Electromagnetic Scattering From Random Media
Electromagnetic Scattering From Random Media
Electromagnetic Scattering From Random Media
ON PHYSICS
SERIES EDITORS
Timothy R. Field
McMaster University
1
3
Great Clarendon Street, Oxford OX2 6DP
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Auckland Cape Town Dar es Salaam Hong Kong Karachi
Kuala Lumpur Madrid Melbourne Mexico City Nairobi
New Delhi Shanghai Taipei Toronto
With offices in
Argentina Austria Brazil Chile Czech Republic France Greece
Guatemala Hungary Italy Japan Poland Portugal Singapore
South Korea Switzerland Thailand Turkey Ukraine Vietnam
Oxford is a registered trade mark of Oxford University Press
in the UK and in certain other countries
Published in the United States
by Oxford University Press Inc., New York
c Oxford University Press 2009
The moral rights of the author have been asserted
Database right Oxford University Press (maker)
First Published 2009
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press,
or as expressly permitted by law, or under terms agreed with the appropriate
reprographics rights organization. Enquiries concerning reproduction
outside the scope of the above should be sent to the Rights Department,
Oxford University Press, at the address above
You must not circulate this book in any other binding or cover
and you must impose the same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Field, Timothy R.
Electromagnetic scattering from random media/Timothy R. Field.
p. cm — (International series of monographs on physics; no. 144)
ISBN 978–0–19–857077–6
1. Stochastic processes. 2. Random fields. 3. Mathematical physics. 4. Electromagnetic
waves–Scattering. I. Title.
QC20.7.S8.F54 2008
530.14 1—dc22 2008042115
Typeset by Newgen Imaging Systems (P) Ltd., Chennai, India
Printed in Great Britain
on acid-free paper by
Biddles Ltd., King’s Lynn, Norfolk
ISBN 978–0–19–857077–6
1 3 5 7 9 10 8 6 4 2
To my family
This page intentionally left blank
PREFACE
vii
viii PREFACE
in terms of correlation functions and higher order statistics of various spatial and
temporal physical scattering data. Under certain assumptions concerning the na-
ture of the scattering population, it is possible to deduce that the scattered field
intensity consists of two components, a Rayleigh or Gaussian speckle pattern,
modulated by the local scattering cross-section, which fluctuate over two inde-
pendent correlation timescales. This is the so-called compound K-distribution
model, which has been the extant model for sea-surface radar scattering since its
inception over two decades ago. Since then, although a prominent and intensive
area of research, the problem of understanding the dynamics of the scattering
process has remained unsolved. This problem is of vital significance, since a de-
scription of the scattering dynamics (in terms of propagators) implies the form of
all correlation functions and higher order statistics, and provides a valid prior for
the power spectral density.1 An appropriate dynamical model of these scattering
processes has recently been discovered. From very primitive physical assumptions
it has been possible to provide a detailed description of the stochastic dynamics
of the electromagnetic scattering process that captures all the physical degrees of
freedom in the problem. This discovery and its implications for modelling elec-
tromagnetic scattering and propagation phenomena are the basis of the current
monograph.
The major features of the book can be summarized as follows: this pioneer-
ing study describes electromagnetic scattering from a population whose size is
fluctuating in space and time. Its key findings are to characterize the dynami-
cal laws governing the time evolution of the scattered electromagnetic field; the
physical model enables extraction of all spatio-temporal correlation information
and higher order statistics; this enables the results of radar and laser scatter-
ing experiments to be interpreted, and has implications for real time anomaly
detection; this work is significant also because it illustrates how ideas in the
Black–Scholes theory of financial option pricing can be applied to a physical
problem in which non-Gaussian noise processes play an essential role, revealing
scope for cross-fertilization between these disparate areas.
The primary readership for the book is intended as researchers in academia
and industry, in the areas of electromagnetic scattering, SDEs, diffusion mod-
elling, radar systems engineering, signal processing, applications of non-Gaussian
noise processes, quantum optics, laser propagation, population processes, radar
systems engineering, stochastic volatility models in the context of financial math-
ematics, and those with a general interest in applications of stochastic methods
to physics. The book should also benefit advanced graduate students working in
these areas. Practitioners in these fields should find the book a useful resource in
the design and implementation of the various models in both simulation and ex-
perimental data analysis. Lastly, we have attempted to access a broad readership
1 The reader should compare the remarks in Jakeman and Tough (1988), p. 508: “A full
analysis of the temporal correlation properties of the variables x and z implicit in (5.24) would
require knowledge of its fundamental solution or propagator, which is, as yet, unknown. ...”
et seq.
PREFACE ix
We are grateful for academic support from the Departments of Electrical and
Computer Engineering and Mathematics at McMaster University and the Brain
Body Institute. Amongst those colleagues who deserve special thanks are John
Bienenstock, Alexander Bain, Simon Haykin FRSC, Thomas Hurd, Eric Jakeman
FRS, Vikram Krishnamurthy, John McWhirter FRS, Jose Principe, David Sher-
rington FRS, and Kon (Max) Wong FRSC, and my graduate students Patrick
Fayard and Tao (Stephen) Feng. We especially thank Robert Tough for collabo-
ration at the early stages and inspiring much of this scientific development, and
for contributing the majority of Chapter 11. His constant mentoring and encour-
agement has been an essential part of the success of establishing this research
area. We also thank former colleagues Brian Bramson, Richard Glendinning,
Stephen Luttrell, and John O’Loghlen for valuable discussions, and Samantha
Lycett and Kevin Ridley for supplying experimental data studied in Chapter 12.
The author acknowledges the award of a Discovery Grant from the Natural
Sciences and Engineering Research Council of Canada.
x
CONTENTS
I STOCHASTIC CALCULUS
1 Heat equation and Brownian motion 3
1.1 Einstein’s 1905 derivation 3
1.1.1 Green’s function solution 5
1.2 Brownian motion 5
2 Ito calculus 7
2.1 Ito stochastic integral 7
2.1.1 Ito isometry 9
2.2 Ito differential 9
2.2.1 Ito’s formula 10
2.2.2 Ito product rule 10
3 Stochastic differential geometry 12
3.1 Diffusions on manifolds 12
3.2 Kinematics of diffusion 16
4 Examples of stochastic differential equations 20
4.1 Geometric Brownian motion 21
4.2 Bessel process 21
4.3 Solution to Laplace’s equation 23
xi
xii CONTENTS
8 Dynamics of K-scattering 52
8.1 Stochastic dynamics of K-amplitude process 53
8.1.1 Intensity 55
8.1.2 Phase 57
8.2 Geometry of K-amplitude fluctuations 59
8.3 Asymptotic behaviour 60
8.3.1 Equilibrium distribution 60
8.3.2 Detailed balance 62
8.4 Correlation and spectra 63
8.4.1 Intensity autocorrelation 63
8.4.2 Power spectral density 64
8.5 Interpretation and implications 67
9 Models of weak scattering 69
9.1 Weak scattering amplitudes 70
9.2 Stochastic dynamics 71
9.3 Geometry of amplitude fluctuations 74
9.4 Asymptotic behaviour 79
9.4.1 Detailed balance 85
10 Scattering from general populations 88
10.1 Extended random walk model 88
10.2 Generalized dynamics 91
The part introduces the concepts and techniques of stochastic calculus that en-
able us to describe dynamical (scattering) processes in terms of stochastic dif-
ferential equations (SDEs). Chapter 1 begins with an historical account of the
heat (diffusion) equation in terms of Einstein’s 1905 derivation and the corre-
sponding Green’s function solution. This leads naturally to the concept of the
Wiener process, which describes the trajectory of an individual particle amongst
a cloud whose density evolves according to the heat equation. Equipped with
the Wiener process as an essential ingredient for the random driving force, in
Chapter 2 we introduce the reader to the basic concepts and techniques of the
Ito (and, to some extent, Stratonovich) stochastic calculus that we shall require
in the description of scattering, and highlight its key aspects that differ from
notions of the classical calculus.
Since we shall be concerned with physical scattering processes that are rep-
resented as vectors, and also in part for mathematical edification, we develop
the subject further from the point of view of stochastic differential geometry in
Chapter 3. The part concludes in Chapter 4 with an account of some specific
examples of SDEs and their applications, which serve to illustrate the various
concepts and techniques of the stochastic calculus introduced in the previous two
chapters.
1
This page intentionally left blank
1
HEAT EQUATION AND BROWNIAN MOTION
This chapter explores the interrelationship between the theory of the heat pro-
cess, in terms of the probability density function for a cloud of diffusing particles,
and the underlying stochastic process pertaining to individual particles, which
generates such an ensemble average behaviour that is called Brownian motion.
dn = nφ(∆)d∆. (1.1)
The heat equation concerns the evolution of the density ρ in space and time. To
this end we write
∂ρ
ρ(x, t + τ ) = ρ(x, t) + τ + o(τ ). (1.2)
∂t
On the other hand, we can express the left-hand side as an integral over space,
via the following lucid reasoning. The events that a Brownian particle finds itself
2 The reader may also be interested to compare the (broadly) contemporaneous work of Louis
Bachelier on random walks, in the context of finance, which appears in Bachelier’s doctoral
thesis (orig. Bachelier 1900) and thus predates Einstein’s famous 1905 paper on the Brownian
movement.
3
4 HEAT EQUATION, WIENER PROCESS
where, to be precise, we have used the following three facts: any equivalent
pair of events, such as those identified above, have equal probabilities; distinct
choices of ∆ yield a set of disjoint events and so the probability of their union is
additive (reflected in integral on the right-hand side above); the displacements
are independent of space-time location, and so the joint probability of these
attributes factorizes. Now, equating the two expressions for ρ(x, t + τ ) above is
the essence of Einstein’s argument. Expanding (1.3) in powers of ∆ we obtain
∞
∂ρ 1 2 ∂ 2 ρ
ρ(x, t + τ ) = ρ−∆ + ∆ + o(∆ 2
) (x,t) φ(∆) d∆. (1.4)
∂x 2 ∂x2
−∞
Terms of o(∆2 ) can be neglected since we assume that φ decays rapidly for
|∆| > , 1 and so ∆n φ ≈ 0 (since ∆n ≈ 0 for |∆| < ), n > 2. Also, ∆φ
vanishes since φ is
assumed to be an even function, and since it is a probability
density function φ = 1. Thus we are left with
∂2ρ
ρ(x, t + τ ) = ρ(x, t) + τ D (1.5)
∂x2
in which we define the diffusion coefficient D as
1
D := lim ∆2 φ. (1.6)
τ →0 2τ
Remark. In this derivation of the diffusion coefficient we can see both the
mathematical and physical origins of the identity for the quadratic variation of
the Wiener process, namely the non-vanishing integral of dWt2 = dt, that is
familiar in the modern Ito calculus.
Now comparing (1.5) with (1.2) and taking the limit as τ → 0 we obtain the
heat equation
∂ρ ∂2ρ
= D 2. (1.7)
∂t ∂x
A generalization of Einstein’s argument that incorporates a drift (correspond-
ing to a displacement density φ that is non-symmetric) and retains all higher
derivatives in space is the Kramers–Moyal expansion (see e.g. Risken 1989) that
we shall encounter later in the context of population models in Section 7.1.
BROWNIAN MOTION 5
where s < t ≤ u < v (so the intervals (s, t), (u, v) are non-overlapping) and such
that the distribution of Wt − Ws is the Gaussian N (0, |t − s|), and is subject to
the initialization W0 = 0.
The autocorrelation function of the Wiener process E [Ws Wt ], s ≤ t, can be
calculated by writing the second factor under the expectation as (Wt − Ws ) + Ws
and then applying the above independent increments property, which leads to
E [Ws Wt ] = s. (1.12)
Therefore, the Wiener process, although it has constant mean zero, fails to be
wide sense stationary.
6 HEAT EQUATION, WIENER PROCESS
This process can also be constructed from the continuum limit of a discrete
random walk, as follows. Consider a discrete random walk in discrete time, start-
ing at the origin, with step size and time step δt. Suppose that the displacement
at each step is drawn from the discrete set {, −} with equal probability. Let
the step size and time step scale relative to each other according to (cf. the dis-
cussion of Einstein’s derivation above) 2 = δt, for all δt > 0. Then, from the
central limit theorem applied to the displacement after n steps in the interval
[0, t], where t is fixed and n → ∞, δt → 0, we find that this is N (0, t) distributed
and, by construction, satisfies the independent increments property. The process
so constructed therefore coincides (in probability) with the above Wiener process
defined in the standard (continuum) manner.
2
ITO CALCULUS
This chapter develops the basic tools of the Ito calculus that flow from the devel-
opment of the Brownian motion or Wiener process in Chapter 1. We emphasize
those particular techniques that are relevant to the detailed analysis of the scat-
tering processes encountered in the chapters that follow.
where s < t ≤ u < v, so the intervals (s, t) and (u, v) are non-overlapping;
second, the distribution of the increment Wt − Ws is the Gaussian with mean
zero and variance |t − s|, i.e.,
for some integrable function f (s) (e.g. a continuous one), we shall consider re-
placing the measure ds by the stochastic differential dWs . In general, the inte-
grand will be considered to be a stochastic process σs , the so-called (stochastic)
volatility. As we shall see, in the sections that immediately follow, this concept
of stochastic integration leads to corresponding modifications in the classical no-
tion of derivative, that will be quite essential to our development of scattering
dynamics.
7
8 ITO CALCULUS
Fig. 2.1. Ito stochastic integral – the volatility σs is evaluated at the extreme
left point of each subinterval, before multiplying by the Wiener measure dWs .
wherein the measure of integration (the increment of the Wiener process) is itself
random. Such an integral differs in nature from the classical Riemann integral,
and is illustrated in Fig. 2.1. It is important (unlike in the case of the classical
Riemannian integral, where it is irrelevant) that in the discrete sum approxi-
mating the Ito integral, the integrand is evaluated at the left-most point of each
subinterval. Thus, in the Ito interpretation,
b j
max
where jmax = (b − a)/δs − 1. A detailed comparison of the above Ito integral and
the (intimately related) Stratonovich integral, which evaluates the integrand at
the midpoint of each subinterval, is provided in Appendix B.
Remarks. On expectations of stochastic integrals. Owing to the distributive
property of the expectation functional E over addition, the definition (2.5), and
the symmetry of the fluctuations δWt with respect to sign (cf. the associated
symmetry of the displacement distribution function φ(∆) discussed in Section
1.1), the expectation of a stochastic integral is zero, i.e.
ITO DIFFERENTIAL 9
b
E σs dWs = 0. (2.6)
a
in which ξt satisfies dξti dξtj = σ ij dt. The meaning of this ‘differential’ statement,
which is infinitesimal in nature, can be understood as a mathematical shorthand
for a corresponding integral equation, in which the integrals that arise, where
relevant, are understood in the Ito sense, as explained in Section 2.1.
2.2.1 Ito’s formula
It is often natural and relevant to consider prescribed functions of a stochastic
process such as in (2.9), together with their stochastic differentials. Accordingly,
let us define a new related stochastic process
Ft := f (Xti , t) (2.10)
wherein the function f is known. For example, in the description of random
phasor dynamics in Chapter 6, it is the complex exponential function that is
relevant. Ito’s formula then provides us with a convenient explicit expression for
the stochastic differential of Ft in terms of f and dXt . The key to the result, which
constitutes a departure from the classical rules of calculus, is that the underlying
process Xt has positive quadratic variation, i.e. that its squared fluctuations
integrate to a non-zero amount. Consistently, retaining second-order terms, we
write
∂f ∂f 1 ∂2f
dFt = dt + i
dXti + dXti dXtj + o(dt). (2.11)
∂t ∂X 2 ∂X i ∂X j
Substituting (2.9) into the above we deduce that
∂f ∂f 1 ij ∂ 2 f ∂f
dFt = + bi + σ dt + dξ i (2.12)
∂t ∂X i 2 ∂X i ∂X j ∂X i t
which statement is Ito’s formula.
2.2.2 Ito product rule
The origin of the Ito product rule is similar in essence to the existence of a
quadratic variation that enters into the Ito formula, where now we consider a
pair of processes. Essentially, the result says that we must take into account
product differentials when calculating the differential of a product, and thus the
result constitutes a modification of the classical Leibnitz rule. Explicitly, for the
product of a pair of stochastic processes Ut Vt we write the incremental product
δ(Ut Vt ) = Ut δVt + Vt δUt + δUt δVt + o(δt) (2.13)
which takes the infinitesimal form
d(Ut Vt ) = Ut dVt + Vt dUt + dUt dVt (2.14)
in which it is essential to observe that the third term on the right-hand side
above is non-zero.
ITO DIFFERENTIAL 11
Remarks. Observe that in the product expression dUt dVt = E[dUt dVt ], i.e.
the expectation E is superfluous; in the case that the Wiener processes driving
U and V are independent this product is zero, corresponding to a zero (U, V )
entry in the diffusion tensor σ ij .
3
STOCHASTIC DIFFERENTIAL GEOMETRY
12
DIFFUSIONS ON MANIFOLDS 13
and has the simplifying property that it annihilates the diffusion tensor. In this
geometry it is natural to work with the Ito drift of the process, whose transfor-
mation properties are tensorial with respect to (M, σij ). The situation in this
regard is contrasted with the case of the Kolmogoroff forward drift obtained by
regarding the process as existing on Rn (Karatzas and Shreve 1988). The cur-
rent treatment has the advantage over previous work (cf. Nelson 1967a, 1985)
that each geometrical operation may be expressed in the coordinate free abstract
index notation, as elucidated for example, in Penrose and Rindler (1984).
Consider a continuous time diffusion process Xti on a manifold M of dimen-
sion n taken to satisfy the (time independent) Ito stochastic differential equation
(SDE)
dXti = β i (Xti )dt + i
(Xti )dWt
(j)
σ(j) (3.1)
j
(j)
where {Wt } are a collection of n independent Wiener processes.4 The (con-
travariant) diffusion tensor σ ij is then determined by dXti dXtj = σ ij dt, so that
i j
consistently r σ(r) σ(r) = σ ij . We distinguish between two notions of drift on
the manifold as follows (cf. Nelson 1985). Regarding Xti as a process on a coor-
dinate patch of Rn we have the Kolmogoroff mean forward drift (Karatzas and
Shreve 1988) defined as
i
i Xt+δt − Xti
β = lim Et (3.2)
δt→0 δt
and the corresponding mean backward drift
i i
X t − X t−δt
β̃ i = lim Et (3.3)
δt→0 δt
for δt > 0 in a coordinate chart {xi } that is normal with respect to the Euclidean
metrical geometry δij of Rn (e.g. Nelson 1967b). Denoting the Levi-Civita con-
nection of δij by ∂i the probability density ρE , with respect to the Euclidean
volume measure on Rn , satisfies the FPE (e.g. Risken 1989)
∂ρE ∂ 1
+ i (ρE β i ) = ∂i ∂j (σ ij ρE ) (3.4)
∂t ∂x 2
and the corresponding equation for the backward drift
∂ρE ∂ 1
+ i (ρE β̃ i ) = − ∂i ∂j (σ ij ρE ). (3.5)
∂t ∂x 2
Now, for a non-degenerate diffusion we write σ ij σjk = δki , with σ := det{σij } =
0. In this way the inverse σij supplies the natural metrical geometry of M from
4 The status of the parenthesized index is conceptually distinct from that of a tensor or its
coordinate representation, and the index serves merely to label an individual member amongst
the collection of Wiener processes.
14 STOCHASTIC DIFFERENTIAL GEOMETRY
∇i V j = ∂i V j + Γijk V k (3.6)
below have equal divergences. Since this holds for arbitrary ρ independently of b, β, σ the
identity follows.
DIFFUSIONS ON MANIFOLDS 15
We then define the trace of the Christoffel symbol (on its lower indices)
The vector β i = δii β i then becomes a genuine tensorial object on M, but one
whose definition via (3.2) is tied to the choice of a normal coordinate chart
with respect to the Levi-Civita connection ∂i of the Euclidean metric δij on Rn .
Henceforth in abstract index expressions, indices will be raised and lowered with
σ ij , σij , respectively in accordance with the Einstein summation convention.
This geometrical characterization of the diffusion tensor will be relevant to
the treatment of the amplitude fluctuations for the various scattering processes
we consider in Part III. In expressions that arise hereafter we shall use abstract
index or coordinate basis notation as appropriate to the context and clarity of
exposition, and the summation convention shall be assumed throughout.
for an arbitrary function f (t, x). Since Xt is taken to satisfy (3.1) we deduce
from Ito’s formula applied to f (t, Xt ) that these equations may be expressed in
6 We recommend that the interested mathematical reader consult Hughston (1996) for an
eloquent and thorough exposition of the abstract index notation in the context of stochastic
differential geometry.
KINEMATICS 17
operator form as
∂ 1 ij
D= + bi ∇i + σ ∇ i ∇j ,
∂t 2 (3.24)
∂ 1 ij
D̃ = + b̃i ∇i − σ ∇i ∇j .
∂t 2
1 i
v i := (b + b̃i ), (3.25)
2
1 i
ui := (b − b̃i ). (3.26)
2
1 ij
ui = σ ∇j log ρ. (3.27)
2
It is illuminating to study the situation for pure diffusion (the heat equation) in
the context of these kinematical quantities. The pure diffusion process has vanish-
ing (forward) drift, as we have seen from the elements of Einstein’s derivation.
