2.2
2.0
2nd Eigenvalue
1.8
1.6
1.4
1.2
1.0
4 12 20 28 36 44 52
µ
Figure 5.4: The 2nd eigenvalue relative to noise variance is shown as a function of
the subarray size (µ).
λ1
This result is very close to that given in [7]. We have evaluated as a
σ 2η
function of the subarray size for two perfectly correlated equal power waves
1
incident at angles θ 0 and θ1 such that sin θ 0 − sin θ1 = on a ULA with
µ
64 sensors and array snr =10. The results are shown in fig. 5.4.
v HM −1a = 0 (5.19)
d
− j 2 π sin θ p
where zp = e λ
. Consider a polynomial
[ ]
T
v HM −1 1, z, z 2 ,..., z M −1 =0 (5.20)
whose roots are indeed z0 , z1 ,..., z P −1 . In the complex plane all roots lie on a
unit circle. The angular coordinate of a root is related to the DOA. For
example, let the pth root be located at ϕ p . Then ϕ p = πsin θ p . This method
was first suggested by Pisarenko [3], many years before MUSIC was invented !
Root Music: The Pisarenko’s method has been extended taking into account the
noise space spanned by two or more eigenvectors [8]. The extended version is
often known as root Music. Define a polynomial,
M −1
S(z) = ∑ v [1, z, z ,..., z ]
m= P
H
m
2 M −1 T
(5.21)
1
The roots of S(z) or D(z) = S(z)S( ) lying on the unit circle will
z
correspond to the DOAs of the sources and the remaining M-P roots will fall
inside the unit circle (and also at inverse complex conjugate positions outside
the circle).
In the minimum norm method a vector a is found which is a solution
of a S f a = min under the constraint that a a = 1 . It turns out that the
H H
Thus, the minimum norm method belongs to the same class of direct methods
of DOA estimation initiated by Pisarenko [3]. An example of DOA obtained by
computing zeros on the unit circle under the ideal condition of no errors in the
spectral matrix is shown in table 5.2. Two equal power signal wavefronts are
assumed to be incident on eight sensor ULA at angles 10o and 14o . The DOA
estimates from the error-free spectral matrix are exact (columns one and two in
table 5.2), but in the presence of even a small error in the estimate of the
spectral matrix, a significant error in DOA estimation may be encountered (see
columns three and four in the table).
The zeros of the polynomial defined in any one of the above methods
are important from the point of DOA estimation. In particular, the zeros which
fall on the unit circle or close to it represent the DOAs of the incident
wavefronts. These zeros are often called signal zeros and the remaining zeros,
located deep inside the unit circle, are called the noise zeros. In MUSIC, a peak
in the spectrum represents the DOA of the wavefront. But the height of the
peak is greatly influenced by the position of the signal zero. The peak is
infinite when the zero is right on the unit circle but it rapidly diminishes when
the zero moves away from the unit circle. The peak may be completely lost,
particularly in the presence of another signal zero in the neighborhood but
closer to the unit circle. Hence, it is possible that, while the spectral peaks
remain unresolved, the signal zeros are well separated. The shift of the signal
zero may be caused by errors in the estimation of the spectral matrix from finite
data. Let a signal zero at zi be displaced to ẑi . The displacement, both radial
and angular, is given by
where we have assumed that δ u << 1 . Perturbation analysis reported in [9] for
time series shows that the mean square error in δ r and δ u are related
{ }
E δ ri
2
= 2N
cos( θ i
)
2N
≤ 2
E{ δθ }
(5.22)
2 ωd π
c
i
where N stand for the number of time samples (or snapshots). From (5.22) it
follows that E δri { } >> E{ δθ } , particularly when the wavefronts are
2
i
2
I 0 σ η1
H 2
A A
S f̃ = S f + 2 (2.57)
AΓ AΓ 0 I σ η2
S f 1 f 1 S f 1 f 2 AS f A AS f Γ H A H
H
S f̃ = =
S f 2 f 1 S f 2 f 2 AΓS f A
H
AS f A H
Note that S f 1 f 1 is the spectral matrix of the upper sensor outputs and S f 2 f 2 is
the spectral matrix of the lower sensor outputs. S f 1 f 2 is the cross-spectral
matrix between the upper sensor and lower sensor outputs. From the
S0f 1 f 1 v = γS f 1 f 2 v (5.23)
[ ] [
v H S0f 1 f 1 − γS f 1 f 2 v = v H AS f A H − γAS f Γ H A H v ]
[ ]
= v H AS f I − γΓ H A H v
(5.24)
=v H
AS [ I − γΓ ]A
f
H H
v
= v H Qv = 0
{ }
−j ∆.δ i
pencil S0f 1 f 1 , S f 1 f 2 are γ i = e c , i = 0,1,..., P − 1 from which we can
estimate θi ,i = 0,1,..., P − 1 . This method of DOA estimation is known as
ESPRIT (Estimation of signal parameters via rotation invariance technique)
[10]. An example of DOA obtained via ESPRIT is in table 5.3. The
eigenvalues of the pencil matrix (5.23) under the ideal condition of no errors in
the spectral matrix as well as with errors in the spectral matrix are shown in
table 5.3. Two equal power signal wavefronts are assumed to be incident on
eight sensor ULA at angles 10o and 14o. The eigenvalues estimated from error
free spectral matrix are exact (columns one and two in table 5.3), but in the
presence of a small error in the estimation of the spectral matrix, a significant
error in DOA estimation may be encountered (see columns three and four in the
table).
The generalized eigenvector v i , corresponding to eigenvalue γ i ,
possesses an interesting property. Let us write (5.24) in expanded form
where Ŝ f represents the source matrix where the ith column and the ith row are
ω0
−j ∆.δ k
deleted. Since γ i ≠ e c , i ≠ k , the diagonal matrix will be full rank and
also Ŝ f is full rank by assumption, we must have v iH a k, = 0 for all k ≠ i .
{ 0
Thus, the generalized signal eigenvector of the pencil matrix S f 1 f 1 , S f 1 f 2 is }
orthogonal to all direction vectors except the ith direction vector. We shall
exploit this property for signal separation and estimation later in chapter 6.
L −1
where ã = ∑ α l e − jω c δt l a(θ + δθl ) where α l , δtl and δθl are for the lt h
l=0
ray, complex amplitude, time delay and direction of arrival with respect to a
direct ray from the main source, respectively. Assume that there are L rays
reaching the sensor array. For the direct ray we shall assume that α 0 = 1,
δt0 = 0 and δθ 0 = 0 . The covariance matrix of the array (ULA) output is
given by (see eq. 1.70)
where we have assumed that the scatterers are uniformly distributed in the
angular range ±∆ . The Q matrix for the assumed uniform distribution is
given by
d
sin 2π (m − n)∆ cosθ
{Q}mn = λ
d
2π (m − n)∆ cosθ
λ
and
j 2 λπd sin θ j
2 π( M −1)d
sin θ
D = diag 1, e ,..., e λ
[ ]
v ηH c f − σ 2ηI v η = 0 (a matrix M-r x M-r zeros) (5.27)
r −1
σ 2s Lσ α2 ∑ λ m v ηH D(θ)e m e mH D H (θ)v η = 0 (5.28)
m=0
r −1
where we have used eigendecomposition of Q≈ ∑λ
m=0
e e . Here λ m
H
m m m
2
v ηH D(θ)e m = 0 (5.29)
for all m<r. The azimuth of the center of the cluster can be estimated using the
orthogonality property demonstrated in (5.29). The eigenvectors e m , discrete
prolate spheroidal sequence (DPSS), are obtained by eigendecomposition of the
Q matrix. But in Q there is an unknown parameter pertaining to the width of
the cluster, namely, ∆ cos θ which has to be estimated. We have in chapter 1
(page 59) indicated that the rank of the covariance matrix (noise free) is closely
related to this unknown parameter.
5.1.5 Adaptive Subspace: Adaptive methods for the estimation of the
eigenvector corresponding to the minimum (or maximum) eigenvalue have
been described by many researchers[12-14]. The method is based on inverse
power iteration [15]
v̂ 0 = [1,0,...,0]
T
ṽ k +1 = S−1
f (ω)v̂
k
(5.30)
k +1 ṽ k +1
v̂ =
ṽ k +1 ṽ k +1
T
k
where v̂ is an eigenvector at kth iteration and the last equation is meant for
normalization. To estimate the eigenvector corresponding to the largest
till time N , that is, the eigenvectors of the spectral matrix, Ŝ Nf (ω) . When a
new sample arrives the eigenvectors are updated using the following recursive
equation followed by Gram-Schmidt orthonormalization (GSO),
Ŝ Nf +1v̂ mN = g mN +1
(5.31)
v̂ mN +1 = GSO{g mN +1}, m = 1,2,... P
The choice of initial eigenvectors to start the recursion is important for rapid
convergence. A suggested choice in [14] is to orthonormalize P initial
frequency snapshots and use them as the initial eigenvectors. The starting value
of the spectral matrix is chosen as
Ŝ Pf = (I + F 0 F 0H )
I k ⊥F r for r < k
2
I k = 1 for all k (5.32)
I 3 = γ 12 F1 + γ 22 F 2 + γ 32 F3
... (5.33)
I P = γ 1P F1 + γ 2P F 2 +...+γ PP F P
T(ω )S(ω )T H (ω ) =
(5.34)
T(ω )a(ω , θ )S0 (ω )a H (ω , θ )T H (ω ) + σ η2 T(ω )T H (ω )
where the transformation matrix T(ω ) must possess the following properties:
where bw stands for the signal bandwidth. In (5.36) we observe that the signal
power spread over a band of frequencies has been focused at one frequency,
namely, ω 0 . This idea of focusing of energy has been actively pursued in [16,
17] for DOA estimation.
In the direction vector there is a parameter, namely, sensor spacing
ωd
which may be changed according to the frequency such that remains
c
constant. This will require the sensor spacing at frequency ω i should be equal
ω 0 d0
to di = , i ∈bw . In practice the required change in the physical
ωi
separation is difficult to achieve, but resampling of the wavefield through
interpolation is possible. This approach to focusing was suggested in [18]. In
chapter 2 we introduced the spatio temporal covariance matrix (STCM) which
contains all spatio temporal information present in a wavefield. It has been
extensively used for source localization [19]. We shall probe into the signal and
noise subspace structure of STCM of ULA as well as UCA and show how the
structure can be exploited for source localization.
5.2.1 Wideband Focusing: Consider P uncorrelated point sources in a far field
and a ULA for DOA estimation. As in (4.12b) the spectral matrix of an array
signal may be expressed as S f (ω) = A(ω, θ)S 0 (ω)A H (ω, θ) where the
columns of A(ω, θ) matrix are the direction vectors. We seek a transformation
matrix T(ω) which will map the direction vectors at frequency ω into
direction vectors at the preselected frequency ω 0 as in (5.35). There is no
unique solution but a least squares solution is possible,
[ ]
−1
T(ω) ≈ A(ω 0 ,θ) A H (ω,θ)A(ω,θ) A H (ω,θ) (5.37)
The transformation matrix given by (5.37) depends upon the unknown azimuth
information. However, it is claimed in [20] that approximate estimates obtained
through beamformation or any other simple approach are adequate. As an
example, consider a single source case. For this, A(ω,θ) =
Clearly, using (5.38a) we can also achieve the desired transformation. In (5.38a)
let ω 0 =0; the transformation matrix becomes
which we shall use as a filter on the array output. The filtered output is given
by
∞
1
f 1 (t) = ∫
2π − ∞
T(ω)F(ω)e jωt dω
(5.39)
= col f (t), f (t − sin θ),..., f (t − (M − 1)sin θ)
d d
c c
to zero frequency. When there is more than one source we should compute a
series of steered covariance matrices over a range of azimuth angles. Whenever a
steering angle matches with one of the directions of arrival the steered
covariance matrix will display a strong ‘dc’ term, equal to the power of the
source.
The main drawback of wideband focusing is the requirement that the
DOAs of incident wavefronts must be known, at least approximately, but the
resulting estimate is likely to have large bias and variance [20]. To overcome
this drawback an alternate approach has been suggested in [22]. Let A and B
be two square matrices and consider a two sided transformation, TBT H , which
is closest to A . It is shown in [22] that this can be achieved if T = v A v B
H
T(ω 0 ,ω i ) = v s (ω 0 )v sH (ω i ) (5.40)
T(ω 0 ,ω i )S f (ω i )T H (ω 0 ,ω i )
= v s (ω 0 )v sH (ω i )S f (ω i )v s (ω i )v sH (ω 0 )
= v s (ω 0 )λ(ω i )v sH (ω 0 )
where
are the eigenvalues of the noise free spectral matrix, S f (ω i ) . Next, we average
all transformed matrices and show that
= v s (ω 0 )λv sH (ω 0 )
where S0 (ω ) and S1 (ω ) are spectra of the first and the second source
respectively. We compute the spectral matrix at two different frequencies, ω 0
and ω1 , with different sensor spacings, d0 and d1 , where
ω0
d1 = d0
ω1
and form a sum.
S f (ω 0 ) + S f (ω1 ) =
a(ω d0 sin θ )S (ω )a H (ω d0 sin θ )
0
c
0 0 0 0
c
0
+ σ ηI
2
d d
+a(ω 0 0 sin θ1 )S1 (ω 0 )a H (ω 0 0 sin θ1 )
c0 c
S f (ω 0 ) + S f (ω1 ) =
a(ω d0 sin θ )S (ω )a H (ω d0 sin θ )
0
c
0 0 0 0
c
0
+ σ ηI
2
d d
+a(ω 0 0 sin θ1 )S1 (ω 0 )a (ω 0 0 sin θ1 )
H
c0 c
a(ω d1 sin θ )S (ω )a H (ω d1 sin θ )
1
c
0 0 1 1
c
0
+ + σ ηI
2
(5.42)
+a(ω1 d1 sin θ1 )S1 (ω1 )a H (ω1 d1 sin θ1 )
c c
d d
= a(ω 0 0 sin θ 0 )[ S0 (ω 0 ) + S0 (ω i )]a H (ω 0 0 sin θ 0 )
c c
d d
+a(ω 0 0 sin θ1 )[ S1 (ω 0 ) + S1 (ω i )]a H (ω 0 0 sin θ1 ) + 2σ 2ηI
c c
Let us assume that N virtual arrays, each with different sensor spacing, have
been created by a resampling process. We can generalize (5.42),
1 N −1 d0 1 N −1 d
∑
N i=0
S f (ω i ) = a(ω 0
c
sin θ 0 ) ∑
N i=0
S0 (ω i ) a H (ω 0 0 sin θ 0 )
c
d0 1 N −1 d
+ a(ω 0 sin θ1 ) ∑ S1 (ω i ) a H (ω 0 0 sin θ1 ) + σ 2ηI
c N i=0 c
(5.43)
(x − md0 )
sin ω max
h(m) = c , m = 0, ±1,..., ±∞ (5.44)
(x − md0 )
ω max
c
CSTCM =
∞
In this subsection we shall show how the STCM can be used for the direction
of arrival estimation of a broadband source. For this it is necessary that the
eigenstructure, in particular, the rank of STCM, will have to be ascertained.
This has been done for a ULA in [23] and for a UCA in [24, 4]. We shall
assume that the source spectrum is a smooth function and that it may be
approximated by a piecewise constant function
where
1 1 1
rect(ω − ∆ωl) = for ∆ω(l − ) ≤ ω ≤ ∆ω(l + )
∆ω 2 2
=0 otherwise
and S0l (S0,l = S0,−l ) is the average spectrum in the lth frequency bin, ∆ωl .
Using (5.45) in (2.67) we obtain
1
∆ω(l − )
l=L 2
1 1
CSTCM = ∑ S0l
2π l = − L ∆ω ∫ h(ω,ϕ 0 )h H (ω,ϕ 0 )dω (5.46)
1
− ∆ω(l − )
2
1
∆ω(l − )
2
1
2π∆ω ∫ h(ω,ϕ
1
0 )h H (ω,ϕ 0 )dω = g l Qg lH
− ∆ω(l − )
2
where
gl =
e − jωτ 0 ,e − jωτ1 ,...,e − jωτ M −1 ;e − jω( ∆t + τ 0 ) ,e − jω( ∆t + τ1 ) ,...,
diag e − jω( ∆t + τ M −1 ) ;...;e − jω(( N −1)∆t + τ 0 ) ,e − jω(( N −1)∆t + τ1 ) ,...,
− jω(( N −1)∆t + τ M −1 )
e ω = ∆ωl
and
∆ω
[Q]α,α' = sin c ((n − n' )∆t + τ m − τ m' )
2
l=L
CSTCM = ∑S
l=−L
0l g l Qg lH (5.47)
The rank of CSTCM is determined by the rank of matrix Q . It follows from the
results in [11, 24] that 99.99% energy is contained in the first
(( N − 1)∆t + 2a ) ∆ω + 1 eigenvalues of Q where [x] denotes the next
c 2π
integer greater than x. A comparison of the theoretical and numerically
determined rank of CSTCM is given in fig. 5.5 for a circular array. For a linear
array the corresponding number is given by
(( N − 1)∆t + − ϕ ∆ω
+ 1 [19]. Note that the dimension of the
(M 1)d cos
)
c 2π
signal subspace is approximately equal to the time bandwidth product.
When the size of the observation space, that is, dimensions of
2a ∆ω
CSTCM , is larger than the rank of CSTCM , NM > (( N − 1)∆t + ) + 1 ,
c 2π
there exists a null subspace, a subspace of the observation space, of dimension
equal to
2a ∆ω
Dim{v null } = NM − ((N − 1)∆t + ) + 1
c 2π
60
50
Rank
Exp
40
Theory
30
20
0 10 20 30 40 50 60 70
Radius / λ min
Figure 5.5: The effective rank of signal only STCM as a function of the radius of a
circular array. The drop in the rank at radius 30λ min is due to a transition from
smooth spectrum to line spectrum. We have considered a sixteen sensor equispaced
circular array with other parameters N=4, B=0.8 (normalized bandwidth) and one
source.(From [4]. c 1994, With permission from Elsevier Science)
∞
1
2π −∫∞
v iH CSTCM v i = v iH h(ω,ϕ 0 )h H (ω,ϕ 0 )v i S f 0 (ω)dω
(5.48)
∞
1 2
= ∫
2π − ∞
v iH h(ω,ϕ 0 ) S f 0 (ω)dω ≈ 0
Assume that S f 0 (ω) > 0 over some frequency band. For (5.48) to hold good
we must have in that frequency band
2
v iH h(ω,ϕ 0 ) ≈ 0 i ∈null space (5.49)
Equation (5.49) is the basic result used for estimation of ϕ 0 . For the purpose
of locating the null we can define a parametric spectrum as in narrowband
MUSIC.
