Saichev 2013
Saichev 2013
Alexander I. Saichev
´
Wojbor A. Woyczynski
Distributions
in the Physical
and Engineering
Sciences, Volume 2
Linear and Nonlinear Dynamics in
Continuous Media
Applied and Numerical Harmonic Analysis
Series Editor
John J. Benedetto
University of Maryland
College Park, MD, USA
Mathematics Subject Classification (2010): 31-02, 31Axx, 31Bxx, 35-02, 35Dxx, 35Jxx, 35Kxx, 35Lxx, 35Qxx,
70-02, 76-02, 76Lxx, 76Nxx, 76Sxx
The Applied and Numerical Harmonic Analysis (ANHA) book series aims to pro-
vide the engineering, mathematical, and scientific communities with significant
developments in harmonic analysis, ranging from abstract harmonic analysis to
basic applications. The title of the series reflects the importance of applications
and numerical implementation, but richness and relevance of applications and
implementation depend fundamentally on the structure and depth of theoretical
underpinnings. Thus, from our point of view, the interleaving of theory and appli-
cations and their creative symbiotic evolution is axiomatic.
Harmonic analysis is a wellspring of ideas and applicability that has flour-
ished, developed, and deepened over time within many disciplines and by means
of creative cross-fertilization with diverse areas. The intricate and fundamental re-
lationship between harmonic analysis and fields such as signal processing, partial
differential equations (PDEs), and image processing is reflected in our state-of-
theart ANHA series.
Our vision of modern harmonic analysis includes mathematical areas such as
wavelet theory, Banach algebras, classical Fourier analysis, time–frequency anal-
ysis, and fractal geometry, as well as the diverse topics that impinge on them.
For example, wavelet theory can be considered an appropriate tool to deal
with some basic problems in digital signal processing, speech and image pro-
cessing, geophysics, pattern recognition, biomedical engineering, and turbulence.
These areas implement the latest technology from sampling methods on surfaces
to fast algorithms and computer vision methods. The underlying mathematics of
wavelet theory depends not only on classical Fourier analysis, but also on ideas
from abstract harmonic analysis, including von Neumann algebras and the affine
group. This leads to a study of the Heisenberg group and its relationship to Gabor
systems, and of the metaplectic group for a meaningful interaction of signal de-
composition methods. The unifying influence of wavelet theory in the aforemen-
tioned topics illustrates the justification for providing a means for centralizing and
disseminating information from the broader, but still focused, area of harmonic
analysis. This will be a key role of ANHA. We intend to publish the scope and
interaction that such a host of issues demands.
v
vi ANHA Series Preface
The above point of view for the ANHA book series is inspired by the history
of Fourier analysis itself, whose tentacles reach into so many fields.
In the last two centuries, Fourier analysis has had a major impact on the de-
velopment of mathematics, on the understanding of many engineering and sci-
entific phenomena, and on the solution of some of the most important problems
in mathematics and the sciences. Historically, Fourier series were developed in
the analysis of some of the classical PDEs of mathematical physics; these series
were used to solve such equations. In order to understand Fourier series and the
kinds of solutions they could represent, some of the most basic notions of analy-
sis were defined, e.g., the concept of “function”. Since the coefficients of Fourier
series are integrals, it is no surprise that Riemann integrals were conceived to deal
with uniqueness properties of trigonometric series. Cantor’s set theory was also
developed because of such uniqueness questions.
A basic problem in Fourier analysis is to show how complicated phenomena,
such as sound waves, can be described in terms of elementary harmonics. There
are two aspects of this problem: first, to find, or even define properly, the harmon-
ics or spectrum of a given phenomenon, e.g., the spectroscopy problem in optics;
second, to determine which phenomena can be constructed from given classes of
harmonics, as done, e.g., by the mechanical synthesizers in tidal analysis.
Fourier analysis is also the natural setting for many other problems in engi-
neering, mathematics, and the sciences. For example, Wiener’s Tauberian theo-
rem in Fourier analysis not only characterizes the behavior of the prime numbers,
but also provides the proper notion of spectrum for phenomena such as white
light; this latter process leads to the Fourier analysis associated with correlation
functions in filtering and prediction problems, and these problems, in turn, deal
naturally with Hardy spaces in the theory of complex variables.
ANHA Series Preface vii
Nowadays, some of the theory of PDEs has given way to the study of Fourier
integral operators. Problems in antenna theory are studied in terms of unimodu-
lar trigonometric polynomials. Applications of Fourier analysis abound in signal
processing, whether with the fast Fourier transform (FFT), or filter design, or the
adaptive modeling inherent in time–frequency-scale methods such as wavelet the-
ory. The coherent states of mathematical physics are translated and modulated
Fourier transforms, and these are used, in conjunction with the uncertainty princi-
ple, for dealing with signal reconstruction in communications theory. We are back
to the raison d’être of the ANHA series!
This book continues our multivolume project that endeavors to show how
the theory of distributions, also often called the theory of generalized func-
tions, can be used by a theoretical researcher or graduate student working in
the physical and engineering sciences or applied mathematics as well as by
advanced undergraduate students. Our general goals, the intended audience,
and the philosophy we are pursuing here are already described in detail in
the introduction to Volume 1, which covers the distributional and fractal
(fractional) calculus, the integral transform, and wavelets. However, given
the long time that has elapsed since publication of the first volume, for the
benefit of the reader of the present volume, we are repeating the main points
below.
xi
xii Introduction to Volume 2
while a physicist, not generally disposed to follow the same logic, might say:
Wait a second, let’s check the number 10100 , which is bigger than
most physical quantities—I know that the number of atoms in
our galaxy is less than 1070 . The iterated logarithm (in base 10)
of 10100 is only 2, and this seems to be pretty far from infinity.
This little story illustrates the sort of psychological difficulties that one en-
counters in writing a book such as this one.
Finally, it is worth mentioning that some portions of the material, es-
pecially the parts dealing with the basic distributional formalism, can be
treated within the context of symbolic manipulation languages such as Maple
or Mathematica, where the package DiracDelta.m is available. Their use in
student projects can enhance the exposition of the material contained in this
book, both in terms of symbolic computation and visualization. We have
used them successfully with our students.
Part III. Potentials, Diffusions and Waves contains an analysis of the three
basic types of linear partial differential equations: elliptic, parabolic,
and hyperbolic;
Part IV. Nonlinear Partial Differential Equations contains chapters on one-
and multidimensional first-order nonlinear partial differential equations
and conservation laws, generalized solutions of first-order nonlinear par-
tial differential equations, Kardar–Parisi–Zhang (KPZ) and Burgers’
equations, Korteweg-de Vries (KdV) equations, the equations of gas
dynamics, and flows in porous media.