Nevertheless, it has non-zero current, v = 0. This (stochastic) current corre-
sponds to the intuitive notion of a ‘flow’ of diffusing particles which, as we shall
see, is quite distinct from the notion of forward drift. The explicit form of this
current can be derived from the Green’s function solution (1.8) and the osmotic
equation (3.27) in combination with the defining equation for the stochastic cur-
rent (3.25). Since the drift vanishes, the current and osmotic velocities are equal
and opposite. According to the osmotic equation and Green’s function solution,
we find
x
v= , t>0 (3.28)
2t
and hence
T [ui ] = ui
(3.30)
T [v i ] = −v i .
In addition, we observe that the pair of (covariant) FPEs (3.9) can be obtained
from each other immediately by invoking the time-reversal operation T (since ρ,
∇, σ ij are each invariant under this map, while ∂/∂t reverses sign).
Remarks. On detailed balance. In a general context, there exist two distinct
notions of stochastic equilibrium, namely that of ordinary equilibrium and the
stronger condition of detailed balance. Ordinary ‘equilibrium’ is attained when
the probability density occurring in the FPE is time invariant. If, in addition,
the stochastic current vanishes, then the condition of ‘detailed balance’ is said
to hold. Clearly, from the above discussion, detailed balance implies equilibrium,
but not conversely.
The interested reader may wish further to study the application of this
stochastic differential geometry to the dynamics of quantum mechanical dif-
fusion on general manifolds, as discussed in Field (2003). This paper explores
non-degenerate diffusion processes on an arbitrary manifold, the dynamics of
which arise from a principle of least action for a Lagrangian consisting of a ki-
netic term quadratic in the forward drift of the process and a local potential.
The equation governing the action emerges as a stochastic Hamilton–Jacobi con-
dition and is expressed in terms of the geometry determined by the Levi-Civita
7 Note that, in the definition of the backward drift (3.3), the differentiation is still forwards
in time, the distinction between the forward drift arising from the subtle difference in the
positioning of the time increment with respect to the conditioning.
KINEMATICS 19
connection of the diffusion tensor. It is argued that there are essentially two dy-
namical structures for the rate of change of the drift in the presence of a local
potential, consistent with the requirement of time-reversal symmetry. In both
cases a conserved energy is identified. An alternative wave function and associ-
ated operator description reveal a complex structure in the dynamical equations,
thus extending the earlier results of Edward Nelson on the stochastic treatment
of the Schrödinger equation (Nelson 1967a, 1985).
4
EXAMPLES OF STOCHASTIC DIFFERENTIAL EQUATIONS
We begin the chapter with some general remarks concerning the physical mo-
tivation for the application of stochastic differential equation (SDE) theory to
problems in the mathematical description of scattering from random media. Such
scattering problems are characterized by two essential ingredients. First, the ex-
istence of a random population, whose size fluctuates in time. A measure of the
size of this population (in the continuum limit) is what is referred to as the scat-
tering cross-section and its statistical characteristics are independent of those of
any external electromagnetic radiation that might be used to indirectly probe
the properties of the population. From an experimental point of view, the cross-
section is usually the object of primary interest and various experiments are
designed to infer its behaviour from the properties of electromagnetic radiation
that scatters from or propagates through the random medium.
Second, the electromagnetic component consists of a known transmitted wave
that interacts with the random population, and is subsequently forward- or
backscattered. The random elements that enter here are the fluctuations in mean
power as determined by the cross-section, together with microscopic phase ran-
domization effects that occur for each component of the scattered field that is
essentially an electromagnetic property. The resulting electromagnetic energy
therefore entails two degrees of randomness, inherited from the scattering popu-
lation size and its microscopic details that affect the degree of coherence amongst
the component scattered waves associated with each member of the population.
Thus, we are confronted with a coupled system that has an inherent two-fold
randomness, namely, the scattering population and the associated electromag-
netic radiation. Both of these components of the state of the system evolve
temporally as well exhibiting spatial correlation. Thus, the scattering processes
that we shall consider are fundamentally continuous time-random entities. It is
therefore appropriate, and indeed necessary, to employ a stochastic methodology
that is able to capture the dynamics of such processes in as complete a manner
as possible, beyond eliminating the inherent randomness in the system.
From this point of view, conventional statistical descriptions, though they
may be established and correct, do not represent a complete physical descrip-
tion, capturing only certain ensemble average characteristics, such as distribu-
tion, correlation, power spectra and certain higher order statistics. SDE theory
in partnership with the Ito calculus, on the other hand, provides a rigourous
mathematical framework that pertains to random states of the system that are
separated in time on an infinitesimal scale. The theory is therefore able, in prin-
ciple, to yield all higher order statistical properties of such a physical system.
20
GEOMETRIC BROWNIAN MOTION 21
The identification of the parameters in SDEs that make this possible, in this
monograph, is achieved through both experimental and independent theoretical
reasoning.
The following sections provide some common illustrative examples in SDE
theory, so that the reader can become familiarized with the basic concepts and
techniques that we shall require in the study of scattering processes.
in which the novice to the Ito calculus should carefully observe the presence of
the second-order term on the right-hand side. Then substituting (4.1) we find
1
d (log St ) = µ − σ 2 dt + σdWt (4.3)
2
which has the merit of a pure differential on the left-hand side and eliminating the
presence of St on the right-hand side. The exact (random) solution is therefore
1
ST = S0 exp µ − σ 2 T + σWT . (4.4)
2
Now, the exponent is Gaussian with mean µ − 12 σ 2 T and variance σ 2 T . Hence
ST has lognormal behaviour, which example is of central importance in the
Black–Scholes theory of option pricing (cf. Appendix D).
(n)
for integer n > 1. Our aim is now to calculate the SDE satisfied by Bt , accord-
ing to the rules of Ito calculus. From Ito’s formula we have
n −1/2 n
1 (i) 2 (j)
(n)
dBt = W d Wt 2
2 i=1 t j=1
3 −3/2 n 2
1
n
−
(i)
Wt 2 d Wt 2 .
(j)
(4.6)
2 i=1 j=1
2
Now let us examine the term of the form d W occurring above in isolation.
Via Ito’s formula and the identity dW 2 = dt we can express each term in the
summation as
(j) (j) (j)
d Wt 2 = 2Wt dWt + dt (4.7)
and thus
n n
(i)
= 2 Wt 2 dW̃t + ndt
(j) 2
d Wt (4.8)
j=1 i=1
in which we invoke a new (related) Wiener process W̃t (for n > 1).8 Combining
these results we deduce that the Bessel process satisfies the SDE
(n) 1
dBt = (n)
(n − 1)dt + dW̃t (4.9)
2Bt
(n)
for Bt = 0.9
(n)
Remarks. Concerning positivity. Directly from its definition, Bt is every-
where positive (or zero); this is reflected in its SDE by the divergence of the drift
as zero is approached, which property (in spite of the persistence of the Wiener
fluctuating term) ‘repels’ the process from the origin. (The reader should com-
pare the discussion of natural boundaries in Chapter 5.)
8 The way in which the components combine to yield a new Wiener process arises from the
∇2 φ = 0 (4.10)
10 The existence of a ‘hitting’ time at which this occurs is ensured by the continuity of the
Brownian path.
This page intentionally left blank
Part II
Dynamics of the scattering process
25
26 SCATTERING DYNAMICS
11 The situation as regards natural boundaries should be compared with the account of the
27
28 DIFFUSION MODELS
We observe that from a physical point of view that only the volatility func-
tions are observable and explain this feature in the context of Girsanov’s theorem
which effects a change of probability measure on the space of all possible paths
of the process. We are thus led to an infinite dimensional family of stochastic
volatility models for the scattering process such that the discriminating features
depend only on the instantaneous observed volatilities and are thus drift inde-
pendent. In this connection, we remark that similar techniques involving change
of measure and stochastic volatility have been studied in the context of financial
mathematics, in particular, in the Black–Scholes theory of option pricing (see
Appendix D).
(Jakeman and Tough 1988). The usual treatment of the K-distributed intensity
process then makes an assumption that the scattering cross-section x is governed
by a birth–death–migration (BDI) population process such that it is asymp-
totically Γ-distributed and, thereby, one obtains the joint distribution P(x, z).
Integrating out the x variable leads to the K-distribution Kν (z) for the (uncon-
ditioned) intensity
√
2z ν/2 Kν (2 z)
Kν (z) = (5.2)
Γ(ν + 1)
where Kν is a modified Bessel function of the second kind (see e.g. Jeffreys and
Jeffreys 1966).
However, the physical scattering observations relate directly to the scattered
radiation, giving rise to the complex-valued amplitude process and correspond-
ing electromagnetic intensity. On the other hand, no direct measurements of the
scattering cross-section are made. Thus from a physical point of view the be-
haviour of the scattering cross-section should be inferred rather than postulated
in advance. From large time averages of measurements of the scattered radia-
tion, we can deduce that the intensity is, to a close approximation, K-distributed.
The additional assumption of a Rayleigh distributed intensity given a δ-function
distributed cross-section (population size) is a direct consequence of a uniform
distribution in the phases ϕn of the contributions to the scattered electric field
amongst a population of independent identically distributed (i.i.d.) scatterer am-
plitudes (this conclusion follows readily from the central limit theorem applied
to a component of the scattered electric field). More precisely if we decompose
STATISTICAL MODELS 29
the total scattered electric field into the sum of contributions from individual
scatterers, thus
E= an exp(iϕn ), (5.3)
n
then the form factors an are taken to be i.i.d. random variables (with any distri-
bution, not necessarily Gaussian) and the phases to be distributed uniformly (cf.
Jakeman and Tough 1988). These two features of the electromagnetic scattering,
which are physically motivated, then imply the Γ distribution for the scattering
cross-section, as seen from the following result.
Lemma 5.1 For an intensity with P(z|x) and P(z) given respectively by the
Rayleigh distribution Rx (z) and the K-distribution Kν (z), the scattering cross-
section necessarily has a gamma distribution Γν (x). Accordingly the joint distri-
bution P(x, z) is equal to Rx (z)Γν (x).
Proof Suppose,
for some unknown distribution
in the cross-section P(x), that
Kν (z) = Rx (z)P(x)dx, i.e. Kν (z) = [exp(−z/x)/x]P(x)dx for some value of
ν. Since this Kν is known to be generated by Γν , we have
−z q(x)
exp dx ≡ 0, ∀z, (5.4)
x x
where q = P − Γν , so that existence of the above integral is ensured
(observe also
that q = 0). Setting u = 1/x the above integral becomes exp(−uz)Q(u)du,
where Q(u) = −q(1/u)/u, and this integral must vanish for all values of z.
Regarding z as a Laplace transform variable, we see from the Laplace transform
inversion theorem that Q ≡ 0. Thus P = Γν . 2
Observe that the same principle applies to different distributions for the intensity,
e.g. if the modulus amplitude is governed by the Weibull distribution, which is
sometimes used in radar clutter modelling. More precisely, if the intensity has a
given (observed) distribution, and the Rayleigh assumption for the conditional
distribution P(z|x) is preserved, then a corresponding argument leads to the
conclusion that the distribution of the cross-section is uniquely determined.
In summary therefore, as a consequence of the various physical assumptions
and an observed K-distribution for the intensity, the joint distribution for x, z
is given by
xν−1 exp(−x − z/x)
P∞ (x, z) = (5.5)
Γ(ν + 1)
in which ν ≥ −1 is the so-called ‘shape parameter’ as discussed in the radar con-
text in Chapter 13 (equal to α − 1 of Chapter 7). We can now proceed to classify
(vector) diffusion processes in the variables x, z that possess the above joint dis-
tribution. To enable the classification we first observe the following elementary
result (cf. also Wong 1963 for a discussion of sufficient conditions for a diffusion
to have certain asymptotic distributions, governed by Pearson’s equation).
30 DIFFUSION MODELS
Lemma 5.2 Given the Fokker–Planck equation (FPE) ∂P/∂t = −∂x (βP) +
∂x2 (ΣP) the relations between the drift, volatility and asymptotic distribution P∞ ,
if such exists, can be summarized as
x
K β
P∞ = exp (5.6)
Σ Σ
β = Σ∂x log(ΣP) (5.7)
x
k βP
Σ= + . (5.8)
P P
Thus any two of {P, β, Σ} implies the other (modulo k in the case of Σ).12
Proof The existence of an asymptotic distribution P∞ implies that ∂x [−βP∞ +
∂x (ΣP∞ )] ≡ 0. Thus the square-bracketed expression is some function f (t),
which must vanish given the decay of P∞ (x) as x → ∞. Therefore βP∞ =
∂x (ΣP∞ ), whereupon the required expression for β is immediate, and integration
yields those for P∞ , Σ. 2
Proposition 5.3 The FPE for K-scattering processes takes the parametric form
2
∂P ∂ ∂ 1−ν z ∂φ
=A (φP) + − 2 +1 φ− P
∂t ∂x2 ∂x x x ∂x
2
∂ ∂ ψ ∂ψ
+B (ψP) + − P (5.9)
∂z 2 ∂z x ∂z
for squared volatility functions φ(x, z, t), ψ(x, z, t). The corresponding non-linearly
coupled SDEs for the vector diffusion process (xt , zt ) are
ν−1 z ∂φ (1)
dxt = A + 2 −1 φ+ dt + (2Aφ)1/2 dWt, (5.10)
x x ∂x
∂ψ ψ (2)
dzt = B − dt + (2Bψ)1/2 dWt , (5.11)
∂z x
(1) (2)
where Wt , Wt are independent Wiener processes.
Proof For equilibrium, set the left-hand side of (5.9) equal to zero. Since A, B
are independent constants we require that the expressions in square brackets on
the right-hand side vanish separately. The form of the joint distribution (5.5)
together with (5.7) imply the forms taken by the drift coefficients in (5.9), given
the squared volatility free functions φ, ψ. 2
The reciprocals of the (characteristic frequency) constants A, B represent inde-
pendent correlation timescales for the respective scattering processes xt , zt . In
tensor notation (Risken 1989) we introduce the operator L̂ = −∂i β i + 12 ∂i ∂j σ ij
in terms of which the FPE can be expressed as ∂ρ/∂t = L̂ρ. The detailed balance
condition, which is stronger than the stationarity (i.e. equilibrium) condition,
requires that each individual transition in the master equation is perfectly bal-
anced. This condition can be expressed as the operator equation L̂ρ = ρL̂† acting
on all functions f (Risken 1989), whereby the choice f = 1 implies stationar-
ity. Alternatively, in terms of the ‘current’ v i = β i − 12 ρ−1 ∂j (σ ij ρ) the FPE
can be written ∂ρ/∂t + ∂i (ρv i ) = 0 and the detailed balance condition becomes
v i = 0. Since the constants A, B in (5.9) are independent and arbitrary, the
detailed balance condition is satisfied for any choice of volatility functions φ, ψ.
In typical physical situations, such as those we shall describe in Chapter 12, zt
de-correlates much more rapidly than its companion xt , so that A B. The
frequency A is inherent to the scattering population and is independent of the
incident electromagnetic field; in contrast, on dimensional grounds B is propor-
tional to the wave-number of the illuminating radiation and thus B ∼ c|k|.13
The (reciprocals of the) constants A, B can be determined experimentally from
the scattered electromagnetic intensity time series {zt } by measuring the period
between successive peaks over the two characteristic long and short time scales,
13 The ∼ symbol here indicates an (approximate) scaling relation.
32 DIFFUSION MODELS
respectively (see e.g. §6 in Jakeman and Tough 1988); cf. also §IIC in Field and
Tough (2003b) for a detailed theoretical account of the observability of these
characteristic frequency constants.
and likewise for ψ(z) with remainder term Rψ (z). The Taylor theorem
x with
remainder (see e.g. Jeffreys and Jeffreys 1966) states that Rφ (x) = 0 (x − u)
φ(2) (u)du and that this remainder satisfies the inequalities
1 2 1
x inf [φ(2) ] ≤ Rφ (x) ≤ x2 sup[φ(2) ] (5.14)
2 [0,x] 2 [0,x]
x inf [φ(2) ] ≤ Rφ (x) ≤ x sup[φ(2) ] (5.15)
[0,x] [0,x]
with corresponding inequalities for ψ(z). Thus we deduce the limiting behaviour
of the drifts b(x) , b(z) in the respective component SDEs (5.10), (5.11). We find
(x) (1) Rφ (1) 1
lim [b ] = νφ (0) + z lim + zφ (0) lim . (5.16)
x→0 x→0 x2 x→0 x
The second term on the right-hand side is bounded by virtue of (5.14). Thus
necessary and sufficient conditions for the natural boundary requirement
14 We suppress the (z, t) and (x, t) arguments in the sets of functions {φ, R }, {ψ, R }
φ ψ
respectively for notational clarity.
SCATTERING DYNAMICS 33
x
0 < limx→0 b(x) ≤ ∞ are φ(1) > 0, or else φ(1) (0) = 0 with 0 (x − u)φ(2) (u)du >
0, and sufficient that inf [0,x] [φ(2) ] > 0. Correspondingly for b(z) , we find
so that a necessary and sufficient condition on the component drift for the natural
boundary to exist with respect to z is 0 < ψ (1) (0) ≤ ∞.
As a special case of (5.9) (Jakeman and Tough 1988; Tough 1987) when
φ(x, z, t) = x and ψ(x, z, t) = z we find
2
∂P ∂ ∂ z
=A (xP) + x − ν − P
∂t ∂x2 ∂x x
2
∂ ∂ z
+B (zP) + − 1 P , (5.18)
∂z 2 ∂z x
where P = P(x, z, t). Observe that the natural boundary condition is satisfied
for this choice of volatility functions. The associated non-linearly coupled SDEs
(5.10), (5.11) for the vector diffusion process (xt , zt ) become
ν−x+z (1)
dxt = A dt + (2Ax)1/2 dWt (5.19)
x
1−z (2)
dzt = B dt + (2Bz)1/2 dWt . (5.20)
x
This special case can be motivated by the corresponding situation for the Rayleigh
(Gaussian speckle) process as described in Section 6.1.
In respect of classification of electromagnetic scattering processes, the novelty
of the present approach rests on the following key observation.
The argument for the local observability of the volatility functions can also be
understood from the point of view of Girsanov’s theorem (Oksendal 1998 and
Appendix D) which states that, via a change of measure on the space of paths,
the process dW̃t = dWt + γt dt can be regarded as a pure Brownian motion. More
34 DIFFUSION MODELS
Observe from Ito’s formula that dMt = −Mt γt dWt so that Mt is manifestly a
martingale with respect to the original measure P, i.e. EPs Mt = Ms for s ≤ t.
In respect of our electromagnetic scattering processes we may re-express (5.10)
and (5.11) in the Q(i) ‘drift-neutral’ measures as
(1)
dxt = (2Aφ)1/2 dW̃t , (5.24)
(2)
dzt = (2Bψ)1/2 dW̃t . (5.25)
t
15 We require γt to satisfy Novikov’s condition, namely that exp( 12 0 γs2 ds) < ∞.
CORRELATED VECTOR SCATTERING 35
cross-correlation has a timescale inherited from the (fluctuations in the) scattering cross-section
and thus depends on the constant A.
36 DIFFUSION MODELS
system
(x)
Pb = 1
2 [∂x (P Σxx ) + ∂z (P Σxz )] ,
(5.32)
P b(z) = 1
2 [∂x (P Σzx ) + ∂z (P Σzz )] .
Before solving this system let us recall briefly the 1-dimensional case. Consider
that we are prescribed the asymptotic distribution P and we find the drift for
an arbitrary volatility function. Then the system above reduces simply to the
scalar relation b = 12 Σ∂x [log(ΣP )] – thus knowledge of P uniquely establishes
the relationship between the drift and volatility (diffusion) coefficients (cf. Wong
1963). In the two-dimensional case of the vector scattering process (x, z) that we
are concerned with here, the asymptotic (joint) distribution is indeed prescribed
according to (5.29) and thus we follow a similar approach to the one-dimensional
case. First, let us recall some simple identities for the partial derivatives of the
joint distribution with respect to x, z:
∂P z ν−1
= P −1 + + ,
∂x x2 x
(5.33)
∂P = − P .
∂z x
Similarly for the z-drift a straightforward calculation using (5.35) and arranging
terms according to A and B dependence, yields
z(ν + 1 − x)
b(z) = A + B(x − z) (5.38)
x
in accordance with the analysis of the random walk model as developed in Chap-
ter 8. This completes our discussion of the vector K-scattering process in the
case that detailed balance is imposed.
Now, we turn to consider the case of classifying K-scattering models if there
is ‘weak’ equilibrium, i.e. an asymptotic limiting distribution exists but the is
a non-zero current for the (xt , zt ) as described in terms of the abstract fluid
flow above. We consider the density-weighted immersion of the two-dimensional
flow determined by the vector field v into Euclidean 3-space, co-ordinatized by
(X, Y, Z),18 thus
v → V = P v̆ (5.39)
in which we define v̆ := (vx , vz , 0). The (weak) equilibrium condition can then
be expressed (locally) as
V =∇×A (5.40)
∂X AY = ∂Y AX . (5.42)
Thus, without loss of generality, we may set A = (0, 0, A) for some scalar po-
tential function A ≡ A(X, Y ) and hence V = (∂Y , −∂X , 0)A. Observe here that
A has the status of an arbitrary scalar function (with suitable decay at ‘spatial’
infinity). This enables us to construct a general expression for the forward drift
of the process as
consisting on the right-hand side of a detailed balance part and current contri-
bution respectively, wherein the former is obtained by setting A = 0. In respect
of the latter, we identify the current in terms of a gauge freedom in choice of this
scalar function. Observe that the relation P bD.B. = 12 ∂· (Σ· · P ) is unaffected by
the presence of a gauge term involving A. We therefore summarize the scheme for
classifying the dynamics of asymptotic equilibrium as follows: we begin with a
prescribed asymptotic joint distribution (e.g. that appropriate to K-scattering),
choose the diffusion coefficients Σ(·,·) and gauge potential A as free (indepen-
dent) functions, and then construct the drift according to (5.43) – this provides
a complete classification of the diffusion dynamics, incorporating both detailed
balance and current contributions. Functionally, we can summarize this situation
by the statement
In Chapter 8, we shall then see how specific choice of the functions φ, ψ, χ occur-
ring in the (x, z) representation of Σ(·,·) may be obtained, from first principles,
via considerations of a random (phasor) walk model. It emerges that detailed
balance is satisfied for such a model and so we may assume A = 0 in (5.44).