M− p N
1 M
∑ i
λ̂
M − p i = p+1
Γ(λ̂ p+1 , λ̂ p+ 2 ,..., λ̂ M ) = ln M (5.50a)
∏ i λ̂
i = p+1
where N stands for number of snapshots. The log likelihood ratio is then
compared with a threshold γ. Whenever
the test is said to have failed. Modified forms of the sphericity tests have been
suggested in References [27, 28].
The choice of γ is subjective. Alternate approaches which do not
require a subjective judgment have been proposed [29]. The number of
wavefronts is given by that value of p in the range 0 to M − 1 where, either
or
where pmin stands for the position of the minimum of (5.50) and HP denotes
the hypothesis that the true number of signal sources is P. The probability of
error for the MDL criterion monotonically goes to zero for a large number of
snapshots or high snr. In contrast, the probability of error for the AIC criterion
tends to a small finite value. However, for a small number of snapshots or low
snr the performance of AIC is better than that of MDL.
(a)
800
600
Likelihood Ratio
400
200
0
0 2 4 6 8 10 12 14
p
1000
(b)
800
Measure
600
400
200
0 2 4 6 8 10 12 14
p
Q−1
f (t) = ∑ α k e − j (ω 0 + k∆ω )t (5.51)
k =0
where α k is the complex amplitude of kth sinusoid. The output of the ith
sensor is given by f i (t) = f (t + ∆τ i ) + ηi (t) , where ∆τ i is time delay at
the ith sensor with respect to a fictitious sensor at the center of the array. Each
sensor output is tapped at N time instants and arranged in a vector form,
The output data vector of the ith sensor may be written in a compact form
shown below:
where
H = [h 0 ,h1 ,...hQ−1 ]
h k = col{1,e jω k ,...e j ( N −1)ω k }
ω k = ω 0 + k∆ ω
jω Q−1 ∆τ i
A i = diag{e jω 0 ∆τ i ,e jω 1 ∆τ i ,...e }
ε = col[α 0 ,α1 ,...α Q−1 ]
Next, we shall stack up all data vectors into one large vector F of dimension
MNx1. It may be expressed as
F = TDE + η (5.53)
where
T = diag{H,H,...H} ,(NMxMQ),
D = diag{a(r0 )A 0 , a(r1 )A1 ,...a(r M −1 )A M −1} , (MQxMQ) and
T
E = [εT, ε T, ε T, ... ] , (MQx1)
ri r r
0
S ϕ
S0
i
(x i , y i )
Figure 5.7: A circular array of sensors and a source inside the circle. The range (r) and
azimuth (ϕ) of the source and all time delays are measured with respect to the center
of the array as shown.
c f = TDΓ 0 D H T H + σ 2ηI
{ }
where Γ 0 = E EE H = 1 ⊗ Γ where 1 is a square matrix of size MxM whose
elements are all equal to 1, Γ = diag{γ 0 , γ 1 ,... γ Q−1} and γ 0 , γ 1 ,...γ Q−1 are
powers of the random sinusoids. Symbol ⊗ stands for Kronecker product. We
will assume hereafter that the noise variance σ 2η is known or has been
estimated from the array output when there is no signal transmission or by
averaging the noise eigenvalues of the covariance matrix and that it has been
subtracted from the covariance matrix.
Let us consider the structure of the mth column of the STCM. By
straightforward multiplication it can be shown that the mth column of the
covariance matrix is given by
[ ]
T T T T
c m = c 0m ,c1m ,...,c Q−1
m (5.54)
jω Q−1 µ i
A i ΓA 0H = diag{γ 0 e jω 0 µ i , γ 1e jω 1µ i ,...γ Q−1e } (5.55)
{ }
jω Q−1 (µ i − m)
(5.57)
Hcol γ 0 e jω 0 (µ i − m) , γ 1e jω 1 (µ i − m) ,...γ Q−1e
{
H # Cim = a(ri )a(r0 )col γ 0 e jω 0 (µ i − m) , γ 1e jω 1 (µ i − m) ,...γ Q−1e
jω Q−1 (µ i − m)
}
We now define a matrix D whose columns are H # Cim , i=0, 1, 2, ... M-1,
a(ri )
and bi = , i=0,1,...,M-1. Since the location information is present in
a(r0 )
µ 0 ,µ1 ,...µ M −1 , our aim naturally is to estimate these from the columns of
G. Each column may be considered as a complex sinusoid whose frequency can
be estimated, provided ∆ωµ i ≤ π . For a boundary array, specifically a
circular array, this limitation can be overcome. In chapter 2 we have given an
algorithm to estimate time delays or the location of the source from the phase
estimates.
As a numerical example we have considered two sources each emitting
eight random tones. The frequencies emitted by the first source are (0, 200,
400, 600, 800, 1000, 1200, 1400 Hz) and those by the second are (1600, 1800,
2000, 2200, 2400, 2600, 2800, 3000 Hz). An eight sensor UCA of radius 100
meters surrounds both sources. Sixteen delayed snapshots (taps) were used
making the size of STCM as 128x128. The results are displayed in table 5.4.
5.3.2 BPSK Signal: Binary phase shift keying signal consists of square pulses
of fixed width with amplitude given by a Bernoulli random variable, an
outcome of a coin tossing experiment where head equals +1 and tail equals -1.
A typical BPSK sequence is shown in fig. 5.8. The analytic representation of
BPSK signal is given by
L −1
c0 (t) = ∑ c0,n hT c (t − nTc ) (5.59a)
n=0
-1
0 127 255
Time units
Figure 5.8: BPSK waveform. A train of rectangular pulses with amplitude alternating
between +1 and -1.
Q
xm (t) = ∑ pk am (θ k )∑ bk,l ck (t − lT s − τ k ) + ηm (t)
k =0 l (5.59b)
m = 0,1,2,... M − 1
where
am (θ k ) : response of mth sensor to a signal coming from kth user
There are two different approaches to DOA estimation with BPSK coded signal.
In the first approach the usual covariance matrix is computed and in the second
approach, due to the fact that the code used by the user of interest is known, the
received signal is first cross-correlated with that code. The output of the cross-
correlator is then used to compute the covariance matrix. Both approaches yield
similar results; though in the second approach the interference from the users of
no interest is reduced by a factor proportional to the code length.
Precorrelation Covariance Matrix: The outputs of the sensor array, after
removing the carrier, are arranged in a vector form,
Q
f(t) = ∑ pk a(θ k )∑ bk,l ck (t − lT s − τ k ) + η(t) (5.60a)
k =0 l
where f(t) , a(θ k ) , and η(t) are Mx1 vectors. The precorrelation covariance
matrix is given by
2 Q Q
η ∑ ∑ pk pk ′ a(θ k )a(θ k ′ )
σ I +
k =0 k′=0
cf = (5.60c)
∑
l
∑l bk,l ck (t − lTs − τ k )bk ′,l ′ck ′ (t − l ′Ts − τ k ′ )
′
We shall assume that the data bits coming from different users are independent
and codes assigned to different users are uncorrelated. As a result, (5.60c)
reduces to
But the quantity inside the inner curly brackets is always equal to 1; hence we
obtain
Q 2
c f = ∑ pk a(θ k )a(θ k ) + σ 2ηI
k = 0
[ ]
c f = a(θ 0 ),a(θ1 ),...,a(θQ−1 ) diag p02 , p12 ,..., pQ−1
2
{ }
[a(θ ),a(θ ),...,a(θ )] + σ I
H 2
0 1 Q−1 η (5.60d)
= A diag{ p , p ,..., p }A + σ I
2
0
2
1
2
Q−1
H 2
η
(l +1)T s
1
g 0 (l) =
Ts ∫ f(t)c (t − lT )dt
lT s
0 s (5.61)
where f(t) stands for array signal vector without carrier. Using the signal
model given by (5.60a), eq. (5.61) may be written as
g 0 (l) =
(l +1)T s ∞
1
p0 a(θ 0 ) ∫ [ ∑ b0 j c0 (t − jT s )]c0 (t − lT s )dt
Ts lT s j =−∞
(l +1)T s ∞
1 Q−1
+ ∑ pk a(θk )
T s k =1 ∫ [∑ bkj ck (t − jT s + τ k )]c0 (t − lT s )dt
−∞
(5.62)
lT s
(l +1)T s
1
+
Ts ∫ η(t)c (t − lT )dt
lT s
0 s
τk τk
1 Q−1
∑
T s k =1
pk a(θ k ) bkl −1 ∫ ck (t − T s − τ k )c0 (t)dt + bkl ∫ ck (t − τ k )c0 (t)dt
0 0
T
1 s
T s ∫0
ηl = η(t + lT s )c0 (t)dt
1 Q 2
E{b b } = 2 ∑ pk a(θ k )a(θ k ) H ρ1
1 1H
l l (5.64b)
T s k =1
where
{ τk
ρ1 = E [ ∫ ck (t + T s − τ k )c0 (t)dt]2
0 }
and the sources are assumed to be uncorrelated. Using the representation of
BPSK signal (5.59a) we obtain
ρ1 =
τ k L −1 L −1
2
τ k hT c (t − m0 Tc )
2
1 T s L −1 L −1
ρ1 = ∫ ∑ ∑ ∫0 dτ k (5.65a)
Ts hT c (t + T s − mk Tc − τ k )dt
m0 mk
0
where τ k , the arrival time of the signal from the kth user, is assumed to be
uniformly distributed over 0 to T s , the symbol duration. The integral over τ k
may be expressed as a sum over L integrals, one for each chip. When hT c is a
rectangular function the integral with respect to τ k in (5.65a) can be evaluated
in a closed form. After some algebraic manipulations we obtain
L2 3
ρ1 = Tc (5.65b)
3T s
1 Q 2
∑
H
E{b1l b1l } = pk a(θ k )a(θ k ) H (5.66a)
3L k =1
2 2H
Evaluation of E{bl bl } proceeds on the same lines as above. In fact the
result is identical, that is,
1 Q 2
∑
H H
E{bl2 bl2 } = pk a(θ k )a(θ k ) H = E{b1l b1l } (5.66b)
3L k =1
1 Ts Ts
E{ηl ηlH } = ∫ ∫ η(t + lT )η (t + lT )c (t )c (t )dt dt
H
E 1 s 2 s 0 1 0 2 1 2
T s2 0 0
L −1 E {η(t + lT )η (t + lT )} H
∑∫ ∫
1 Tc Tc 1 c 2 c
=
T s2 m1 = 0
0 0 hT c (t1 − m1T c )hT c (t2 − m1T c )dt1dt2
σ 2η
= I (5.67)
L
2 Q 2 σ 2
c g0 g0 = p a(θ 0 )a (θ 0 ) +
2
0
H
∑
3L k =1
pk a(θ k )a H (θ k ) + η I
L
(5.68)
[ ]
c g0 g0 = a(θ 0 ),a(θ1 ),...,a(θQ−1 ) diag p02 ,
3L
2 2
p1 ,...,
2 2
3L
pQ−1
σ 2η
[a(θ ),a(θ ),...,a(θ )]
H
0 1 Q−1 + I
L
2 2 H ση
2
2 2
= Adiag p0 , 2
p1 ,..., pQ−1 A + I (5.69)
3L 3L L
τ τ
E f − (t + )f +H (t − )
2
1
c αf (τ) = ∑
T t=− T 2 2
2
T
τ τ
E f(t + )f H (t − )e − j 2 παt
2
1
= ∑
T t=− T 2 2
2
T→∞
Let us show how to evaluate the (k,l)th element of the matrix c αf (τ) .
[c (τ)]
α
f kl
T
1 ∗
2 d τ d τ − j 2 παt
= ∑ E f 0 (t + k sin θ 0 − ) f 0 (t + l sin θ 0 + )e
T t=−
T c 2 c 2
2
T→∞
d
d jπα(k +l ) sin θ 0
α
= c (τ + (k − l) sin θ 0 )e
f0
c
c
(5.71a)
[ ]
jω c ( k −l ) sin θ 0 jπα(k +l ) sin θ 0
c αf (τ) ≈ c αf 0 (τ)e c
e c
, which may be further
kl
[c (τ)]
jω c (k −l ) sin θ 0
α α
approximated as f ≈ c (τ)e
f0
c
for ω c >> 2πα . Using
kl
the narrowband approximation and the assumption ω c >> 2πα we can
express the cyclic covariance matrix as
The above relation is quite similar to (4.12b), which was the starting point in
the subspace algorithm, for example, in MUSIC. Naturally, based on (5.71b), a
subspace algorithm known as Cyclic MUSIC has been proposed in [37].
Although, in deriving (5.71b) we have assumed a single source, it holds good
even in the presence of multiple signals with different cyclic frequencies and
any type of stationary noise.
Let us consider the diagonal terms of the cyclic covariance matrix.
From (5.71a) the diagonal terms, k=l, are given by
[ ]
j 2 παl sin θ 0
c αf (τ) = c αf 0 (τ)e c
(5.72a)
k =l
{[
c̃ αf (τ) = col c αf (τ) ]k =l = 0
[
, c αf (τ) ] k =l =1
[
,..., c αf (τ) ]
k =l = M −1 }
j 2 πα dc sin θ 0 j 2 πα( M −1) sin θ 0
d
a(α,θ 0 ) = col 1,e ,...,e c
and for P uncorrelated sources, but with the same cyclic frequency, we obtain
c αf 0 (τ)
c̃ αf (τ) = {a(α,θ 0 ),a(α,θ1 ),...,a(α,θ P −1 )}... (5.72c)
c α (τ)
f P−1
Note that in deriving (5.72) we have not used the narrowband approximation
which was earlier used in deriving (5.71b). The computer simulation results
reported in [38] showed improved results obtained by using (5.72c) over (5.71b)
τ τ
c αfg (τ) = ∑ E f ∗ (t − )g(t + )e − j 2 παt
1 2
T t=− T 2 2 (5.73)
2
T→∞
vanishes for all τ and α is equal to the cyclic frequency either of f(t) or g(t).
To demonstrate this property let us consider f (t) = s(t)cos(ω1t + θ1 ) and
g(t) = s(t)cos(ω 2 t + θ 2 ) and substituting in (5.73) we obtain
c αfg (τ)=
T cos((ω + ω )t − (ω − ω ) τ + θ + θ )
1 2 1 2 1 2
2
1 2 − j 2 παt
cs (τ) ∑ e
2T t = − T + cos((ω1 − ω 2 )t − (ω1 + ω 2 ) τ + θ1 − θ 2 )
2 2
T→∞
0.20 (a)
Cyclic cross-correlation
0.15
0.10
0.05
0.00
0.0 0.1 0.2 0.3 0.4
Frequency in Hz
0.25
(b)
0.20
Cyclic cross-correlation
0.15
0.10
0.05
0.00
0.150 0.151 0.152
Frequency in Hz
Figure 5.9: Cyclic cross-correlation function at zero lag as a function of (a) cyclic
frequency and (b) carrier frequency in Hz.
d (M-1)d
- jω sin θ 0 - jω sin θ 0
α φ(θ 0 ) = diag{α 0 ,α1e c
,...α M −1e c
}
d (M-1)d
- jω sin θ 0 - jω sin θ 0
= diag{1,e c
,...e c
}col{α 0 , ...α M −1}
= φ d (θ 0 )α c
α c = φ dH (θ 0 )v s v sH φ d (θ 0 )α c (5.75a)
P −1
Q = ∑ φ dH (θ p )v s v sH φ d (θ p ) (5.75b)
p= 0
Note that since the source is known, its azimuth and elevation are known a
priori. The unknowns are γ m ,ε m , m = 0,1,..., M − 1 . (Note
ξ m , m = 0,1,..., M − 1 are dependent on γ m ,ε m .) To solve for a pair of
unknowns we would need one more source, say, with different azimuth and
elevation, (θ1 ,ϕ1 ) . Once again the largest eigenvector may be related to the
direction vector of the second source.
We shall, for the sake of simplicity, assume that the deformed array is in the x-
y plane and also place the calibrating source in the x-y plane. Then, a single
source is enough for estimation of γ m and ε m which now take a form
m
γ m = ∑ cosα i , m = 0,1,..., M − 1
i=0
(5.78)
m
ε m = ∑ sin α i , m = 0,1,..., M − 1
i=0
λc
2π
[
∠{v 0 }m − ∠{v 0 }m −1 + λ c n = ] (5.79)
d cosα m cosϕ 0 + d sin α m sin ϕ 0 = ∆xm cosϕ 0 + ∆ym sin ϕ 0
where n is an unknown integer and ∆xm and ∆ym are x,y coordinates of mth
sensor relative to m-1st sensor. The ambiguity, arising out of the unknown
integer, may be resolved through geometrical considerations. Note that the mth
sensor must be on a circle of radius d centered at
n=2
d
m-1st p2
n=1
n=0
Figure 5.10 The ambiguity in (5.77) is resolved by requiring that the line it
represents must intersect the circle of radius d drawn at m-1st sensor. There are two
intersection points. The sensor may be at any one of the two intersections.
m-1 st sensor. Further the position of the sensor must satisfy (5.79), which,
incidentally, is an equation of a straight line [41]. For illustration let the line
corresponding to n=1 intersect the circle at two points p1 and p2 (see fig. 5.10).
The sensor can either be at p1 or p2. This ambiguity is resolved by choosing a
point which results in minimum array distortion [42].
(a)
Array
Source
(b)
Array
Source
(c)
5.5.1 Single Reflecting Surface: Assume that the source emits a stationary
stochastic signal. A vertical sensor array is used to receive the signal radiated by
the source. The array output in frequency domain may be written in terms of
the radiated signal as follows:
τ1 is the delay of the reflected signal relative to the direct arrival and w1 stands
for reflection coefficient. The spectral matrix of the array output is easily
derived from (5.80)
[
S f (ω) = [a 0 , a1 ] 1, w1e − jωτ1 ] [1, w e ][a , a ]
T − jωτ 1 H
1 0 1 S0 (ω) + σ 2ηI
(5.81)
[
ã = [a 0 , a1 ] 1, w1e − jωτ1 ]
T
Define = a 0 + w1e − jωτ1 a1 and rewrite (5.81) as
0
150
1200
100 1100
He 1000
igh 50 900 ge
t Ran
Figure 5.12: Parametric spectrum for a source situated above a reflecting surface
(100 λ ) and 1000 λ away from an eight element vertical ULA. The array center is at
1 6 λ above the reflecting surface.