Introduction to Volume 2 xv
Finally, the third and last volume of this series, Distributions in the Phys-
ical and Engineering Sciences: Random and Fractal Signals and Fields (in
preparation), will contain the following:
each section to reduce clutter, but outside the section in which they appear,
referred to by three numbers. For example, formula (4) in Sect. 3 of Chap. 1
will be referred to as formula (1.3.4) outside Sect. 1.3, but only as formula (4)
within Sect. 1.3. Sections and chapters can be easily located via the running
heads.
Finally, a Springer Extras Appendix C: Distributions, Fourier Transform,
and Divergent Series, freely available at http://extras.springer.com, pro-
vides a compact version of the foundational material on distributions (gen-
eralized functions) contained in Volume 1 of this book series. It explains the
basic concepts and applications needed for the development of the theory of
linear and nonlinear dynamics in continuous media and its diverse applica-
tions in physics, engineering, biology, and economics as presented in Volume
2. The goal is to make the present book more self-contained if the reader
does not have easy access to the first volume. We aimed at a compression
level of about 30 %, a compromise that permits the reader to obtain suf-
ficient (for our purposes) operational acquaintance with the distributional
techniques while skipping more involved theoretical arguments. Obviously,
browsing through it is no replacement for a thorough study of Volume 1.
The structure of Appendix C roughly mimics that of Volume 1, with chap-
ters replaced by sections, sections by subsections, etc., while in some cases
several units were merged into one.
About the Authors: Alexander I. Saichev received his B.S. in the Ra-
dio Physics Faculty at Gorky State University, Gorky, Russia, in 1969, a
Ph.D. from the same faculty in 1975 for a thesis on kinetic equations of non-
linear random waves, and his D.Sc. from the Gorky Radiophysical Research
Institute in 1983 for a thesis on propagation and backscattering of waves
in nonlinear and random media. Since 1980 he has held a number of fac-
ulty positions at Gorky State University (now Nizhny Novgorod University)
including senior lecturer in statistical radio physics and professor of math-
ematics and chairman of the mathematics department. Since 1990 he has
visited a number of universities in the West, including Case Western Reserve
University, the University of Minnesota, New York University, and the Uni-
versity of California, Los Angeles. He is coauthor of the monograph Nonlinear
Random Waves and Turbulence in Nondispersive Media: Waves, Rays and
Particles and has served on the editorial boards of Waves in Random Media
and Radiophysics and Quantum Electronics. His research interests include
mathematical physics, applied mathematics, waves in random media, nonlin-
ear random waves, and the theory of turbulence. In 1997 he was awarded the
Russian Federation’s State Prize and Gold Medal for research in the area of
nonlinear and random fields. He is currently professor of mathematics at the
Radio Physics Faculty of the Nizhny Novgorod University and a Professor
in the Department of Management, Technology, and Economics at the Swiss
Federal Institute of Technology (ETH) in Zurich, Switzerland.
W.A.W.
Contents
Introduction to Volume 2 xi
Notation xxiii
xix
xx Contents
Index 405
Notation
xxiii
xxiv Notation
Lp (A) Lebesgue space of functions f with A |f (x)|p dx < ∞
N Nonnegative integers
φ = O(ψ) φ is of order not greater than ψ
φ = o(ψ) φ is of order smaller than ψ
PV Principal value of the integral
R Real numbers
Rd d-dimensional Euclidean space
Re z The real part of z
ρ Density
sign (x) = 1 if x > 0, −1 if x < 0, and 0 if x = 0
sinc ω = sin πω/πω
S Space of rapidly decreasing smooth functions
S x to S, space of tempered distributions
Dual
S(x) = 0 sin(πt2 /2) dt, Fresnel sine integral
T, S Distributions
T [φ] Action of T on test function φ
Tf Distribution generated by function f
T̃ Generalized Fourier transform of T
z∗ Complex conjugate of z
Z Integers
∇ Gradient operator
→ Fourier map
→ Converges to
⇒ Uniformly converges to
∗ Convolution
[[ . ]] Physical dimensionality of a quantity
∅ Empty set
End of proof, example
x·y Dot (also called inner) product of x and y
√
|x|
= x · x, the norm of vector x
The integral over the whole space
Part III
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 3
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 1,
© Springer Science+Business Media New York 2013
4 Chapter 9. Potential Theory and Elliptic Equations
in spaces of dimension n = 1, 2, 3.
1-D case. Consider the 1-D case (n = 1), that is, the equation
d2
G0 (x) = δ(x). (5)
dx2
We shall assume symmetry of the solution in both directions on the x-axis
and look for an even Green’s function. A solution
d2
G0 (x) = 0
dx2
is also a fundamental solution, and that a difference of two fundamental solu-
tions is a solution of the homogeneous problem. Hence, the general solution
of (5) is of the form
G0 (x) = |x|/2 + Ax + B,
where A and B are arbitrary constants. A similar comment is applicable to
the 2-D and 3-D cases considered below.
2-D case. We shall show that the fundamental solution of a two-
dimensional Poisson equation
2
∂ ∂2
ΔG0 = + G0 (x) = δ(x) (7)
∂x21 ∂x22
ΔG0 ≡ 0 (10)
1
As in Volume 1, φ stands, usually, for an infinitely differentiable test function with
compact support contained in V (in this case).
6 Chapter 9. Potential Theory and Elliptic Equations
x1 = ρ cos ϕ, x2 = ρ sin ϕ.
The Laplacian in the polar coordinate system is expressed by the formula
1 ∂ ∂ 1 ∂2
Δ= ρ + 2 2.
ρ ∂ρ ∂ρ ρ ∂ϕ
the last equality being justified by the definition of the distributional deriva-
tive. Indeed, let us cut out from the whole plane of integration R2 a disk of
radius ε, and write
2 2
G0 Δφ d x = G0 Δφ d x + G0 Δφ d2 x. (12)
ρ<ε ρ≥ε
Furthermore, note that the contribution from the first integral on the right-
hand side tends to zero as ε → 0. This follows from the boundedness of all
the derivatives of any test function (implied by each derivative’s continuity
and compact support via the Weierstrass theorem), so that |Δφ| < K < ∞,
and as a result,
ε
1
2
G0 Δφ d x < K ρ ln(ρ)dρ ∼ −K ε2 ln ε → 0.
ρ<ε 0 2
Therefore, asymptotically,
G0 Δφ d x ∼
2
G0 Δφ d2 x, ε → 0. (13)
ρ≥ε
9.1. Poisson Equation 7
where, in the 2-D case being considered, the integral on the right-hand side
is a line integral along the circle ρ = ε. The integral of the first term, like the
first integral on the right-hand side of (12), vanishes as ε → 0. Since
d 1
n · ∇G0 = − G0 = − , (14)
dρ 2πρ
the integral of the second term,
1
− φ(n · ∇G0 )dl = φ dl,
ρ=ε 2πε ρ=ε
equals the mean value of function φ(x) on the circle ρ = ε. In view of the
continuity of φ, this mean value tends to φ(0) as ε → 0. This completes the
proof of formula (11).