Thus, a physical – as opposed to parametric data driven – model emerges as
a special case of the general structure above. The latter is constructed merely
to preserve the joint (x, z) asymptotic distribution, while the former is a spe-
cific instance of this with the essential property that, in addition, it specifies all
correlation information and higher order statistics.
6
RAYLEIGH SCATTERING
Ψt = It + iQt (6.1)
40
QUADRATURE COMPONENTS 41
and thus the mean value of z is what we expect from (6.6). In this way the
analysis of the component I, Q processes using the Ornstein–Uhlenbeck process
gives insight into the origin of the square root volatility for the intensity zt . It is
interesting to note that the FPE describing the Rayleigh process also encodes the
factorization properties of the intensity distribution, which result from its being
the sum of squares of two Gaussian processes, each with its own characteristic
factorization properties. If we change variables to z = αz/2, t = 2αt, the FPE
(6.10) becomes
min(m,n)
n!m!
z(t)n z(0)m = n!m! exp(−rt). (6.13)
r=0
(n − r)!(m − r)!(r!)2
In much the same way, it can be shown that the expansion of the Green’s func-
tion of the FPE (6.5) in terms of Hermite polynomials captures the factorization
properties of the constituent Ornstein–Uhlenbeck processes. In each case the lin-
earity of the deterministic part of the underlying SDE establishes the exponential
decay in the correlation function characteristic of a Markov process.
The constant volatility in the Ornstein–Uhlenbeck case does not impose a
natural boundary at the origin, so that the process can take positive and nega-
tive values. Conversely, the square root volatility emerging in the Rayleigh case
establishes a natural boundary that maintains the positivity of zt . Observe that
the forms taken by the SDEs (6.3) and (6.4) inherently capture these fundamen-
tal properties of the physical processes they describe. The special cases (5.18),
(5.19) and (5.20) will be significant in the derivation of K-scattering dynamics
from a random walk model in Chapter 8 and in the experimental analysis of
Chapter 12.
Remarks. On random motion in a potential. The situation of stochastic volatil-
ity discussed here should be compared with the more familiar case of constant
RANDOM WALK 43
N
& '
(N ) (j)
Et = exp iϕt (6.16)
j=1
for constant population size N . Since Maxwell’s equations for the electromag-
netic field possess U (1) gauge invariance with respect to duality rotations, i.e.
multiplication by exp(iΛ) for (constant
) Λ (cf. Penrose and Rindler 1984), the as-
sumption of independence of ϕ(j) implies that these
&phases'are uniformly dis-
(j)
tributed. Accordingly, in (6.16) the phase factors exp iϕt are independent
and uniformly distributed on the unitcircle in the complex plane C. Our (phase)
(j)
diffusion model therefore takes ϕt as a collection of (displaced) Wiener pro-
cesses evolving on a suitable timescale,
(j) 1 (j)
ϕt = ∆(j) + B 2 Wt , (6.17)
( )
with the random initializations ∆(j) chosen as a set of independent random
variables uniformly distributed on the interval [0, 2π). The effect
& of' these ini-
(j)
tializations is to render each component phasor process exp iϕt , and thus
the resultant amplitude obtained from their coherent addition, stationary (cf.
Remarks On stationarity below). The component phase dynamics is simply
(j) 1 (j)
dϕt = B 2 dWt , (6.18)
(j) 2
with squared volatility dϕt = Bdt.
44 RAYLEIGH SCATTERING
N
& '
(N ) (j) 1 (j) 2 (j)
dEt = idϕt − dϕt exp iϕt . (6.19)
j=1
2
N & '
(j) (j)
The first term j=1 idϕt exp iϕt on the right-hand side of (6.19) consists
of a sum of independent randomly phased Wiener processes, with variance equal
to BN dt, while the second term is independent of the scatterer label j. Thus
from (6.19) we can write
(N ) 1 (N ) 1
dEt = − BEt dt + (BN ) 2 dξt , (6.20)
2
* (j)|dξt | = dt,
2
where ξt is a complex Wiener process satisfying dξt2 = 0. The
process ξt is adapted to the filtration F (ϕ)
= j F , where F is the filtration
(j)
(j)
appropriate to the component scatterer
& phase
' ϕt . The amplitude process Ψt is
(N ) 1
then defined by Ψt = limN →∞ Et /N̄ 2 and satisfies the SDE
1 1
dΨt = − BΨt dt + (Bx) 2 dξt , (6.21)
2
where the continuous valued random variable x, the average scattering
+ , ‘power’,
arises from an asymptotically large population via x = limN →∞ N/N̄ . Observe
that the process exhibits mean reversion (towards the origin), is of constant
volatility, and has an asymptotic stable Gaussian distribution.
Remarks. On geometry of phase wrapping. The effect of the exponentiation
in (6.16) is to ‘phase wrap’ the process onto the unit circle in the complex
plane. Thus, a coordinate discontinuity in the phase process at the boundary of
a coordinate interval of length 2π is mapped to a continuous behaviour on the
unit circle. The manifold of the phasor process is thus a cylinder, with symmetry
axis corresponding to time, on which the sample paths of the phasor process are
continuous (with probability 1).
Remarks. On stationarity. With regard to stationarity, observe primarily that,
given the random uniform initializations (6.17) the distribution of phase remains
uniform over the unit circle at all times – the initial distribution is uniform due
to the nature of ∆(j) and this coincides with the asymptotic distribution. The
RANDOM WALK 45
46
BDI PROCESSES 47
In the BDI case we evaluate (7.15) in this limit, as follows. For notational
clarity we shall drop the prime and ‘nor’ suffix on t, x, respectively. From (7.6)
the summation of (7.15) contains
∂
− [ν(1 − x)P] (7.16)
∂x
For n ≥ 3 the sum contains derivatives of D(n) /N̄ n−1 which, for first-order
transitions, is equal to
CONTINUUM LIMIT 49
N+2
(2) (2)
GN RN+2
N+1
(1) (1)
GN RN+1
N
(1) (1) (2)
GN–1 RN (2)
GN–2 RN
N–1
N–2
It is of some interest that, even for a first-order master equation, the expan-
sion (7.15) is of infinite order generically, and accordingly the fluctuations in the
process can not be driven by the usual Wiener process (cf. Jakeman et al. 2003).
For higher order transitions the situation is depicted in Fig. 7.1. We can express
the evolution of the probability density in a similar fashion as
∞
(i)
dPN (i) (i) (i)
= {GN −i PN −i − (GN + RN )PN + RN +i PN +i }, (7.23)
dt i=1
(i) (i)
where GN −i = RN = 0 for i > N , thus maintaining N ≥ 0. Correspondingly
the Kramers–Moyal coefficients are given by
∞
ln n (i) n (i)
D =
(n)
i (G + (−1) R ) . (7.24)
n! i=1
The order of the Kramers–Moyal expansion for higher order transition processes
can be summarized by the following result.
Proposition 7.2 The Kramers–Moyal expansion (7.15) is finite order as N̄ → ∞
if and only if for all integer m > 2
∞
im [G(i) + (−1)m R(i) ] = o(N m−1 ) (7.25)
i=1
CONTINUUM LIMIT 51
52
K-AMPLITUDE 53
Remark. In many respects this result, which is derived from very fundamental
physical and mathematical principles, represents the central theoretical advance
within the entire body of the monograph.
In the above expression, A and B are independent constants with the di-
mension of frequency, and they may take arbitrary values. In most situations
of interest, however, such as those reported in Field and Tough (2003a) and
developed in Chapter 12, the wavelength of the illuminating radiation is such
that the two corresponding reciprocal correlation timescales satisfy A B. The
description of Rayleigh scattering (i.e. constant scattering cross-section) is then
recovered when A = 0. Theorem 8.1 implies the following result.
Corollary 8.2 The squared volatility of the amplitude process Ψt is given by
Azt
|dΨt |2 = Bxt + dt. (8.13)
2xt
It is the linearity of the right-hand side above in zt that, in part, enables the
anomaly detection mechanism described in Section 12.2 (orig. Field and Tough
(2003a)).
8.1.1 Intensity
The stochastic differential of the intensity process zt can be expressed in terms
of the amplitude via the identity
The terms involving dξt above can be combined in terms of a real-valued Wiener
(ϕ)
process Wt according to
12
2zt
γt∗ dξt γt dξt∗
(ϕ)
+ ≡ dWt . (8.16)
xt
1
Azt (α − xt ) 2Azt2 2 (z)
dzt = B(xt − zt ) + dt + 2Bxt zt + dWt (8.18)
xt xt
(z) (x)
in which Wt is correlated with Wt of (8.9), and satisfies
12 12
Azt2 (z) 1 (ϕ) A (x)
Bxt zt + dWt = (Bxt zt ) 2 dWt + zt dWt . (8.19)
xt xt
(ϕ) (j)
The filtration of Wt arises from the constituent phases ϕt in the random
(x)
walk according to (8.16), while that of Wt stems solely from the fluctuations
in the endogenously specified population model – such (scattering) model is thus
sometimes referred to as the endogenous model. Observe that, if B += 0, ,(8.19)
implies W (z) = W (x) , and from (8.4) γt is constant, so |γt |2 = E |γt |2 = 1.
Accordingly, zt = xt is a solution of (8.18), as required by (8.3). From Proposition
8.3 we obtain the following result.
Corollary 8.4 The squared intensity volatility is determined by
2Azt2
dzt2 = 2Bxt zt + dt. (8.20)
xt
which, from (8.12), leads to the above expression for dzt2 . Observe that for A B
the dominant contribution to the squared intensity volatility is proportional to
the instantaneous value of the intensity. Thus for a Rayleigh timescale B −1 , over
which xt remains approximately constant, the time series for dzt2 and zt should
exhibit strong correlation. This feature has been experimentally verified in a case
of optical scattering, which is described in Section 12.1 (orig. §4(a) of Field and
Tough (2003a)).
√
In terms of the square-root intensity Rt = zt , an application of Ito’s formula
to (8.20) yields the following result.
Corollary 8.5 The squared volatility in the modulus amplitude is determined by
1 Azt
dRt2 = Bxt + dt. (8.22)
2 xt
K-AMPLITUDE 57
8.1.2 Phase
The complex amplitude process can be expressed in polar form Ψt = Rt exp(iθt )
and thus, writing iθt = log(Ψt /Rt ), we deduce from Ito’s formula that
2 2
dΨt 1 dΨt dRt 1 dRt
idθt = − − + . (8.23)
Ψt 2 Ψt Rt 2 Rt
Since the left-hand side is purely imaginary we can express dθt in terms of Ψt
alone as
2 ∗ 2
As in the derivation of the SDE for the intensity, we can express the terms
(θ)
involving ξt in (8.27) as a distinct real-valued Wiener process Wt according to
12
1 ∗ zt
(γ dξt − γt dξt∗ ) ≡
(θ)
dWt . (8.28)
2i t 2xt
The results above lead to expressions for the frequency constants A and B,
(j)
as follows. With respect to an average over the phase fluctuations dϕt there
exists a residual constant term in the squared phase volatility, i.e.
& ' 1
E dθt2 /dt|F (ϕ) = B. (8.37)
2
In principle, this enables the Rayleigh constant B to be deduced from scattering
data (alternatively an estimate of the Rayleigh correlation timescale B −1 can
GEOMETRY OF FLUCTUATIONS 59
be found from measuring the time difference between successive peaks in the
intensity time series {zt }). Expression (8.30) implies that xt = 2B−1 zt dθt2 /dt
and so the instantaneous values of the cross-section xt , and therefore rt , are
observable through the squared phase fluctuations. Consequently, the constant
A can be deduced from the square of (8.11)
drt2 1
= A (8.38)
dt 2
(cf. §3 in Field and Tough (2003a) for an account of the observability of the
squared volatilities for discretely sampled time series data).
is an ellipse whose major axis lies in the instantaneous radial direction defined
by Ψt . Degeneracy occurs only in the Rayleigh case, A = 0, for which S is a
circle, i.e. the fluctuations in Ψt are isotropic.
We remark, in general, that the random variables It , Qt possess a joint prob-
ability distribution that is U (1) symmetric, i.e. given by a surface of revolution
about the perpendicular axis to the origin in the I, Q-plane. Nevertheless It ,
Qt are correlated in general, and become independent only in the Rayleigh case,
A = 0, for which the surface of revolution is Gaussian. In this case the component
It , Qt processes can be described by the pair of (uncoupled) Ornstein–Uhlenbeck
processes determined as the real and imaginary parts of (6.21).
∂P 1
=− ∂i (bi P) + ∂i ∂j (σ ij P). (8.42)
∂t i
2 i,j
ASYMPTOTICS 61
From (8.9), (8.18), (8.19) the components of the diffusion tensor in the x, z
coordinate representation are given by
ij Ax Az
σ =2 (8.43)
Az Bxz + Az 2 /x
Proof The derivation of (8.45) follows immediately from (8.42), (8.43), and
(8.44), while the following identities for the derivatives of the joint distribution
P
∂z P = −
x z
∂z (zP) = 1 − P
x
2 z
∂z2 (zP) = − + 2 P
x x
α−1 z
∂x P = + 2 −1 P
x x
2 2 3
α−1 z α − 1 2z
∂x P =
2
+ 2 −1 − − 3 P (8.47)
x x x2 x
∂P
+ ∂i (Pv i ) = 0 (8.48)
∂t i
1
v i = bi − P −1 ∂j (σ ij P). (8.49)
2 j
where the functions β · (·), Σ· (·) are determined from (8.4), (8.9) and are indepen-
dent of A, B. The equilibrium condition in the (x, u) representation, obtained by
setting the left-hand side of (8.42) equal to zero, implies detailed balance, since
this condition holds for arbitrary values of the constants A, B. Consequently,
v i in the (x, z) representation also vanishes, since v i transforms homogeneously
(i.e. tensorially) under coordinate transformations (see e.g. Risken 1989).
CORRELATION AND SPECTRA 63
√ (u)
dut = (1 − ut )dt + 2ut dWt , (8.52)
√
where γt dξt∗ + γt∗ dξt = 2ut dWt . The propagator (i.e. Green’s function for the
(u)
√
1 u + u0 exp(−t) 2 exp(−t/2) uu0
P (u, t|u0 ) = exp − I0
1 − exp(−t) 1 − exp(−t) 1 − exp(−t)
(8.53)
where Iα denotes the modified Bessel function (e.g. Jeffreys and Jeffreys 1966).
In a similar manner, the propagator for (8.9) is given by
(α−1)/2
1 x exp(At) (x + x0 exp(−At))
P (x, t|x0 ) = exp −
1 − exp(−At) x0 1 − exp(−At)
√
2 exp(−At/2) xx0
× Iα−1 . (8.54)
1 − exp(−At)
∞
n!
P (x, t|x0 ) = xα−1 exp(−x) exp(−Ant)Lα−1
n (x)Lα−1
n (x0 ), (8.55)
n=0
Γ(n + α)
n
x−α exp(x) d
Lα
n (x) = (xα+n exp(−x)) (8.56)
n! dx
(cf. Wong 1963 for corresponding derivations). Combining (8.53), (8.55) leads to
the following result.
64 K-SCATTERING
P (z, x, t|z0 , x0 )
(α−1)/2
1 x exp(At)
=
x(1 − exp(−t))(1 − exp(−At)) x0
z/x + z0 exp(−t)/x0 x + x0 exp(−At)
× exp − exp −
1 − exp(−t) 1 − exp(−At)
4 √
2 exp(−t/2) zz0 2 exp(−At/2) xx0
× I0 Iα−1 . (8.57)
1 − exp(−t) xx0 1 − exp(−At)
Thus a general two-point correlation function can be expressed as the integral
∞
F1 (xt , zt )F2 (x0 , z0 ) = dx dz dx0 dz0 F1 (x, z)F2 (x0 , z0 )P (x, z, t|x0 , z0 )
0
xα−2 exp(−z0 /x0 − x0 )
× 0
. (8.58)
Γ(α)
In particular, we deduce the following important consequence.
Corollary 8.15 The intensity autocorrelation function is determined by the fol-
lowing expression:
zt z0 =
ut u0
xt x0
= α(α + exp(−At))(1 + exp(−t)). (8.59)
where S(ω) =
Ψ5 ∗
t Ψ0 .
√ Γ(α + 1)
lim
xt x0 = = α =
x. (8.65)
t→0 Γ(α)
Corollary 8.17 The power spectral density of the K-distributed noise process
characterized by (8.12) is given by
Γ(α + 12 )2 ∞ 1 1 t
S(ω) = 2 2F1 − , − , α, exp(−At) exp − cos((ω − ω0 )t).
Γ(α)2 0 2 2 2
(8.67)
S(ω)
3F2 (−1/2, −1/2, (1/2 + i(ω − ω0 ))/A; α, 1 + (1/2 + i(ω − ω0 ))/A; 1)
= 2
1/2 + i(ω − ω0 )
Γ(α + 12 )2
× . (8.68)
Γ(α)2
These calculations illustrate how the compound representation of the ampli-
tude (8.3) facilitates the analysis of the associated FPE (8.45). In terms of the
constituent spectra for the two component factors in (8.3) observe that, since
these component processes are independent, the autocorrelation of the resul-
tant amplitude factorizes into that of the components. Therefore, according to
the Wiener–Khintchine and convolution theorems, the power spectrum of the
resultant amplitude is equal to the convolution of the component spectra.
The explicit dependence on the electromagnetic frequency scale B in the
above expressions may be restored most simply on dimensional grounds.23 The
frequency scales A and B, which satisfy A B (for typical carrier frequencies),
may be found from an experimentally observed autocorrelation by fitting these
parameters according to the theoretical expression (8.66). From the series expan-
sion of the hypergeometric function, the radar cross-section (RCS) component of
the resultant amplitude autocorrelation (8.66) may be written as a sum of terms
proportional to exp(−nAt) (Fayard 2008), whose spectra are therefore Cauchy
or ‘Lorentzian’, while the Rayleigh spectral component is also Cauchy.24 Since
the Cauchy distribution is stable (see Appendix A), via the Fourier convolution
S (ψ) = S (r) ∗ S (γ) and taking the leading (n = 1) term in the hypergeomet-
ric expansion, it follows that the spectrum of the resultant amplitude ψ is also
(approximately) Cauchy, with FWHM equal to 2A + B. (The DC part of the
RCS spectrum, equal to αδ(ω) and reflecting merely the fact that the RCS has
a constant non-zero mean value of α, has been removed for clarity.)
Observe with regard to the above analysis the special case that the cross-
section is constant (A = 0). Then the amplitude autocorrelation consists solely
for example, a term exp(− 12 |t|) becomes exp(− 12 B|t|) in general units of time, etc..
23 Thus,
24 Anautocorrelation function R(τ ) = exp(− 12 kt) has associated power spectrum S(ω) =
2k/π(k2 + 4ω 2 ) with full-width-half-maximum (FWHM) therefore equal to k.
INTERPRETATION AND IMPLICATIONS 67
of an exponential decay term and thus the spectrum is exactly Lorentzian, i.e. a
Cauchy distribution (see Appendix A).
We have recalled in this chapter that the random walk model with step num-
ber fluctuations, due to Jakeman (see Jakeman 1980; Jakeman and Tough 1988),
accounts for certain statistical properties of K-scattering. In addition, we have
provided the extension to a complete dynamical description, in terms of contin-
uous time diffusion processes. This dynamical extension is generalized further
in Chapter 9, where we demonstrate how to include the effect of weak scatter-
ing in superposition with a coherent offset signal, in a corresponding stochastic
dynamical framework.
9
MODELS OF WEAK SCATTERING
It will be necessary for the reader to review the results of Chapter 8 (orig.
Field and Tough 2003b), which are essential in the present context for the treat-
ment of weak scattering.
1
Scaling by 1/N , 1/N 2 for the respective terms under the summation, in the
xt -continuum limit (N → ∞) this becomes
ΨR
t = a + γt . (9.2)
9.1.0.2 Homodyned K The situation here is the same as for K-scattering with
the superposition of a constant offset t = a that does not fluctuate with Nt .
In the continuum limit this amounts to adding a constant to the K-amplitude,
thus
ΨHK
t = a + ψt . (9.3)
STOCHASTIC DYNAMICS 71
ΨGK
t = axt + ψt (9.5)
in the continuum limit. Observe with respect to scaling in the continuum pop-
ulation limit that, in each case, we have divided by the (unique) length scale
factors, appropriate to the relevant terms in s(j) separately, which yield finite
non-zero resultant amplitudes.
In combination with (9.2), (9.3), (9.5), (8.9) and the results of Corollary 8.9
these identities enable us to derive
the SDEs satisfied by Zt , Θt in terms of
(r) (θ) (x)
the component Wiener processes Wt , Wt , Wt encountered in Chapter 8.