P −1
αp
f m (t) = ∑ f 0 (t − τ pm ) + ηm (t) (5.83)
p= 0 Rp
(b)
Figure 5.13: (a) Parametric spectrum at 1600 Hz is shown as a function of range and
depth. A broadband acoustic source is assumed at 2000m (range) and 25m (depth) in
a channel of depth 100m. Twenty frequency snapshots were used to compute the
spectral matrix. Next, the parametric spectra computed at 50 frequencies, equispaced
in the band 1000Hz to 2000Hz, were averaged. The averaged spectrum is shown in
(b). (From [47]. c 1999, With permission from Elsevier Science)
distance to the pth image and α p is the reflection coefficient for the ray arriving
from the pth image. ηm (t) is noise received by the mth sensor. It is easy to
show that
d
τ pm = τ p + (m − 1) sin θ p (5.84)
c
P −1 d
j ( m −1)2 π sin θ p
dFm (ω) = ∑ w p dF0 (ω)e λ
+ dΝ m (ω) (5.85)
p= 0
α p jωτ p
where wp = e [44]. Let us express (5.85) in a compact matrix
Rp
notation by defining
1 ... 1
jω d sin θ d
jω sin θ P−1
e c 0 ... e c
A=
...
jω d ( M −1) sin θ d
jω ( M −1) sin θ P−1
e c 0
c
... e
§5.6 Exercises:
1. The line represented by (5.79) intersects the circle at two points p1 and p2
(see fig. 5.10). Let two adjacent sensors be on the x-axis. Show that one of the
points will be at the intersection of the x-axis and circle. Where will be the
second point?
2. In chapter 4 (4.13a) it was shown that A H Q = 0 where Q was defined in
terms of partitions of the spectral matrix. Is Q itself the noise subspace?
Remember that in obtaining Q no eigendecomposition was required.
3. Show that, taking into account the variation in the sensitivity of the
sensors, equation (5.72b) takes the form
where
2 2 2
α 2 (θ 0 ) = diag{α 0 (θ 0 ) , α1 (θ 0 ) ,..., α M −1 (θ 0 ) }.
Consider a ULA with M sensors and P uncorrelated wavefronts that are incident
on the array. Let v s be a matrix whose columns are the signal eigenvectors of
A
the array spectral matrix. Show that Jv s = G −1 where matrices A
AΓ
M
( × P) and G (P × P) are as defined in (5.6c) and the Γ (P × P) is as
2
defined in (2.57). This result provides an alternate approach to the ESPRIT
algorithm described in §5.1. It can be used to extend the concept of subspace
rotation to multiple subarrays as described in [49].
References
Source Estimation
and S0 (ω) is the cross-spectral vector between the desired output and the array
output. We had derived a specific result for a single wavefront with white
background noise. It was shown that the filter function is given by
S0 (ω)
σ 2η
where Q=
S (ω)
1+ M 0 2
ση
F̂(ω ) = H WH (ω , θ 0 )F(ω )
= Qa 0H (ω , θ 0 )[a 0 (ω , θ 0 )Ξ0 (ω ) + Ν(ω )] (6.1b)
= Ξ0 (ω ) + Qa 0H (ω , θ 0 )Ν(ω )
Note that the signal component in the output remains undisturbed and the noise
1
variance is reduced by a factor of . The filter response to an incident plane
M
wave coming from a different direction with a direction vector a(ω , θ ) is
given by Qa 0 (ω , θ 0 )a(ω , θ ) . The response of the Wiener filter will be
H
σ 2s 2
It is easy to show the output snr as a 0 (ω) . By definition, the array gain
ση2
2
is equal to a 0 (ω) which, for ideal sensor array with omnidirectional and unit
response sensors, is equal to M (the number of sensors).
6.1.2 Two Source Case: We shall now consider two wavefronts incident on an
array (ULA) in presence of white background noise. The sources are assumed to
be uncorrelated between themselves as well as with the background noise. The
aim is to extract a first source signal while we suppress the signal from the
second source as well as the background noise. Corresponding to this model the
spectral matrix and the cross-spectral vector in the Wiener filter equation take
the following form,
where
ω ω ω
−j d sin θ 0 −j 2 d sin θ 0 −j ( M −1)d sin θ 0
a 0 (ω , θ 0 ) = col[1, e c
,e c
, ... e c
]
and
where θ 0 and θ1 are DOA angles of the first and the second wavefront,
respectively. The filter response for extraction of the first source signal is given
by
−1
a 0 (ω,θ 0 )a 0H (ω,θ 0 ) +
H W (ω,θ 0 ) = S1 (ω) σ 2η a 0 (ω,θ 0 ) (6.3)
a1 (ω,θ1 )a1 (ω,θ1 ) S (ω) + S (ω) I
H
0 0
The inverse of the quantity inside the square brackets in (6.3) has been derived
in chapter 4 (p. 234). Using that result we obtain
V −1 − S1 (ω) ×
S0 (ω)
H W (ω,θ 0 ) = V −1a1 (ω,θ1 )a1H (ω,θ1 ))V −1 a 0 (ω,θ 0 ) (6.4a)
S (ω) H
1 + 1 a1 (ω,θ1 )V −1a1 (ω)
S0 (ω)
where
S0 (ω )
V −1 =
ση 2 [
I − Qa 0 (ω , θ 0 )a 0H (ω , θ 0 ) ] (6.4b)
where
S 0 (ω )
σ η2
Q=
S (ω )
1+ M 0 2
ση
Though the filter is tuned to receive the first wavefront some amount of energy
from the second wavefront will leak into the filter output. This is known as
cross talk. Ideally, the cross talk should be zero. Let a1 (ω)Ξ1 (ω) be the
Fourier transform of the signal emitted by the second source. The output of the
filter tuned to the first source is given by
S0 (ω ) a 0H (ω , θ 0 )a1 (ω , θ1 )
a 0H (ω , θ 0 )V −1a1 (ω , θ1 ) = (6.5b)
σ η2 S (ω )
1+ M 0 2
ση
and
The first term is the desired signal power. The remaining terms represent the
interference. Of these three, the magnitude of the first is much lower than that
of the second and third terms. The magnitude of the first is proportional to
4
a1H (ω,θ1 )a 0 (ω,θ 0 )
while the magnitude of the second and third terms is
M4
2
a1H (ω,θ1 )a 0 (ω,θ 0 )
proportional to . Hence, we drop the first term from
M2
the interference expression. The cross talk, defined as a ratio of the power leaked
from the second source and the actual power in the second source, is given by
2
a 0H (ω,θ 0 )a1 (ω,θ1 )
cross talk ≈ ×
M2
1 2 M(1 − 0 0 1 1
)
σ η M2
1 − 2 (6.8)
S (ω) a H
(ω,θ )a (ω,θ )
2
1 + 1 2 M(1 − 0 0 1 1
)
ση M2
2
a1H (ω,θ1 )a 0 (ω,θ 0 )
Note that represents the square of cosine of the angle
M2
between two direction vectors a 0 (ω,θ 0 ) and a1 (ω,θ1 ) . When
2
a1H (ω,θ1 )a 0 (ω,θ 0 )
a 0 (ω,θ 0 ) = a1 (ω,θ1 ) , =1 the cross talk =1, as
M2
2
a 0H (ω,θ 0 )a1 (ω,θ1 )
expected. But when a 0 (ω,θ 0 )⊥a1 (ω,θ1 )) , =0 and
M2
cross talk is zero. Aside from these two extreme situations the cross talk may
MS1 (ω)
be reduced to zero if >>1, that is, the array signal-to-noise ratio
σ 2η
(asnr) for the second source is very large. The cross-talk is shown in fig. 6.1.
6.1.3 Linear Least Squares Estimate (LLSE): As previously stated the signal
model is
ξ 0 (t)
ξ1 (t)
f(t) = [a 0 ,a1 ,...a P −1 ] + η(t)
(6.9)
ξ P −1 (t)
= Aξ(t) + η(t), t = 0,1,..., N
2
f(t) − Aξ(t) = min (6.10)
0.9
0.8
0.7
Cross talk
0.6
0.5
0.4
0.3
0.2
0 1 2 3 4 5 6
Angle in degrees
Figure 6.1: Wiener Filter: Cross talk as a function of angular distance between two
sources. The first source is on broadside (DOA=0o ). Array signal-to-noise ratio
(asnr)=10 and eight sensor ULA is assumed.
Differentiating (6.10) with respect to ξ(t) and setting the derivative to zero we
obtain
[] A f(t)
−1
ξ̂(t) = A H A H
(6.11)
= ξ(t) + [A A] A
−1
H H
η(t)
We observe that the signal term is extracted without any distortion but the
[A A]
−1
noise term, given by
H
A H η(t) , behaves differently; for example, the
noise becomes correlated. The output noise covariance matrix is given by
[ ]
−1
Cη̂ = A H A σ η2 (6.12)
H
When A A is singular or close to being singular, that is, with a large
eigenvalue spread, the noise in the output may get amplified. Consider a case of
80
8
Amplitude
.8
.08
.008
0 1 2 3 4 5
Angle in degrees
Figure 6.2: The noise amplification factor as a function of DOA of source #2. Source
#1 is held fixed at 0o and source #2 is moved. Array size M=8.
a 0H a1
2
[ H
]
det A A = M 1 −
M
2
(6.13)
Using (6.13) in (6.12) we obtain the variance of the noise in all array outputs
1
σ η̂2 = σ η2 (6.14)
a 0H a1
2
M 1 −
M
We have plotted in fig. 6.2 the noise amlpification factor, the factor
multiplying σ η , as a function of the angular separation between two sources.
2
H W (ω) = S−1
f (ω)AS 0 (ω)
circular array,
The interest is to find out how a circular array fares in comparison with a linear
array with respect to cross talk capability. For a given number of sensors (say,
(M − 1)λ
M) the maximum aperture of an ULA is fixed at but the aperture of
2
a circular array can be very large, at least in principle. Since the array aperture
is the main factor deciding the cross talk it is expected that the circular array
ought to perform better in terms of lower cross talk. Using the same number of
sensors, cross talk may be reduced when arranged over a large circle. This is
shown in fig. 6.3 for an eight sensor UCA. The cross talk has been reduced
considerably when the array radius is increased from four to thirty-two
wavelengths. A linear array of eight sensors will have aperture of 3.5λ. An
UCA having an aperture of 3.5λ was found to show much higher cross talk
than that of the ULA shown in fig. 6.1. Thus, performance of an UCA is
significantly better only when the radius is increased considerably.
6.1.4 Effects of Errors in DOA : Waveform estimation requires a knowledge of
the direction of arrival of wavefronts. But the DOA estimates are subject to
0.8
4λ
0.6
Cross talk
0.4 32 λ
0.2
0
0 1 2 3 4 5 6
Angle in degrees
Figure 6.3: Cross talk as a function of angular separation of two sources. Eight
sensors are uniformly spaced over a circle of radius 4λ (solid curve) and 32λ (dashed
curve).
errors largely on account of finite data length used in their estimate. In this
section we investigate the effects of errors in DOA estimates on interference
from signals of no interest. A comprehensive analysis of the effects of model
errors including the errors in DOA may be found in [2]. Let θ̂ p = θ p + ∆θ p ,
p = 0,1,..., P − 1 be the estimated DOAs, where θ p is correct DOA and
∆θ p is estimation error. We shall rewrite (6.11) showing explicitly the
dependence on the estimated DOAs
[ ]
−1
ξ̂(t) = A H (θ̂)A(θ̂) A H (θ̂)f(t) (6.16)
[ ]
Let us express A(θ̂) = B c where B is a matrix whose columns are the
direction vectors of all interfering sources (signals of no interest) and c is a
vector representing the direction vector of the signal of interest (soi).
[A ]
−1
H
(θ̂)A(θ̂) may be simplified [2] as
−1
B H B BH c
= H
c B cHc
−1 −1
H B H cc H B H B H cc H B B H c
B B − 2 − B B − 2 c2
= c c
H (6.17)
−[c c − c Pc] c B1 [c c − c Pc]
H −1 H H H −1
−1
where P = B(B B) B
H H
and B1 = (B H B)−1 B H . Using (6.17) in (6.16)
and simplifying we obtain
[A ]
−1
H
(θ̂)A(θ̂) A H (θ̂)
c H (I − P)
ξ̂ soi (t) = f(t) = d H (θ̂)f(t) (6.19)
c (I − P)c
H
We now expand d H (θ̂) in a Taylor’s series and retain the first derivative term
only, d(θ̂) ≈ d(θ) + d1 (θ)∆θ , where d1 (θ) is the first derivative of
d(θ) with respect to θ . The estimated signal of interest reduces to
P −1
[
∆ξ soi (t) = ∑ β H ξ(t) p ∆θ p ] (6.21)
p= 0
where [ ]
∆ξ soi (t) stands for the error in ξ̂ soi (t) , β = A H d1 (θ) and [.] p
stands for the pth element of a column vector.
The errors in the calibration of an array can cause serious errors in the
waveform estimation [3]. It is, therefore, imperative to carefully calibrate the
array response, particularly when subspace methods are used for localization or
waveform estimation. One simple solution to this problem of high sensitivity
to calibration errors is to reduce the degrees of freedom by using a low rank
approximation of the spectral matrix [4, 5]. In this approximation a spectral
matrix is simply approximated by its signal subspace, that is,
S f (ω) = v s α s v sH + v ηα ηv ηH
(6.22)
≈ v s α s v sH
1
H = (S−1
f Γ)H (6.23a)
λ
where
∆θ
θ+
2
1
Γ= ∫ a(ω,φ)a
H
(ω,φ)dφ (6.23b)
∆θ ∆θ
θ−
2
H
H
= (S−1 H
f a(ω,θ)a (ω,θ)H (6.24)
H SfH
1
HHS f H =
a (ω,θ)S−1
H
f a(ω,θ) (6.25)
S−1a(ω,θ)
H cap = H f −1 (6.26)
a (ω,θ)S f a(ω,θ)
1
and the output power is equal to H The optimum filter
a (ω,θ)S−1
f a(ω,θ)
given in (6.26) is also known as the Applebaum filter [6].
6.2.1 Extended Source: For ULA the Γ matrix takes the following form
∆θ
sin(k )
ωd
[Γ ]m,n = ∑ ε k Jk ((m − n) )cos(k θ) ∆θ + 2
k = 0,2,4... c k
2
∆θ
sin(k )
ωd
j ∑ 2Jk ((m − n) )sin(k θ) 2
k =1,3,5... c ∆θ
k
2
where ε 0 = 1 and ε n = 2 for all k. To evaluate the integral in (6.23b) we
have used the result derived in (1.53b). For UCA the Γ matrix takes the
following form:
where
m = 0,1,... M − 1
2 πkm
j
{W}km = e M
k = 0, ±1,...∞
π
ωa ωa − j ( + ϕ 0 )( k −l ) ∆θ
[Θ]k,l = Jk ( )Jl ( )e 2 sin c((k − l) )
c c 2
valid for all values of ∆θ . We have used a series expansion of the steering
vector for a circular array (2.51).
6.2.2 Single Point Source Case: Let there be a single source emitting a plane
wave with a direction vector a 0 (ω,θ 0 ) in presence of white noise. The
spectral matrix is given by S f = a 0 (ω,θ 0 )a 0H (ω,θ 0 )Sξ (ω) + σ 2ηI , and its
inverse is given by
S−1
f =
1
ση
2 [
I − Qa 0 (ω , θ 0 )a 0H (ω , θ 0 ) ] (6.27)
where
Sξ (ω)
σ 2η
Q=
S (ω)
1+ M ξ 2
ση
S ξ (ω )
For large array signal-to-noise ratio, >> 1 , we may approximate
σ η2
1
Q≈ . Using (6.27) in (6.26) we obtain
M
H cap
1
=
0 [
a(ω , θ ) − Q a H (ω , θ )a(ω , θ ) a (ω , θ )
0 0 0
] (6.28)
M
2
a (ω , θ )a 0 (ωθ 0 )
H
1− Q
M
F̂(ω ) = H cap
H
[a 0 (ω , θ 0 )Ξ0 (ω ) + Ν(ω )] (6.29a)
H
H cap a 0 (ω,θ 0 )Ξ0 (ω)
H
a (ω,θ)a 0 (ω,θ 0 ) 1 − QM
= 2 Ξ 0 (ω) (6.29b)
M a (ω,θ)a 0 (ωθ 0 )
H
1 − QM
M2
= Ξ0 (ω) when a H (ω,θ) = a 0 (ω,θ 0 )
H
H cap Ν(ω)
1
=
[0 0 ] 0 0 Ν(ω)
a(ω,θ) − Q a H (ω,θ )a(ω,θ) a (ω,θ )
(6.29c)
M
2
a H (ω,θ)a 0 (ωθ 0 )
1 − QM
M 2
a 0H (ω,θ 0 )
= Ν(ω) when a H (ω,θ) = a 0 (ω,θ 0 )
M
The variance of the noise in the filter output (6.29c), when
σ N2
a H (ω , θ ) = a 0 (ω , θ 0 ) , turns out to be . The response of the filter was
M
computed from (6.29a) for different steering directions. The DOA of the
incident wavefront was assumed to be at 0o. An ULA with eight sensors was
considered. The response of the filter depends upon the signal-to-noise ratio
(snr). For snr=10 (or asnr=80) the response becomes extremely sharp as
demonstrated in fig. 6.4. For comparison we have also plotted a Wiener filter
response (dashed curve) where we have assumed the signal-to-noise ratio equal
to ten while other parameters remain the same as for the Capon filter.
Wiener filter
0.8
0.6
Response
snr=10
0.2
0
0 0.5 1 1.5 2 2.5
Angle in deg.
Figure 6.4: Response of Capon filter as a function of the angle of the steering
vector. An ULA of 8 elements was assumed. Equivalent Wiener filter response is also
shown by the dashed curve.