3-D case. In a similar fashion, one can prove that the fundamental so-
lution of the Poisson equation (4) in the 3-D case (n = 3) has the form
1
G0 (x) = − . (15)
4π|x|
We would now like to make the reader aware that compared to the 1-D
and 2-D problems, the above 3-D problem has additional interesting aspects.
Observe that the Laplace operator, as well as other operators expressing fun-
damental laws of physics, enjoys various symmetry properties. The Laplace
operator is invariant under translations, reflections, and rotations of space.
This means, in particular, that if the origin of the Cartesian coordinate sys-
tem is shifted and if that operation is followed by a rotation of the whole
space, then in the new coordinate system, the expression for the Laplace
operator will not change. The fundamental solution of the Poisson equation
ΔG(x, y) = δ(x − y)
8 Chapter 9. Potential Theory and Elliptic Equations
thus depending only on the distance between the “observation” point x and
the “source” point y. This gives rise to the representation
1 1
δ(x − y) = − Δ (16)
4π |x − y|
of the shifted Dirac delta on 3-D space, which is often encountered in the
physics literature.
The presence of boundary surfaces, or as mathematicians say, boundary
conditions, can destroy—completely or partially—the above symmetry of
fundamental solutions, and therefore make the search for them much more
difficult. However, sometimes the boundaries themselves have symmetries
that makes it possible to use the knowledge of fundamental solutions in the
whole space.
1 1
G0 (x, y) =
4π (x1 − y1 )2 + (x2 − y2 )2 + (x3 + y3 )2
1
− ,
(x1 − y1 )2 + (x2 − y2 )2 + (x3 − y3 )2
(x3 , y3 > 0).
The above reflection method can also be used in the case of regions more
complex than the half-space, such as a wedge-shaped region with reflecting
boundaries at the angle α = π/n. The corresponding Green’s function can
be easily written analytically with the help of a schematic diagram, which,
in the case of n = 3, is shown in Fig. 9.1.1.
9.2. 1-D Helmholtz Equation 9
FIGURE 9.1.1
Illustration of the reflection method in the case of a wedge region.
A solid dot marks the position of the real source, while hollow
dots indicate positions of imaginary sources. Signs + and − signify
whether the sign of the imaginary source is the same as that of the
real source or the opposite.
G ∼ G0 , (x → 0). (2)
2
Recall that G0 appears here as a function of the Euclidean distance, so it is also a
function of two variables.
10 Chapter 9. Potential Theory and Elliptic Equations
Now, let us return to the 1-D Helmholtz equation (1). Initially, we shall
solve it in the region x > 0, where δ(x) ≡ 0, and the equation becomes the
homogeneous equation
d2
G(x) + k 2 G(x) = 0.
dx2
Its general solution has the form
has a form identical to (6). Substituting U (x, t) (7) for E(x, t), and W (x, t)
(6) for D(x, t) in (5), and then canceling the common factor exp(iωt), we
arrive at the Helmholtz equation
d2
u(x) + k 2 u(x) = w(x), (8)
dx2
which describes the complex amplitude u(x) of a monochromatic wave. The
constant
ω
k= ,
c
where ω is the wave frequency and c its speed, has a clearcut physical mean-
ing and is called the wavenumber. Its dimension is reciprocal length. Alter-
natively,
k = 2π/λ,
where λ is the wavelength.
We are now ready to formulate the radiation condition. It postulates that
the only wave present in space is the wave radiated by the source. Physically,
it is obvious that any source is of finite spatial extent. As a mathematical
condition, this translates to compactness of the support of the function D(x).
Without loss of generality we can assume that the support is contained in the
negative half-line. This implies, in particular, that to the right of the source,
where D(x) ≡ 0, the solution has a traveling waveform E(x − ct), and the
corresponding complex monochromatic wave is proportional to
exp(iωt − ikx).
Comparing this radiating monochromatic wave with the general solution (3),
we see that the radiation condition implies A = 0.
The assumed evenness of Green’s function (or equivalently, the radiation
condition on both sides of the source) gives
G(x) = B exp(−ik|x|).
The first term on the right-hand side is infinitely differentiable and does not
contribute to the singularity. However, as x → 0,
iB sin(k|x|) ∼ iBk|x|,
12 Chapter 9. Potential Theory and Elliptic Equations
so that the second term behaves like the Green’s function of the 1-D Poisson
equation (9.1.6). Taking into account the asymptotic relation (4), we obtain
B = i/2k. As a result, the Green’s function of the 1-D Helmholtz equation
has the form
1
G(x) = − exp(−ik|x|). (9)
2ik
ΔG + k 2 G = δ(x), (1)
FIGURE 9.3.1
The spherical coordinate system.
and the Helmholtz equation (1) for a spherically symmetric (i.e., independent
of angles θ and ϕ) function G becomes the ordinary differential equation (2).
A substitution v(r) = G(r)r reduces (2) to the equation
v + k 2 v = 0,
1 −1 −ikr
G(r) = − r e . (3)
4π
FIGURE 9.4.1
Cylindrical coordinate system.
is the zero-order Hankel function of the second kind. Thus, the Green’s func-
tion of the 2-D Helmholtz equation can be written in the form
i (2)
G(ρ) = H0 (kρ). (2)
4
9.4. 2-D Helmholtz Equation 15
The minus sign in front of ikρ reflects the fact that the above fundamental
solution of the Helmholtz equation in the plane satisfies the radiation condi-
tion that postulates that the only wave present in space is the wave radiated
by the source.
ρ2 G + ρG + k 2 ρ2 G = 0.
z 2 u + zu + (z 2 − ν 2 )u = 0,
16 Chapter 9. Potential Theory and Elliptic Equations
which has numerous physical applications. The two linearly independent real
solutions of the Bessel equation, which can be found in mathematical tables
or generated by computer packages, are traditionally denoted by Jν (z) and
Nν (z), and are called the Bessel functions and Neumann functions of order ν,
respectively.
In our problem, they arise from the effort to find solutions of the
Helmholtz equation that are not radially symmetric and that are of the
form
Gn = gn (ρ)einϕ .
Substituting this expression into the Helmholtz equation, we note that it
becomes an identity (for every ϕ and every ρ > 0) if the function un (z) =
gn (z/k) satisfies the Bessel equation of order ν = n. Since the Bessel equation
is linear, its general solution U (ρ, ϕ) is, for ρ > 0, a superposition of the above
particular solutions, i.e.,
∞
U (ρ, ϕ) = An Jn (kρ) + Bn Nn (kρ) einϕ .
n=−∞
u + u = 0.
The complex Hankel functions that satisfy the radiation condition are related
to the Bessel and Neumann functions through the equality
FIGURE 9.4.2
Graph of the Bessel function J0 (z).
0.02
Imaginary I Part
0
–0.02
G(ρ,z)
Real Part
–0.04
–0.06
–0.08
–10 –8 –6 –4 –2 0 2 4 6 8 10
z
FIGURE 9.5.1
Plots of the real and imaginary parts of the Green’s function G(y, z)
illustrating their evenness with respect to the variable z. Here,
ρ = 1, and κ = 1/4.