The dynamics are simplest for Rice scattering owing to the differential of (9.2).
(In the context of radar applications, the Rice scattering model is referred to
as a ‘Swerling zero target in Rayleigh clutter’, where the ‘target’ strength is
72 WEAK SCATTERING
represented by the signal (assumed constant over the timescale of interest) and
the Rayleigh process γt represents background ‘clutter’.) More care is required
in the calculations for the homodyned and generalized K-scattering processes
owing to certain cross-terms that arise. Nevertheless the strategy is the same
for each case, and we are led to the dynamical characterizations of the vector
scattering process St = (xt , Zt , Θt )tr according to the scheme
(no summation over i) for a collection of Wiener processes {Wti | ∀i} (not nec-
essarily independent) with respective drift and diffusion coefficients β i , Σij de-
termined by
E[dSti ]
βti = ,
dt (9.9)
dSti dStj = Σij
t dt.
The corresponding Fokker–Planck equation (FPE) (e.g. Risken 1989) for the
joint probability density ρt (x, Z, Θ) is then
∂t ρ + ∂i (ρV i ) = 0 (9.10)
1
V i = β i − ρ−1 ∂j (Σij ρ). (9.11)
2
9.2.0.4 Rice The amplitude dynamics of the Rice process is identical to that
of the Rayleigh process and the cross-section is constant and equal to unity, as
evident from (9.2). We deduce from the identities above that, in terms of the
geometry of the underlying Rayleigh process, the resultant intensity satisfies the
SDE
& 1
' 1
& 1 '
(r) (θ)
dZt = B 1 − ut − aut2 cos θt dt + (2B) 2 (ut2 + a cos θt )dWt − a sin θt dWt .
(9.12)
( + , )
a sin Θ A (Z 1/2 + ax cos Θ)/Z 1/2 − (α − x + 12 ) − Bx /2Z 1/2
(9.22)
where αt , βt are real-valued Ito differentials and φt is chosen so that their Ito
product αt βt vanishes, i.e. the Wiener components of αt , βt are statistically inde-
pendent (see e.g. Karatzas and Shreve 1988). Comparing the two decompositions
1
of dΨt above, it follows that (neglecting terms of o(dt 2 ))
Therefore
(αt2 − βt2 ) sin 2φt = 2Rt dRt dΘt
(9.26)
(αt2 − βt2 ) cos 2φt = dRt2 − Rt2 dΘ2t
Also, from (9.25) we find αt2 + βt2 = dRt2 + Rt2 dΘ2t so that
(α) (β) (Ψ,Ψ∗ )
Σt + Σt = Σt . (9.28)
Lemma 9.4
<
(α) (Ψ,Ψ∗ ) (Ψ) (Ψ∗ )
Σt = 1
2 Σt ± Σ t Σt
< (9.30)
(β) (Ψ,Ψ∗ ) (Ψ) (Ψ∗ )
Σt = 1
2 Σt ∓ Σ t Σt .
with ± corresponding to the major/minor axes of the error surface of the resul-
tant amplitude, respectively.
76 WEAK SCATTERING
Observe that (9.29), (9.30) have the appropriate symmetry under interchange
α ←→ β. The angle φt represents a rotation in the geometry of the resultant
amplitude fluctuations relative to the case of pure K-scattering, for which φ = 0.
From (9.27) this angle is determined as follows.
Lemma 9.5 The phase rotation φt , that yields an orthogonal dyad (see Fig.
9.1) associated with independent Wiener increments in the resultant amplitude
process Ψt , satisfies the geometrical identity
(Z,Θ)
4Zt Σt
tan 2φt = (Z) (Θ)
. (9.31)
Σt − 4Zt2 Σt
Equivalently, in terms of the resultant complex amplitude process, we have the
geometrical identity
+ ,
Ψ2t dΨ∗2
t
tan 2φt = − , (9.32)
[Ψ2t dΨ∗2
t ]
where , denote the real and imaginary parts, respectively.
Before applying this geometry to the weak scattering processes described
earlier, as a preliminary we give a result which provides the relationship between
the structure of the diffusion tensor that arises in the cases of homodyned and
generalized K-scattering.
Proposition 9.6 The transformation a
→ −axt maps the homodyned to the
generalized K-scattering diffusion tensor of the vector scattering process
(xt , Zt , Θt ).
Proof Choose an arbitrary instant of time, labelled t = 0. Define
(GK)
Ψt = axt + ψt
(9.33)
= −ax0 + ψt
(HK)
Ψt
for all t ≥ 0, coincident at t = 0. Thus ψ0 = 2ax0 + ψ0 and otherwise ψt ,
ψt are considered independent K-scattering processes. The
result is equivalent
(·)
to the corresponding (complex-valued) vector processes xt , Ψt having the
same diffusion tensor, at the chosen instant. The amplitude components are best
computed using the complex polarization, i.e.
2
iī dΨt dΨt dΨ∗t
Σ dt = . (9.34)
... dΨ∗2t
The results of Corollary 8.9 and the above relation between ψt , ψt at t = 0 imply
(·)2 (·)
that dΨ0 are identical. Likewise the expressions for |dΨ0 |2 coincide, by virtue
of the cosine rule applied to P RR of Fig. 9.1. The same method shows that
(·)
dxt dΨt are identical at the chosen instant. 2
The image point R has the physical interpretation of a fluctuating canceling
beam, π out of phase with the original t (R). A result corresponding to Proposi-
tion 9.6 does not hold for the vector scattering drift, as evident from comparing
Propositions 9.2 and 9.3.
AMPLITUDE FLUCTUATIONS 77
C f
C
c
Q u
R'
I
– 0 R
D D
half plane in Fig. 9.1) and Σ(Z,Θ) <, > 0 according as P lies in the upper/lower
half plane, while the opposite situation holds for the complement D̄.
9.4.0.10 Rice Noting that ψt is a complex Gaussian process, we see that the
familiar Rice process (Rice 1954) emerges as the model for weak scattering. If
we write the amplitude and phase of the scattered field as (E, Θ) their joint
distribution takes the form
E exp − E 2 + a2 − 2Ea cos Θ
P (E, Θ) = . (9.41)
π
From this we can derive the familiar result for the marginal p.d.f. of the field
amplitude, the Rice distribution,
P (E) = 2E exp − E 2 + a2 I0 (2Ea) , (9.42)
where I0 is the modified Bessel function of the first kind. The phase distribution
associated with the Rice scattering model can be obtained from (9.41) by inte-
gration over E. The result can be expressed in a reasonable closed form in terms
of the error function,
80 WEAK SCATTERING
PRice (冟a)
1.4
1.2
0.8
0.6
0.4
0.2
–3 –2 –1 1 2 3
∞ ∞
exp −a2 sin2 Θ 2
P (Θ|a) = P (E, Θ) dE = E exp − (E − a cos Θ) dE
π
o 0
∞
exp −a2 sin2 Θ
= (E + a cos Θ) exp −E 2 dE
π
−a cos Θ
1
a cos Θ
= exp −a2 + √ (1 + erf (a cos Θ)) exp −a2 sin2 Θ . (9.43)
2π 2 π
Figure 9.2 shows the behaviour of this function, for differing values of a, whose
square can be interpreted as a signal to noise power ratio.
Ebα α−2 %
P (x, E, Θ) = x exp(−bx) exp − E 2 + a2 x exp(2Ea cos Θ/x) .
πΓ (α)
(9.44)
The field amplitude p.d.f. associated with the homodyned K-scattering model
cannot be rendered in a simple closed form for general values of the modified25
shape parameter α. Its compound representation takes the form
25 Modified in the sense that ‘shape parameter’ often refers to the quantity ν = α − 1, in a
radar context.
ASYMPTOTIC BEHAVIOUR 81
∞
2EbαE %
P (E) = xα−2 exp(−bx) exp − E 2 + a2 x I0 (2Ea/x) dx. (9.45)
Γ (α)
0
The asymptotic phase distribution for the homodyned K-scattering model can-
not be evaluated in closed form. The compound representation of the process
indicates that the phase p.d.f. can be written as
∞
bα
P (Θ|a, b, α) = P (Θ|a, x) exp(−bx) xα−1 dx, (9.46)
Γ (α)
0
where we define
2 2 2
1 −a a cos Θ a cos Θ −a sin Θ
P (Θ|a, x) = exp + √ 1 + erf √ exp .
2π x 2 πx x x
(9.47)
This can be recast in the form
P (Θ|a, b, α)
2 α/2 √ ab α2 − 14 cos Θ a2 sin2 Θ 2 + 4
α 1
=
a b
= Kα 2 ba +2 √ Kα−1/2 2 ba2 sin2 Θ
πΓ (α) πΓ (α)
α+1 α+1 1 <
a b 2 2
cos Θ
+2 Kα−1 2 ba2 sin2 Θ + t2 cos2 Θ
πΓ (α)
0
2 α−1
× sin Θ + t2 cos2 Θ 2 dt
(9.48)
by using the integral representation of the error function
√ 1
√ 2a x cos Θ
erf a x cos Θ = √ exp −s2 a2 x cos2 Θ ds. (9.49)
π
0
PHK (冟1,a, a)
1.4
1.2 a = 0.1
0.8 a = 1.0
0.6
a = 10.0
0.4
0.2
–3 –2 –1 1 2 3
Fig. 9.3. Phase p.d.f.s derived from the homodyned K-scattering model,
α = 0.1, 1.0, 10.0.
PHK (冟a,1,1)
2
a = 0.5
1.5
0.5
a = 2.0
–3 –2 –1 1 2 3
Fig. 9.4. Phase p.d.f.s derived from the homodyned K-scattering model,
a = 0.5, 1.0, 1.5, 2.0.
(Here we have introduced the scale parameter b to relax the condition that
the power in the complex Ornstein–Uhlenbeck process is taken as unity.) This
provides us with the ‘compound’ representation of the generalized K-scattering
process in accordance with (8.3). This is to be contrasted with the corresponding
result for the homodyned K-scattering process above. Thus by integration we
have the field amplitude p.d.f. given by
4Ebα =
P (E) = I (2Ea) K 2E a2+b (9.51)
(α−1)/2 0 α−1
Γ (α) (a2 + b)
ASYMPTOTIC BEHAVIOUR 83
which is essentially the result obtained in Jakeman and Tough (1987) using the
method of characteristic functions. The calculation of the asymptotic phase dis-
tribution for Rice scattering can be extended straightforwardly to the generalized
and homodyned K-scattering models, essentially by exploiting the compound
representation (8.3). Thus using (9.47) we construct
∞
bα
P (Θ|a, b, α) = P (Θ|ax, x) exp(−bx) xα−1 dx. (9.52)
Γ (α)
0
This consists of three terms; two are straightforward while the third can be
expressed in terms of a hypergeometric function. To this end we have
∞ α
bα 1 1 b
xα−1 exp − b + a2 x dx = ,
Γ (α) 2π 2π b + a2
0
∞
bα 1 1
√ a cos Θ xα− 2 exp − b + a2 sin2 Θ x dx (9.53)
Γ (α) 2 π
0
a cos Θ Γ (α + 1/2) 1
= < 2Γ (α) √π 1 + a2 sin2 Θ%bα .
2
b + a2 sin Θ
The third term can be evaluated by substituting (9.49) and integrating over x,
thus
∞
√ √
exp − a2 sin2 Θ + b x xα−1 a x cos Θerf a x cos Θ dx
0
1
2Γ (α + 1) 2 −(α+1) (9.54)
= √ a cos2 Θ b + a2 sin2 Θ + a2 s2 cos2 Θ ds
π
0
2Γ (α + 1) a2 cos2 Θ 1 3 −a2 cos2 Θ
= √ α+1 2 F1 , α + 1; ; 2 2 .
π a2 sin2 Θ + b 2 2 a sin Θ + b
PGK (冟1, a, a)
a = 10
0.5
0.4
0.3
0.2
a = 0.1
0.1
–3 –2 –1 1 2 3
Fig. 9.5. Phase p.d.f.s derived from the generalized K-scattering model,
α = 0.1, 1.0, 10.0.
to which the above result reduces when the (modified) shape parameter α takes
integer values. Figure 9.5 shows the phase p.d.f. derived from the generalized K-
scattering model. We have chosen a = 1,
x = 1 and α = 0.1, 1, 10. Noise with
larger ‘spikes’, associated with lower values of α, results in a broader distribution
of phase. In Fig. 9.6 we show the variation in the phase distribution with the
parameter a, keeping the mean noise power
x = 1 and α = 1. The phase
distribution becomes narrower as the parameter a increases. Comparison with
Fig. 9.2 shows that, while the mean noise power is the same in each, the more
appreciable spikes in the character of the noise is manifest in a broader phase
distribution.
The most marked difference between the phase p.d.f.s derived from the ho-
modyned and generalized K-scattering models is evident at small values of α
(i.e. less than unity), where a singular behaviour is observed at the origin. This
can be seen quite clearly in Fig. 9.4. When α takes larger values, a behaviour
more reminiscent of that seen in Fig. 9.2 emerges, as the noise becomes more
Gaussian in character. In the case where α = 1, the phase p.d.f. displays a
cusp at the origin, irrespective of the value of a; this can be seen in Fig. 9.4.
The differences between the phase p.d.f.s derived from the homodyned and gen-
eralized models can be understood qualitatively in terms of the ‘signal’ fluc-
tuating with xt in the latter, but remaining constant in the former. Jakeman
and Tough (1987) discuss the implications of this difference between the mod-
els in some detail, without making explicit reference to the asymptotic phase
p.d.f.s.
ASYMPTOTIC BEHAVIOUR 85
PGK (冟a,1,1)
1
a = 2.0
0.8
0.6
0.4
0.2 a = 0.5
–3 –2 –1 1 2 3
Fig. 9.6. Phase p.d.f.s derived from the generalized K-scattering model,
a = 0.5, 1.0, 1.5, 2.0.
1
bî = Piî bi + Pijî Σij , Σîĵ = Piî Pjĵ Σij , (9.56)
2
where Piî denotes the transition matrix of partial derivatives ∂xî /∂xi , with a
corresponding notation for second derivatives. Attention should be paid to the
non-tensorial nature of the second term in the drift transformation, which is
characteristic of the Ito calculus. The probability density transforms as ρ̂J = ρ
where J is the Jacobian of P , i.e. εî1 î2 ...în J = εi1 i2 ...in Piî11 Piî22 ...Piînn . Using the
ĵ
identity ∂j log J ≡ Pĵk Pjk and the relation (9.11) we deduce the vector current
transformation V î = Piî V i , i.e. the current transforms as a tensor. The equivalent
holomorphic/anti-holomorphic representation follows, via a complex change of
coordinates. (The reader should compare orig. Field (2003) and Section 3.2 for an
86 WEAK SCATTERING
We conclude the chapter with some general remarks. The K-distribution pro-
vides a useful model of the non-Gaussian statistics of strongly scattered radiation
with a uniform distribution of phase. In recent work (Field and Tough 2003a,b) a
description of the K-scattering process in terms of SDEs has been developed that
makes direct contact with a simple underlying random walk model of scattering.
In this chapter we have extended this analysis to models of weak scattering, in
which the distribution of phase is non-uniform.
The K-scattering process can be derived from an isotropic random walk
with a fluctuating number of steps. To incorporate a non-uniform distribution of
phase, we consider a random walk on which a preferred direction or bias has been
imposed. In the case where the random walk has a large, but fixed, number of
steps, the Rice process emerges as a model for weak scattering. We have analysed
the phase distribution associated with this model, and established the connection
between its random walk formulation and a description in terms of SDEs.
The extension of the Rice scattering model to the non-Gaussian regime is
effected when we allow the number of steps in the biased random walk to fluc-
tuate. We have shown how this leads to the generalized K-scattering process
discussed in Jakeman and Tough (1987) and have made contact between this
model and the homodyned K-scattering process. In each case we have char-
acterized the associated distribution of phase in detail, and have developed a
description in terms of SDEs and their equivalent FPEs. This complements the
earlier dynamical description of K-scattering (Field and Tough 2003a,b).
The results of this chapter have implications for detection schemes where the
signal behaviour (represented by the coherent offset in the resultant amplitude)
can, to a reasonable extent, be modelled in the context of ambient K-distributed
noise (cf. the results reported in §4 of Field and Tough 2003a). The results of
Section 9.3 indicate a method for anomaly detection based on departures in
ASYMPTOTIC BEHAVIOUR 87
the geometry of the resultant amplitude fluctuations from that expected in the
pure K-scattering case. The results should find application in adaptive imaging
problems, in the de-noising of optical images (signal separation from noise, i.e.
extraction of t from Ψt ) and anomaly detection in radar back-scatter where a
(coherent) reflection contribution is involved (cf. Jakeman and Tough 1987).
10
SCATTERING FROM GENERAL POPULATIONS
In this chapter we derive the stochastic dynamics for scattering of a wavelike field
from a large population of scatterers whose dynamics is arbitrary. This leads to
a result concerning the observability of the scattering cross-section in terms of
the resultant phase fluctuations that is independent of the population dynamics.
An emergent concept is a certain notion of an ideal filter. The diffusion-based
model of K-scattering arises, encountered in Chapter 8, as a special case. The
experimental implications of the results in a variety of contexts are discussed
later in Part III.
Motivated by the possible application of the recent results on electromagnetic
scattering from random media to more general situations (e.g. medical imaging,
wireless communications) than those encompassed by the K-distribution model
(Field and Tough 2003a,b), the current chapter focuses on a significant new re-
sult in connection with inference of the scattering cross-section or population,
in local time. A special case of the result was reported previously in Field and
Tough (2003b) [see paragraph above eqn (2.35) therein], stating that the instan-
taneous values of the cross-section are deducible through the phase fluctuations
in the scattered field. This was demonstrated theoretically in the context of K-
scattering. Intriguingly, the same result holds for an arbitrary population. More
precisely, given the structure of a random walk model, component phase, and
step number fluctuations, the result holds for an arbitrary specification of popu-
lation dynamics. In this sense, the result is a geometrical feature of (the dynam-
ical extension of) Jakeman’s random walk model with step number fluctuations
(Jakeman 1980), and as such should apply to a large number of experimental
situations involving interference effects of wavelike fields arising from random
populations. From a filtering point of view, the result represents an improve-
ment on Kalman/particle filtering methods, since an exact expression for the
‘hidden’ state (the population level) in terms of the additional phase degrees of
freedom can be derived.
88
EXTENDED RANDOM WALK 89
s(j)
N 7 89
& ':
(N )
(j)
Et = aj exp iϕt (10.1)
j=1
with (constant) population size N , random phasor step s(j) . Notice here, for the
purpose of generality, the inclusion of the component amplitude ‘form factors’ aj ,
i.e. the strength of the individual scattering components may vary throughout
the medium. Also in this chapter we shall allow for the possibility that the
component phases ϕ(j) are correlated for short times, as may occur by virtue of
some special initial condition. As such, we refer to this random walk construction
as the ‘extended random walk model’ by comparison with the simpler model of
the previous chapter.
In the aforementioned references, all components of this expression are in
effect considered at a given instant of time, thus not addressing the question
of continuous time evolution properties or ‘dynamics’. This extra structure is
supplied by a (phase) diffusion model (Field and Tough 2003b) which takes
(j)
the component phases {ϕt } to be a collection of (displaced) Wiener processes
(j) 1 (j)
evolving on a suitable timescale. Thus ϕt = ∆(j) + B 2 Wt , with component
(j)
random initialization ∆ uniformly distributed on the interval [0, 2π). In situa-
tions where the component phasors s(j) are aligned initially (e.g. for the received
‘T 2’ signal arising in magnetic resonance imaging, described e.g. in Ernst et al.
1987 and discussed in Chapter 14) (cf. also Field and Bain 2008), the initial-
izations ∆(j) are identical for all j, whereas in some cases (e.g. the statistical
description of radar scattering from the sea surface, cf. §4(b) in Field and Tough
2003a) it is more appropriate to draw ∆(j) independently. In any case, these
primitive assumptions enable us to derive the dynamics of Rayleigh scattering,
essentially from first principles. The stochastic differential of (10.1), according
to Ito’s formula (e.g. Oksendal 1998; Karatzas and Shreve 1988), is given by
N
& '
(N ) (j) 1 (j) 2 (j)
dEt = aj idϕt − dϕt exp iϕt . (10.2)
j=1
2
If we write dζt for the first term on the right-hand side above, then for t ≥ T ,
(j)
where T is the ‘phase decoherence’ time such that {ϕt | t ≥ T } have negligible
1/2
correlation, we have |dζt |2 = j aj Bdt, and therefore dζt =
2 2
j aj B 1/2 dξt
where ξt is a complex-valued Wiener process (satisfying |dξt |2&= dt, dξt2' = 0).
(N ) 1
Defining the (normalized) Rayleigh amplitude by γt = limN →∞ Et /N 2 leads
to the resultant dynamics (cf. Field and Tough 2003b).
Proposition 10.1 For sufficiently large times t ≥ T the dynamics of Rayleigh
scattering is given by the complex Ornstein–Uhlenbeck equation
90 GENERAL POPULATIONS
1 1 1
dγt = − Bγt dt + B 2
a2 2 dξt . (10.3)
2
If {∆(j) } are assumed independent then the result holds for arbitrarily small
times.