−1 S (ω) −1 H
V a1 (ω,θ1 )a1 (ω,θ1 )V −1
num = V − 1 a 0 (ω,θ 0 )
S0 (ω) 1 + S1 (ω) a H (ω,θ )V −1a (ω,θ )
1 1 1 1
S0 (ω)
(6.30a)
M(1 + 2 1 − )
ση M2
den=
S0 (ω)
(1 − QM)
σ 2η S1 (ω) a1H (ω,θ1 )a 0 (ω,θ 0 )
2
(1 + 2 1 − QM )
ση M2
(6.30b)
Ξ0 (ω)
output = H (6.31)
+H Cap (ω,θ 0 )a1 (ω,θ1 )Ξ1 (ω) + H Cap (ω,θ 0 )Ν(ω)
H
Notice that the signal term is equal to the actual signal. But this was not so in
the case of the Wiener filter where we had to assume a large array snr in order to
arrive at this result.
The contribution of the second source is represented by the second
term in (6.31). The cross talk is then given by
2
a 0H (ω,θ1 )a1 (ω,θ 0 )
M
= 2 (6.32)
S (ω) a1H (ω,θ1 )a 0 (ω,θ 0 )
2
1 + 2 1 −
1
σ M 2
η
The cross talk is plotted in fig. 6.5 for an ULA and in fig. 6.6 for an UCA.
Compare these two figures with figures 6.1 and 6.3 where we have plotted the
cross talk for the Wiener filter. Clearly, the Capon filter performs better for
both types of array geometries. An expression for snr which agrees with (6.32)
is derived in [7].
Finally, we shall evaluate the noise term, that is, the leftover noise in
the array output,
0.8
0.6
Cross talk
0.4
0.2
0
0 1 2 3 4 5 6
Angle in deg
Figure 6.5: Capon Filter: Cross talk as a function of angular distance between two
sources. The first source is on broadside (DOA=0o ). Array signal-to-noise ratio
(asnr)=10. Continuous curve: 8 sensor ULA and Dashed curve: 16 sensor ULA.
H
H Cap (ω,θ 0 )Ν(ω)
S (ω) V −1
a (ω,θ )a H
(ω,θ )V −1
a 0H (ω,θ 0 ) V −1 − 1 1 1 1 1
Ν
S0 (ω) 1 + S1 (ω) a H (ω,θ )V −1a (ω,θ )
1 1 1 1
S0 (ω)
= (6.33)
den
a 0H (ω,θ 0 )V −1Ν − ψa1H (ω,θ1 )V −1Ν
=
den
where the denominator term, den, is given in (6.30b) and
S1 (ω) H
a 0 (ω,θ 0 )V −1a1 (ω,θ1 )
S0 (ω)
ψ=
S (ω) H
1+ 1 a1 (ω,θ1 )V −1a1 (ω,θ1 )
S0 (ω)
0.8
4λ
0.6
Cross talk
0.4
32 λ
0.2
0
0 1 2 3 4 5 6
Angle in deg
Figure 6.6: Capon Filter: Cross talk as a function of angular separation of two
sources. Eight sensors are uniformly spaced over a circle of radius 4λ (solid curve)
and 32λ (dashed curve).
The variance of the leftover noise may be computed from (6.33). We obtain
after simplification
H
Var H Cap {
(ω,θ 0 )Ν(ω) }
σ 2η a1H (ω,θ1 )a 0 (ω,θ 0 )
= 1 − ψ
M M
S1 (ω) a1H (ω,θ1 )a 0 (ω,θ 0 )
2
` (6.34a)
2
1 + M(1 − )
σ ση 2
M
= η
M a (ω,θ1 )a 0 (ω,θ 0 )
H 2
S (ω)
1 + 1 2 M(1 − QM 1 )
ση M
When the array snr is large, then QM ≈ 1 and (6.34) reduces to a simple form
M −1
f̂ (t) = ∑a
m=0
m f m (t − m τ) (6.35a)
P −1
F̂(ω) = ∑ (a
m=0
m e − jmτω )F(md,ω)
π
(6.35b)
1
= ∫
2π − π
A(u)F * (u,ω)du
where
P −1
A(u) = ∑ ã
m=0
m e − jmu
ãm = am e − jmτω
θ0
x
Figure 6.7: Position of a null and two poles in its immediate neighborhood but
slightly inside the unit circle.
P −1
F(u,ω) = ∑ F(md,ω)e
m=0
− jmu
(z − z0 )
A(z) = (6.36)
(z − z+ )(z − z− )
The response of the notch filter given by (6.36) is shown in fig. 6.8. See
exercise 3 for another type of null steering filter independent of frequency.
6.3.2 Beam Steered Adaptive Array: In Capon filter (6.26) let us replace the
inverse of the spectral matrix by its eigenvalue eigenvector representation,
1.0
(a)
0.8
Magnitude
0.6
0.4
0.2
0.0
-90 -60 -30 0 30 60 90
Angle (deg.)
(b)
1
Phase (radians)
-1
-2
-3
-90 -60 -30 0 30 60 90
Angle (deg.)
Figure 6.8: Wavenumber response of the notch filter (a) amplitude response and (b)
phase response.
aH (ω,θ){v λ v + v λ v }a(ω,θ)
s
−1 H
s s η
−1 H
η η
We assume that the look direction vector a(ω,θ) lies in the signal subspace;
hence a(ω,θ)⊥v η . Hence (6.37a) simplifies to
H cap =
{v λ v }a(ω,θ)
s
−1 H
s s
(6.37b)
aH (ω,θ){v λ v }a(ω,θ)
s
−1 H
s s
The Capon filter has an interesting property, that is, when the look direction
coincides with one of the signal directions, for example, a(ω,θ) = a(ω,θ 0 )
it is approximately orthogonal to all direction vectors,
a(ω,θ m ), m = 1,2,..., P − 1. To show this, consider
In (6.38a) the inverse of the signal eigenvalues may be approximated, for large
σ 2η
snr or large array size, by λ−1m ≈ α −1
m − . Therefore, (6.38a) may be
α 2m
expressed as
− m=0 αm
a (ω,θ 0 ){v s λ s−1v sH }a(ω,θ 0 )
H
We shall show that the first term inside the brackets is indeed zero. Let us
consider the signal term alone from (4.12b) and compute the pseudoinverse
(denoted by #) on both sides of the equation
H
A H S #f A = A H A # S0−1A # A = S0−1 (6.40)
H
H cap a(ω,θ m ) ≈ 0 m = 1,2,..., P − 1 (6.41)
Ss# (ω)a(ω,θ)
H̃ cap = (6.42)
a H (ω,θ)Ss# (ω)a(ω,θ)
#
where Ss (ω) is the pseudoinverse of the signal-only spectral matrix. It is
shown in [10] that the modified Capon filter is robust against look direction
errors. It was shown in chapter 5 (page 274) that the generalized eigenvector
corresponding to a given direction of arrival (DOA) is orthogonal to all other
direction vectors.
6.3.3 Adaptive Capon filter: In chapter 3 we have derived two different filters to
extract a wavefront coming from a specified direction. It may be recalled that
the Wiener filter minimizes the mean square error between the filter output and
the desired signal and the Capon filter minimizes the output power while
maintaining unit response in the desired direction. The Wiener filter is given by
H W (ω) = S−1 f (ω)S 0 (ω) and the Capon filter is given by
S−1
f (ω)a(ω,θ)
H Cap (ω) = where a(ω,θ) is the direction vector
a (ω,θ)S−1
H
f (ω)a(ω,θ)
in the desired direction. Note that S0 (ω) = a(ω,θ)S0 (ω) where S0 (ω) is
power from the desired direction. In the Wiener filter this power is assumed to
be a constant but in the Capon filter the power is estimated
[Ŝ (ω)]
−1
N +1
f =
[ ]
−1
where z = Ŝ Nf (ω) F N +1 (ω) and µ ∈(0,1) , which is a free parameter to
be chosen depending upon how fast the estimated spectral matrix changes from
snapshot to snapshot. For stationary process where the change is small, µ ≈ 1.
Using (6.43a) and (6.43b) we can recursively compute the Wiener and Capon
filters. We will do this for the Capon filter. Multiply both sides of (6.43b) by
a(ω,θ)
. Noting the definition of the Capon filter and
[ ]
−1
a H (ω,θ) Ŝ Nf (ω) a(ω,θ)
assuming that the power from the desired direction does not change much from
snapshot to snapshot we obtain
N +1 1 N a(ω,θ)
H Cap = H Cap − βzz H (6.43c)
[ ]
−1
µ H
a (ω,θ) Ŝ Nf (ω) a(ω,θ)
where
(1 − µ)
β=
[ ]
−1
µ (1 − µ)F HN +1 (ω) Ŝ Nf (ω) F N +1 (ω) + µ
a(ω,θ)
zH = F HN +1 (ω)H Cap
N
[ ]
−1
H N
a (ω,θ) Ŝ (ω)f a(ω,θ)
N +1 1 N
H Cap = H Cap − βzF HN +1 (ω)H Cap
N
(6.43d)
µ
It is shown in [11] that (6.43d) yields a stable estimate of the Capon filter.
Q
σ 2η
∑
2
C z 0 z 0 = p02 a(θ 0 )a(θ 0 ) H + pk2 a(θ k )a(θ k ) H + I (5.68)
3L k =1
L
The first term on the right hand side is of interest as we would like to estimate
the direction vector of the user of interest. The second term represents the co-
channel interference from all other sources. Notice that this term will be small
for large L.
6.4.2 Estimation of All Direction Vectors: We like to estimate the direction
vectors of all users in the same cell. For this we shall compute the
postcorrelation covariance matrices for all users. Thus, we will have Q
equations of the type given by (5.66)
Q
σ 2η
∑
2
C g k g k = pk2 a(θ k )a(θ k ) H + pi2 a(θ i )a(θ i ) H + I
3L i=0
L (6.44)
k ≠i
k = 0,1,..., Q − 1
σ 2η -1
Θ = £-1 C + £ II (6.46)
Ts
The error term in (6.46) may be expressed as a product of a diagonal matrix and
a column unit matrix. The elements of the diagonal matrix are equal to row
-1
sums of £ . Thus, the noise covariance matrix in the estimated direction
matrix remains diagonal. The variance of the noise may be estimated from the
eigenvalues of the direction matrix. The power, transmitted by each user, that
2
is, pi , can also be estimated from the largest eigenvalue of the direction
matrix.
6.4.3 Simulation Results: The estimated direction vector of a user is compared
with the known direction vector. A dot product between the two vectors is
computed as a measure of similarity,
â H (θl )a(θl )
εl =
â H (θl ) a(θl )
Note that 0 ≤ ε l ≤ 1 , the lower limit represents the worst estimate and the
upper limit represents the best estimate. We have computed the mean and the
where c αf is the cyclic covariance matrix defined in (5.71a) for zero lag. The
filter coefficients are chosen to maximize the cross-correlation (6.47) or its
magnitude square. Further, we require that w + and w − are unit norm vectors.
2
w − H c αf (τ)w + = max, w + H w + = 1, w − H w − = 1 (6.48a)
The solution of (6.48a) is given by left and right singular vectors corresponding
α
to the largest singular value of c f (τ) . Also w − and w + are, respectively, the
eigenvectors corresponding to the largest eigenvalues of c αf (τ)c αf (τ) H and
c αf (τ) H c αf (τ) [5].The cyclic covariance matrix appearing in (6.48) is in
practice replaced by the cyclic autocorrelation function defined in terms of time
average,
1 2
ĉ αf = ∑ f − (t)f +H (t) (6.48b)
T −T
2
When the carrier frequencies of different sources are sufficiently apart and the
signal duration T is large the cyclic autocorrelation matrix given by (6.48)
approaches the cyclic covariance matrix of a single source which is a rank one
matrix as shown in (5.71b) whose left and right singular vectors corresponding
1
to the largest singular value are equal to a 0 (for ULA). Thus, we have an
M
interesting result
1
w+ = w− = a0 (6.48c)
M
0.8
0.6
0.4
Amplitude
0.2
-0.2
-0.4
0 20 40 60 80
Time
Figure 6.9: A source emits a waveform (sinc function) shown by solid curve. The
signal reaches a sensor via four paths as in equation (6.49). The sensor output is
shown by dashed curve.
medium. For the time being we shall not consider such a situation. We shall
assume that all copies of the signal remain correlated. Even in this simplified
model of multipath propagation the waveform received by a sensor may have
no resemblance with the signal emitted by the source. To appreciate the kind of
deterioration a signal may suffer let us consider a source emitting a signal
sin(x)
described by sinc function (= ) and a multipath channel having four
x
paths. The relative amplitudes and delays are assumed to be random numbers as
given below:
hs − ha hs + ha
tan θ 0 = and tan θ1 = − (6.51)
l l
and ha is the depth of the array (midpoint) from the surface. In deriving (6.51)
we have used the method of images as outlined in chapter 1. For this channel
the direction vectors and weight vector, respectively, are A = a 0 ,a1 where [ ]
− j 2 π d sin(θ 0 ) − j 2λπ 2 d sin(θ 0 ) −j
2π
( M −1)d sin(θ 0 )
a 0 = col 1,e λ ,e ,...,e λ
− j 2 π d sin(θ1 ) − j 2λπ 2 d sin(θ1 ) −j
2π
( M −1)d sin(θ1 )
a1 = col 1,e λ ,e ,...,e λ
w = col{1,r}
(a)
Source
Array
(b)
Figure 6.10: Two types of simple channels for which the Aw vector may be
estimated from a knowledge of source location and channel characteristics.
of the direction vectors of the direct wavefront from the source and the one from
the image of the source.
Aw = [a 0 + ra1 ] (6.52)
First, let us try the linear least squares estimate (LLSE) of F0 (ω) . We shall
transform (6.11) into frequency domain, leading to
[ ]
−1
F̂0 (ω ) = w H A H Aw w H A H F(ω )
(6.53a)
[ ]
−1
= F0 (ω ) + w H A H Aw w H A H Ν(ω )
The quantity inside the square brackets in (6.53a) may be evaluated using
(6.52). It is given by
r H a1H a 0 + ra 0H a1 (6.53b)
= 2M 1 +
2M
The variance of the noise power in the filtered array output is given by
σ η2
σ η̂2 =
[w H A H Aw]
σ η2 (6.54)
=
r H a1H a 0 + ra 0H a1
M 1 + r +
2
M
Compare (6.54) with (6.14), which was derived for two uncorrelated sources. In
the present case both sources are correlated (the second source is an image of the
primary source). The variance of the noise is reduced by a factor four, when r=1
and a1 = a 0 . The multipath propagation has indeed helped to improve signal
estimation.
Next, we shall try the Capon filter to estimate the waveform in the
presence of interference, another source at known location, and the usual
background white noise. The position of the sources and the receiving array are
shown in fig. 6.11. We shall assume that the sources are uncorrelated. The
directions of arrival of the direct and the reflected wavefronts are given by
hs − ha hs + ha
tan θ 00 = , tan θ10 = −
l l
hs + ∆h − ha h + ∆h + ha
tan θ 01 = , tan θ11 = s
l l
For simplicity we assume that the coefficient of reflection r is the same for
both sources. The direction vectors are given by
A 0 w = [a 0 + ra1 ]
A1w = [a 01 + ra11 ]
s1
s0
Figure 6.11: Two sources at the same range but at different heights. A vertical array
of sensors is assumed.
To compute the Capon filter given by (6.26) we need the spectral matrix of the
array output. Since the sources are assumed uncorrelated and the background
noise is white, the spectral matrix may expressed as
S−1 Aw
H cap = H H f 0 −1 (6.57)
w A 0 (ω , θ )S f A 0 w
w H A 0H S−1
f A1w w H A 0H S−1Ν(ω )
H F(ω ) = F0 (ω ) + H H −1
H
cap F1 (ω ) + H H f−1 (6.58)
w A0 S f A0 w w A0 S f A0 w
0.8
0.6
Cross talk
0.4
0.2
0
50 55 60 65 70
Separation in meters
Figure 6.12: Cross talk as a function of separation between two sources. The range
is the same for both sources. Solid curve: range=1000m and dashed curve=5000m.
The first source is 50 meters above the surface.
While the signal from the zeroth source has been fully extracted there are two
terms in (6.58) interfering with the desired signal. Here we shall consider the
term representing the interference from the source #1. The second term, that is,
the noise term, will not be considered as it follows the same approach used
previously in connection with the single source case. The interference due to
second source will be measured in terms of cross talk as defined previously. In
the present case the cross talk is given by
(w H A1H S−1 H H −1
f A 0 w)(w A 0 S f A1w)
cross talk = 2 (6.59)
w H A 0H S−1
f A0 w
The cross talk as a function of source separation for two uncorrelated sources is
plotted in fig. 6.12. A 16 sensor vertical ULA was assumed. Notice that for a
short range the second source has little influence on the first source. But this
influence grows rapidly as the range increases.
6.5.2 Partially Known Channel: In many real life problems the channel
characteristics are never fully known as it is impossible to measure the micro
level variations causing path length variations on the order of a fraction of
wavelength. Such variations are known to affect the performance of source
localization algorithms, particularly those belonging to a high resolution class
[13]. On the other hand we may have fairly good knowledge about the general
Aw
Es = , Aw ≠ 0 (6.60)
Aw
In (6.60) the weight vector w is dependent on the channel parameters and the
columns of A matrix on the source position. Notice that the weight vector w
in (6.60) occurs linearly while the source parameters in A occur nonlinearly.
A least squares method of estimating the nonlinear parameters by first
eliminating the linear parameters followed by minimizing the norm of the error
vector was first suggested by Guttman et al.[14] and applied to a signal
processing problem by Tuft and Kumaresan [15]. We shall exploit here this
approach. In (6.60) we assume that the source position is approximately known
and write the equation in terms of the unknown w vector,
w
Es = A = Aw̃ (6.61)
Aw
#
Let A be the pseudoinverse of A. The least squares estimate of w̃ will be
given by
˜ˆ = A # E s
w
Substitute back into (6.60) and obtain an estimate of E s . The mean square
error is given by
⊥
where P A is the orthogonal projection complement of matrix A . The mean
square error is now minimized with respect to the source location parameters.
We define a parametric spectrum as,
1
S(R 0 , Zs) = H ⊥
E s P A ES (6.62)
˜ˆ = A max
w #
Es (6.63)
# #
where A max is A evaluated where the parametric spectrum is maximum.