Δu + k 2 u = 0. (3)
u(y, z) z→0+
= u0 (y), (4)
lim Gz (y) = 0,
z→0
and the integral Gz (y)d2 y is evaluated thus:
∞
2 ρ dρ
Gz (y)d y = z = 1, z > 0.
0 (ρ + z 2 )3/2
2
Here, we used the fact that the function Gz (y) is radially symmetric. Hence,
the scalar field (2), which satisfies the radiation condition in the upper half-
plane z > 0, gives a rigorous solution to the diffraction problem.
where
∂ z
g(y − p, z) = 2 G(y − p, z) = (1 + ikR)e−ikR (9)
∂z 2πR3
and
R= z 2 + (y − p)2 = r2 + p2 − 2(y · p), r= z2 + y2. (10)
separating the real part, we will find the scalar field radiated by the antenna
which in the acoustic case, represents pressure.
To analyze this field, or the complex amplitude (8), physicists distinguish
a number of zones wherein the complex amplitude is described by qualita-
tively different asymptotics. In the first zone, kz < 2π, called the near-field
zone, g(y, z), see (9), practically coincides with the expression (6), and weakly
converges, for z → 0, to the Dirac delta δ(y). In physical terms, the first zone
is determined by the inequality z < λ, which explicitly includes the length
λ = 2π/k of the radiated wave.
The domain in which the opposite inequality z λ is satisfied is called
the wave zone. Here we can neglect the first term on the right-hand side of
(9), and write
ikz −ikR
g(y − p, z) ≈ e . (11)
2πR2
In turn, the wave zone can be decomposed into several further zones. We
shall just mention the one that turns out to be important for the purpose
of understanding the antenna’s functioning. Assume that the antenna has
a finite size, say 2a. In other words, u0 (y) ≡ 0, for |y| > a. Moreover, the
characteristic scale of the function u0 (y) is usually equal to a and much
larger than the wavelength: a λ. In addition, as will become clear from
the arguments provided below, the field radiated by the antenna is negligibly
small far from the z-axis, and in the analysis of the radiated wave one can
assume that the condition
is satisfied. Inside this domain one can, with high accuracy, replace R in the
denominator of (11) by r, and rewrite it in the form
ikz p2 2(y · p)
g(y − p, z) ≈ exp −ikr 1 + 2 − . (13)
2πr2 r r2
Observe that the right-hand side of the formula is now split into two fac-
tors. The first depends only on the radial coordinate r, while the second
depends only on the angular coordinates. The radial part is proportional to
the Green’s function (9.3.3) of the point source located at the origin. This
means that the expression (21) satisfies the radiation condition and decays
at the rate 1/r, as r → ∞. Sometimes, the above-mentioned conditions are
written in the form of the asymptotic relationships
∂u/∂r + iku = o (1/r) ,
(r → ∞). (22)
u = O (1/r)
s1 = ρ cos φ, s2 = ρ sin φ,
Then the inner product appearing in the integral (19) can be written as
(p · s) = ργ cos(φ − ψ), and the 2-dimensional Fourier transform of (19) takes
the form
π ∞
˜ 1
f (ρ, φ) = 2 dψ f (p) exp(−iργ cos(φ − ψ))γ dγ. (25)
4π −π 0
In what follows we shall also assume that the integrand f (p) is radially
symmetric, i.e. f (p) = f (γ). In this case the right-hand side of (25) splits
into the product of two integrals, and the integral with respect to ψ can be
calculated via the formula (9.4.4), which yields
π
exp −iργ cos(φ − ψ) dψ = 2πJ0 (ργ). (26)
−π
In the radially symmetric case considered here, the result of the integration
in (25) is independent of the angle φ, and (25) takes the form
∞
˜ 1
f (ρ) = f (γ)J0 (ργ)γ dγ. (27)
2π 0
For convenience, let us introduce a new function
This is the integral Bessel transform, mapping an original function f (γ) to its
Bessel image F (ρ). Similarly, calculating the inverse 2-dimensional Fourier
transform of the radially symmetric function f˜(ρ), one arrives at the formula
for the inverse Bessel transform
∞
f (γ) = F (ρ)J0 (ργ)ρ dρ. (30)
0
On the other hand, Bessel functions of different orders are connected by the
following recurrence relations:
d n+1 Jn+1 (z) d Jn (z)
n+1
z Jn (z) = z Jn+1 (z), =− . (32)
dz zn dz zn
To conclude this subsection, we provide three examples of calculations of
directional diagrams for radially symmetric antennas. For this purpose, let
us rewrite the expression (23) in the form
1
D(θ) = 2
cos2 θ|U0 (k sin θ)|2 , (33)
4π
where ∞
U0 (γ) = u0 (ρ)J0 (γρ)ρ dρ (34)
0
is the Bessel image of the complex amplitude of the wave in the antenna
plane z = 0.
The above integral can be evaluated using the first recurrence relation in
(32). Indeed, for n = 0, this relation is of the form
d
zJ0 (z) = zJ1 (z). (39)
dz
Substituting the right-hand side into the integral (38), one obtains
a
U0 (γ) = J1 (γa),
γ
and the corresponding directional diagram (33) is described by the function
4
D(θ) = D cot2 θJ12 (ka sin θ). (40)
k 2 a2
Here, as above, the constant D denotes the value of the directional diagram
at θ = 0, and we have taken into account the fact that
1 z n
Jn (z) ∼ , (z → 0). (41)
n! 2
100
10–1
10–2
D(θ)
10–3
10–4
10–5
10–6
0 0.5 1 1.5
θ
FIGURE 9.5.2
Directional diagram D(θ) of a disk antenna for which the field u0 (y)
is constant at all points. The value D = D(θ = 0) is assumed to be
equal to 1.
is satisfied. In this case, even for a small angle θ, the argument ka sin θ can
assume large values for which the factor U0 (ka sin θ) is negligibly small. This
is the essence of antenna directionality: the wave radiated by the antenna
propagates primarily along the z-axis. The shape of the directional diagram
is illustrated in Fig. 9.5.2. The graph shows the function (40), for ka = 8π,
in which case the antenna’s radius is equal to four wavelengths.
Observe that the directional diagram (40) decays very slowly as the an-
gle θ increases. This phenomenon is related to the slow decay of the Bessel
function as its argument increases, and can be demonstrated via the method
of stationary phase discussed in Chap. 5 of Volume 1. Indeed, this method
permits us to pass from the integral (9.4.4) to the asymptotic formula
2 π π 1
Jn (z) = cos z − n − +O √ (z → ∞).
πz 2 4 z z
In practical terms, the slow decay observed above signifies that the an-
tenna’s directionality is not very sharp. An analysis of the asymptotics of the
Fourier images provided in Chap. 4 of Volume 1 gives a deeper understanding
of the reason for this poor directionality: it is due to the jump of the field
u0 (y) at the antenna’s edge. The next example shows that the directionality
would improve if such jumps were absent.