Remarks. On the asymptotic distribution of phase. If the phases are initialized
at some value, say the real direction in the complex plane, then the distribution
of the resultant Rayleigh amplitude is Gaussian, of non-zero mean, by virtue of
the central limit theorem applied to the i.i.d. collection of random phasors (with
a square root scaling). Observe that, for small times, the major axis of its error
surface S – an ellipse – is oriented in the direction of the imaginary axis, since
the magnitude of the fluctuations (for each component phasor) in that direction
is dominant. As relaxation occurs, i.e. the asymptotic equilibrium distribution is
approached, the surface S stretches in the real direction, and tends to a circular
geometry in the limit t → ∞. At the same time, the mean value – geometrically
the centre of gravity of the ellipse – tends to the origin of the complex plane.
Accordingly, the resultant asymptotic distribution of phase is uniform.
If we re-scale the (Rayleigh) amplitude according to γt
→
a2 −1/2 γt , then
the re-scaled field satisfies (10.3) with the form factors equal to unity. In what
follows we shall therefore assume the field to be scaled in this way, i.e.
a2 =
1. In the case of a fluctuating number of steps N
→ N + t in (10.1),
, we define
the (continuous-valued) cross-section as xt = limNt →∞ Nt /N̄ . The resultant
+ ,
(normalized) amplitude ψt = limN →∞ E (Nt ) /N̄ therefore has the compound
representation
1
ψt = xt2 γt , (10.4)
& 1
'
(N )
where γt = limN →∞ Et t /Nt2 , and in which xt and γt are independent pro-
cesses. The intensity zt has the compound representation zt = xt ut , where
ut = |γt |2 is the instantaneous intensity of the component (unit power) Rayleigh
process (cf. the analysis of asymptotic behaviour and propagators in §s III, IV
of Field and Tough 2003b).
It is worth clarifying at this point the precise meaning and definitions of
the various amplitudes that have occurred in the exposition of the random walk
model. We begin with E, as the superposition of N random phasors, where N is
fixed. For a large population, N → ∞, and the root mean square (r.m.s.) of E
tends to infinity. Thus, to obtain a finite resultant in the limit of an asymptot-
ically large number of scatterers, we define a ‘normalized’ Rayleigh amplitude
1
γ, by dividing through by the r.m.s. value N 2 . (Equivalently, we could absorb
this normalization into the form factors aj .) The term ‘Rayleigh’ refers to the
fact that the number of scatterers is fixed. In the general case that the scatter-
ing population fluctuates in time, we define the ‘normalized’ amplitude ψ as in
1
the case of Rayleigh scattering, dividing E by the r.m.s. value N̄ 2 , where now
GENERALIZED DYNAMICS 91
the number of terms Nt in the random walk fluctuates. Re-arranging the result-
ing expression produces the compound representation of the resultant amplitude
(10.4), for a general scattering process. The Rayleigh amplitude is then recovered
if the cross-section is unity.
A corresponding dynamical situation for ‘weak’ scattering processes, i.e. where
the field ψt lies in (weak) superposition with a coherent offset signal t , should
be possible for a general type population as a generalization of the results de-
scribed in Chapter 9 (orig. Field and Tough 2005), although this extension is
not addressed in the current monograph.
derived according to the scheme outlined at the end of the previous section. For
arbitrary γt , xt an application of Ito’s formula to (10.4) yields
This enables the resultant amplitude dynamics to be calculated under the as-
sumption that γt is a unit power Rayleigh process according to (10.3), with unit
form factors.
Proposition 10.2 The generalized resultant amplitude dynamics is given by
1 1
dψt bt Σt 1 AΣt 2 (x) B2
= A − 2 − B dt + dWt + dξt . (10.7)
ψt 2xt 4xt 2 2x2t γt
Observe that ∂/∂B acting on the drift/volatility parameters in (10.7) yields ex-
pressions that are independent of bt , Σt , as expected from the endogenous spec-
ification of population dynamics (10.5). Using the vanishing of the Ito products
(x)
dξt2 , dξt dWt , and the property |dξt |2 = dt, the above result yields the squared
amplitude fluctuations as follows.
Corollary 10.3
2
dψt AΣt
= dt, (10.8)
ψt 2x2t
AΣt zt
|dψt | =
2
+ Bxt dt. (10.9)
2x2t
The generalized intensity dynamics can be computed from Proposition 10.2 and
the identity dzt ≡ ψt dψt∗ + ψt∗ dψt + dψt dψt∗ .
Proposition 10.4 The generalized intensity SDE is given by
bt zt 1 zt (x) 1 (r)
dzt = A + B(xt − zt ) dt + (2AΣt ) 2 dWt + (2Bxt zt ) 2 dWt
xt xt
(10.10)
where
12
2zt
(γt∗ dξt + γt dξt∗ ) ≡
(r)
dWt . (10.11)
xt
dψt 1 dψt
dθt ≡ − , (10.13)
ψt 2 ψt
where denotes the imaginary part. Since the right-hand side of (10.8) is real-
valued, only the first term on the right-hand side of (10.13) contributes to dθt ,
in respect of which
1
dψt B 2 dξt dξt∗
= − ∗ . (10.14)
ψt 2i γt γt
Observe that, in contrast to the situation for the resultant amplitude and inten-
sity SDEs (10.7) and (10.10), this is functionally identical to the corresponding
result in K-scattering (i.e. independent of the population parameters bt , Σt ),
the essential difference lying in the evolutionary structure of the processes xt , zt .
Observe from (10.11), (10.16) that the radial and angular fluctuations in the re-
sultant amplitude are statistically independent in the general case. The squared
phase volatility obtained from (10.15) leads to the central result of the chapter.
Theorem 10.6 The instantaneous values of the scattering cross-section are ob-
servable through the intensity-weighted squared phase fluctuations according to
2 dθt2
xt = zt (10.17)
B dt
if xt is an Ito process, not necessarily a diffusion, and throughout space and time.
The (experimental) significance of this result is that the relation (10.17) is
exact and moreover independent of the dynamics of xt . The result resembles (but
is distinct from) the minimal variance of the intensity-weighted phase derivative
94 GENERAL POPULATIONS
Fig. 10.1. Geometry of random walk for generalized scattering process – gener-
ically (a.e.) each component phasor point lies on two circles.
discussed in Jakeman et al. (2001), for differentiable processes. In the present sit-
uation however, the processes considered are not differentiable, and instead the
squared phase differential arises. Since the elements of the random walk model
(10.1), preserved by the generalized framework given in the exposition of this
section for dealing with general populations, are the only essential ingredients
involved here, the result is geometrical in nature (cf. Fig. 10.1). With regard to
this geometry, the derivations above show that the dynamics of the resultant
field are not affected if the radii for each component phasor are drawn indepen-
dently from an arbitrary probability distribution. (The result of the theorem was
anticipated from a physical point of view previously in Field and Tough 2003a;
see discussion following Prop. 4.1. therein.)
A slight complication is posed in the computation of dθt2 from experimental
data, owing to the discontinuous-valued behaviour of θt at coordinate intervals
of 2π. This is resolved by instead using the (continuous-valued) ‘phase-wrapped’
process wt = exp(iθt ), whose stochastic differential is dwt = exp(iθt )[idθt − 12 dθt2 ],
which enables the squared phase fluctuations to be computed from the single-
valued process wt via |dwt |2 = dθt2 . In respect of discrete-time implementation,
we remark that if Wt is a Wiener process, then δWt = Wt+h − Wt is normally
distributed as N (0, h), so that its square is a ‘chi-squared’ χ2 (1) variable. The
sum of n such variables is therefore distributed as χ2 (n), from which an esti-
mate of dqt2 from δqt can be obtained (via the weak law of large numbers) by
GENERALIZED DYNAMICS 95
In this part we develop simulation techniques for scattering processes and de-
scribe some experimental tests that verify our theoretical development. This de-
velops the interrelationships between theoretical prediction, real, and simulated
data, and is illustrated in a specific case of scattering from a general population.
Chapter 11 provides a detailed account of the simulation of K-scattering
processes, including the influence of Doppler, target returns, and improved ac-
curacy second-order algorithms for numerical integration of the gamma process.
In Chapter 12 we provide two independent experimental tests of the proposed
stochastic theory, and provide a scheme for anomaly detection. The first in Sec-
tion 12.1 is a study of data gathered from electromagnetic scattering from a
random phase screen, in the absence of anomalies. A study is made of the (for-
ward) scattered intensity which is observed to be very closely K-distributed. The
volatility function of the intensity is measured and observed to correlate very
strongly with the instantaneous square root intensity, thus providing a calibra-
tion of the stochastic model. This calibration setting is applied to an independent
set of (complex-valued) radar return data in Section 12.2. The radar scattering
is from a region of the ocean surface in which a tethered anomalous reflecting
body is placed and whose location is known. It is previously known that such data
typically contains a large number of spurious ‘clutter spike’ anomalies and that
the phase behaviour is an important indicator of a genuine anomaly (Luttrell
2001). Accordingly, we derive a complex-valued stochastic differential equation
(SDE) for the amplitude of the scattered radiation, under the assumption that
the phase behaviour is uniform. We exhibit a strong correlation between the ob-
served versus the theoretical prediction of the volatility for the complex-valued
process in the clutter domain. At the location of the anomaly this correlation be-
comes weak, which provides a means of detection, isolating the anomaly to a high
degree of accuracy. With the theory firmly established from these experiments,
Chapter 13 provides a discussion of the non-linear character of the stochastic
dynamics of radar sea clutter, from independent experimental and theoretical
points of view.
The part concludes in Chapter 14 with some simulations that illustrate the
concepts of scattering from general populations as developed earlier in Chapter
10. The implications of these results for tracking populations in various physical
contexts are discussed.
97
This page intentionally left blank
11
SIMULATION OF K-SCATTERING
99
100 K-SCATTERING SIMULATION
It is natural to develop the drift F and volatility σ in Taylor series, which yields
F (n) σ (n)
t t
n n
∆xt = (∆xu ) du + (∆xu ) dWu . (11.4)
n
n! n
n!
0 0
This simple stochastic analogue of the Euler integration algorithm forms the
basis of many widely used Brownian dynamics algorithms that incorporate hy-
drodynamic interactions (Ermak and McCammon 1978). Taking the first four
terms in the Taylor expansions in (11.4) gives us
t t t
(1) 1 2
∆xt = F t + σWt + F (1)
∆xu du+σ ∆xu dWu + F (2) (∆xu ) du
2
0 0 0
t t t
1 2 1 3 1 3
+ σ (2) (∆xu ) dWu + F (3) (∆xu ) du+ σ (3) (∆xu ) dWu + · · · .
2 6 6
0 0 0
(11.6)
t
(1)
σ ∆xu dWu . (11.8)
0
Not all these are provided by the lowest order solution (11.5) when it is intro-
duced into (11.1). Nonetheless, it does generate terms of orders t, t3/2 , t2 . These
in turn should be re-inserted into (11.1) to produce extra terms in the iteration.
Thus, from (11.5) alone, we have
t2
∆xt = F t + σWt + F F (1)
2
t t
3/2 (1)
(1)
+F σ Wu du O t + Fσ udWu O t3/2
0 0
t
+σ (1) σ Wu dWu (O (t))
(11.9)
0
t t
1 1
+ F (2) σ Wu2 du O t2 + σ (2) σ 2 Wu2 dWu O t3/2
2 2
0 0
t t
σ (3) σ 3
+σ (2)
σF uWu dWu O t2 + Wu3 dWu O t2 .
6
0 0
t
(1)
σ σ Wu dWu (11.10)
0
t
1 (2) 2
σ (∆xu ) dWu . (11.11)
2
0
t u
F (1) σ (1) σ du Ws dWs (11.12)
0 0
102 K-SCATTERING SIMULATION
and
t u
σ (2) σ (1) σ 2 dWu Wu dWs Ws . (11.13)
0 0
2
Each of these is of O(t ) and so need be iterated no more.
We now turn our attention to (11.8); including all terms up to O(t3/2 ) ob-
tained thus far we generate the following new terms
t u t u
(1) 2 (1) (1)
σ σ dWu Ws dWs + σ F σ dWu Ws ds
0 0 0 0
t u t u
1
+ σ (1) F σ dWu s dWs + σ (1) σ (2) σ 2 dWu Ws2 dWs . (11.14)
2
0 0 0 0
The last three of these terms are of order t ; the first, however, is O(t3/2 ) and 2
This process could be carried on to higher order; the number of terms rapidly
becomes very large. Perhaps the most striking qualitative difference between this
and the corresponding result obtained in the additive noise case is the presence
of a term with a linear time dependence in the stochastic part of the increment.
Now that we have developed an iterative solution we calculate the expecta-
tion value of the increment in x occurring over the time t. All terms in (11.16)
containing odd powers of W average to zero. Among those containing even pow-
ers of W the majority of the expectation values go to zero as a consequence of
the Ito rule (2.6)
t
u
E dWu dWs g (u, s) = 0. (11.17)
0 0
t u t
Wt2 − t
dWu dWs = dWu Wu = ; E Wt2 = t,
2
0 0 0 t (11.18)
u
E dWu dWs = 0.
0 0
t
(1) t 12
E (∆xt ) = F t + F F + F (2) σ E Wu2 du
2 2
0
t2 1
= F t + F F (1) + F (2) σt2 . (11.19)
2 4
This is identical with the corresponding result obtained in the additive noise
case, to second order in time.
We now wish to evaluate the mean square of the stochastic part of the incre-
ment, i.e.
t2
∆x̂t = ∆xt − F t − F F (1) (11.20)
2
again retaining only terms up to second order in t. (We have introduced the caret
notation because superscript notation has already been used in this case.) Thus
104 K-SCATTERING SIMULATION
we need only consider the terms in (11.16) up to and including those that are
O(t3/2 ),
t t
(1) (1)
∆x̂t =σWt + F σ Wu du + F σ u dWu
0 0
t t
(1) 1
+σ σ Wu dWu + σ (2) σ Wu2 dWu + O t2 . (11.21)
2
0 0
0
t
+ 2F (1) σ 2 E Wt Wu du
0
t t
+ 2F σ (1) σE Wt u dWu + σ (2) σ 2 E Wt Wu2 dWu + O t2 .
0 0
(11.22)
The various expectation values can be evaluated, using the Ito calculus. We recall
the familiar relation E[Wt2 ] = t. The Ito isometry (see Section 2.1.1) allows us
to evaluate the next contribution:
2
t t
t2
E Wu dWu = E Wu2 du = . (11.23)
2
0 0
The relationship between this SDE and other descriptions of the intensity of a
Gaussian speckle process is discussed in Tough (1987). To make contact with the
general results above, we make the identifications
F (z0 ) = (1 − z0 ) ,
F (1) (z0 ) = −1,
(2)
F (z0 ) = 0
√ (11.29)
σ (z0 ) = 2z< 0,
(1)
σ (z0 ) = 2 z20 ,
1
<
σ (2) (z0 ) = − 1 23 .
4 z 0
t2 1
E (∆z) = F t + F F (1) + F (2) σt2 + O t3
2 4
(11.30)
t
= (1 − z0 ) t 1 − + O t3 .
2
The mean square value of the stochastic part of the increment in z can be
evaluated as follows
28 This can be derived by calculating the stochastic differential d(W 4 ), integrating, and taking
t
the expected value.
106 K-SCATTERING SIMULATION
2 2
2 t
E (∆ẑ) = σ 2 t + σ (1) σ + 2F (1) σ 2 + 2F σσ (1) + σ (2) σ 3 + O t3
2
3
= 2z0 t + (1 − 3z0 ) t + O t .
2
(11.31)
In the special case of the Rayleigh process described by (11.28) further
progress can be made, albeit by eschewing the iterative approach. Instead, we
identify the FPE that is stochastically equivalent to this SDE and solve explicitly
for its propagator. Using this we can evaluate expectations of the form
+ ,
E [(zt − zo )] , E (zt − zo )2 (11.32)
for arbitrary time intervals t. By making a suitable small time expansion we
can check the validity of our more general analysis in this case. The statistics
of this increment are of interest in the context of improving the precision in
the stochastic volatility analyses of Chapter 12; the results below that hold for
arbitrary times identify precision corrections to the O(t) treatment that we have
in practice applied to the experimental data analysis, which serves to illustrate
the theory sufficiently well. However, in situations that present themselves where
the sample time is not as small as one would like, the higher order correction
analysis we provide here should be useful.
We recall that the FPE equivalent to (11.28) is
∂2 ∂ ∂
(zP (z, t|z0 )) + ((z − 1) P (z, t|z0 )) = P (z, t|z0 ) (11.33)
∂z 2 ∂z ∂t
whose fundamental solution or propagator satisfies the initial condition
P (z, 0|z0 ) = δ (z − z0 ) . (11.34)
(Further background information on this topic can be found in Tough 1987;
our notation follows that of this reference quite closely.) The propagator can be
expanded in terms of the Laguerre polynomials as
∞
P (z, t|z0 ) = exp(−z) Ln (z) Ln (z0 ) exp(−nt) (11.35)
n=0
n
1 d
Ln (z) = exp(z) (exp(−z) z n ) .
n! dz
Using this expansion, and some basic properties of these polynomials, we find
that
∞
E (zt − z0 ) = (z − z0 )P (z, t|z0 ) dz
0
= (1 − z0 ) (1 − exp (−t))
t2 t3
= (1 − z0 ) t − + + O t4 (11.36)
2 6
RAYLEIGH, GAMMA PROCESSES 107
∞
2 2
E (zt − z0 ) = (z − z0 ) P (z, t|z0 ) dz
0
2 2
= (1 − exp(−t)) + z0 (4 exp(−t) − 2) (1 − exp(−t)) + 2 (1 − exp(−t))
z02
13
= 2z0 t + 2 − 5z0 + z0 t − 2 − z0 + z0 t3 + O t4 .
2 2 2
(11.37)
3
Higher order moments can, in principle, be determined from the characteristic
function, which is most conveniently derived using an alternative form for the
propagator
√
1 (z + z0 θ) 2 θzz0
P (z, t|z0 ) = exp − I0 ,
1−θ 1−θ 1−θ
θ = exp(−t) . (11.38)
Here I0 is a zeroth-order modified Bessel function of the first kind. Using this
we find that
∞ (−s)n E ((z − z )n )
t 0
E(exp(−s (zt − z0 ))) =
n=0 n!
∞ √
1 θ 1 2 θzz0
= exp z0 s + exp −z s + I0 dz
1−θ 1−θ 1−θ 1−θ
0
sz0 (1 + s) (1 − θ)
exp
1 + s (1 − θ)
= . (11.39)
1 + s (1 − θ)
Expanding this characteristic function in a power series and equating coefficients
of powers of s verifies the results (11.36) and (11.37). These in turn allow us to
verify the results obtained to order t2 by iteration. In particular, the mean square
of the stochastic part of the increment can be calculated as
2 2
E ∆ẑ 2 = E (zt − z0 ) − (1 − z0 ) t2 + O t3
= 2z0 t + (1 − 3z0 ) t2 + O t3 . (11.40)
So far in this chapter we have examined the iterative solutions of some SDEs
that occur in the description of clutter processes. We have considered the case
where the SDEs incorporates multiplicative noise, and so has been interpreted
within the Ito convention. A corresponding simpler treatment holds for the case
of iterative solutions of ordinary differential equations and Langevin equations
incorporating additive noise. The greater complexity of the multiplicative noise
calculations has limited us to results valid up to second order only, whereas the
corresponding Langevin case would naturally yields results valid up to third order
in time. This fact highlights some of the problems encountered in the analysis of
systems with multiplicative, as opposed to additive, noise.
The general results obtained have been specialized to the relatively simple
SDEs that describe the Rayleigh and K-scattering processes. As these incorpo-
rate multiplicative noise, we have generated results valid up to second order in
time. Recent experimental work (Field and Tough 2003a) has focussed attention
of the behaviour of increments in the intensity of the Rayleigh and K processes
that occur over time intervals that are very short in comparison to the correlation
timescale of the process concerned. Accordingly, this analysis can be carried out
using a first-order expansion in time. The results we present here are useful in
assessing the validity of this approximation in cases where the relevant sample
time interval, for example as dictated by experimental design, is a significant
proportion of the correlation time. In the special case of the Rayleigh process
we are able to make significantly more progress, by adopting the Fokker–Planck
formulation of the problem. The results we derive specialize to provide a verifi-
cation of the analysis for more general processes, and should be of relevance to
analyses of empirical data with shorter correlation timescales and constraints on
the available sample frequency ranges.
In addition to providing some insight into the temporal development of
the system described by SDEs incorporating multiplicative noise, the results
of this iterative analysis also allow us to simulate their behaviour numerically, in
effect providing numerical solution to the SDEs concerned. One way to achieve
this, which has been applied to Langevin type equations (Greenside and Helfand
1981) is very much in the spirit of the Runge–Kutta algorithm, frequently ap-
plied to the solution of ordinary differential equations. These workers develop
a method of stepping forward in time, combining several contiguous values of
the dependent variables in such a way that agreement with a Taylor series de-
velopment of the solution is maintained to some stated order in the time step
(Whittaker and Watson 1969). When applying this method to SDEs, both the
deterministic and stochastic parts of the solution have to be controlled. Given the
difficulties encountered in characterizing the stochastic part of the solution, it is
reasonable to require agreement to second order in the time increment; typically
Runge–Kutta algorithms applied to deterministic equations seek agreement to
fourth or even higher order. We shall apply these methods to the solution of the
SDEs discussed in Section 11.3, and so provide a more accurate method for the
simulation of K-scattering processes.