˜ˆ is the least mean square estimate of w̃ . An example of parametric
Thus, w
spectrum is shown in fig. 6.13. A low power source (-10dB) is assumed at
range 4600m and depth 50m in a Pekeris channel of depth 200 meters. A
vertical ULA is placed at a depth of 70m. For the same channel the reflection
coefficients were computed (6.63) from the eigenvector corresponding to the
largest eigenvalue. The reflection coefficients are normalized with respect to
Aw , which may be obtained from the fact that the weighting coefficient
corresponding to the direct path is by definition equal to one; hence its actual
observed value must be equal to Aw . In table 6.3 the estimated and the actual
reflection coefficients for the first eight images out of twenty multipaths used
in computation are listed for two different array lengths.
In computer simulation we have found that, for good results, the
number of sensors has got to be many more than the lower limit given in [16].
The least mean square error in the estimated reflection coefficients for different
number of sensors is shown in fig. 6.14. Here the number of significant images
Figure 6.13: The parametric spectrum for a single source located at range 4600 m and
depth 50 m. The source power is -10dB (relative to the background noise). Under
water channel is 200m deep Pekeris channel (soft bottom, speed: 1600m/s and
relative density: 2.0). A vertical ULA consisting of 40 sensors is placed at a depth
of 70m from the surface. Range scan is from 1000 m to 8000 m in steps of 200 m.
Depth scan is from 10 m to 180 m in steps of 20 m.
-40
-30
mse in dB
-20
-10
0
10 20 30 40 50 60 70
No of sensors
Figure 6.14: The role of array size on the mean square error in the estimated
reflection coefficients is shown above. Twenty multipaths were assumed.
is twenty (P=20) and hence, according to the lower limit (> 2P+ 2)
the minimum number of sensors ought to be more than 42. We observe that an
array of sixty sensors appears to be optimum.
In addition to unknown reflection coefficients we have background
noise which is likely to be both spatially and temporally correlated. But, since
it is uncorrelated with the signal, it occurs in a linear combination; as such it
may be estimated using the approach used for the estimation of the reflection
coefficients. Indeed, such an approach has been used by Boehme [17] for the
estimation of the noise spectrum first using an approximate knowledge of the
channel. In the next step, the previously estimated noise spectrum is used in
the expression for the likelihood ratio which is then maximized with respect to
the unknown channel parameters and the source position. The maximization of
the likelihood ratio is, however, highly computation intensive [18].
§6.6 Exercises:
1. Consider P uncorrelated wavefronts incident on an M sensor ULA. Let the
background noise be uncorrelated with all signals. Show that the Wiener filter
in the frequency domain is given by
H W (ω) = S−1
f (ω)AS 0 (ω)
2. Show that the Wiener filter (also Capon filter) can be written as
H W = v s α s−1v sH a 0 S0
where a 0 is the direction vector of the wavefront from the first source and S0
is signal power, (Hint: Use (3.38b) and the property given in (4.14b).) Such a
filter is robust against calibration error (see p.342).
3. In §6.1 and §6.2 we have seen that in estimating a waveform in the presence
of interference there is always some cross talk, that is, leakage of power from
the interfering signal to the desired signal. It is possible to devise a set of
filters, Hm (ω), m = 0,1,..., M − 1 , which will null the interference without
M −1
distorting, that is, ∑H
m=0
m (ω) = 1, the desired signal (but with no noise).
The array has been steered to receive the desired signal. The DOA of the desired
signal and that of the interference are known. Show that
M −1
M − e jωτ m ∑ e − jωτ i
Hm (ω) = i=0
2
M −1
M −
2
∑e
i=0
− jωτ i
where τ i is the time delay of the interference at ith sensor. The interference is
nulled except when ωτ i is equal to an integer multiple of 2π [4].
References
1. B. Otterstein, R. Roy, and T. Kailath, Signal waveform estimation in sensor
array processing, Asilomar Conf, pp. 787-791, 1989.
2. B. Friedlander and A.J. Weiss, Effects of model errors on waveform
estimation using the MUSIC algorithm, IEEE Trans., SP-42, pp. 147-155,
1994.
3. R. T. Compton, Pointing accuracy and dynamic range in steered beam
adaptive array, IEEE Trans., AES-16, pp. 280-287, 1980.
4. M.T. Hanna and M. Simaan, Absolutely optimum array filters for sensor
array, IEEE Trans., ASSP-33, pp. 1380-1386, 1985.
5. Q. Wu and K. Wong, Blind adaptive beamforming for cyclostationary
signals, IEEE Trans., SP-44, pp. 2757-2767, 1996.
Tomographic Imaging
363
source
y
ds
t
x
B
Object
sensor
the object (see fig. 7.1). Let N in be the number of photons incident on the
object at point A and Nd be the number of photons coming out at point B
within the time interval of measurement. N in and N d are related as below
B
N d = N in exp[− ∫ f ( x, y)ds ] (7.1a)
A
Nd
Define attenuation as negative of log , which turns out to be equal to the
Nin
integral of f (x, y) along path AB,
B
Nd
Nin ∫A
p = − log e = f (x, y)ds (7.1b)
x cos(θ) + ysin(θ) = t
Parallel
beam
A
ds
t θ
x
B
(a)
Scanning
sensor
y
Fan
beam
α
t θ x
Scanning
(b) sensor
B
N
pθ (t) = − log e d = ∫ f (x, y)ds
Nin A
(7.2)
∞ ∞
pθ (t) is known as Radon transform (see page 21). For a fixed θ and variable t,
we obtain p θ (t) , a continuous function of t, known as parallel projection
which may be generated by illuminating an object with a parallel beam and
scanning the output with a receiver (see fig. 7.2a)
Taking Fourier transform on both sides of (7.2) we obtain
Pθ (ω)
∞
= ∫ p (t)exp(− jωt)dt
−∞
θ
∞ ∞ ∞
= ∫∫
−∞ −∞
f (x, y)dxdy ∫ δ(x cos(θ) + ysin(θ) − t)exp(− jωt)dt
−∞
∞ ∞
+∞
1
∫ ∫ F(u,v)e
j (ux + vy)
f (x, y) = dudv
4π 2 −∞
(7.4)
∞ 2π
1
= 2 ∫ sds ∫ F(s,θ)e js( x cos θ+ y sin θ) dθ
4π 0 0
F(s, π + θ) = F(−s,θ)
u
F(u,v)
Figure 7.3: The Fourier transform of the object function is scanned along a series of
radial lines, one for each projection.
π ∞
1 1
f (x, y) = ∫ ∫
2π 0 2π − ∞
F(s,θ) s e js( x cos θ+ y sin θ) ds dθ
(7.5)
π ∞
1 1
= ∫ ∫
2π 0 2π − ∞
F(s,θ) s e jst ds dθ
Note that, for a fixed θ , F(s, θ ) is equal to Pθ (ω) given in (7.3). Thus, the
quantity inside the curly brackets in (7.5) may be obtained from the parallel
projection by simply filtering it with a filter having a transfer function,
H(s) = s .
π
1
2π ∫0
f (x, y) = f̃ (x cosθ + ysin θ)dθ (7.6)
where
1 ∞
∫ θ
jω( x cos θ+ y sin θ)
f̃ (x cosθ + ysin θ) = P (ω) ω e dω (7.7)
2π − ∞
δx δ
δ mn
to n th sensor
Figure 7.4: A target is divided into many square cells (or cubes in 3D). The wave
speed is assumed to be constant in a cell. The path delay introduced by mth cell is
δ mn
given by where δ mn is path length in mth cell of a ray going to nth sensor.
cm
wm,n represent a weighting coefficient which when used along with the
maximum delay gives a delay contributed by mth cell to nth sensor. In terms of
path length of a ray in the mth cell the weighting coefficient is given by
δ mn
wm,n = (see fig 7.4). Thus, the total delay observed at the nth sensor is
2δ
given by
M −1
tn = ∑w
m=0
τ
m,n m n=0, 1,...,N-1 (7.8a)
where we have assumed that there are M cells and N sensors. In matrix form (1)
may be expressed as
t = wτ (7.8b)
τ = (w T w)−1 w T t (7.9)
T
provided w w is nonsingular. The question of the rank of the weight matrix
w has no quantitative answer but we can give some qualitative guidelines:
i) The ray path lengths in different cells must be quite different so that
there is correlation among weight coefficients. In fig.7.4, this is more likely
to happen with a fan beam than with a parallel beam.
ii) The weight matrix is fully determined by the sensor array geometry.
For example, if sensors are too closely spaced all rays will travel through the
same group of cells and each cell will contribute roughly the same delay. The
weight matrix will then tend to be more singular.
iii) The sensor and source arrays must be so designed that the rays pass
through different cells in different directions. More on this possibility will be
discussed in the next subsection on borehole tomography.
It may be noted that since the weight matrix is generally a large sparse matrix,
efficient techniques have been designed for fast and economical (in terms of
memory requirements) inversion of the weight matrix. This is, however,
beyond the scope of this book. The interested reader may like to review an article
by S. Ivansson [2].
Borehole Tomography: The use of a source array often improves the rank
condition of the weight matrix. Consider a P source array arranged in some
unspecified form. Equation (7.8b) may be used to express the output as
where t p is the array output due to pth source and w p is the weight matrix
corresponding to the position of pth source. Next, we stack up all array vectors
into a single vector. Note that τ is independent of the source position.
Equation (7.10) reduces to
t̃ = w̃τ (7.11)
where t̃ = {t 0 ,t1 ,...,t P −1 } and w̃ = {w 0 , w1 ,..., w P −1 } . The solution of
(7.11) may be expressed as
Source Sensor
array (a) array
Source
array (b)
Figure 7.5: Source and sensor arrays in borehole tomography. (a) A target lies
between two boreholes. In borehole# 1 sources are arranged as a ULA of sources and
in borehole# 2 sensors are arranged as ULA of sensors. (b) In another arrangement
the sensor array is on the surface and the source array is in the borehole.
Figure 7.6: A target consisting of 16 cells (unit cells) lies between two boreholes
separated by 16 units. An eight sensor array is located in the right hole and the
source array (up to five sources) is in the left hole. The center of sensor array and
source array is at the average depth of the target.
τ = ( w̃ T w̃)−1 w̃ T t̃ (7.12)
We shall now consider a specific example of source and sensor arrays used in
borehole tomography in geophysical exploration [3]. A typical arrangement in
borehole tomography is shown in fig. 7.5. The sources are fired sequentially
and the transmitted signals are recorded for later processing.
Source/Sensor Array Design: For successful reconstruction of wave speed
T
variations the primary requirement is that w̃ w̃ in (7.12) must be invertible.
Since w̃ is entirely determined by the source and sensor array geometry, it is
possible to come up with a proper design for the source and sensor arrays which
would make the rank of w̃ equal to the number of cells. A simple numerical
T
example is worked out to show how the rank of w̃ w̃ depends on the number
of sources, source spacing, and sensor spacing. The interesting outcome of this
exercise is the fact that the number of sensors need not be greater than the
number of cells. One may achieve by using multiple sources what could be
achieved by using more sensors.
The source and sensor array geometry along with the target location are
shown in fig. 7.6. The sensor spacing (d units) and source spacing (s units) are
the variable parameters and the rest of the geometry remains fixed. From each
source eight rays (straight lines) were drawn toward eight sensors. The line
intercept in each cell was found and the weight coefficient was computed as
described on page 370 (also see fig. 7.4). The weight matrix w̃ is first
T
computed and then the rank of w̃ w̃ , whose inverse is used in (7.12). The
T
results are shown in table 7.1. To achieve the full rank property for w̃ w̃ we
must have three to five sources and the sensor spacing should be around 1.5
units. Note that we have considered only eight sensors which is half the
number of cells. We have compensated for this deficiency by using three to five
sources. When the source array is close to the target the angular width of the
illuminating beam becomes large, which in turn requires a large aperture sensor
array to capture the wide illuminating beam. However, indefinite increase of the
sensor spacing will not help. There exists a range of sensor separation over
T
which not only w̃ w̃ is full rank but it is also stable as shown in fig 7.7
whereas for sensor separation between 1.5 and 3.0 units the eigenvalue spread
T
of w̃ w̃ is low and the matrix becomes singular outside the range 1.0 to 4.0.
The above findings are specific to the geometry of source and sensor arrays and
the target; nevertheless similar behavior is expected in other situations.
10 3
10 2
0 1 2 3 4 5
Sensor separation
Figure 7.7: Eigenvalue spread as a function of sensor separation. Five sources are
spaced at interval 1.5 units. The target used is same as in fig. 7.6.
et al. [5] who based their work on Wolf's work [6] on the inverse scattering
problem assuming the first order Born approximation. A good review of
diffraction tomography may be found in [1, 7].
Linear Array: An object is illuminated from various directions with a diffracting
source of radiation such as acoustic waves whose wavelength is comparable to
the scale of inhomogeneities. The incident wave energy is scattered in all
directions by the diffraction process within the object. A long linear array of
sensors facing the incident wave field is used to record the forward scatter (see
fig. 7.8). In §1.6 we derived an expression for the scattered field in the x-y
plane due to a plane wave traveling in z direction and illuminating a three
dimensional object (spherical). A similar result for a two-dimensional object
(cylindrical) was also given. For the sake of simplicity we shall talk about
tomographic imaging of a two-dimensional object. Consider an arrangement
wherein a cylindrical object is illuminated with a plane wave traveling at right
angle to the axis and a linear array of sensors located on the opposite side as
shown in fig. 7.8. The Fourier transform of the scattered field, which is
measured at a set of discrete points by the sensor array, may be obtained from
(1.82) where set u0=0 and v0=k0,
Object
Plane wave
Figure 7.8: A basic experimental setup for tomographic imaging. A linear array,
ideally of infinite aperture, is used to receive the forward scatter from the object.
j k 2 −u 2 l
jk 2 e 0
P0 (u) = 0 ∆c̃(u, k02 − u 2 − k0 )
2 k0 − u
2 2
(7.13a)
u ≤ k0
k0 2k0
Figure 7.9: As u varies from - k0 to +k0 P0 (u) traces a cross section of ∆c̃(u, v)
along a semicircular arc (thick curve).
j k 2 −u 2 l
jk02 e 0 u cosϕ + ( k02 − u 2 − k0 )sin ϕ,
Pϕ (u) = ∆c̃
2 k02 − u 2 −usin ϕ + ( k 2 − u 2 − k )cosϕ
0 0 (7.13b)
π π
u ≤ k0 and − ≤ ϕ ≤
2 2
The center of the circular arcs will all lie on a circle of radius k0 centered at
(−k0 sin ϕ , − k0 cos ϕ ) (see figure 7.10).
7.2.1 Filtered Backpropagation Algorithm: This is an adaptation of the filtered
backprojection algorithm developed for nondiffracting radiation to diffracting
radiation. The essential difference is that the sampling paths are now arcs of a
circle instead of radial lines in nondiffracting radiation. We start by expressing
the Fourier integral in polar coordinates (see (7.5-7.7))
π ∞
1 1
2π ∫0 2π −∫∞
js( x cos ϕ + y sin ϕ)
f (x, y) = F(s,ϕ) s e ds dϕ
2k0
Figure 7.10: The entire 2D Fourier transform, ∆c̃(u, v) , is sampled along a series of
semicircular arcs by rotating the object keeping the transmitter and array fixed.
Because u must lie within - k0 to +k0 the radius of the disc spanned by the
semicircular sampling arcs is equal to 2k0 .
π ∞
1 1
2π ∫0 2π −∫∞
= F(s,ϕ) s e jst
ds dϕ
π
1
2π ∫0
= f̃ (x cosϕ + ysin ϕ)dϕ
where
1 ∞
f̃ (x cosϕ + ysin ϕ) = ∫ F(s,ϕ) s e js( x cos ϕ + y sin ϕ) ds
2π − ∞
π
1
2π ∫0
δc̃(x, y) = ˜ cosϕ + ysin ϕ)dϕ
δc̃(x (7.14a)
where
k0
2 − j k 2 −u 2 l
k02 − u 2 e 0
1 2
δc̃(x
˜ cosϕ + ysin ϕ) = ∫ jk0 (7.14b)
2π − k 0 ×P (u) u e ju( x cos ϕ + y sin ϕ) du
ϕ
According to (7.14b) the scattered field measured by the linear array is filtered
with a filter whose transfer function is given by
− j k 02 −u 2 l
H(u) = k02 − u 2 e u (7.14c)
− j k 2 −u 2 l
The term of interest in the filter transfer function is e 0
which
represents backward propagation of the wavefield from the plane of observation
to the target (see §1.2.4). For this reason the reconstruction algorithm described
above is called the filtered backpropagation algorithm. Except for this difference
the algorithm is quite similar to the filtered backprojection algorithm.
7.2.2 Multisource Illumination: There are many situations where it is not
possible to turn an object around for multiple illuminations nor is it possible
to turn around the source-array configuration, keeping the object fixed as the
space around the object may not be accessible as is the case in geophysical
imaging, nondestructive testing, remote monitoring, etc. In such a situation it
is recommended to employ an array of sources, often arranged as a linear
equispaced array. A typical example is borehole tomography which we have
already considered in the previous section in the context of algebraic
reconstruction. In this section we shall reconsider the same in the context of the
backpropagation algorithm [8]. But, first, let us look at a simpler system, a
circular array of transceivers of interest in medical imaging (see fig. 7.11).
Circular Array: The back scatter is lost in a linear array tomographic system;
naturally, some potentially useful information is also lost, in the sense that
only a part of the object spectrum lying within a disc of radius equal to
2k0
Forward
Scatter
χ0
χ
x
Backward
Scatter
Transmitter/ receiver
v
(b)
2k0 Direction of
illumination
χ0 0
45
u
Figure 7.11: (a) An experimental setup of transceivers for tomographic imaging. (b)
∆C(u, v) is now sampled over a circumference of a circle of radius k0 as shown. The
angle of incidence of the plane wave illumination is 45o.
Ps (R, χ, χ 0 ) =
π
k02 j ( k 0 R+ 4 ) 2 (7.15)
e ∆C(k0 (cos χ − cos χ 0 ), k0 (sin χ − sin χ 0 )
4 πk0 R
The left hand side is simply the observed scattered field on a large circle. The
right hand side is a Fourier transform of the object function which is evaluated
on a circle of radius k0 and centered at kx = −k0 cos χ 0 and ky = −k0 sin χ 0
(see fig. 7.11).