Example 3. Disk antenna, field vanishing on the boundary. This time,
consider the field of a radiated wave of the form
2
a − |y|2 , f or |y| < a,
u0 (y) = (44)
0, f or |y| > a.
100
10–1
10–2
D(θ)
10–3
10–4
10–5
10–6
0 0.5 1 1.5
θ
FIGURE 9.5.3
Directional diagram D(θ) of a disk antenna with surface field u0 (y)
vanishing toward the antenna’s boundary according to (44).
a2
U0 (γ) = 2 J2 (γa). (46)
γ2
64 cos2 θ 2
D(θ) = D J (ka sin θ). (47)
k 4 a4 sin4 θ 2
which differs from (1) only by the location of the source, x2 = x1 . Let us
multiply (1) by G(x, x2 ), and (3) by G(x, x1 ), and then subtract the second
resulting equation from the first to obtain
G(x, x2 )ΔG(x, x1 ) − G(x, x1 )ΔG(x, x2 )
(4)
= G(x, x2 )δ(x − x1 ) − G(x, x1 )δ(x − x2 ).
Integrating both sides of the above equality over the domain V , applying the
second Green’s formula (9.1.9), and taking into account the probing property
of the Dirac delta, we get
G(x, x2 )(n · ∇G(x, x1 )) − G(x, x1 )(n · ∇G(x, x2 )) dS
S
where n denotes the external normal to the surface S. Also note that in view
of the boundary condition (2), the surface integral is equal to zero, which
gives us the reciprocity theorem in the following form:
Remark 1. The reciprocity property means that the field will not change
if the source point x1 is interchanged with the observation point x. If two
observers look at each other through a complex system of mirrors and optical
guides, then if the first can see the eye of the second, the second can see the
eye of the first.
Remark 3. The Green’s function inside the closed surface S provides the
solution of the interior boundary problem. In a physical context it describes,
for example, the acoustic or electromagnetic fields in resonators. However, the
reciprocity theorem is also valid for the analogous exterior boundary problem,
in which one has to find the Green’s function outside the surface S.
9.6. Helmholtz Equation in Inhomogeneous Media 31
in the form of the integral (9.1.3) involving a given f (x) and G(x, x ). Such a
solution satisfies the homogeneous boundary conditions such as the Dirichlet
condition
u(x)|S = 0. (10)
In a situation such that (9) is augmented by an inhomogeneous boundary
condition
u(x)|S = g(x), (11)
where g(x) is a given function on a surface S, the sought solution can be
found by splitting u into a sum of two functions
The solution of the auxiliary boundary value problem (14) is possible via the
previously used Green’s formula (9.1.9), which we will rewrite in a form more
suitable for our purposes:
(ψLφ − φLψ) d3 x = ψ(n · ∇φ) − φ(n · ∇ψ) dS, (16)
V S
where
L = Δ + k2 (17)
32 Chapter 9. Potential Theory and Elliptic Equations
In this case, in view of (14), the integral of the second term on the left-hand
side of (16) vanishes,
3
φLψ d x = G(x, x )(Δ + k 2 )w(x) d3 x = 0,
V V
Taking into account the boundary condition (14) and the corresponding ho-
mogeneous boundary condition (2) for the Green’s function, we arrive at the
sought solution of the boundary value problem (14):
We shall solve the boundary value problem (1)–(2) via the method of
separation of variables. Its essence will be better understood if we begin by
explaining the intuitions underlying it; (1) is too complex to be solved in
one big swoop. So, as a first step, we shall try to find a solution u(y, z) of a
simpler homogeneous equation,
∂ 2u
Δ⊥ u + + k 2 (z)u = 0. (3)
∂z 2
Although this equation is still quite complex, we can find its particular so-
lutions by separating the variables, yielding an even simpler equation. This
is accomplished by assuming that the solution of (3) can be written in the
product form:
y(y, z) = Y (y)Z(z), (4)
with the first function, Y (y), depending only on the variable y, and the
second function, Z(z), depending only on the variable z. Substituting (4)
into (3) and dividing the result by Y Z gives
Δ⊥ Y Z + k 2 (z)Z
=− , (5)
Y Z
where the primes denote ordinary differentiation with respect to z.
Observe that the two sides of the above equation depend on different
variables: the left-hand side depends only on y, and the right-hand side on
z. The conclusion is that the two sides of (5) must be equal to an identical
constant, independent of the variables y and z. Denote this constant by −μ2 .
Then (5) splits into two equations:
Δ ⊥ Y + μ2 Y = 0 (6)
and
Z + [k 2 (z) − μ2 ]Z = 0. (7)
Let us denote their solutions by Yμ (y) and Zμ (z), respectively. By finding
these, we will find a particular solution of (3). We emphasize once more that
instead of a complex partial differential equation with variable coefficients,
we are now working with two simpler equations: (6) is a partial differential
equation but with constant coefficients, with a solution in the whole un-
bounded 2-D plane, while (7) is an ordinary differential equation, the theory
of which is better understood than that of partial differential equations.
Starting with (6) we recall that in Sect. 9.4, we found a solution of the
equation
Δ⊥ G + μ2 G = δ(y), (8)
9.7. Waves in Waveguides 35
The asymptotic behavior of this integral, for large ρ, can be found by the
method of steepest descent described in Sect. 5.5 of Volume 1. It calls for the
replacement of cosh μ in (11) by the first two terms of its Taylor expansion
in μ, and yields
1
Yμ (y) ∼ exp −|μ|ρ , (ρ → ∞). (12)
8π|μ|ρ
Having found an arbitrary partial solution of (3) via the method of sep-
aration of variables (4), we find ourselves in possession of infinitely many
solutions corresponding to different values of the separation constant μ:
is also
a solution of (3), representing, we hope, a general solution. The no-
tation μ indicates summation (or integration) of particular solutions over
the index μ, and A(μ) is an arbitrary function that is an analogue of arbi-
trary constants entering in the general solutions of the ordinary differential
equations.
The solution (15) is sufficiently rich for the purpose of constructing the
Green’s function for the boundary value problem (1)–(2). In the next step, we
will require not only that the solution (15) satisfy (3) but also that it satisfy
the boundary conditions (2). It is clear that a sufficient condition here is that
similar conditions be satisfied by all functions Zμ (z) appearing in (15). Thus
all Zμ (z) need to satisfy the following boundary value problem:
Now suppose that we have found all solutions of this boundary value
problem and substituted them into (15). The functions Yμ (y) entering into
(15) now satisfy not (6) but an inhomogeneous equation (8). The function
u(y, z) thus produced will not yield zero on the right-hand side after the
substitution in the left-hand side of (3); its value at the point y = 0 is now
significant. Indeed,
∂ 2u
Δ⊥ u + 2
+ k 2 (z)u = A(μ) Zμ Δ⊥ Yμ + Yμ {Zμ + k 2 (z)Zμ } .