COMPOUND K 109
This unit power Rayleigh process can be simulated over arbitrary time steps. To
yield a K-scattering process it should be multiplied by the correlated gamma
INFLUENCE OF DOPPLER 111
process,
generated
by the solution of (8.9) which will typically have accuracy
2
to O (A∆t) characteristic of the entire compound process. Observe that the
procedure we have outlined here provides us with a relatively efficient method
for the simulation of the complex amplitude of the scattered field, arising as the
product according to (8.3).
The SDE associated with the intensity z = p2 + q 2 is obtained using the usual
rules of Ito calculus as
= (z)
dz = β (1 − z) dt + 2βzdWt . (11.51)
∂P (z, t) ∂ ∂2
=β ((z − 1) P (z, t)) + β 2 (zP (z, t)) (11.52)
∂t ∂z ∂z
which has the stationary solution P (z) = exp (−z). The correlation function of
the intensity implicit in this model exhibits a simple exponential decay, falling
off as exp(−βt). Because the SDEs for the COU process are linear they can be
integrated over an arbitrary time step, as explained earlier. Thus we have
112 K-SCATTERING SIMULATION
p0 = p (0), q0 = q (0) ;
∆p = p (t) − p (0); ∆q = q (t) − q (0)
∆p cos ωt − sin ωt 10 p0
= exp(−βt/2) −
∆q sin ωt cos ωt 01 q0
4
(1 − exp (−βt)) gp
+ . (11.53)
2 gq
The g-functions are drawn independently from a zero mean, unit variance Gaus-
sian distribution. These results form the basis of a simple simulation of coherent
clutter, which is discussed in more detail in the next section.
To establish the behaviour of the square modulus @ of the increment
A in the
2 2
complex field Ψ = p + iq, we make an expansion of (∆p) + (∆q) ; here the
angular bracket denotes an averaging over the gs. In this way, we find that
@ A
2 2
(∆p) + (∆q) = 1 − exp(−βt) + p20 + q02 (1 + exp(−βt)
−2 exp(−βt/2) cos ωt) . (11.54)
We can expand this up to second order in t as
@ A
2 2 1 z0 ω 2 z0
(∆p) + (∆q) ∼ βt + β t − +
2 2
+ 2 ,
2 4 β (11.55)
2 2
z0 = p0 + q0 .
To ascertain the effects of the Doppler evolution on the measurement of the
volatility in the complex field (which we can see from (11.55) to be equal to β),
we must introduce estimates of the various terms in this expression.
An independent analysis of the data presented in Section 12.2 (Ward 2002),
for which the time step t is known to be 1 ms, indicates that the intensity
correlation function decays in the order of 10 ms; thus we approximate β ≈ 100.
The Doppler frequency is identified as 100 Hz, i.e. an angular frequency ω =
200π. When these values are introduced into (11.55) we find that
|∆Ψ|2
βt = 1 + 0.1 − 12 + 4π 2 + 14 z0 (11.56)
≈ 1 + 4z0 .
From this we see that, for the data analysed in Field and Tough (2003a) and
Section 12.2, Doppler induced effects are comparable with the diffusive evolution
of the complex field and so can influence the determination of the volatility. It
is particularly striking that the intensity dependence of the Doppler term is the
same as that interpreted as a contribution to the volatility of the type previously
accounted for by the endogenous model. The interaction of the diffusive, Doppler,
and power modulation terms and their effects on the performance of the anomaly
detection can best be assessed by simulation studies. Methods with which these
may be implemented will be discussed in the following.
COHERENT CLUTTER 113
29 A Mathematica notebook that implements this and other simulation procedures is available
1.5
0.5
–1.5
–2
1.5
0.5
1.5 1 1 1.5
–1.5
–2
1.5
0.5
20 40 60 80 100
Fig. 11.4. Comparison of the intensity (solid) and measured field volatility (dot-
ted) of a COU process with β = 1.0, ω = 6.0, ∆t = 0.1.
SECOND-ORDER ALGORITHMS 117
20 40 60 80 100
Fig. 11.5. Comparison of the intensity (solid) and measured field volatility (dot-
ted) of a COU process with β = 1.0, ω = 0, ∆t = 0.1.
1.5
0.5
20 40 60 80 100
Fig. 11.6. Comparison of the intensity (solid) and measured intensity volatility
(dotted) of a COU process with β = 1.0, ω = 0, ∆t = 0.1.
Appendix D.
118 K-SCATTERING SIMULATION
–1
–2
Fig. 11.7. An IQ plot of a target plus clutter return; the target power is one
quarter that of the clutter, while the target Doppler is twice that of the
clutter.
given by Klauder and Petersen (1985), whose results have been re-derived re-
cently by Quiang and Habib (2000). In essence, these workers determined the
mean and variance of the increment in a process described by a SDE, to second
order in the time step and expressed in terms of the drift and volatility functions
and their derivatives.
A simple simulation of the gamma process proceeds as follows. Integrating
(8.9) we obtain the simple recursion
=
x (n + 1) = x (n) + A (ν − x (n)) ∆t + 2Ax (n) ∆tg (n) . (11.57)
Here ∆t is the time step over which the process evolves and the g(n) are indepen-
dent Gaussian distributed random variables, with zero mean and unit variance.
This simple procedure reproduces the mean and variance of the increment in the
process to linear order in ∆t, and is frequently referred to as an Euler algorithm.
SECOND-ORDER ALGORITHMS 119
∂ ∂ ∂2
P (x, t|x0 ) = ((x − ν) P (x, t|x0 )) + 2 (xP (x, t|x0 )) . (11.58)
A∂t ∂x ∂x
The propagator of this can be constructed in a reasonable closed form32 as
(ν−1)/2
1 x exp (At) (x + x0 exp (−At))
P (x, t|x0 ) = exp −
1 − exp (−At) x0 1 − exp (−At)
√
2 exp (−At/2) xx0
× Iν−1 . (11.59)
1 − exp (−At)
In the absence of the stochastic driving term, (8.9) reduces to
32 Here I
ν−1 is a modified Bessel function of the first kind (see Section 9.6 in Abramowitz
and Stegun 1970).
120 K-SCATTERING SIMULATION
dxD
= A (ν − xD ) (11.60)
dt
which can be solved to yield
xD = ν − (ν − x0 ) η,
(11.61)
xD ≡ xD (t) , x0 ≡ x (0) , η = exp(−At) .
Here, the subscript D denotes the deterministic part of x. We can characterize
the stochastic part of the development of x by evaluating the generating function
∞
C (s) = exp(−s (x − xD )) P (x, t|x0 ) dx. (11.62)
0
@
(x − xD )A = 0,
2
(x − xD ) = (1 − η) [(1 − η) ν + 2ηx0 ] , (11.66)
@ A
3 2
(x − xD ) = 2 (1 − η) [(1 − η) ν + 3ηx0 ] .
Finally, we introduce the above expression for η into these results, and expand
up to cubic order in the time step, which leads us to
(x − xD ) = 0,
@ A
2 2 7x0 3
(x − xD ) = 2x0 At + (ν − 3x0 ) (At) + − ν (At) + · · · ,
3
@ A
3 2 3
(x − xD ) = 6x0 (At) + 2 (6x0 − ν) (At) + · · · . (11.67)
It is interesting to note that the mean cube of the stochastic part of x has a
quadratic time contribution; in earlier work (Klauder and Petersen 1985; Quiang
SECOND-ORDER ALGORITHMS 121
and Habib 2000; Pusey and Tough 1982) this term has not been analysed explic-
itly for SDEs with multiplicative noise, and has been assumed to be third order
in time.
If we now expand xD to second order in time as
xD = x0 + (ν − x0 ) At (1 − At/2) (11.68)
we see that we can construct an algorithm similar in structure to (11.57) that
nonetheless generates increments whose mean and variance are maintained to
quadratic order in the time step:
A∆t
x (n + 1) = x (n) + A (ν − x (n)) ∆t 1 −
2 (11.69)
<
2
+ 2Ax (n) ∆t + (ν − 3x (n)) (A∆t) g (n) .
One of the problems faced when generating a gamma Markov process by
simulation of the solution of the SDE (8.9) is that presented by a particularly
large random term causing x to take a negative value precluded by the natural
barrier inherent in the SDE itself. This potential inconsistency can be removed
by working with the logarithm of x, which can legitimately take negative values.
It is straightforward to construct the SDE satisfied by y = log x; application of
Ito’s formula yields
dy 1 d2 y 2
dy = dx + (dx)
dx 2 dx2
1 √ 1 1
= (ν − x) dt + 2x dW − · 2x dt
x 4 2 x2
ν−1 2
= − 1 dt + dW
x x
√ −y
= ((ν − 1) exp (−y) − 1) dt + 2 exp dW. (11.70)
2
While this identification of the appropriate SDE is sufficient for us to construct an
Euler type algorithm, essentially by inspection, we might wish to include terms
of higher order in the time step, much as in the direct simulation of the gamma
variate itself. To calculate the required higher order statistics of the increment
in the log gamma process we again make use of the analytic form (11.59) of the
propagator of the gamma Markov process. The generating function approach
adopted earlier has to be modified slightly. An appropriate moment generating
function can be constructed as follows:
s
x
exp(s (log (x) − log (x0 ))) =
- s . x 0
x 2
@ A
= 1 + s
(log (x) − log (x0 )) + s2 (log (x) − log (x0 )) · · · .
2
x0
(11.71)
122 K-SCATTERING SIMULATION
The integration over the modified Bessel function can be effected analytically
(eqn 11.4.28 in Abramowitz and Stegun 1970) to give us
- s . s
x (1 − η) ηx0 Γ (ν + s) x0 η
= exp − 1 F1 ν + s; ν; . (11.73)
x0 xs0 1−η Γ (ν) 1−η
η = exp(−A∆t) . (11.77)
We now make an expansion in both s and ∆t from which we can extract the
mean and mean square values of the increment in the logarithm. Thus we have
SECOND-ORDER ALGORITHMS 123
- s .
x ν−1 3ν − ν 2 − 2 (ν − 1) 2
=1+s − 1 A∆t + + (A∆t) · · ·
x0 x0 x20 x0
2
A∆t ν − 5ν + 5 3 − 2ν 1 2
+s2
+ + + (A∆t) · · · + · · · .
x0 2x20 2x0 2
(11.78)
From this we can read off
ν−1 3ν − ν 2 − 2 (ν − 1) 2
(log (x) − log (x0 )) = − 1 A∆t + + (A∆t)
x0 x20 x0
@ A
2 A∆t ν 2 − 5ν + 5 3 − 2ν 2
(log (x) − log (x0 )) =2 + + + 1 (A∆t)
x0 x20 x0
(11.79)
and calculate the variance of the logarithmic increment as
@ A A∆t 4 − 3ν + x0
2 2 2
(log (x) − log (x0 )) −
(log (x) − log (x0 )) = 2 + (A∆t) .
x0 x20
(11.80)
The linear terms in time in the mean and variance of the increment are those that
one would expect from (11.70) and might incorporate into an Euler algorithm.
The quadratic correction terms can be accounted for just as was done in (11.69)
for the case of the gamma, as opposed to the log-gamma, process. In a final
step we regenerate the gamma process by exponentiation. Thus, the log-based
algorithm can be summarized as follows:
− 1) + ((ν
y (n) = y (n −1) exp(−y (n − 1))− 1) A∆t
3ν − ν 2
− 2 exp(−2y (n − 1))
+ (A∆t)
2
+ (ν − 1) exp(−y (n − 1))
B
(4 − 3ν) exp(−2y (n−1)) 2
+ 2A∆t exp(−y (n−1)) + (A∆t) g (n)
+ exp(−y (n−1))
x (n) = exp(y (n)) .
(11.81)
The final component of the clutter simulation within the compound model is
the generation of a gamma process, described by a SDE. Simple Euler algorithms
are almost invariably used to effect such simulations. Algorithms of higher order
accuracy have been discussed in the literature; their generality unfortunately
renders them rather forbidding. However, in the case of the gamma process,
significant analytic progress has been made and simple algorithms, of an accuracy
equivalent to that of these more general methods, have been constructed for the
generation of the gamma variate and its logarithm.
12
EXPERIMENTAL TESTS
In this chapter we study two independent examples of data gathered from elec-
tromagnetic scattering from random media. The first example involves scattering
from a uniform randomizing optical phase screen, which is applied as a calibra-
tion tool for the intensity volatility in our diffusive model of K-distributed noise.
The second example consists of ocean surface radar scattering data that contains
K-distributed sea clutter, a multitude of ‘clutter spike’ anomalies, and a genuine
target for which our methodology is used to isolate the anomaly with respect to
range.
In both cases we shall study the electromagnetic scattered amplitude stochas-
tic processes and compare the empirically observed volatilities with the theoreti-
cal predictions from our stochastic model. In the absence of anomalies we expect
a strong correlation between these two quantities, via the standard statistical
correlation measure
(X − X̄)(Y − Ȳ )
c(X, Y ) = . (12.1)
(VarX VarY )1/2
Observe that this standard measure, which satisfies −1 ≤ c(X, Y ) ≤ 1, is zero for
independent variables, unity for ‘perfectly correlated’ variables, and is invariant
under (positive) affine transformations of the random variables X
→ λX + µ,
λ > 0.
Since the data were supplied in discrete time, an important issue arises con-
cerning the instantaneous observability of the stochastic volatility explained in
Lemma 5.4. Although the relations (5.24), (5.25) are exact within the framework
of the Ito calculus, these differential relations are idealizations of the incremen-
tal properties that can be physically measured. In practice, physical observations
occur at a discrete set of times and, in accordance with the fact that the physi-
cal processes exist in continuous time, the significance of the Ito calculus is that
sample times remain unspecified. For discrete time sampling of an Ito process
qt , satisfying the stochastic differential equation (SDE) dqt = µt dt + σt dWt , we
may write
where nt is drawn from a normal distribution with zero mean and unit variance
N (0, 1), so
n2t = 1. It follows that δqt2 = σt2 n2t δt + o(δt). Given the Ito inter-
pretation of the (discrete) SDE (12.2), the independent increments property of
Wt implies that σt and δWt = nt δt1/2 are independent (this independence also
124
OPTICAL WAVELENGTH 125
33 We assume that the data are sampled at a sufficiently high frequency for the o(1) term in
(12.3) to be negligible so that δt1/2 δt and the δWt contribution to (12.2) dominates the
drift over a sample period, as evidenced, for example, in Fig. 12.3.
126 EXPERIMENTAL TESTS
–2
–4
Logarithmic probability distribution
–6
–8
–10
–12
–14
–16
0 5 10 15
Normalized intensity
Fig. 12.1. Random phase screen data, the logarithmic probability distribution
for the normalized forward-scattered intensity, showing comparison of the
nearest fit theoretical K-distribution (solid line) and the experimentally ob-
served histogram. (The discrepancy in the tail of the K-distribution is well
known and a more refined model would be necessary to model such extreme
intensities.)
0.95
0.9
Correlation function
0.85
0.8
0.75
0.7
0.65
0.6
0.55
0.5
50 3
45 2.5
40 2
35 1.5
30 1
Smoothing parameter/0.1 ms Volatility exponent
Fig. 12.2. Correlation of zt and |dzt |p for random phase screen data. The values
p ≈ 2, smoothing window parameter ∆ ≈ 40 give an optimal correlation
of approximately 95%. For visualization purposes, the correlation function
shown is M − (M − c)1/2 where M = sup(c). The surface c is approximately
flat near its maximum (i.e. the absolute value of its derivatives are close to
zero). This is a desirable property from the point of view of the sensitivity of
the technique to the precise values chosen for the volatility exponent p and
smoothing parameter ∆, which are inevitably subject to empirical errors.
2.5
1.5
0.5
0
5500 6000 6500 7000 7500
Sample number at 10 kHz
Fig. 12.3. Random phase screen data – comparison between raw intensity and
|dz|p (both normalized by their respective means over the time shown). The
figure shows a strong correlation for p ≈ 2.
100
200
Range cell at resolution 0.3 m
300
400
500
600
700
800
900
1000
500 1000 1500 2000 2500 3000
Sample number at PRF 1 kHz
Proposition 12.1 The squared volatilities of the amplitude process Ψt are given
in terms of the intensity and phase squared volatilities by
Bσ(z)
2
dΨ2t = exp(2iθt ) − 2
zt σ(θ) dt, (12.8)
2zt
Bσ(z)
2
|dΨt |2 = 2
+ zt σ(θ) dt. (12.9)
2zt
Observe that Proposition 12.1 does not involve drift quantities and only the
instantaneous volatilities in the intensity and phase feature explicitly in (12.8),
(12.9).
In the pure Rayleigh case, the I, Q processes can be described by the pair
of (uncoupled) Ornstein–Uhlenbeck processes (6.3), (6.4) from which we deduce
2
(e.g. via Ito’s formula applied to arctan(I/Q)) that σ(θ) = 2/z thus exhibiting
singular behaviour at zeroes of the intensity. This suggests that such singular be-
haviour should occur also in the K-distributed case which constitutes a Rayleigh
process with a modulated cross-section. The phase volatility can be calculated
from the normalized amplitude process via the identity |d exp(iθt )|2 = dθt2 , which
can be applied directly to the data and avoids difficulties arising from the discon-
tinuity in θt at 2π. A comparison of the intensity and (squared) phase volatility
is shown in Fig. 12.5, which reveals singular behaviour in σ(θ) near zeroes of
the intensity, characteristic of the Ornstein–Uhlenbeck (Rayleigh) process, and
approximate constant behaviour when the intensity lies above some threshold.
In respect of the singular behaviour at zeroes of the intensity, our results
are consistent with studies of the phase derivative, for a differentiable amplitude
process, reported in Jakeman et al. (2001). In contrast, however, in the differen-
tiable case it is the intensity, rather than amplitude-weighted phase derivative
that has minimal variance. This distinction between the differentiable and dif-
fusion models for the amplitude Ψt should enable one to select which type of
model is appropriate in different physical situations (see orig. §IIC in Field and
Tough 2003b for a detailed theoretical account of the phase behaviour in the
non-Gaussian case).
2
From a physical point of view the behaviour of σ(θ) is anticipated as follows.
The amplitude Ψ = I + iQ is generated from (some component of) the complex-
valued electric field E. One therefore expects a contribution to σ(θ) that varies
inversely with R, since |δE(θ) | ∼ Rδθ and
|δE(θ) | contains an ‘intensive’ part
that is independent of the total field intensity Etot
2
. This observation is consistent
with the Ornstein–Uhlenbeck description of the Rayleigh process given above. In
N
the random walk model (Jakeman and Tough 1988) with Etot ∼ n=1 exp(iϕn ),
there is a fluctuation term that relates only to the random phases ϕn and which
is reflected in W (2) of the intensity SDE (5.20). Simultaneously, in the case of a
time varying scattering cross-section, there exist fluctuations in θt induced by an
RADAR WAVELENGTH 131
7
phase volatility
6
Intensity/(squared) phase volatility
intensity
5
0
0 50 100 150 200 250 300 350 400
Sample number at 1 kHz
2
(12.9). The analysis of the phase shows that σ(θ) is proportional to the constant
B and varies inversely with z and linearly with x in accordance with the general
scaling arguments adduced above. Consequently for A = 0, i.e. the K-distributed
case, the Wiener increments in the I, Q components of the field are correlated.
The special relation above for independence of the quadrature components is
recovered in the Rayleigh case for which A = 0.
12.2.2 Anomaly detection
As a consequence of Proposition 12.1 and the invariance of c(X, Y ) under linear
transformations, therefore, we obtain the following result.
Corollary 12.2 For a K-distributed35 process described by (5.19), (5.20) the in-
stantaneous correlation between the observed and theoretical |dΨt | squared volatil-
ities reduces to
cΨ = c(|dΨt |2 , zt ). (12.10)
The results of applying this correlation measure at each range cell are shown
in Fig. 12.6, which reveals high correlation at all ranges except the target loca-
tion. The persistence of this local spatial de-correlation over time is shown in
Fig. 12.7. Using these correlation results, plots were generated of the time series
for the theoretical and observed |dΨ|2 squared volatilities at both the maximum
and minimum correlation ranges as discerned from Fig. 12.6. This was carried
out in both real and permuted time, as explained above. The results shown in
Figs. 12.8–12.11 illustrate the instantaneous nature of the correlation, and how
this is destroyed at the target location.
The plots of the inverse predictive functions shown in Figs. 12.9 and 12.11
in fact provide the approximate cumulative distribution functions (c.d.f.) for the
respective random variables, when the vertical axis of the inverse function is
re-scaled to the interval [0, 1]. This feature would be useful in obtaining a mea-
sure of discrimination between the corresponding distributions, since the Fisher
information geodesic distance between two distributions pi can be approximated
directly in terms of the discrete c.d.f.s ci as cos θij = α [δci (α)δcj (α)]1/2 . In
the present approach, however, we assume that, in the clutter domain, the un-
derlying distributions are identical and apply the statistical correlation measure
(12.1) to the distinct random variables. It may however be possible to exploit
the Fisher information discrimination measure through a corresponding SDE
description for the target (cf. Wootters 1981).
The approach to anomaly detection described in the present section should
be compared with Haykin (1999) in which real empirical data are used to build
the predictive sea clutter model.
It is of some (theoretical) interest at this point to consider the possibility of
an anomaly detection scheme based on the same type of principle of comparison
35 This 2
result would not hold for the pure Rayleigh process because in that case σ(z) ∝ z,
2 ∝ 1/z and the right-hand side of (12.9) is then constant.