By changing the angle of incidence of the wavefront, χ 0 , it is
possible to cover the Fourier plane with a series of circles spanning a disc of
radius equal to 2k0 (see fig. 7.12). The increased coverage has been possible
because we captured the back scatter as well. Note the important difference is
that the scattered field measured with a circular array, being in the far field
region, directly yields the object Fourier transform. On the contrary, with a
linear array we need to Fourier transform the observed field. This important
result is first verified against the measured scattered field.
Verification of Fourier Diffraction Theorem: First we shall verify the FDT
through an example where an exact scattered field as well as its object Fourier
transform are known. Consider a liquid cylinder in water and assume that its
λ
refractive index is slightly above that of the water, δn ≤ where δn is
4a
2 k0
~ J ( k a)
O( k 0 (s i n χ − s i n χ 0),k 0 (cos χ − cos χ 0 )) = 2 πδ n a 1
k
where
2 2
k= (s i n χ − s i n χ 0 ) + (cos χ − cos χ 0) k0
kx k
cos χ − cos χ 0 = , sin χ − sin χ 0 = y (7.16)
k0 k0
Solving the above equations for χ 0 , we get the following inverse mapping
functions:
2 2
−ky − kx −1 ky − kx −1
−1 p −1 p
χ 0 = tan χ = tan (7.17)
−kx + ky 2 2
−1 kx + ky −1
p p
kx2 + ky2
where p = . Equations (7.16) and (7.17) together give a set of
2k02
transformation equations that can be used to map from the k-plane into the χ -
plane. Every point in the k-plane is mapped onto the χ -plane. The values
of (χ, χ 0 ) thus obtained may not correspond to any of those points where the
0.03
Field in arbitrary units
0.02
0.01
0.00
-0.01
0 90 180 270 360
Angle in degrees
F igure 7.13: The scattered field measured by a circular array is now compared with
the Fourier transform of a uniform cylindrical object evaluated on a circle centered at
−k0 −k0
( , ) as shown in fig. 7.11. The mean square error is 1.759x10-5.
2 2
O(χ, χ 0 ) = ∑ ∑ O(χ , χ )h (χ − χ )h (χ
i j
i j 1 i 2 0 − χj)
χ χ
where h1 (χ) = 1 −
, for χ ≤ ∆χ otherwise =0 and h2 (χ 0 ) = 1 − 0 ,
∆χ ∆χ 0
for χ 0 ≤ ∆χ 0 otherewise =0. Here ∆χ and ∆χ 0 are the sampling intervals.
Once the values of the Fourier transform are obtained over a rectangular grid in
(kx, ky) space, the inverse two dimensional Fourier transform can be computed
to obtain the object profile. The above algorithm is essentially an adaptation of
the frequency domain interpolation algorithm, which is known to be very fast
[1].
7.2.4 Imaging with a Circular Array: Since a circular array captures the entire
diffracted energy, that is, both forward and backward scattered energy, a greater
part of the object spectrum is utilized, indeed twice that of forward-scatter-only
(linear array) setup. Consequently, we expect a better resolution of small
inhomogeneities. To demonstrate this, we have c arried out the following
0.004
Relative refractive index
0.000
0.012
(b)
0.008
0.004
0.000
-16 -8 0 8 16
Distance in λ
F igure 7.14: An example of improved performance of the circular array over linear
array. A cylinder of radius 2λ with a small inhomogeneity of radius 0.25λ embedded
in it (see inset in (b)) is used as a test target. (a) Reconstruction using a linear array
and (b) using a circular array.(From [18] with permission.)
.005
.000
Relative refractive index
-.005
.015
c d
.010
.005
.000
-.005
-4 -2 0 2 4 -4 -2 0 2 4
Distance in λ
Figure 7.15: A comparison of performance of a linear array of finite size with that of
a circular array. The number of illuminations in all cases was 64. (From [18] with
permission.)
considered a linear array of the same length and one transmitter located on the
broad side, and the scattered field was calculated using the object Fourier
transform over semicircular arcs. For a circular array, however, the scattered
field was computed using the exact solution given in [16]. The reconstruction
(a central slice) of the target is shown in fig. 7.14. The reconstruction obtained
using a linear array is shown in fig. (7.14a) and that obtained using a circular
array is shown in fig. (7.14b). Clearly the circular array outperforms the
equivalent linear array as the small inhomogeneity is more accurately located.
Next, we would like to emphasize the role of the array size on object (fig. 7.15)
χ 0
0 45
kx
Circle of radius k 0
Figure 7.16: (a) For narrowband the scattered field is proportional to the object
Fourier transform evaluated on a circle. (b) For finite band, the scattered field is
proportional to the object Fourier transform evaluated inside a crescent shaped
region. (From [18] with permission.)
reconstruction. When using a linear array, it is necessary that the array output
be Fourier transformed before it is used for reconstruction. Consequently, the
errors in the Fourier transformation due to finite size of the array will degrade
the reconstruction. This effect is demonstrated in fig. 7.15. The first three
figures (7.15a, b, c) were obtained using a linear array of three different sizes,
namely, 64, 128, and 512 receivers spaced at 2λ and 100 λ away from the
object, and a cylinder of radius 1λ with a refractive index contrast of 0.01. The
scattered field was computed using the exact solution given in [16]. The
reconstruction shown in fig. (7.15d) was obtained using a 64 element circular
array (radius=100λ). The reconstruction obtained with the circular array is
superior to that obtained with a linear array of a much larger size (512
receivers). Notice that the side lobes have practically disappeared .
f u2 − f l2
r= (7.18)
4 f u2
limited number of views, often covering a finite angular interval, are likely to
be available, leaving large gaps in the Fourier plane. As shown earlier,
broadband illumination can help to reduce the gaps. Signal processing tools
have also been suggested for extrapolation of the observed Fourier transform
into the missing gaps. A priori information about the object, such as a limit
on the support of the object function either in space or frequency domain, does
a
0.010
0.000
-0.010
0.010
Relative refractive index
b
0.006
0.002
-0.002
0.015
c
0.010
0.005
0.000
-0.005
-16 -8 0 8 16
Distance in λ
Figure 7.18: The effect of the lower cut-off frequency on the reconstruction. (a)
1500Hz, (b) 500Hz and (c) 250Hz. The upper cut-off frequency is 3000 Hz. (From
[18] with permission.)
not help to uniquely reconstruct the object function from a limited number of
samples on an algebraic contour in the Fourier plane [19]. Extrapolation
outside the frequency domain, where the observations are available, has been
attempted using the principle of maximum entropy [21, 22] which is familiar
to the signal processing community as it is extensively used to extrapolate the
covariance function for spectrum estimation [20].
u u u u
F1 (u1 , u2 ) = 2k02 ∆c̃( 1 + 2 , k02 − ( 1 )2 − k02 − ( 2 )2 )
d d d d
u u
− j k02 − ( 1 )2 La − j k02 − ( 2 )2 Lb (1.91)
e d e d
×
u u
k02 − ( 1 )2 k02 − ( 2 )2
d d
λ0 2u 2u
Let d = , u1′ = 1 , and u2′ = 2 . For π ≤ u1 ,u2 ≤ −π it turns out that
2 λ0 λ0
k0 ≤ u1′, u2′ ≤ −k0 . A point in the Fourier plane, (u1′, u2′ ) , would correspond
to a point (u,v) in the Fourier plane of ∆c̃ where
u = u1′ + u2′
{ }
(7.19)
v=± k02 − (u1′)2 − k02 − (u2′ )2
[v ± ] + (u − u′)
2
k02 − (u1′)2 1
2
= k02 (7.20)
For a given value of u1′ equation (7.20) describes a circle with radius k0 and
centered at (u1′, ± k02 − (u1′)2 ) , for example; for u1′ = 0 the two circles are
A'
- k0 o k0 u
B'
B (a)
A'
B'
B
(b)
Figure 7.19: (a) The object Fourier transform lying inside two circular disks is
scanned along a series of semicircular arcs. (b) As an example, consider four plane
wavefronts A, A´, B, B´ emitted by distant sources in the source array.
Corresponding circular arcs are shown in (a). The object Fourier transform is
scanned on semicircle AO by wavefront A and on OB by wavefront B. Wavefronts A´
and B´scan the object Fourier transform on A´OB´.
O'
A k0
u
- k0 o k0
O''
u'2 = 0
Figure 7.20: For a fixed u1′ , the object Fourier transform is scanned over a circle A.
For different values of u2′ , that is, for different directions of the beam, a series of
circles will intersect the circle A over an arc O´´OO´. The object Fourier transform
will be sensed over this arc only.
centered on the y-axis at ±k0 . In fig. 7.19 we show different circles (arcs) for
different values of u1′ ; in particular, the thick arcs are for u1′ = 0 and
k
u1′ = −k0 and the thin arcs are for u1′ = ± 0 . Similarly, by eliminating u1′
2
from (7.19) we obtain
[ ] + (u − u′ )
2
v ± k02 − (u2′ )2 2
2
= k02 (7.21)
which describes a circle with radius k0 and centered at (u2′ , ± k02 − (u2′ )2 ) .
For a fixed u2′ the object Fourier transform is scanned over a circle, for
example, A in fig. 7.20.
Let the receiving array (ULA) be beamed to receive the wavefield in
some particular direction, that is, for some fixed u2′ . For u2′ = 0 the circle
described by (7.21) will intersect circle A at two points, namely, O ′ and O (see
fig. 7.20). The object Fourier transform is sensed only at these points of
intersection. For different values of u2′ , that is, for different directions of the
beam, we can draw a series of circles which will intersect the circle A over an
0.05
0.04
0.03
0.02
0.01
-0.01
0 10 20 30 40 50 60 70
Figure 7.21: Maximum entropy reconstruction, dashed line (--) filtered object and
solid line (__) maximum entropy reconstruction. There is only a marginal
improvement in the maximum entropy reconstructed object.
π
k 2 j ( k 0 R+ 4 ) 2
Ps (R, χ 0 ) = 0 e ∆C (−2k0 cos χ 0 , −2k0 sin χ 0 ) (7.25)
4 πk0 R
where R now stands for distance to the center of the object whose size is
assumed to be much smaller than R, so that the far field approximation hold
good. The experimental setup is shown in fig. 7.22a. The backscatter
coefficient at a fixed frequency, Ps (R, χ 0 ) , is proportional to the object
Fourier transform at spatial frequency (−2k0 cos χ 0 , −2k0 sin χ 0 ) where χ 0
is the angle of illumination (see fig. 7.22b). If we now use a broadband signal
for illuminating the object (keeping χ 0 fixed) and Fourier decompose the
χ0
χ0
u
x
Backscatter
Transceiver
(a) (b)
F igure 7.22: (a) A single transceiver is used to illuminate an object and receive the
backscatter. A plane wavefront is assumed to be incident at an angle χ 0 . (b)
Backscatter coefficients are proportional to the object Fourier transform over a radial
line as shown.
received signal we shall obtain the object Fourier transform over a radial line at
an angle χ 0 (see fig. 7.22b). By illuminating the object repeatedly over 360o,
either by physically taking the transceiver around the object or rotating the
object around its axis, we can cover the entire Fourier plane. This commonly
used experimental setup was suggested by [13]. The above result is akin to the
Fourier slice theorem of nondiffracting tomography (§7.1). Naturally, many of
the reconstruction methods developed for transmission (nondiffracting)
tomography, in particular, backprojection, backpropagation and interpolation
methods, can be used in the present case of reflection tomography.
Additionally, there is further similarity between the transmission tomography
and reflection tomography. As noted in §7.1 the projection of an object is equal
to the line integral of some physical quantity (e.g., absorption in X-ray
tomography) over the ray path. A similar physical insight can be given to
reflection tomography.
7.4.1 Line Integral: Let a broadband plane wavefront be incident on a scattering
object. A series of echoes will be emitted as the wavefront penetrates the
object. At any time instant the receiver will receive echoes from all scattering
elements which lie on a surface (see fig. 7.23). Let f(x,y) be the reflectivity
function
Different variables appearing in (7.26) are illustrated in fig. 7.24 The limits on
integrals suggest that the object is of infinite size. But in practice the object is
finite and the transceiver is placed outside the object. To overcome this
difficulty we shall assume that the reflectivity is zero outside the domain of the
object.
For fixed χ 0 we have a waveform which is a function of time or ρ
ρ
(=ct). We then compute a Fourier transform of the waveform ps ( , χ 0 ) . After
c
simplification we obtain
Ps (k, χ 0 ) =
∞ 2π
(7.27)
∫ ∫ f (r,θ)exp(− jk(
0 0
r 2 + R2 + 2rRcos(θ − χ 0 ))rdrdθ
r 2 + R2 + 2rRcos(θ − χ 0 )
r2
≈R+r cos((θ−χ 0 )+ (1−cos2 (θ−χ 0 )))
2R
we obtain
∞ 2π r2
− jk ( (1− cos 2 (θ− χ 0 )))
∫ ∫ f (r,θ)e
− jkR − jkr cos(θ− χ 0 )
Ps (k, χ 0 ) = e 2R
e rdrdθ
0 0
(7.28)
It is easy to see that
Object
Transceiver
Figure 7.23: At any time instant the receiver will receive echoes from all scattering
elements which lie on a surface, s.
s
(r, θ)
θ
x
ρ
R
Transceiver
Figure 7.24: Symbols used in (7.26) are explained in the figure above.
Ps (k, χ 0 ) = e − jωR ×
∞ ∞ ( x sin( χ 0 )− y cos( χ 0 )) 2
− jk (7.29a)
∫∫ e − jk ( x cos( χ 0 )+ y sin( χ 0 )) dxdy
2R
f (x, y)e
−∞ −∞
If we assume that the object size is much smaller than R, the middle term in
(7.29a) may be set to 1, leading to a simple Fourier transform relation,
+∞
1 2k02δc̃e j ( k 0 r − r ′ )
4π ∫− ∫∞ ∫
f 1 (r, t) = f 0 (r' , t)dx ′dy ′dz ′
r − r′
j (k 0 r s − r ′ )
e
f 0 (r' , t) = e − jω 0 t
r s − r′
where r s is source position vector. Since the source and detector are at the
same location, r = r s . Using the point source illumination expression given
above in (1.77) we obtain
+∞
1 − jω 0 t 2k02δ c̃e j 2( k 0 r − r ′ )
f 1 (r, t) =
4π
e ∫− ∫∞ ∫ r − r′ 2 dx ′dy′dz ′ (7.30a)
Object
Figure 7.25: Uniform planar array of transceivers above a three dimensional object.
f 1 (x,y,z=0, ω 0 )=
1
+∞
2k δc̃( x ′, y ′, z ′ )e
2 [
j 2( k 0 ( x − x ′ ) 2 +( y − y ′ ) 2 +( z − z ′ ) 2 ] 2 )
(7.30b)
1
∫ ∫
4π − ∞
∫ 0
which we shall express in a form that enables us to use a result (1.78), derived
in chapter 1 on page 61,
∂ f 1 (x, y, z = 0, ω 0 )
∂ω 0 =
ω 20
1
(7.31)
j
+∞
δc̃( x ′, y ′, z ′ )e [
j 2( k 0 ( x − x ′ ) 2 +( y − y ′ ) 2 +( z − z ′ ) 2 ] 2 )
∂ f 1 (x, y, z = 0, ω 0 )
=
∂ω 0 ω 20
1 + ∞ ∆c̃(u,v, 4k0 − u − v ) j (ux + vy)
2 2 2
∫∫ e dudv (7.32)
2π 2 c 3 − ∞ 4k02 − u 2 − v 2
∆c̃(u,v, 4k02 − u 2 − v 2 ) =
c3 ∂ F1 (u,v, ω 0 )
4k02 − u 2 − v 2 (7.33)
2 ∂ω 0 ω 20
Thus, the Fourier transform of the speed variations is derived from the Fourier
transform of the backscatter measured on a plane surface. It is interesting to
observe that the Fourier transform thus computed actually corresponds to the
Fourier transform of the object on a sphere centered at the origin and with
radius equal to 2k0 (see fig. 7.26). A broadband signal will be necessary to
cover the entire Fourier transform of the speed variation function.
Figure 7.26: The Fourier transform of a reflected signal (echo) corresponds to the
Fourier transform of the object on a sphere centered at the origin and radius equal to
2k 0 .
or 2D Fourier transform of the object. This approach was taken by Milanfar and
co-workers [26] in the context of ray tomography where the input data are
projections of object.
7.5.1 Fourier Transform of Binary Convex Polygonal Object: Consider the
evaluation of the 2D Fourier transform over a p-sided binary and convex
polygonal domain (see fig. 7.27). Take any point inside the polygon and join it
to all corners forming p triangles which lie entirely inside the polygon and
make this as the origin of the coordinate system
∫∫ e
j (ux + vy)
F(u,v) = dxdy
over polygon
p (7.34)
=∑ ∫∫ e
j (ux + vy)
dxdy
n =1 over n th triangle
To evaluate the integral over nth triangle refer to fig. 7.27b where we show the
integration along a narrow strip under the rotated coordinate system such that
the new the x-axis is perpendicular to nth side. Note that this is valid only for
convex objects (for nonconvex objects, it's not possible to drop a perpendicular
from the origin to at least one edge, such that the perpendicular lies entirely
within the object) that are nondegenerate. The triangle is then covered by a
series of strips. Equation (7.34) reduces to
Our goal is to determine(an , bn ) and (an+1 , bn+1 ) from (7.37). This may
be achieved by expressing (7.37) on the k y = 0 and k x = 0 axes. We get
the following equations
th
k triangle x'
y'
ρ
k
φ1k
θk ( a k, b k )
φ
2k
x
e − juan − 1 e − juan+1 − 1
p
u F(u,v = 0) = − ∑ ρn {
2
− }
n =1 an sin θ n an +1 sin θ n
(7.38)
e − jvbn − 1 e − jvbn+1 − 1
p
v 2 F(u = 0,v) = ∑ ρn { − }
n =1 bn sin θ n bn +1 sin θ n
ρn
where Γ n = . From the back scatter due to an illumination at angle
sin(θ − θ n )
π
θ (≠ 0 or ) , we can estimate as described in [20, 27] the coefficients in the
2
exponents of the complex sinusoids. Thus, we get the additional information in
the form of linear combination of the x- and y-coordinates of the corners,
(an cos θ + bn sin θ ) . . n = 1,2, ... p . The steps in the pairing algorithm are
as below:
1) Generate a list of x-coordinates, y-coordinates and the linear combination of
the x- and y-coordinates. It is presumed that the list is not in the same order as
the indexed corners.