∂z μ
Substituting
Δ⊥ Yμ = −μ2 Yμ + δ(y)
gives the equation
∂ 2u
Δ⊥ u + + k 2 (z)u =
∂z 2
A(μ)Yμ Zμ + [k 2 (z) − μ2 ]Zμ + δ(y) A(μ)Zμ (z).
μ μ
9.8. Sturm–Liouville Problem 37
The first sum on the right-hand side vanishes, in view of (16), and the second
sum becomes the right-hand side of the original equation (1), provided that
A(μ) can be selected in such a way that
A(μ)Zμ (z) = δ(z − z0 ). (17)
μ
This issue can be resolved via a detailed analysis of the so-called Sturm–
Liouville problem, which will be the main topic in the next section. We shall
return to the solution of the boundary value problem (1)–(2) in Sect. 9.9.
LZ = λρ(z)Z, (1)
where ρ(z) is a positive continuous function on the interval [0, h], augmented
by the boundary conditions of the third kind
p(z) > 0, q(z) ≥ 0, p(z) ∈ C 1 [0, h], q(z) ∈ C[0, h]. (5)
In plain language, the last two conditions mean that for z contained in the
interval [0, h], the function q(z) is continuous and that p(z) and its first
derivative are continuous as well.
38 Chapter 9. Potential Theory and Elliptic Equations
assume that k 2 (z) is continuous on the interval [0, h], so that in particular,
k 2 < ∞, and define
then (9.7.16) becomes (1), and all the conditions in (5) are satisfied. The
boundary conditions (9.7.16) are a particular case of the conditions (2) with
a1 = 0, a2 = 1, b1 = 1, b2 = 0.
Excluding the trivial solution Zm (z) ≡ 0, it turns out that solutions
of the boundary value problem (1)–(2) exist only for certain special values
of λ, which are called the eigenvalues of the Sturm–Liouville problem. The
corresponding solutions are called the eigenfunctions of the Sturm–Liouville
problem.
We shall say that a function f (z) belongs to the class M (in brief, f ∈ M )
if f (z) is twice continuously differentiable on the interval [0, h] (in brief,
f ∈ C 2 [0, h]) and if it satisfies conditions (2)–(3). An operator L is said to
be self-adjoint if for all functions f, g ∈ M ,
Integrating the last integral by parts twice removes the derivatives of the
function g(z), and we obtain
h h
d
f (z) dz p(z)g (z) dz = d
g(z) dz p(z)f (z) dz
0 0 h
−p(z) g(z)f (z) − f (z)g (z) .
0
The condition (18) is obviously easier to verify than the original linear
dependence condition (16).
To prove the above observation, note that the function Z(z) defined in
(16), differentiated repeatedly, yields the following system of n equalities:
n
k=1 Ck Zk (z) = Z(z),
n
C Z
k=1 k k (z) = Z (z);
... = ... (19)
n (n−1)
k=1 Ck Zk (z) = Z (n−1) (z).
But such a solution must be identically equal to zero, and this means that
the condition (16) is fulfilled and the functions Z1 (z), . . . , Zn (z) are linearly
dependent. The reverse implication is obvious.
To prove this, let f (z) and g(z) be solutions of the boundary value prob-
lem (1)–(2) corresponding to λ1 and λ2 , respectively, λ1 = λ2 . This means
that
Lf ≡ λ1 ρf and Lg ≡ λ2 ρg.
Multiplying the first equation by g(z), the second by f (z), subtracting the
second from the first, and integrating over the interval [0, h] we get
The left-hand side of the above identity is zero in view of the self-adjointness
of the Sturm–Liouville operator. Since λ1 = λ2 , we obtain that (ρf, g) = 0.
At this point we would like to introduce the concept of simple eigenvalue.
The linearity of the problem (1)–(2) implies that there are infinitely many
eigenfunctions corresponding to each eigenvalue. For example, if Z1 (z) is such
an eigenfunction, then for every C = 0, so is Z2 (z) = CZ1 (z). If all eigen-
functions corresponding to an eigenvalue λ are of this form, then λ is called
a simple eigenvalue. Otherwise, such an eigenvalue is called multiple. If every
eigenfunction Z(z) corresponding to λ can be written as a linear combination
(18) of linearly independent eigenfunctions Z1 (z), . . . , Zn (z) corresponding to
λ, then the eigenvalue λ is said to have multiplicity n.
Indeed, let f (z) and g(z) be two eigenfunctions of the boundary value
problem (1)–(2) corresponding to the eigenvalue λ, which means that both
of them satisfy (1). We verify directly that their Wronskian,
f (z) g(z)
W (z) = = f (z)g (z) − g(z)f (z), (22)
f (z) g (z)
42 Chapter 9. Potential Theory and Elliptic Equations
(Lf, f ) ≥ 0. (25)
Indeed, h
d
(Lf, f ) = (q, f ) −
2
f (z) (p(z)f (z)) dz. (26)
0 dz
Integrating by parts yields
h
d
− f (z) (p(z)f (z)) dz = (p, f 2 ) + p(0)f (0)f (0) − p(h)f (h)f (h).
0 dz
Substituting the result into (26) and utilizing condition (2), which is satisfied
for every function from the set M , we finally obtain
a1 2 a2
(Lf, f ) = (q, f 2 ) + (p, f 2 ) + p(0) f (0) + p(h) f 2 (h). (27)
b1 b2
Observe that for b1 > 0, b2 > 0, all the terms on the right-hand side are
nonnegative in view of (3) and (5), and for b1 = 0, b2 = 0, the last two terms
vanish. This concludes the proof of Property 3.
Let us complement this equation with the boundary conditions of the first
kind
Z(0) = Z(h) = 0. (30)
Equation (29) can be written in the simplified form
Z + μZ = 0, (31)
where
λρ − q μp + q
μ= , or, equivalently, λ = . (32)
p ρ
First, consider the case μ < 0. Then the general solution of (31) is of the
form √ √
Z = C1 e− |μ|z + C2 e |μ|z , (33)
where C1 and C2 are arbitrary constants. Substituting this expression into the
two boundary conditions (30), we arrive at the following system of equations
for C1 and C2 :
√C1 + C2 √ = 0,
(34)
− |μ|h |μ|h
C1 e + C2 e = 0.
44 Chapter 9. Potential Theory and Elliptic Equations
1
Δ= √ √1 = 2 sinh |μ|h , (35)
− |μ|h |μ|h
e e
is nonzero for every h > 0 and μ < 0. This means that the system (34) has
only the trivial solutions C1 = C2 = 0. Thus μ < 0 cannot be an eigenvalue,
because it corresponds to a trivial, identically vanishing, solution of (31).
It is not difficult to show that in the case of the boundary conditions (30),
the same argument remains valid for μ = 0 when the general solution of (31)
is of the form
Z = C1 + C2 z. (36)
Now let us consider the case μ > 0. Here the general solution of (31) is
of the form
√ √
Z(z) = C1 cos μz + C2 sin μz. (37)
The boundary conditions (30) imply that
C1 = 0;
√ √ (38)
C1 cos μh + C2 sin μh = 0.