σ(θ)
RADAR WAVELENGTH 133
1
0.9
0.8
0.7
0.6
Correlation
0.5
0.4
0.3
0.2
0.1
0
0 200 400 600 800 1000 1200
Range cell at resolution 0.3 m
between observed and theoretical squared volatilities, but that does not involve
the scattering cross-section xt explicitly. Indeed, such a scheme can be readily
constructed, as follows (albeit up to second order in the time differential). First
recall the squared differential relations for the amplitude and its modulus:
|dΨt |2 = (Bxt + Azt /2xt )dt,
(12.11)
dRt2 = 12 (Bxt + Azt /xt )dt.
Taking linear combinations of these yields
1
|dΨt |2 − dRt2 = Bxt dt
2 (12.12)
2dR2 − |dΨ |2 = Azt dt
t t
2xt
whose product gives the quadratic time differential relation
1
|dΨt |2 − dRt2 2dRt2 − |dΨt |2 = ABzt dt2 . (12.13)
4
In principle, this general relation facilitates an alternative stochastic volatility
based anomaly detection, without concern for the properties of xt which is no-
tably absent in (12.13).36
36 The author is grateful to Dr. Tao (Stephen) Feng for pointing out the existence of such a
scheme.
134 EXPERIMENTAL TESTS
1.6
1.4
1.2
1-correlation
1
0.8
0.6
25
0.4
20
0.2
15
0 10
0
200 400 5 Time window translation/0.1s
600 800 1000 0
1200
Range cell at resolution 0.3 m
Fig. 12.7. 3-dimensional plot showing range and time horizontally and
1 − c(|dΨt |2 , zt ) vertically; each point along the time axis corresponds to
a windowed sample of 1 s, thus covering a total time duration of 3 s of data.
The figure illustrates the persistence of the anomaly detection mechanism
over time – the object of interest disappears out of view for a short period
while it is submerged due to bulk wave motion, and then re-surfaces.
0
200 300 400 500 600 700
Sample number at 1 kHz
4
3.5
3
2.5
2
1.5
1
0.5
3.5
predicted (solid line)/observed
(dotted line) squared volatility
2.5
1.5
0.5
0
0 100 200 300 400 500 600 700 800 900 1000
Sample number at 1kHz
3.5
Predicted (solid line)/observed
(dotted line) squared volatility
2.5
1.5
0.5
0
0 100 200 300 400 500 600 700 800 900
Permuted sample number at 1 kHz
In furtherance of this study one hopes to exploit the freedom in the volatility
functions in the classification of K-distributed processes provided in Chapter 5,
thus exploring different calibrations for σ(z) and the underlying phase volatility
σ(θ) from a simulation or data-driven point of view. Our analysis could be ex-
tended to include non-Markovian models for the behaviour of xt , zt , Ψt via the
introduction of extra variables into a larger number of coupled SDEs (cf. Tough
1987), and in addition, to explore other models for the scattering cross-section,
which do not fall into the general category of Ito processes. The analysis of the
discrete sampled data in this chapter, which integrates the underlying Ito SDEs
to first order only, could be enhanced by higher order numerical integration
schemes (Klauder and Petersen 1985) as described in Chapter 11. The aim of
pursuing these various avenues would be to optimize the correlation between the
predicted and observed volatilities in the predominant K-distributed domain.
The techniques we have described could also likely be adapted to situations that
occur in the modelling of financial time series with a view to detecting anoma-
lous behaviour, and to other situations of propagation of acoustic waves through
random media.
13
NON-LINEAR DYNAMICS OF SEA CLUTTER
We review experimental evidence for the non-linearity of sea clutter and the role
of the z-parameter or Mann–Whitney rank-sum statistic in quantifying this non-
linear behaviour in the context of a hybrid amplitude modulation (AM)/frequency
modulation (FM) model for sea clutter, viewed as a cyclostationary process. An
independent theoretical derivation of the stochastic dynamics of radar scattering
in a sea clutter environment, in terms of a pair of coupled stochastic differential
equations (SDEs) for the received envelope and radar cross-section (RCS), en-
ables the identification of non-linearity in terms of the shape parameter for the
RCS. We are led to conclude that, from both experimental and theoretical points
of view, the dynamics of sea clutter are non-linear with a consistent degree of
non-linearity that is determined by the sea state.37
Haykin et al. (2002) advocated a state-space formalism for the processing
of radar signals in the presence of sea clutter (i.e. radar backscatter from an
ocean surface). Such a model not only accounts for the temporal dimension of
sea clutter in an explicit manner but also its statistical characterization. Basic
to this formalism is whether the underlying dynamics of sea clutter are linear,
or non-linear.
In the detailed experimental study reported in Haykin et al. (2002), it was also
demonstrated that sea clutter is a non-linear dynamic process, with the degree of
non-linearity increasing as the ‘sea state’ becomes higher. The conclusion reached
on the non-linearity of sea clutter was based on two premises, using real-life data
collected with an instrument-quality coherent radar system:
1. The characterization of sea clutter embodies two forms of continuous-wave
modulation:
(i) amplitude modulation, which is linear, and
(ii) frequency modulation, which is non-linear
The latter phenomenon is responsible for the non-linearity of sea clutter.
2. The z-parameter, denoting the Mann–Whitney rank-sum statistic, is less
than the special value −3, which is a strong indicator of non-linearity.
With regard to point 1, it is also noteworthy that in another study that
focussed on the spectral characterization of sea clutter using the Loève trans-
form (Haykin and Thomson 1998), it was discovered for the first time that sea
37 The author acknowledges the input and collaboration of Prof. Simon Haykin with regard
to the material on the z-parameter and the hybrid AM/FM model in this chapter.
138
NON-LINEAR DYNAMICS 139
–2
–4
–6
Z[–]
–8
–10
–12
–14
–16
20 40 60 80
NMAD(df/dt ) [Hz]
model for the (discrete) number of component scatterers. For the scattered ra-
diation the origin of the SDE dynamics lies in the behaviour of the component
phases which are taken to evolve in time according to a Wiener process on a
suitable (Rayleigh) timescale.
Thus we are able to represent the essential ingredients of the radar back-
scatter temporally, in the form of a continuous time stochastic process, say qt ,
which evolves in time according to
dq
= bt + σt Γt (13.2)
dt
in which Γt is the familiar white noise process and Γt has the autocorrelation
property
Γt Γt = δ(t − t ). For our purposes it will be sufficient to understand
and interpret from the dynamical equations for the RCS and the received radar
amplitude that, in a discrete-time setting,
where {ti } is a discrete set of observation times, δt = ti+1 − ti , and {nti } are a
collection of independent N (0, 1) random variables. Then the above properties of
q and its time derivative are evident (see Oksendal 1998 for a detailed rigourous
account).
The essence of the approach taken is therefore to postulate the exact dynam-
ics in continuous time, and then sample at a discrete set of times corresponding
to the physical measurements. This procedure is inevitably more precise than
an attempt at a model that is fundamentally discrete time in nature, since the
physical observables are not quantized in time.
We shall assume the (dynamical extension of the) random walk model for the
resultant back-scattered amplitude or ‘received envelope’
NON-LINEAR DYNAMICS 143
s(j)
N 7
89
& ':
(N ) (j)
Et = aj exp iϕt (13.4)
j=1
with (fluctuating) population size N , random phasor step s(j) , and ‘form factors’
aj . The key result of relevance to our discussion is obtained by taking the (Ito)
stochastic differential of (13.4). This provides the following coupled stochastic
dynamics of the RCS and scattered amplitude/received envelope Ψt .38
Proposition 13.1 The dynamics of the RCS and received envelope for radar
sea clutter, with shape parameter ν = α − 1, are given by the following set of
non-linearly coupled SDEs:
(x)
dxt = A(α − xt )dt + (2Axt )1/2 dWt (13.5)
1/2
dΨt 2(α − xt ) − 1 1 A (x) B 1/2
= A − B dt + dWt + dξt (13.6)
Ψt 4xt 2 2xt γt
in which γt is a unit power Rayleigh process, whose dynamics are obtained by
setting xt equal to a constant of unity and A = 0 in the above system.
(The original proof of this result appears as Proposition 2.1 in Field and
Tough 2003b, and we refer the reader to Chapter 8 herein for a detailed math-
ematical derivation.) Thus, the non-linear SDE for Ψt is derived theoretically
from first principles beginning with the random walk model for the scattered
electric field under the assumption of a uniform phase distribution.39 An imme-
diate consequence of this dynamical equation is the ‘noise-free skeleton’, obtained
by setting the volatility coefficients of the fluctuating Wiener terms equal to zero.
Accordingly, the randomness of the process is eliminated and the residual dy-
namics are deterministic and differentiable. Physically, this corresponds to an
evolution conditioned on the current state of the system and then averaged over
an ensemble.40 The concept of the residual noise-free part is explored further
below.
The constants A, B in (13.5), (13.6) have the physical dimension of frequency,
so that their reciprocals represent correlation timescales for the RCS modulation
and Rayleigh scattered components, respectively. The constant B is electromag-
netic in origin with a value B ∼ c|k| where k is the wave vector of the carrier. In
radar situations, the illuminating radiation is such that A B, with the value
of A being determined as an intrinsic property of the statistics of the scattering
38 We
√
can express Ψt = It +jQt (j = −1), the familiar sum of its ‘in-phase’ and ‘quadrature-
phase’ components.
39 The assumption of a uniform phase distribution can be relaxed and a corresponding de-
tailed dynamical description in terms of SDEs has been given in Field and Tough (2005).
40 In other words, for an Ito process q with SDE dq = b dt + σ dW , the ensemble average
t t t t t
evolution is determined by E[dqt ] = bt dt.
144 NON-LINEAR DYNAMICS
13.3.1 Superposition
In light of the SDE theory, we may argue that the SDE of sea clutter is inde-
pendent of the amplitude profile of a transmitted pulse, provided the transmit
energy is maintained constant. This property, which is derived explicitly in Field
(2005), is related to the fact that the form factors (i.e. the amplitude weight-
ings) in (13.4) may be taken as unity for an asymptotically large population
(cf. also Jakeman and Tough 1988 where the emergent statistical properties are
independent of the choice of form factors).
For a radar pulse of constant amplitude, suppose that the two halves of the
pulse have transmit frequencies ω1 and ω2 . Then we may consider the correlation
between the SDE of sea clutter for the two portions as frequency ω1 increases
relative to ω2 . The transmit frequencies are proportional to the Rayleigh con-
stant B appearing in (13.6) (B ∼ c|k|, c is the speed of light, k is the carrier
wave vector), and the relationship between the two SDEs, for the two different
transmit frequencies, is through (13.6): the two terms involving the constant A
are the same for both SDEs; on the other hand, the terms involving the Rayleigh
constant have different B values corresponding to the two transmit frequencies.
Nevertheless, on physical grounds, the two complex Wiener processes ξt for each
transmit frequency should be considered perfectly correlated. The reason for this
correlation is that the physical origin of the component phase fluctuations φ(j)
is (microscopic) Doppler – the Doppler frequency ratio ω1 /ω2 is a function of
the radial velocity of the j-th member of the population, so the micro-Doppler
phase shift scales with the transmit frequency; the ξt process is the same for
any transmit frequency (assuming these are transmitted simultaneously) as this
depends only on the behaviour of the component scatterer.
In a similar fashion, consider the simultaneous transmission of two pulses of
constant amplitude, with two different frequencies as above, and the resulting
SDE of sea clutter received by a common antenna. Since Maxwell’s equations
of electromagnetism are linear, the resulting Ψ is a linear superposition Ψ =
RADAR PARAMETERS 145
k1 Ψ(1) + k2 Ψ(2) where k1 , k2 are the relative intensities of the two transmit
waveforms, normalized so that k1 + k2 = 1, and Ψ(i) are the constituent complex
amplitude processes, both satisfying the SDE (13.6), with different Rayleigh
constants B corresponding to the two transmit frequencies. Since the beams are
simultaneous, the ξ processes are perfectly correlated, with the remaining parts
of (13.6) involving the constant A being the same for both transmit frequencies.
Thus, the non-linear dynamics do not infringe the principle of superposition
inherent in Maxwell’s equations.41
the population to behave this way, the phenomenological reasons for which we do not describe
here.
146 NON-LINEAR DYNAMICS
3.5
small
3
Radar cross-section
2.5
2 large
moderate
1.5
0.5
0
0 2000 4000 6000 8000 10000 12000
Time
12
2Axt
dxt xt )dt
(x)
= A(1 − + dWt . (13.8)
α
A natural question that arises for the domain of low α (high sea state) is whether,
following a burst, the RCS returns to equilibrium and undergoes a period of
relative quiescence before a successive burst, or whether the high relative variance
is achieved through perpetual bursts of varying magnitude, with the maximum
timescale for their separation (pertaining to the largest deviations) being given by
the reciprocal of the constant A. Inspection of (13.8) reveals the latter situation
to be the case, as setting x ≈ 1 therein removes the drift term but nevertheless
produces a highly volatile behaviour for small values of α. Observe also, in this
regard, the essential difference between the roles of A and α, which determine
the correlation timescale and magnitude of excursions respectively – thus, for
instance, the single point distribution of the RCS is unaffected by A, whereas it
is α that parameterizes this distribution according to the discussion of the shape
parameter above. These features are indeed evident from the simulated data
shown in Fig. 13.2 (for α = 1). Observe also, embedded within the largest peak
to peak timescale O(A−1 ), the presence of sub-scale peaks separated by times
of smaller order, including scales less than the characteristic Rayleigh timescale
O(B−1 ).
Now, in contrast, as the sea state settles down to a low value, the (normalized)
RCS has small fluctuations away from its mean (unity), so that there is no
significant modulation of the Rayleigh scattering time series – in other words,
the scattering is approximately of constant local power.
RADAR PARAMETERS 147
We close this chapter with some general discussion concerning the main as-
pects of the non-linear dynamics of sea clutter. We have described a detailed
analysis of radar sea clutter data, whose primary purpose is to address the pres-
ence of non-linearity, from real experimental data. A natural quantifier for this
non-linearity is the z-parameter or Mann–Whitney rank-sum statistic, which
has been successfully applied in the context of a hybrid AM/FM model for sea
clutter. The SDE dynamical model of radar sea clutter has also been verified pre-
viously to a remarkable degree of accuracy, in terms of real experimental data
(orig. section 4(b) in Field and Tough 2003a). Moreover, an independent theo-
retical account for such a model was provided in Field and Tough (2003b), and
has served as the basis for other significant developments (Field and Tough 2005;
Field 2005). As we have seen in Section 13.2, this stochastic dynamic behaviour
is inherently non-linear, due to the broader timescale fluctuations in the RCS.
The extent of non-linearity arises naturally in the SDE description through the
relative variance or shape-parameter, which encodes the sea state. Thus, from
an SDE dynamical perspective, the non-linear character of radar sea clutter is
firmly established, both theoretically and experimentally.
Calculation of the z-parameter is from real data containing noise, the latter
being akin to the stochastic fluctuating terms present in (13.5), (13.6). However,
z has the stochastic element removed, i.e., it is not a random variable. Accord-
ingly, some ensemble averaging takes place in the calculation of z, and for this
purpose the statistical properties of ergodicity and stationarity are assumed, le-
gitimate over realistic short timescales. In terms of the parameter A of (13.5),
such timescales are short enough that the assumption of constant A is valid. Nev-
ertheless, they should be long enough (of the order of A−1 ) for the fluctuations
in the RCS (or equivalently, as we elucidate below, the frequency modulation
effect) to be appreciable so that non-linearity can indeed be detected.
From an engineering physics perspective, the dynamics of sea clutter are
perhaps more naturally viewed in terms of amplitude (AM) and frequency mod-
ulation (FM). Studies have indicated that the degree of non-linearity is governed
by the extent of FM which, in turn, is more noticeable for higher sea states
(i.e. the shape parameter ν is large). To relate this further to the SDE descrip-
tion of Section 13.2, it is convenient to view the resultant amplitude process
Ψt in the product representation Ψ = x1/2 γ, in which x is the RCS and γ is
a unit power Rayleigh process. Then, the AM consists of the fluctuations of
γt (Rayleigh ‘speckle’) which is ‘frequency’ modulated by the RCS process xt
over a much broader timescale. The FM/AM contributions therefore have char-
148 NON-LINEAR DYNAMICS
arises from the parameters in the scattering population model as described in Chapter 7.
RADAR PARAMETERS 149
set simulated using SDEs, for which the shape parameter is known, and thereby
develop the precise relationship between the z-parameter and shape parameter
quantifiers of non-linearity. It may, indeed, also be possible to relate the two pa-
rameters in purely theoretical terms. We can also generate data with and without
noise, which forms the basis for further experiment. We suggest that these two
lines of enquiry could form the basis of future developments in the investigation
of the non-linear properties of radar scattering dynamics.
14
OBSERVABILITY OF SCATTERING CROSS-SECTION
where i is a discrete time index and {ni } are an independent collection of N (0, 1)
distributed random variables. Applying a smoothing average
·∆ to the left-hand
side (the ‘observations’) of (14.1) with window ∆ = [t0 − ∆, t0 + ∆] yields an
approximation to xt0 , with an error that tends to zero as the number of pulses
inside ∆ tends to infinity and ∆ → 0 (see discussion of χ2 statistic following
Theorem 10.6).44
44 We assume here that the sample paths of x are continuous a.s., which is a consequence of
t
(10.5).
150
EXPERIMENT 151
3
Exact (solid) and inferred (dotted) cross-section
2.5
intensity
time series; intensity as indicated
2
inferred
cross-section
0.5
0
0 0.5 1 1.5 2 2.5 3
Time x 104
Remarks. On finite pulse rate. In the case that the pulse rate is bounded, as in
a real experiment, one should consider the optimization of the smoothing window
– the window must be neither so large that the structure of the temporal variation
in the cross-section is lost, nor so small that an average over the normal random
seeds is no longer effected. Thus a compromise optimal value should exist, and
can indeed be derived (at least for K-scattering). We refer the interested reader
to Fayard and Field (2008) for a detailed account.
On theother hand, smoothing the zt time series, for any choice of param-
eter ∆, does not yield the desired close correlation with xt ; indeed any such
attempt to ‘decorrelate the speckle pattern’ merely produces an intensity profile
with the same general shape as the original zt , with oscillations on a timescale
approximately equal to ∆.
analysis of the inferred cross-section alone, which has hitherto been regarded as
the ‘hidden’ state of the system. The electromagnetic scattering process should
then appropriately be viewed as a secondary exogenous device, whose purpose
is merely to extract the real time behaviour of the underlying scattering cross-
section, where the latter is the object of primary interest.
Our results also suggest the application to the physics of magnetic resonance
(MR) imaging and spectroscopy and experimental NMR (Field and Bain 2008).
The random medium (e.g. brain tissue) lies inside a background magnetic field
B0 , with which the constituent (proton) spin vectors are aligned, in their mini-
mum energy configuration. An applied RF pulse causes resonant absorption to
occur, so that the spins re-align, typically at a ‘pulse flip’ angle of 90◦ to B0 .
Radiation of this absorbed energy gives rise to the received MR signal (the ‘free
induction decay’ or FID), which is detected through the generation of electro-
motive force in a coil apparatus, due to the time varying local magnetic field.45
The MR signal has the usual in-phase (I) and quadrature-phase (Q) compo-
nents familiar from radio theory, and thus corresponds to the amplitude process
ψ = I + iQ for each point in space. For a perfectly homogeneous (total) mag-
netic field throughout the medium, each spin vector precesses at the Larmor
frequency ω0 about the longitudinal axis, where ω0 is given by the Larmor equa-
tion ω0 = γB0 and γ is the (local) gyromagnetic ratio (e.g. Ernst et al. 1987). (In
the radar scattering situation described above, ω0 corresponds to the Doppler
frequency, arising from bulk wave motion in the scattering surface.) However,
the local inhomogeneities in the net magnetic field, due to the local magnetic
properties of the medium, give rise to a process known as ‘spin-spin’ or ‘T 2 re-
laxation’ constituting the (random) exchange of energy between neighbouring
spins. These local perturbations in the total magnetic field can reasonably be
considered as independent for each component spin, so that the dynamics of
each spin vector can be modelled as a phase diffusion process with (transverse)
resultant as in (6.16) and phase initializations {∆(j) } equal. After sufficient re-
laxation time t ≥ T has elapsed, phase decoherence occurs. In principle, this
enables the spin population to be tracked according to Theorem 10.6, and thus
real-time MR images to be generated. This application is explored in greater
depth, and from a less detailed mathematical perspective, in Field (2006). In
the context of experimental NMR these theoretical ideas have been tested and
verified, as reported in Field and Bain (2008).
45 This effect is the result of Faraday’s law, i.e. Maxwell’s curl equation for a time dependent
First, let us remark on the basic properties of sums of independent random vari-
ables. Suppose {Xi } is a collection of independent continuous
random variables
with density functions pXi . Let Z be their sum, Z = Xi . Then the distribution
of Z is given by the n-fold convolution, namely
A.0.0.2 Stability The distribution F is said to be stable if, given {Xi } inde-
pendent and {Xi , X} identically distributed according to F , then for all choices
of constants a, b there exist constants c, d such that
.
aX1 + bX2 = cX + d (A.4)
.
with equality = in distribution.
By expressing these conditions for infinite divisibility and stability separately
in terms of characteristic functions, as explained above, one can deduce that
153
154 APPENDIX A
stability implies infinite divisibility. The converse does not hold, however. Some
pertinent cases are the following:
Gaussian. The Gaussian distribution is stable. This follows from the distribu-
tion properties:
X ∼ N (µ, σ 2 ),
1
pX (x) = √ exp(−(x − µ)2 /2σ 2 ),
2πσ 2
+ ,
ΦX (ω) = exp iµω − σ 2 ω 2 /2 .