2) Take the first element from the x-coordinate list and any one element from
the y-coordinate list and form a linear combination, (a1 cosθ + bn sin θ)
n = 1,2, ... p
3) Compare the result of the linear combination with those estimated with
π
θ (≠ 0 or ) illumination. The best match (within the limits of estimation
2
error) will indicate the correct choice of bn .
4) Take the next element from the x-coordinate list and go to step (2).
For the purpose of illustration we consider a square object of size (6m,6m),
rotated by 30 deg. and shifted away from the origin by (5m,5m). It is
illuminated from three directions, 0, π and π/6. The x- and y-coordinates got
from the noiseless scattered field in the first two directions and their linear
combination (θ=30 deg) are shown in table 7.4 and those estimated from the
scattered field got in the third direction are shown in table 7.5. The application
of the pairing algorithm is illustrated in table 7.6. The best match with the
estimated coefficients is shown in column three in bold figures and the
corresponding y-coordinate is shown in the last column.
We may encounter the problem of repeated x- or y-coordinates or their
projections. The projections of two corners may overlap or come very close to
each other depending upon the orientation of the object. As shown in fig. 7.28,
for a square object depending upon the orientation, the adjacent projections
(e.g., x1 and x2) may come close to each other or overlap. The problem of
Table 7.5: The estimated projections from the scattered field (noise free) are shown
(Source: [27] with permission from IEEE ( c 1998 IEEE).)
y
3
y
2
y
1
x x x x
1 2 3 4
Figure 7.28: A square object and the projections of its corners on the x- and y-axes.
Note that, depending upon the orientation, the adjacent projections (e.g., x1 and
x 2 ) may come close to each other(From [27] with permission from IEEE ( c 1998
IEEE).).
0.08
0.06
Amplitude
0.04
0.02
0.00
0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2
Wavenumber
Figure 7.29: Computed backscatter from a weakly scattering square object (shown in
fig 7.28) illuminated by a broadband signal.(From [27] with permission from IEEE
( c 1998 IEEE).)
π π
number is varied from to π in steps of along three directions,
64 64
namely, x-axis, y-axis and a radial direction at an angle θ =30 degrees. The
backscatter at each wavenumber was computed using the Fourier transform
approach described in [1]. A typical example of backscatter caused by x-axis
illumination is shown in fig. 7.29. To this scattered field sufficient white
Gaussian noise was added so that the snr became equal to a specified figure.
Here the snr is defined as ten times the logarithm (base 10) of the ratio of the
average scattered energy to noise variance. The corners and also the Tn ' were
estimated using the procedure given in [27]. The mean square error (MSE) in
the estimation of coordinates of the corners was studied as a function of snr.
The results, obtained by averaging over fifty independent experiments, are
shown in fig. 7.30a. Notice that MSE rises very rapidly for snr below 8 dB.
This is largely on account of the fact that the projections of two adjacent
corners (e.g., x1 and x4, and x2 and x3 in fig. 7.29) are close to each other; in
this example they are 1.2515 meters apart. For a different orientation, say, at
6o when the separation becomes 0.8362 meters, the MSE rapidly rises for snr
below 15 dB. The estimation error (MSE) also depends upon the number of
corners in a polygonal object. The numerical results are shown in fig 7.30b.
§7.6 Exercises:
1. What is the essential difference between an array of sensors used for DOA
estimation (chapters 2 and 5) or for signal waveform estimation (chapters 3 and
6) and the array used in nondiffracting radiation tomography?
1
mse
.01
3 6 9 12 15
snr (in dB)
.1
.01
mse
.001
.0001
4 5 6 7 8
Number of corners
Figure 7.30: (a) Mean square error (m2 ) in the estimation x- and y-coordinates as a
function snr. (b) Mean square error (m2 ) as a function of the number of corners
(snr=20dB). The dotted lines show the error bounds for 95% confidence.
20m
50m
References
§8.1 Migration:
The interface between two homogeneous layers may be considered as a thin
layer of point sources (scatterers). This forms the basis of the exploding
reflector model [2]. The wavefield observed on the surface of earth can be
extrapolated downward into the earth. The interfaces separating homogeneous
layers reflect or scatter wave energy. Such an interface may be modeled as a
surface with point scatterers whose density is proportional to the impedance
discontinuity. To image an interface, that is, to map the impedance
discontinuity, it is necessary to compute the distribution of the wave energy on
an interface. This problem has been treated as a boundary value problem [3, 4]
or an initial value problem [5]. As a boundary value problem we solve the
wave equation in homogeneous half space with a boundary condition that the
wavefield is given on the surface of the earth (z=0).
us
Measurement plane
eo
ium en
ed g
m omo
H
Backward Extraploation
Figure 8.1: The wavefield measured on a horizontal plane can be extrapolated upward
(forward extrapolation) or downward toward the source (backward extrapolation).
The boundary value problem has been solved in the time domain by
Claerbout [3] by solving a finite difference equation with a time varying
boundary condition and in the frequency domain by Stolt [4] by expressing the
extrapolation as a filtering problem. Extrapolation in the z direction may also
be expressed as propagation in backward time. Note that in the wave equation
the double derivative with respect to z differs from the double derivative with
respect to time only in a scale factor given by -c2. Thus, starting from some
time instant, the wavefield observed at the surface (z=0) is propagated backward
in time, a process that is equivalent to extrapolation of the wavefield to a lower
level (z<0). The boundary value problem may be reformulated as a source
problem with zero boundary condition but driven by an external source which is
given as a time reversed output of each receiver [6]. In another approach
extrapolation is posed as an initial value problem but marching backward [5].
In this approach the time axis is scaled by the wave speed which converts a
recorded seismic section into a wavefield throughout the subspace as it might
have appeared at the latest recording time. Thus the converted wavefield is next
propagated backward in time.
8.1.1 Imaging Conditions: Imaging requires two steps, namely, (i)
extrapolation in space or reverse propagation in time and (ii) an imaging
condition, that is, how to decide when an image has been formed. In optical
imaging convergence of all rays emerging from a point to another point (image
point) is the imaging condition. In seismic or acoustic imaging, the imaging
condition commonly used is when the depropagated field reaches the starting
time which is the time when the scatterer was illuminated or excited. This
information can be found given the wave speed and the distance from the source
[see fig. 8.2]. It is also possible to set the excitation time to zero provided the
scattered wavefront and illuminating wavefront travel along the same path but
Scatterer
in opposite directions. This forms the basis of the popular migration principle
called exploding reflector which we shall describe in detail in the next section.
Qualitatively speaking, imaging is focusing of wave energy. An imaging
condition based on how well the wave energy is focused at a point is also a
likely candidate as an imaging condition. Indeed, in seismic imaging, it has
been suggested that when p-waves and s-waves are focused at the same point, an
image of the point is obtained [7].
8.1.2 Downward Continuation of Sources and Sensors: The source and sensor
arrays are normally placed on the same surface. It is possible to analytically
compute the field when both source and sensor arrays are relocated onto another
plane given the field on the observation plane. For simplicity we shall assume
that both the observation plane and the plane onto which the source and sensor
arrays are relocated are horizontal. We are already familiar with continuation of
wavefield from one plane to another (see chapter 1 and also later in this
chapter). Continuation of a source array requires an additional concept of
reciprocity which states that when the positions of an omnidirectional source
and an omnidirectional sensor are interchanged the observed field remains
unchanged [8]. To extrapolate the source array keeping the sensor array fixed we
need to interchange the source and the sensor arrays and then apply the
wavefield extrapolation algorithm. By virtue of the principle of reciprocity the
result of the above approach will be same as that of actual relocation of the
source array. In actual application the source and the sensor arrays are relocated
alternatively in small steps. As the source and the sensor arrays are continued
downwards towards a reflector at some stage the two arrays will completely
coincide when they reach the reflector after a lapse of time equal to the one way
travel time. Occurrence of such a coincidence of source and sensor may be used
as a condition for imaging. This phenomenon is illustrated in fig. 8.3.
Continuation plane
Continuation plane
Reflector
Figure 8.3: The wavefield measured in one plane is continued to another plane as if
the source and the sensor arrays are located on that plane. It may be recalled that as
the wavefield is continued the wave energy actually propagate along a ray.
ray path
c1 ρ 1
Tiny charges are fired at the
same time instant
z=g(x,y)
c2 ρ 2
Figure 8.4: Exploding reflector model, also known as the Loewenthal model. Tiny
charges are placed on a reflector and fired at the same time instant. The ray paths are
perpendicular to the reflector. The wavefront, at the time of firing, that is, t=0,
coincides with the reflector.
The wavefield in a medium bounded by the earth’s surface above and the
interface below will satisfy the following boundary condition and initial
conditions:
Boundary condition: f (x, y, z,t) z = 0 = f 0 (x, y,t) , that is, pressure field
observed on the surface.
Initial condition: f (x, y, z,t) t = 0 = r (x, y, z)δ(t) , that is, pressure
field generated by the exploding charges on the
interfaces.
The wavefront at time instant t=0 is the interface itself. As the time progresses
the wavefront travels upward toward the surface. The wavefield observed at the
surface acts as a boundary condition, and the shape of the wavefront at t=0 is
the initial condition which, in practice, is not known. The boundary condition
(when given over an infinite plane) is enough to solve the wave equation. The
wavefield thus obtained at all points within the space bounded from above by
the observation plane and from below by the initial wavefront and for all t in
the range 0 ≤ t < ∞ . Observe that the wavefield at t=0 is actually the pressure
field generated by setting off charges on the interface and everywhere else it is
zero. Conversely, if the wavefield observed on the surface is propagated
1 d2 f
∇2 f = + r (x, y, z)δ ′(t)
c02 dt 2
where r(x,y,z) as before stands for the reflectivity function and δ ′(t) is the
derivative of δ(t) . At each point an explosive charge proportional to r(x,y,z) is
set off at t=0. The waves propagate unhindered by other reflecting interfaces (no
multiple reflections). The solution of the inhomogeneous wave equation on the
z=0 surface is given by
f (x, y, z = 0,ω)
+∞
e jk ( x − x ′ ) +( y − y ′ ) +( z ′ )
2 2 2
jω (8.2)
= ∫ ∫
4π − ∞ ∫ r ( x ′ , y ′ , z ′ )
(x − x ′ ) + (y − y ′ ) + ( z ′ )
2 2 2
dx ′dy ′dz ′
+∞
ω R(u,v, w) j (ux + vy)
f (x, y, z = 0,ω) =
4π 2 ∫∫
−∞
w
e dudv (8.3a)
R(u,v, w)
F0 (u,v,ω) = ω (8.3b)
w
w
R(u,v, w) = F0 (u,v,ω) (8.4)
ω
Next we shall show how the same result (that is, (8.4)) can be obtained as a
boundary value problem (see (8.7)).
−∞
1
3 ∫ ∫ ∫ 0
f (x, y, z,t) = F (u,v,ω)e + j k 2 −u 2 − v 2 z − j (ux + vz −ωt )
e dudvdω
8π − ∞
± j k 2 −u 2 − v 2 z
We have chosen the positive sign in e as the wavefield is
propagated from surface to the interface where charges are placed, that is,
propagation is towards the source; hence, as per our convention (chapter 1), +ve
is chosen.
8.3.1 Downward Continuation: The wavefield measured on the surface may be
continued downwards to any depth and for all times (see (1.26)). Using the
initial condition in the exploding charge model the wavefield at time t=0 is
equal to the reflectivity function,
ω
Note that in (8.2) k= where c is the wave speed in the medium above the
c
interface and it is assumed to be known. Further, we relate the temporal
ω s2
frequency to the vertical spatial frequency, w. Since = −w 1 + 2 , k=
c w
ω sc
where s = u + v , we can express w = − 1 − ( )2 for an upgoing
2 2
c ω
c
wave and dω = − dw . Using these results in (8.5) we obtain [10]
s2
1+ 2
w
r(x, y, z) =
∞
c 1 s 2 − j (ux + vy + wz )
8π 3 ∫ −∫∞ ∫
F0 (u,v, −c w 1 + 2 )e dudvdw (8.6a)
s2 w
1+ 2
w
r(x, y, z) =
∞
c 1 c s 2 − j (ux + vy + wz )
16π 3 ∫ −∫∞ ∫
F0 (u,v, − w 1 + )e dudvdw (8.6b)
s2 2 w2
1+ 2
w
2ω sc
where w=− 1 − ( )2 . This result agrees with that given in [9].
c 2ω
Computing the inverse Fourier transform on both sides of (8.6) we obtain
c s2
R(u,v, w) = F0 (u,v, − cw 1 + ) (8.7a)
s2 w2
1+ 2
w
or
1 c c s2
R(u,v, w) = F0 (u,v, − w 1 + 2 ) (8.7b)
s2 2 2 w
1+ 2
w
∞ ∞
1
∫ F (s,ω)e dω ∫ sJ0 (sr)e ± j ( k 2 −s2 z)
f (r, z,t) = 2 0
jωt
ds
4π −∞ 0
8π 2 −∞ 0
(8.8)
where r1 and r2 are reflection coefficients and g1 (x, y) and g2 (x, y) surfaces
separating the two layers. It may be noted that by removing the multiple
reflections we have decoupled the two interfaces; in effect we have linearized the
propagation effects. Extension to the N-layer medium is straightforward when
all layers are decoupled.
For extrapolation it is necessary that the correct propagation speed of
the material in each layer is used. First, extrapolation to the first interface is
carried out using (1.26). The Fourier transform of the wavefield observed on the
surface is multiplied with the propagation filter function
f 2 (x, y, z, t) t = 0
−∞
1
∫ ∫ ∫ F (u, v, ω)e
− j k12 −u 2 − v 2 z1 − j k 22 −u 2 − v 2 ( z − z1 ) − j (ux + vy)
= 0 e e dudvdω
8π 3
−∞
c2 w
−∞ F0 (u, v, ±c2 sgn(ω) w 2 + u 2 + v 2 )
1 w + u2 + v2
2
= ∫∫∫
8π 3
−∞
e
−j { }
k12 −u 2 − v 2 − k 22 −u 2 − v 2 z1
e − j (ux + vy − wz ) dudvdw
= r2 (x, y, z)
(8.11)
d z
where τ0 = sin θ 0 , and t0 = 0 cosθ 0 . Note that z0 is a depth to the
c c
interface below sensor m=0. The Fourier transform of the zero-offset data may
be obtained assuming the array size is very large,
Using (8.12) in (8.6a), that is, its 2D equivalent, we obtain after simplification
z=0
ρ1c 1
z=z1
ρ2c 2
z=z 2
ρ3c 3
(a)
Source/sensor array
z=0
ρ c1
1
∆ z1
ρ2c 2
∆ z2
ρ 3c 3
(b)
Figure 8.5: Wavefield extrapolation in two layer medium. (a) Horizontal layers and
(b) inclined layers. represents transceiver.
2π −∫∞
cos θ1 − j ( − tan θ 0 x + z )w
= cosθ 0 e e dw
z = tan θ 0 x + z0 (8.14)
It is interesting to note that the slope of the interface is not equal to the slope
of the line joining the spike arrivals in the zero-offset seismic data (see
exercise 3 at the end of this chapter).
8.3.4 Depropagation of Wavefield: Extrapolating a wavefield backward in time
is known as depropagation; when continued it is possible to home onto the
starting point. In the exploding reflector model all scatter points are
simultaneously fired. We shall now show that the field observed on the surface
is equal to the field generated at scatter points and then propagated to the
surface. By depropagating, we hope to obtain the field generated at the
exploding reflectors. The zero offset data provides only a blurred image of the
reflectors, but by depropagating the blurring can be reduced [5]. Consider a 2D
model with uniform wave speed with reflecting facets. The wavefield in a
uniform medium is given, in the frequency domain, by
+∞
1
∫ ∫ F(u,ω)e
j ( −ux + k 2 −u 2 z ) jωt
f (x, z,t) = 2 e dudω
4π −∞
+∞
c 1
∫∫ F(u,ω)e − j (ux + vz )e − j u 2 + v 2 ct
= 2 dudv (8.15)
4π −∞ u2
1+ 2
v
Exploding
charges θ1
Figure 8.6: A sloping reflector and source/sensor (transceiver) array on the surface.
By a process of stacking, the seismic field is reduced to a field generated by
hypothetical exploding charges on the sloping interface.
+∞
c 1
∫∫ F(u,ω)e − j (ux + u 2 + v 2 ct )
f (x, z = 0,t) = dudv (8.16)
4π 2 −∞ u2
1+ 2
v
Since the wavefield is by and large vertically propagating, u << v (8.17) may
be approximated as
+∞
z 1
∫ ∫ cF(u,ω)e
− j (ux + 2vz )
f (x, z,t = ) ≈ 2 dudv (8.18)
c 4π −∞
The wavefield observed on the surface (z=0) is mapped into the (x,z) plane by
z
substituting for t. Note that the time axis of the recorded seismic data
c
2
+∞
z 1
∫ ∫ cF(u,ω)e
j ( −ux + 2vz )
f (x, z = 0,t = )≈ 2 dudv (8.19)
c 4π
2 −∞
Thus, from (8.18) and (8.19) we have an important approximate result which
provides an initial wavefield for depropagation,
z z
f (x,0,t = ) ≈ f (x, z,t = ) (8.20)
c c
2
In words, the wavefield observed on the surface with its time axis mapped into
z
a z-axis ( t = ) is approximately equal to the wavefield at the reflector which
c
2
is then propagated to the surface.
z z
Now given f (x,0,t = f (x, z,t = ) we need to
) or
c c
2
depropagate till we reach the point of reflection. To depropagate by ∆t time
units we shall use (8.15) with approximation u << v
c
+∞ z − j (ux − vz ) − jv c∆t
=
4π 2 ∫ ∫ FT f (x,0,t = c e
e dudv
−∞ 2
where the subscript d stands for depropagated field (see fig. 8.7)
8.3.5 Relation to Diffraction Tomography: We shall now show that a close
relation exists between imaging by extrapolation and diffraction tomography
which we have dealt with in the previous chapter. Consider a weakly
inhomogeneous object (speed fluctuations only) illuminated by a point source
and the scattered field being received by a sensor kept close to the source. We
have shown in chapter 7 (pp. 400-402) that the Fourier transform of the speed
variations evaluated on a spherical surface of radius 2k is related to the Fourier
transform of the field received by an array of transceivers (see fig. 7.25). The
relationship (7.33) is reproduced here for convenience:
∆t
t=z/c
ddDvvvv
t ∆t
Figure 8.7: The time axis of the wavefield measured on the surface is mapped onto
the z-axis (the wave speed is assumed to be known). The wavefield thus obtained in
the (x, z) plane is now depropagated in steps of ∆t. The depropagated segment
corresponds to the desired reflectivity function over a segment of depth, ∆t c.
c3 ∂ F1 (u,v, ω 0 )
∆c̃(u,v, 4k02 − u 2 − v 2 ) = 4k02 − u 2 − v 2
2 ∂ω 0 ω 02
It may be observed that the zero offset field measured in seismic exploration is
essentially an output of an array of transceivers which may be considered as a
mathematical model of the zero offset field measurement setup.