1 0 √
Δ= √ √ = sin μh, (39)
cos μh sin μh
vanishes for
πn 2
μn = , n = 1, 2, . . . , (40)
h
so that in view of (32), the eigenvalues of the Sturm–Liouville problem (29)–
(30) are
p πn 2 q
λn = + . (41)
ρ h ρ
The corresponding nontrivial eigenfunctions of the Sturm–Liouville problem
(29)–(30) are
√ πn
Zn (z) = An sin μn z = An sin z , (42)
h
where the constants An are selected so that each of the eigenfunctions has
norm one, i.e.,
h
πn
Zn = (Zn , Zn ) = An
2 2
sin2 z dz = 1. (43)
0 h
9.8. Sturm–Liouville Problem 45
It turns out that there exists an intimate connection between the eigenvalues
and eigenfunctions of the Sturm–Liouville problem and relation (47). More
precisely, if f = Z1 (z) is a function realizing the minimum in (47), then it is
also an eigenfunction of the Sturm–Liouville problem corresponding to the
eigenvalue λ1 = μ. Moreover, μ is necessarily the smallest eigenvalue.
Indeed, consider the functional
J[f ] = (Lf, f ) − μf 2 ≥ 0.
Since this functional attains its smallest value at the function Z1 (z), we have
J[Z1 ] = (LZ1 , Z1 ) − μZ1 2 = 0.
The auxiliary function
ϕ() := J[Z1 (z) + η(z)],
where η(z) is an arbitrary function from the class M , is a differentiable
function of the variable that attains its minimal value at = 0, where
ϕ (0) = 0. On the other hand,
d
ϕ (0) = L(Z1 + η), Z1 + η − μ ρ(Z1 + η), Z1 + η .
d =0
46 Chapter 9. Potential Theory and Elliptic Equations
The self-adjointness of the operator L, (9), and the symmetry of the scalar
product with weight ρ,
imply
h
ϕ (0) = 2(LZ1 − μρZ1 , η) = 2 [LZ1 (z) − μρ(z)Z1 (z)] η(z) dz = 0.
0
(Lf, f )
λn = inf , (49)
f ∈Mn−1 f 2
9.8. Sturm–Liouville Problem 47
where
Mk := {f ∈ M : (ρf, Zj ) = 0, j = 1, 2, . . . , k}.
Moreover, the minima are attained at the corresponding eigenfunctions
(L(1) f, f ) (L(2) f, f )
≥ .
f 2 f 2
Thus the same inequality is preserved for the minimal values of the left and
right sides, i.e., for the first eigenvalues of the operators L(1) , L(2) . In other
(1) (2)
words, λ1 ≥ λ1 .
Now let us keep the functions p(z) and q(z) unchanged and consider
different weight coefficients ρ1 ≥ ρ2 in (1). Then for every f ∈ M ,
h h
f 2ρ1 = f 2 (z)ρ1 (z) dz ≥ f 2 (z)ρ2 (z) dz = f 2ρ2 .
0 0
Consequently,
(L(1) f, f ) (L(2) f, f )
≤ ,
f 2ρ1 f 2ρ2
so that
(1) (L(1) f, f ) (L(2) f, f ) (2)
λ1 = inf ≤ inf = λ1 .
f ∈M f 2ρ1 f ∈M f 2ρ2
48 Chapter 9. Potential Theory and Elliptic Equations
In other words, if the weight function ρ is replaced by a larger one, then the
first eigenvalue gets smaller. The same observation remains valid for other
eigenvalues.
The above two results imply the following fundamental property of the
Sturm–Liouville problem:
To prove this asymptotic result, let us consider, together with the Strum–
Liouville equation
d
(p(z)Z ) − q(z)Z + λρ(z)Z = 0, (50)
dz
the equations
p1 Z + (λρ2 − q1 )Z = 0,
(51)
p2 Z + (λρ1 − q2 )Z = 0,
where p1 , q1 , ρ1 are the maximal values of the functions p(z), q(z), ρ(z) on
the interval z ∈ [0, h], and p2 , q2 , ρ2 are their minimal values. Denote the
eigenvalues of the Sturm–Liouville problems corresponding to (50)–(51) by
λn , λn , λn , respectively. The above comparison theorems imply that λn ≤
λn ≤ λn , and the extreme terms in these inequalities can be easily determined
because of the simplicity of the corresponding equations (51). For the sake of
concreteness we will consider the above equations with boundary conditions
of the first kind (30). In this case, in view of the solutions found in 9.8.3,
we already know the eigenvalues of the Sturm–Liouville problem (51), (30),
so the above inequalities give the following inequalities for the eigenvalues of
the problem (50), (30):
p1 πn 2 q1 p2 πn 2 q2
+ ≤ λn ≤ + . (52)
ρ2 h ρ2 ρ1 h ρ1
to those eigenfunctions. We shall prove this fact assuming that the eigen-
functions Z1 (z), Z2 (z), . . . have already been normalized with respect to the
weight function ρ(z), i.e.,
h
(ρZn , Zm ) = Zn Zm ρ(z) dz = δnm , (53)
0
with coefficients
h
Ck = (f, ρZk ) = f (z)Zk (z)ρ(z) dz, (55)
0
is called the Fourier series of the function f (z) with respect to the orthonor-
mal system Z1 (z), Z2 (z), . . . . The sums
n
Sn (z) = Ck Zk (z) (56)
k=1
is called the remainder term of the series. We shall say that the Fourier series
converges in the mean square to the function f (z) if
2
n
Rn 2ρ = f (z) − Ck Zk (z) −→ 0, (n → ∞). (57)
k=1 ρ
n
f 2ρ = (f, ρf ) = Ck2 + Rn 2ρ . (58)
k=1
To prove this property, observe that in view of the definition of the re-
mainder of the Fourier series,
n
f 2ρ = (Sn + Rn , ρ(Sn + Rn )) = Sn 2ρ + Rn 2ρ = Ck2 + Rn 2ρ , (60)
k=1
n
Ck2 ≤ f 2ρ . (61)
k=1
∞
Ck2 = f 2ρ , (62)
k=1
To verify Steklov’s theorem, let us represent f in the form of the sum (59)
and consider the quadratic functional
(Lf, f ) = (LSn + LRn , Sn + Rn ) = (LSn , Sn ) + (LRn , Rn ) + 2(LSn , Rn ).