X ∼ G(α, β),
xα−1 e−x/β
pX (x) = ; x ≥ 0; α, β > 0,
Γ(α)β α
Γ(z + 1) = zΓ(z),
ΦX (ω) = (1 − iωβ)−α
is infinitely divisible. However, consistently with the above general observations,
it is not stable (as apparent from the characteristic functional form).
Cauchy
X ∼ G(α, β),
1 a
pX (x) = ,
π a2 + (x − b)2
ΦX (ω) = exp(ibω − a|ω|) .
This distribution (also referred to as the ‘Lorentzian’ distribution, of importance
in resonance theory) is stable (and therefore, like the Gaussian, also infinitely
divisible). Observe also that (for b = 0) the distribution of the sample mean µN =
N
i=1 Xi /N where Xi are drawn independently from a Cauchy distribution, has
the same distribution as Xi , so the ordinary central limit theorem drastically
fails.
On the basis of the same type of reasoning the K-distribution encountered
in the description of K-scattering is also infinitely divisible (but not stable). Of
mathematical interest, distributions for discrete multi-valued random variables
are neither infinitely divisible nor stable (as can be seen by considering the
possible sample values that occur under linear combinations as compared to
those drawn from the original distribution).
APPENDIX B
ITO VERSUS STRATONOVICH STOCHASTIC INTEGRALS
in which b(I) , b(S) denote the drifts in the Ito and Stratonovich interpretations
respectively and ‘◦’ is a shorthand that indicates the Stratonovich prescription for
taking the stochastic integral, i.e. that the volatility is evaluated at the midpoint
of each subinterval. Now we can study the Stratonovich stochastic integral term
above and translate it into the Ito interpretation:
1
σ ◦ dWt = σ + (σ∂x σ)dWt dWt + o(dt1/2 ) (B.2)
2
in which the origin of the factor 12 on the right-hand side is that the midpoint of
the subinterval is being applied, in the sense explained above.46 Thus, neglecting
terms of o(dt1/2 ) we deduce that the Stratonovich and Ito stochastic integrands
are related by
1
σ ◦ dW = σdW + σ∂x σdt (B.3)
2
which follows from the relation dWt2 = dt. Hence we find that the two notions
of drift are related by
1
b(I) = b(S) + σ∂x σ. (B.4)
2
155
156 APPENDIX B
b
E φs dWs ≡ 0. (B.7)
a
Thus, Ito stochastic integrals are martingales, unlike their Stratonovich counter-
parts.
Finally, comparing with the development of stochastic differential geometry
in Chapter 3, first recall the relation between the Ito and Kolmogorov drifts on
(M, σ ij )
1
bi = β i + Γi (B.8)
2
where Γi = σ jk Γijk , the trace of the Christoffel symbol. Then in one-dimension
Γ1 = −σ∂x σ. (B.9)
(n)
ordinary differential equation dx(n) /dt = b + σΓ(n) in which Γt is a sequence
(n) t (n)
of processes whose integral Wt = Γs ds converges to the Wiener process
as n → ∞. Then the solution x(n) converges to the Stratonovich solution of the
SDE.
Remarks. On numerical integration. Sometimes it may be appropriate from
a numerical stability point of view to use a numerical integration (discrete time
step) scheme based on using the midpoint of each subinterval, the results of which
therefore converge to the solution of the Stratonovich SDE. When confronted
with an Ito SDE (B.1) with a volatility that is state dependent (and possibly non-
linear) one can, according to the above results, proceed to an accurate numerical
solution via a midpoint method applied to the following (Stratonovich) SDE,
which contains a modified drift according to the transformation (B.4),
1 ∂σ
dXt = b(I) − dt + σ ◦ dWt . (B.11)
2 ∂x
The solution thus generated is the desired approximation to the original Ito SDE.
Es ◦ Et ≡ Es , s≤t (C.2)
47 This follows essentially from the probability rule P[A|B]P[B] ≡ P[A ∩ B].
158
APPENDIX D
GIRSANOV’S THEOREM
and thus Mt is itself a martingale with respect to the measure P (it has vanishing
drift). It follows that the stochastic differential of Yt is given, according to the
Ito product rule, by
dQ = MT dP, (D.5)
159
160 APPENDIX D
ensemble averages are applied that the drift concept enters into the description
and, in real time physics applications such as we have considered, it may be that
no such ensemble average is practicable or indeed available.
In contrast to the drift, the volatility is invariant under changes of measure
on path space. This is the mathematical basis of the stochastic volatility anal-
ysis applied to scattering, and implies that no ensemble average is required to
determine the volatility from a single realization (path) of a stochastic process.
Remarks. On coin tossing. As a useful simple analogy, we could consider toss-
ing a weighted coin – generate a (discrete-time/valued) process by starting at
the origin and moving up/down according as heads/tails is obtained. Suppose
now we observe a single path of this process, and that it tends to move upwards
much more than downwards. Then consider the question of whether, given the
observation of the single path, the coin is biased toward heads (positive drift) or
is a fair coin (no drift). The answer to this question is, of course, indeterminate
– the coin could either be fair and the observed path is an unlikely one, or it
could be a ‘typical’ path for a head-weighted coin. Indeed, one can see this intu-
ition reflected in the explicit expression for the change of measure (D.5) which
suppresses probabilities of paths with large drift.
49 These remarks may be omitted without interrupting the main flow of the development.
They are mentioned for the purpose of highlighting the intimate connections that exist be-
tween mathematical physics and financial mathematics, and hopefully therefore to stimulate
interaction between the two scientific communities.
APPENDIX E
PARTITION FUNCTION SOLUTION TO BDI MODEL
to be compared with the special case of the Poisson variate for which
so α = ν/λ.
√
50 If
we replace the variable z by e −1ω then the characteristic function Φt (ω) is immediately
obtained.
161
162 APPENDIX E
In case µ < λ then T (t) grows progressively larger with time. In all cases,
Πt (1) = 1 so that probability is conserved. The expressions for the mean and
variance of the population remain positive and increase with time, the normalized
variance tending to a limiting value, which we read off as
2
λ − µ2 n + λν
2 . (E.7)
((λ − µ) n + ν)
Observe that this depends on the initial population n. However, the mean value of
the population increases without bound; any value might reasonably be chosen as
n at an arbitrarily chosen ‘initial’ time. The master equation does not establish an
equilibrium population, and accordingly, the population is prevented from being
asymptotically stationary. Therefore, these values of the underlying population
rate parameters can be disregarded as a description of a physical system in
equilibrium, such as we have encountered in practice.
λ+µ + ,
λ−µ e
(λ−µ)t
e(λ−µ)t − 1 , if λ = µ,
var(t) = (E.9)
2λt if λ = µ.
1 − (λt − 1)(z − 1)
Πt (z) = (E.10)
1 − λt(z − 1)
Although these probabilities decay to zero with time for all N ≥ 1, the rate at
which they decay decreases with time, in such a way that the noted behaviour of
the first and second moments is admissible. To see this explicitly, we may extract
the pN (t) from the partition function by writing (E.10) in the form
−1
λt 1 − λt λt
Πt (z) = + z 1− z . (E.12)
1 + λt 1 + λt 1 + λt
The probabilities pN (t) can then be obtained as the coefficients of z N in a bino-
mial series expansion of the above expression for Πt (z), whereby we obtain
N
1 λt
pN (t) = . (E.13)
λt(1 + λt) 1 + λt
Thus,
with respect to N , the population has a negative exponential decay rate
λt
log 1+λt that tends to zero as t → ∞.51
The extinction property is also interesting in terms of a sample path of the
population process Nt . Given a death rate perfectly counterbalanced by an equal
birth rate, one might perhaps intuitively expect the population level to remain
stable, in some sense, fluctuating around some positive value. The apparent para-
dox is resolved by realizing that, at some finite time and with probability 1, the
population level will hit Nt = 0 and cannot regain any positive value subse-
quently, because of the absence of immigration; i.e. a situation of population
extinction. In other words, although the transition rates are precisely symmet-
ric with respect to increasing and decreasing Nt , the property of extinction is
asymmetric in this sense, being represented by the boundary at N = 0.
Special cases of the BDI model where a limiting asymptotic distribution exists
can also be calculated for (limiting) values of the population parameters. This
question is best addressed via analysis of the characteristic or partition function.
We explain the behaviour for various values of the parameter α = ν/λ.52
E.0.1.1 α = 1: In this case we obtain the geometric distribution, and accord-
ing to the results of Section 7.3, this yields the exponential distribution in the
continuum limit.
E.0.1.2 α → 0: (i) if ν → 0, with λ, µ fixed, then Π → 1 so that the distribu-
tion is concentrated at N = 0, i.e. extinction . (ii) if λ → ∞, with ν fixed (so
µ → ∞ to maintain
N > 0); say λ/µ → k ≤ 1, then for all values of k the
partition function tends to unity, so we have extinction as in (i). A case of inter-
est arises if µ = λ + 1 so that
N = ν; nevertheless, Π → 1,53 i.e. asymptotic
extinction.
51 property mt = 1 can be verified from (E.13) using the series identity
∞TheNconstant mean
1 Nζ ≡ ζ/(1 − ζ)2 .
52 Recall that, for a limiting population to exist, N̄ = ν/(µ − λ) for values µ > λ.
53 This can be verified using the fact lim u
u→0 u = 1.
164 APPENDIX E
E.0.1.3 α → ∞: (i) if ν → ∞ with λ fixed, the Fano factor F > 1 and the
average population size
N → ∞; then the partition function Π → 0 (for z = 1)
so that the distribution is negative binomial for all values of ν with pN → 0, ∀N .
(ii) if λ → 0 with ν fixed, then F → 1 and the limiting distribution is Poisson
with mean ν/µ.54
It is shown in the Rayleigh case of a fixed step number that the amplitude
obeys a complex Ornstein–Uhlenbeck equation, and a corresponding stochastic
differential equation (SDE) in the K-distributed case is derived.
for constant population size N . Since Maxwell’s equations for the electromag-
netic field possess U (1) gauge invariance with respect to duality rotations, i.e.
multiplication by exp(iΛ) for constant
( ) Λ (cf. Penrose and Rindler 1984), the
assumption of independence of ϕ(j) implies that these
&phases
'are uniformly
(j)
distributed. Accordingly, in (F.1) the phase factors exp iϕt are indepen-
dent and uniformly distributed
on the unit circle in C. Our (phase) diffusion
(j)
model therefore takes ϕt as a collection of (displaced) Wiener processes
(j) 1 (j)
on a suitable
( ) timescale, ϕt = ∆(j) + B 2 Wt , with the random initializations
(j)
∆ a set of independent random variables uniformly distributed on the in-
(j) 1 (j) (j)2
terval [0, 2π), and thus dϕt = B 2 dWt , dϕt = Bdt. From Ito’s formula (e.g.
Oksendal 1998; Karatzas and Shreve 1988) the Ito differential of (F.1) is
N & '
(N ) (j) 1 (j) (j)
dEt = idϕt − dϕt 2 exp iϕt . (F.2)
j=1
2
N & '
(j) (j)
The first term j=1 idϕt exp iϕt on the right-hand side of (F.2) consists of
a sum of independent randomly phased Wiener processes, with variance equal to
BN dt, while the second term is independent of the scatterer label j. Thus from
(F.2) we can write
(N ) 1 (N ) 1
dEt = − BEt dt + (BN ) 2 dξt (F.3)
2
* (j)|dξt | = dt,
2
where ξt is a complex Wiener process satisfying dξt2 = 0. The
process ξt is adapted to the filtration F (ϕ)
= j F , where F is the filtration
(j)
165
166 APPENDIX F
(j)
appropriate to the component scatterer phase
& ϕt . The' (normalized) amplitude
(N ) 1
process Et is then defined by Et = limN →∞ Et /N̄ 2 and satisfies the SDE
1 1
dEt = − BEt dt + (Bx) 2 dξt (F.4)
2
where the continuous-valued random variable x, the average scattering
+ , power,
arises from an asymptotically large population via x = limN →∞ N/N̄ .
E|x [zt ] = xt . The SDE for ψt can then be derived by applying the Ito prod-
uct formula to (F.7). This requires the SDE for the scattering cross-section
to be specified. In accordance with the birth–death–immigration (BDI) model
(Bartlett 1966), we shall take the re-scaled population variate x
→ αx to satisfy
the SDE
1 (x)
dxt = A(α − xt )dt + (2Axt ) 2 dWt (F.12)
(x)
for an independent Wiener process Wt (Field and Tough 2003a). (In terms
of the underlying population parameters of the BDI model, α = ν/λ, the ratio
of the immigration to birth rate, the birth and death rates coinciding for an
infinite-sized population.) Accordingly, xt has an asymptotic Γ-distribution
xα−1 exp(−x)
Γα (x) = (F.13)
Γ(α)
F.2.1 Amplitude
Theorem F.1 The K-amplitude is governed by the SDE
1 12
dψt 1 B2 2(α − xt ) − 1 A (x)
= Bdt + dξt + A dt + dWt . (F.14)
ψt 2 γt 4xt 2xt
F.2.2 Intensity
Proposition F.2 The K-intensity SDE is given by
1
Azt (α − xt ) 2Azt2 2 (z)
dzt = B(xt − zt ) + dt + 2Bxt zt + dWt (F.15)
xt xt
(z) (x)
in which Wt is correlated with Wt of (F.12), and satisfies
12 12
2Azt2 (z) 1 (r) 2A (x)
2Bxt zt + dWt = (2Bxt zt ) 2 dWt + zt dWt (F.16)
xt xt
(r)
and Wt is a real-valued Wiener process defined by
12
2zt
γt∗ dξt γt dξt∗
(r)
+ ≡ dWt . (F.17)
xt
168 APPENDIX F
F.2.3 Phase
Proposition F.3 The resultant phase θt of the K-amplitude process satisfies
the SDE
12
Bxt (θ)
dθt = dWt (F.18)
2zt
(θ)
where the distinct (real-valued) Wiener process Wt is defined according to
12
1 ∗ 2zt
(γ dξt − γt dξt∗ ) ≡
(θ)
dWt . (F.19)
i t xt
(z)
Σt 4Azt3
= 4zt2 + (F.26)
(θ)
Σt Bx2t
which exceeds the quotient obtained in the Rayleigh case, A = 0. These rela-
tions can be used to characterize the geometry of the K-scattering amplitude
fluctuations as follows. The real and imaginary parts of the resultant amplitude
I, Q are the usual and quadrature-phase components, respectively.
Proposition F.6 In the K-distributed case, A = 0, the amplitude diffusion
tensor is non-degenerate, and the fluctuations in the and quadrature phase com-
ponents δIt , δQt are correlated. The (comoving) error surface S of δψt is an
ellipse whose major axis lies in the instantaneous radial direction defined by ψt .
Degeneracy occurs in the Rayleigh case A = 0, for which S is a circle, i.e. the
fluctuations in ψt are isotropic.
α ∂F α α
∂σβα
F,β = ; σβ,γ = . (G.1)
∂xβ ∂xγ
We then have
t t
α α
∆x = F (x (u))du + σβα (x (u)) dWuβ . (G.3)
0 0
t
1
α
∆x = tF + α
σβα Wtβ + t2 F γ F,γ
α
+ στµ F,µ
α
Wsτ ds
2
0
t t t
1
+ F µ σβ,µ
α
sdWsβ + στµ σβ,µ
α
dWsβ Wsτ + στµ σσν F,µν
α
Wsτ Wsσ ds
2
0 0 0
t s t
1
+ στµ σσν σγ,µν
λ α
σβ,λ dWsβ dWuν Wuτ Wuσ + στµ F ν σβ,µν
α
sWsτ dWsβ
2
0 0 0
170
APPENDIX G 171
t s t s
+ στµ σσ,µ
ν
F,να ds dWuσ Wuτ + στµ σσν σγ,ν
λ α
σβ,µλ dWsβ Wsτ dWuγ Wuσ
0 0 0 0
t t s
1
+ στµ σσν σγλ σβ,µνλ
α
dWsβ Wsσ Wsτ Wsγ + στµ σσ,µ
ν α
σβ,ν dWsβ dWuσ Wuτ
6
0 0 0
t s t s
+ στµ F,µ
ν α
σβ,ν dWsβ Wuτ du + F µ στ,µ
ν α
σβ,λ dWsβ udWuτ
0 0 0 0
t
1
+ στµ σγν σβ,µν
α
dWsβ Wsτ Wsγ
2
0
t s u
+ στµ σσ,µ
ν λ
σγ,ν α
σβ,λ dWsβ dWuγ dWvσ Wvτ . (G.4)
0 0 0
The calculation of the expectation values for the increments and their products
proceeds in much the same way as the scalar case; some care is needed in the
ordering of the various indices. Thus we write the stochastic part of the vector
increment as
1
∆x̂α = ∆xα − tF α − t2 F γ F,γ
α
. (G.5)
2
In a similar way to the scalar case, we have
t2 µ ν α
E (∆x̂α ) = σ σ F (G.6)
4 τ τ ,µν
while the covariance matrix is given by
t2 λ β α t2 λ β α
E ∆x̂α ∆x̂λ = tσβα σβλ + σγ σγ F,β + σγα σγβ F,β λ
+ σγ F σγ,β + σγα F β σγ,β
λ
2 2
t2 t2
+ σγλ σβσ σβµ σγ,σµ
α
+ σγα σβσ σβµ σγ,σµ
λ
+ σγβ σγσ σν,β
α λ
σν,σ .
4 2
(G.7)
Extension of these results beyond second order in time in the general multi-
plicative noise case, although proceeding along similar lines, is rather computa-
tionally heavy and is not considered here.
APPENDIX H
OPEN PROBLEMS
172
APPENDIX H 173
174
APPENDIX I 175
The text gives an authoritative account of our current understanding of radar sea
clutter from an engineering perspective. The authors pay particular attention to
the compound K-distribution model, which they have helped develop over the
past two decades. Evidence supporting this model, including a detailed review
of the calculation of EM scattering by the sea surface, its statistical formulation,
and practical application to the specification, design and evaluation of radar sys-
tems are included. The calculation of the performance of practical radar systems
is emphasized. This book provides a less mathematically sophisticated treatment
in the radar context which will serve as an excellent complement to the current
monograph for specialist radar engineers, and also be of interest to the wider
applied physics and mathematics academic communities.
This page intentionally left blank
REFERENCES
177
178 REFERENCES
level detector area, sample time and count rate. J. Phys. A, Math. and Gen.
6, 1327.
Hughston, L. P. (1996). Geometry of stochastic state vector reduction Proc. R.
Soc. London, Ser. A 452, 953.
Jakeman, E. (1980). On the statistics of K-distributed noise. J. Phys. A 13,
31-48.
Jakeman, E., Hopcraft, K. I. and Matthews, J. O. (2003). Distinguishing popu-
lation processes by external monitoring. Proc. R. Soc. Lond. A, 459, 623–639.
Jakeman, E. and Tough, R. J. A. (1987). The generalised K-distribution: a
statistical model for weak scattering. J. Opt. Soc. Am., A4, 1764.
Jakeman, E. and Tough, R. J. A. (1988). Non-Gaussian models for the statistics
of scattered waves. Adv. Phys. 37, 471.
Jakeman, E., Watson, S. M. and Ridley, K. D. (2001). Intensity-weighted phase-
derivative statistics. J. Opt. Soc. Am. A 18 (9), 2121–2131.
Jeffreys, H. and Jeffreys, B. S. (1966). Methods of Mathematical Physics, 3rd
Edn., Cambridge University Press.
Karatzas, I. and Shreve, S. E. (1988). Brownian Motion and Stochastic Calcu-
lus. Berlin: Springer.
Klauder, J. R. and Petersen, W. P. (1985). Numerical integration of
multiplicative-noise stochastic differential equations. SIAM J. Numerical Anal-
ysis 22, 1153–1166.
Kramers, H. A. (1940). Physica 7, 284.
Luttrell, S. P. (2001). Private communication.
Moyal, J. E. (1949). J. Roy. Statist. Soc. 11, 1358.
Nelson, E. (1967a). Dynamical Theories of Brownian Motion, Princeton Uni-
versity Press.
Nelson, E. (1967b). Tensor Analysis, Princeton University Press.
Nelson, E. (1985). Quantum Fluctuations, Princeton University Press.
Oksendal, B. (1998). Stochastic Differential Equations – An Introduction with
Applications, 5th Edition, Springer.
O’Loghlen, J. W. (2001). Private communication.
Penrose, R. and Rindler, W. (1984). Spinors and Space-time, Vol. 1, Cambridge
University Press.
Pusey, P. N. and Tough, R. J. A. (1982). Langevin approach to the dynamics
of interacting Brownian particles. J. Phys. A, 15, 1291–1308.
Quiang, J. and Habib, S. (2000). A second order stochastic leapfrog algorithm
for multiplicative noise Brownian motion. arXiv:physics/9912055v2 , 4th Jan.
2000.
Rice, S. O. (1954). Mathematical analysis of random noise. Bell. Syst. Tech.
J., 23, 24, reprinted in Selected Papers on Noise and Stochastic Processes, N.
Wax, Ed., Dover, New York.
Ridley, K. D., Watson, S. M., Jakeman, E. and Harris, M. (2002). Heterodyne
measurements of laser light scattering by a turbulent phase screen. Applied
Optics 41, No. 3, 532–542.
180 REFERENCES
181
182 INDEX