To relate ∆c̃(u,v, 4k02 − u 2 − v 2 ) to the reflectivity, we shall use (8.7a)
where we have shown the relationship between the Fourier transform of the
reflectivity and the wavefield on the surface (zero offset). It can be shown that
the partial derivative appearing in (7.33) may be given by
∆c̃(u,v, 4k02 − u 2 − v 2 )
c3 ∂ R(u,v, w) (8.23)
=− w2 + s2
4 ∂w w 2 + s 2 w
After carrying out the required differentiation we obtain the following result:
∆c̃(u,v, w)
c 3 R(u,v, w) R(u,v, w) Rw (u,v, w) (8.24)
= + −
4 (w 2 + s 2 ) w2 w
where the subscript r refers to the receiver coordinate. In the first step the
sources are held fixed on the z=0 plane (observation plane) and sensor array is
displaced downward. Next we interchange the source array with the displaced
receiver array. By virtue of the reciprocity theorem the wavefield at the new
position of the sensor array due to the source array also at its new position will
be equal to the previously downward continued field. In the second step we
downward continue the previously continued field but the continuation is done in
source coordinate space. The transfer function for downward continuation is
given by
Source/Sensor
array outputs
Figure 8.8: M2 outputs source/receiver ULAs are arranged over a square grid. All
outputs in a row are known as source gathers and all outputs in a column are known
as sensor gathers.
(a)
reflector
(b)
Figure 8.9: Common depth point (CDP) setup. (a) The source and receiver arrays are
arranged in such a manner that a signal is always reflected from the same reflecting
element. (b) Source-receiver pair.
x2
T x2 = T02 + (8.26)
c2
2l
where x stands for separation between the source and receiver , T0 = , and l
c
is depth to the reflecting element (see fig 8.9b). A plot of T x vs x is very
useful for it enables us to estimate T 0 and the wave speed.
Let us now consider a sloping reflecting element but the source and
receiver arrays are, as before, on a horizontal surface (see fig. 8.10). Let the
slope of the reflecting element be α . To compute the travel time to a
downslope or an upslope sensor we consider the image of the source and
compute the distance from the image to the sensor. For a downslope sensor,
(ir+ )
Tx+ =
c
(8.27)
(x+ )2 + 4l 2 − 4lx+ cos(∠isr+ )
=
c
2
= + + T02 + 2T0 + sin(α)
2 x x
T (8.28)
x+
c c
2
T x2− = − + T02 − 2T0 − sin(α)
x x
(8.29)
c c
The upslope and downslope travel times are shown in fig. 8.11.
i image
Figure 8.10: Inclined reflector. There are two sensors, one sensor, r+ , is downslope
and another sensor, r- , is upslope. To compute the travel time we consider an image
of the source. The time of travel to the downslope sensor is equal to (i r+ ) /c and
similarly for the upslope sensor.
7.5
downslope
7
6.5
Time in sec
5.5
upslope
5
4.5
0 10 20 30 40 50 x150
Distance in meters
Figure 8.11: Downslope and upslope travel times (eqs. 8.21 & 8.22) in seconds as a
function of distance in meters. Wave speed =1500m/s, T0 =5sec. Solid curve is for
horizontal reflector (8.19).
2 2
T x2± = T0 ± ± sin(α) ± ± cos(α)
x x
(8.30)
c c
Let the downslope and the upslope sensor be equidistant from the source, that
is, x+ = x− = x . The average of the squares of downslope and upslope travel
times turns out to be independent of the slope of the reflector,
T x2+ + T x2− 2
= + T02
x
(8.31)
2 c
Similarly, the difference of the squares of downslope and upslope travel times is
given by
T x2+ − T x2− x
= T0 sin(α) (8.32)
4 c
8.4.2 Layered Medium: We now consider a layered medium overlying a
reflector. The overlying medium is modeled as a stack of uniform layers (see
fig. 8.12). The round trip travel time T x and the receiver position x are the two
parameters of interest. They are given by
N
2∆zk
Tx = ∑ (8.33a)
k =1 ck (1 − p 2 ck2 )
N
2∆zk ck
x = p∑ (8.33b)
k =1 (1 − p 2 ck2 )
reflector
Figure 8.12: Layered medium overlying a reflector. Each layer is homogeneous with
constant thickness and speed. The ray path consists of a series of linear segments.
At an interface between two layers the Snell’s law (see p. ) must be satisfied.
sin(θ k )
where p is the ray parameter ( p = ) where θ k is the angle of
ck
incidence in the kth layer (see page for more information on the ray parameter)
and N stands for the number of layers. Let us expand T x as a function of x in
Taylor’s series,
d 2T x x 2 d 4T x x4
T x = T0 + + + (8.34)
dx 2 x=0
2! dx 4 x=0
4!
x2
T x ≈ T0 + N (8.35)
2 ∑ ∆zk ck
k =1
Upon squaring on both sides of (8.35) and retaining only the second order terms
we obtain
∑ ∆z c
k =1
k k
x2 x2
T x2 ≈ T02 + = T 2
+ (8.36)
1 N ∆zk 2 0 2
crms
∑ k
T0 k =1 ck
c
1 N ∆zk 2
where
2
crms = ∑ ck is known as a root mean square speed of the
T0 k =1 ck
layered medium. Indeed we may replace a layered medium by a uniform medium
having a speed equal to crms . For the purpose of focused beamformation we
may use the root mean square speed.
8.4.3 Focusing: To form a focused beam we must have many CDP gathers
which are obtained by means of a specially designed source-receiver array. For
example, consider a linear array where every location is occupied either by a
source or a sensor. The array is fired as many times as the number of locations.
From every reflecting element we obtain a number of gathers equal to the
number of receivers. This is illustrated in fig. 8.13. A ULA of four sensors is
headed by a source. In position (A) the source is fired and the reflected signal is
received by sensor #1. The entire receiver-source array is moved laterally by half
sensor spacing as in position (B) and the source is fired once again. The
reflected signal is received by sensor #2. This procedure is continued as in (C)
and (D). Thus, we get four CDP gathers.
Let T1 ,T 2 ,...T N be the round-trip travel time from the source to sensor #1,
#2,...#N, respectively. Let f 0 (t) be the signal transmitted by the source and
f i (t), i = 1,2,... N be the reflected signals received by four receivers. Since
these are the delayed versions of f 0 (t) we can express them as
T̂1 , T̂ 2 ,..., T̂ N which are computed using (8.26) for an assumed depth to the
reflector,
2lˆ ˆ d
where T̂0 = , l is assumed depth to the reflector and ∆t = . Recall that
c c
d stands for sensor spacing. We assume that the wave speed c is known. The
CDP gathers are coherently summed after correction for the delays computed
from (8.38). As before we shall assume that the source emits a broaband signal.
The coherently summed output may be expressed as follows:
∆tn 2 ∆tn 2
∞
1 1 N −1 − jω T 0 (1+( ) ) − T̂ 0 (1+( ) )
2π −∫∞
∑e
T0 T̂ 0
g(t) = F(ω) e jωt dω
N n=0
(8.39a)
∞
1
2π −∫∞
= F(ω)H N (ω)e jωt dω
where
∆tn 2 ∆tn 2
1 N −1 − jω T 0 (1+( ) ) − T̂ 0 (1+( ) )
H N (ω) = ∑ e
T0 T̂ 0
(8.39b)
N n=0
0.8
0.6
Response
0.4
0.2
0
3 4 5 6 7
Round-trip time in sec.
the aperture size the depth of focus rapidly deteriorates. The minimum depth of
focus appears to be independent of depth to the reflector; however, the required
array aperture increases rapidly as the reflector depth increases. The dependence
of the depth of focus on frequency is significant, as shown in fig. 8.16. Notice
that the depth of focus becomes very narrow beyond about 50Hz.
8.4.5 Inclined Reflector: We consider a sloping reflector. A small segment of
the reflector is illuminated at different angles by means of a source-receiver
array similar to the one used for a horizontal reflector. As the array is laterally
shifted the point of reflection changes, but only slightly, depending upon the
slope (see fig. 8.17). When the slope is zero all CDP gathers emanate from the
same point on the reflecting element. For gentle slope the spread of the
reflecting points is assumed to be small, small enough to allow the assumption
of a constant slope.
1.5
Depth of focus
1.0
0.5
0.0
1 10 100 1000
Array Aperture
Figure 8.15: Depth of focus in units of round-trip propagation time (sec) as a
function of array aperture, also measured in units of propagation time (sec).
0.7
0.6
0.5
Depth of focus
0.4
0.3
0.2
0.1
0.0
0 200 400 600 800 1000
Angular Frequency
Figure 8.16: Depth of focus vs angular frequency. The reflector is at a depth of 5
seconds (round-trip travel time). The array aperture is held fixed at 20 seconds
(propagation time).
l'
Figure 8.17: Common depth point (CDP) gathers from a sloping reflector. The
2l'
reflecting element is at a depth of l' (round trip travel time= ). Two positions of
c
source-sensor array are shown; position #1: circles and position #2: squares. Notice
the displacement of the reflecting point.
The response of the focused beamformer as a function of the slope and the
2l'
return travel time is shown in fig. 8.18. We have assumed that = 5 and
c
x
array aperture, in units of propagation time, is equal to = 20 seconds. There
c
are 200 CDP gathers spaced at an interval equal to 0.1 sec. The result shown in
fig. 8.18 is perhaps the best one can expect as we have assumed the optimum
array size.
8.4.6 Relation Between Focusing and Downward Extrapolation: Focusing
appears to achieve what the downward extrapolation does in migration. Indeed
both are related in the sense that focusing is a simplified version of
extrapolation. In order to see this relationship let us examine the impulse
response function of the downward extrapolation filter whose transfer function
is given by
ω
H(ω,u,v) = exp( j∆z (( )2 − u 2 − v 2 )) (8.40a)
c
0.8
0.6
Response
0.4
0.2
0
6
Ro 5.5 14
un 12
dt 5
rip 10
tim eg
e i 4.5 8 e in d
ns op
ec 4 6 Sl
Figure 8.18: A mesh plot of the response function of the focused beamformer as a
function of slope of the reflector and round-trip time. Array aperture is 20 seconds
(propagation time). Angular frequency assumed is equal to 100 radians.
exp( j ω r) ω
j r +1
1 ∂ c 1 c ω
h(r,ω) = = exp( j r) (8.40b)
2π ∂z r 2π r 3
c
ω ρ2 ∆z 2 ρ2
exp( j r) = exp( jω 2 + 2 ) = exp( jω T02 + 2 ) (8.40c)
c c c c
extrapolation filter are identical to those used in the focused beamformer. The
difference, however, lies in the amplitude term. In place of a variable amplitude,
we use a constant amplitude in the focused beamformer.
8.4.7 Focused Beamformation for Imaging: A focused beam enables us to
estimate the amount of scattered or reflected power from a scattering volume
element or a reflecting surface element located at a point in space. Consider a
volume element illuminated by a point source with the scattered waves being
received by an ULA (see fig. 8.19). To form a focused beam we need to
compute the travel time from the source to the volume element and from the
volume element to a sensor. For simplicity let us assume that the background
wave speed is constant so that ray paths are straight lines. In any realistic
problem, however, the ray paths are more likely to be curved. It will be
necessary to trace the rays before we are able to compute the travel times. Such
a situation was considered in [14]. Let f n (t), n = 1,2,... N be the sensor
outputs. The focused beam output is given by
1 N
g(t) = ∑ f n (t − t(so) − t(orn ))
N n =1
where t(so) stands for the travel time from the source to the volume element
and t(orn ) stands for the travel time from the volume element to the nth
∫ g(t)
2
sensor. The scattered power from the volume element is given by dt
Ts
where T s stands for signal duration. The process is repeated at every point in
the space (x-z plane). The resulting map provides an image of the scattering
strength.
The question of resolution needs to be answered. The size of the
volume element is controlled, along the x-axis, by the array aperture (inversely
proportional to array aperture), along the z-axis, by depth of focus and finally
along the y-axis, by the bandwidth of the signal (inversely proportional to
bandwidth) [14]. An interesting application of the focusing by means of back
propagation of wavefield to detection of an acoustic source in a room without
considering the reflections from the walls is reported in [15]. An array of
microphones is deployed in a 3D space surrounding the source(s) of sound
energy. The output from each microphone is back propagated to a point in
s r1 r2
z
o
Figure 8.19: The array output is focused to a point o where a scattering volume
element is presumed to be located. The power output is proportional to the
scattering strength of the volume element.
x12 − x22
c= (8.42)
T x21 − T x22
1 ∆zk 2
N
2
crms N
=
T0 N
∑
k =1 ck
ck
N +1
1 ∆zk 2
2
crms N +1
=
T0 N +1
∑
k =1 ck
ck
2
crms T
N +1 0 N +1
− crms
2
T
N 0 N
cN +1 = (8.43)
∆z N +1
Now consider an interval containing p layers. Again from (8.36) we obtain after
subtraction
∑c
k =1
N +k ∆z N + k = crms
2
T
N+p 0 N+p
− crms
2
T
N 0 N
(8.44a)
T0 N + p − T0 N
N 0 N
By definition, the right hand side is the interval rms wave speed in the interval
containing p layers
1 p
∆z N + k 2
c2
rms p =
T0 N + p − T0 N
∑
k =1 c N + k
cN + k
(8.44b)
2
crms T
N+p 0 N+p
− crms
2
T
N 0 N
=
T0 N + p − T0 N
xi2
T̂ x i = T02 + 2
+ τi (8.45)
crms
{
T̂ x = col T̂ x1 , T̂ x 2 ,..., T̂ x N }
x2 x2 x2
T m = col T02 + 21 , T02 + 22 ,..., T02 + 2N
crms crms crms
τ = col{τ i , i = 1,2,..., N }
2 2
pdf (T̂ x crms ) pdf (crms )
pdf (c 2
rms T̂ x ) = (8.46)
pdf (T̂ x )
2
pdf (T̂ x crms ) = pdf (τ)
(8.47a)
2
1 1 (c 2 − c 2 )
pdf (c ) =2
exp(− rms rms ) (8.47b)
σ
rms
2πσ 2
−1 (crms
2
− crms
2
)
2
2
Differentiate (8.48) with respect to crms and set the derivative to zero to obtain
∂T m H −1 ∂T m
2
(crms − crms
2
)
3T̂ Hx C−1 − 3T C − 2 =0 (8.49)
∂(crms ) ∂(crms ) σ
2 m 2 2
where
[ ]
− crms
2 2
3 H −1 H −1 (crms )
T C x − T̂ C x = (8.51)
σ
4 m x 2
4crms
N
xi2 T̂ x i =0
∑ 2
i =1 σ τ i
1 −
x2
(8.52)
T02 + 2i
crms
which is equal to the minimum mean square error estimate [20]. Numerical
simulations presented in [20] indicate that the minimum mean square error
estimate is very close to the maximum likelihood estimate which is
2
computationally far more involved. Since the unknown quantity crms in (8.52)
occurs inside the square root term, it is not possible to explicitly solve for it.
x2
To overcome this, we shall introduce an approximation T02 << 2
, which
crms
1 xi2
enables us to replace the square root term by T0 (1 + ) and then solve
2 T02 crms
2
N T̂ x i xi2 xi2
∑ 3 2
1 i =1 T0 σ τ i
2
ĉrms ≈ (8.53)
2 N x 2 T̂ x
∑ 2
i
i =1 σ τ i
i
− 1
T0
∆tn 2 ∆tˆn 2
1 N −1 − jω T 0 (1+( ) ) − T̂ 0 (1+( ) )
H N (ω) = ∑ e
T0 T̂ 0
(8.54)
N n=0
∆x
where ∆tˆ = , ĉ is the assumed speed and ∆x is the basic unit of source
ĉ
∆x
and receiver separation. Note that ∆t = is the actual propagation time.
c
0.5
0.4
0.3
0.2
0.1
0
1400 1450 1500 1550 1600
Wave speed in m/s
Figure 8.20: Response (magnitude square) of focused beamformer as a function of
wave speed. Maximum source sensor separation: 8.5 km (solid line) and 4.5 km
(dashed line). Angular frequency=62.84.
5.4
5.3
5.2
Peak position in sec
5.1
5.0
4.9
4.8
4.7
4.6
1450 1470 1490 1510 1530 1550
Wave speed in m/s
Figure 8.21: Focused beam position as a function of wave speed in m/s. All other
parameters are as in fig. 8.20. The maximum sensor separation is 4.5 km.
§8.6 Exercises
1. A point scatterer is located at a depth l unit below the observation surface. A
transceiver, placed on the surface, records the round trip-travel time. Show that
this travel time is exactly the same as in the CDP experiment with a reflector
replacing the scattering point.
2. Common depth point (CDP) seismic gathers containing an echo from a
horizontal reflector are cross-correlated. Show that the maximum of the cross-
correlation function lies on a hyperbolic time-distance curve. This property is
the basis for a technique of seismic speed estimation known as velocity
spectrum [21].
3. Show that the slope of the line joining all reflected signals in a zero-offset
seismic data over an incline reflector is related to the slope of the reflector
through the following relation, tan α = sin θ 0 where α is the slope of the
line joining reflected signals and θ 0 is slope of the reflector.
References