The last term on the right vanishes in view of Property 7 of the remainder
term, and evaluation of the first term yields
n n
n
n
(LSn , Sn ) = L Ck Zk , Ck Zk = λ k Ck Zk , Ck Zk =
k=1 k=1 k=1 k=1
n
n
n
λk Ck Ci (ρZk , Zi ) = λk Ck2 .
k=1 i=1 k=1
The above two equalities yield
m
(LRn , Rn ) = (Lf, f ) − λk Ck2 ≤ (Lf, f ). (63)
k=1
We shall use this inequality to obtain an estimate for the last term in the
equality (58), keeping in mind that the remainder Rn belongs to the set of
functions
Mn = {f ∈ M : (ρf, Zj ) = 0, j = 1, 2, . . . , n},
and as such, satisfies the inequality
(Lf, f ) (Lf, f )
λn+1 = inf ≤ .
f ∈Mn f ρ
2 Rn 2ρ
This inequality and (63) imply
1
Rn 2ρ ≤ (Lf, f ).
λn+1
Since we showed earlier that λn+1 → ∞ as n → ∞, we get
lim Rn ρ = 0.
n→∞
and moreover, the Fourier series on the right-hand side converges absolutely
and uniformly.
52 Chapter 9. Potential Theory and Elliptic Equations
Indeed, if we take the scalar product of both sides with the test function φ,
we arrive at (64), which shows the validity of the equality (65) in the weak
distributional sense.
is of the form
∞
δ(z − z0 ) = Zk (z0 )Zk (z), (67)
k=1
∞
(δ(z − z0 ), φ) = Dk Zk (z).
k=1
9.9. Waves in Waveguides Revisited 53
∂ 2G
Δ⊥ G + + k 2 (z)G = δ(y)δ(z − z0 ) (1)
∂z 2
and the homogeneous boundary conditions
∂ ∂
G(y, z, z0 ) = 0, G(y, z, z0 ) = 0. (2)
∂z z=0 ∂z z=h
∂ 2u
2
Δ⊥ u + + k (z)u = δ(y) A(μ)Zμ (z). (4)
∂z 2 μ
The functions Yμ (y) and Zμ (z) enter into (3) and (4). The former are given
by integrals (9.7.9) and (9.7.11), while the latter solve the boundary value
problem
From now on, to simplify our notation we will drop the index μ. In view of
Sect. 9.8, problem (5) reduces to the Sturm–Liouville problem (9.8.1)–(9.8.4).
Indeed, it suffices to take there
Here k 2 is the maximal value of the coefficient k 2 (z) inside the ocean layer
z ∈ [0, h].
It follows from the results of the preceding section that the Sturm–
Liouville problem (5)–(6) has an infinite number of solutions (eigenfunctions)
Z1 (z), Z2 (z), . . . corresponding to an increasing sequence of eigenvalues
54 Chapter 9. Potential Theory and Elliptic Equations
In this context, it is clear that an abstract index μ in solution (3) can be re-
placed by an index labeling the eigenfunctions, and (3) itself can be rewritten
in the form
u(y, z) = An Yμ(n) (y)Zn (z), (10)
n
where
μ2 (n) = k 2 − λ2n (11)
expresses the old parameter μ via the new index of summation n correspond-
ing to the numbering of the eigenfunction. In the new notation, (4) takes the
form
∂ 2u
Δ⊥ u + 2 + k 2 (z)u = δ(y) An Zn (z). (12)
∂z n
Comparing its right-hand side with the distributional equality (9), we con-
clude that (12) is identical with (1), and that if we select An = Zn (z0 ), then
equality (10) provides the solution of the boundary value problem (1)–(2).
Consequently, the Green’s function of the acoustic waves inside the ocean
layer is of the form
G(y, z, z0 ) = Yμ(n) (y)Zn (z0 )Zn (z). (13)
μ(n)
The explicit form of the Green’s function will become available once we find
all the eigenfunctions Zn (z) and substitute them into (13). In this calculation
let us restrict our attention to the case of the homogeneous ocean, where
k 2 (z) = k 2 = const.
9.9. Waves in Waveguides Revisited 55
Z + λZ = 0, λ = k 2 − μ2 . (14)
where
π
μ2 (n) = k 2 − k02 (2n + 1)2 , .k0 = (21)
2h
Now it is clear that μ2 (n) is a decreasing function of the integer-valued vari-
able n, whose the maximum value is
The analysis of the function Yμ (y) carried out in Sect. 9.7 shows that if μ20 is
negative, then all Yμ (y) decay exponentially as |y| increases. In physics jar-
gon, one often speaks of such Green’s functions (and thus the hydroacoustic
wave described by it) as being localized. In particular, at large distances (as
|y| → ∞), one can with good accuracy approximate (20) by its first term,
k0 2
G(y, z, z0 ) ≈ exp −|μ0 |r cos(k0 z) cos(k0 z0 ). (23)
π π|μ0 |r
Here we used the asymptotic formula (9.7.12) and introduced the radial co-
ordinate
r = y12 + y22 . (24)
If μ20 > 0 (k > k0 ), the Green’s function behaves in a completely different
fashion. In that case, there exists
1 2
N= ( − 1) ≥ 0 (25)
2
such that for all n ≤ N , we have μ2 > 0. In (25) we use the floor function
notation x (which means the greatest integer not exceeding x), and also
introduce the dimensionless parameter
k 2
= = hk. (26)
k0 π
It follows from (9.7.10) that for μ > 0, the function Yμ (y) decays relatively
slowly as r increases. Thus, the terms in the sum (20) can be split into two
types: the slowly decreasing terms with indices n ≤ N , and the rapidly ex-
ponentially decreasing terms with n > N . For large values of r, the exponen-
tially decreasing terms can be dropped, giving the following approximation
for (20):
4
N
G(y, z, z0 ) ≈ k0 Yμ(n) (y) cos zk0 (2n + 1) cos z0 k0 (2n + 1) . (27)
π n=0
Physicists call the summands in the above formula modes, and say that N +1
modes are propagating in the waveguide.
The final intrinsically consistent asymptotic (as r → ∞) expression for
the Green’s function can be obtained by substituting asymptotics (9.7.10)
into (27):
k0
N
G(y, z, z0 ) = f (r, n) cos zk0 (2n + 1) cos z0 k0 (2n + 1) , (28)
π n=0
9.10. Exercises 57
where
2
f (r, x) = exp −iμ(x)r . (29)
πiμ(x)r
9.10 Exercises
1. Find the Green’s function satisfying the equation G = δ(x − y) for
0 < x, y < l and the homogeneous boundary conditions G(x = 0, y) =
G(x = l, y) = 0.
2. Find the solution of the equation G + k 2 G = δ(x − y) for x, y > 0 that
satisfies the boundary condition G(x = 0, y) = 0 and the radiation
condition.
3. Solve the following boundary problem for the 2-D Laplace equation:
Δu = 0, x1 ∈ R, x2 > 0; u(x1 , x2 = 0) = f (x1 ).
4. What boundary value problem in the half-space x2 < 0 has the solution
found in Exercise 3?
We begin with a study of the classic 1-D diffusion equation (also called heat
equation) and its self-similar solutions. This is the simplest example of a lin-
ear parabolic partial differential equations. Well-posedness of an initial value
problem with periodic data is then discussed. Subsequently, the exposition
switches to the complex domain, and we introduce a simple version of the
general Schrödinger equation. This makes it possible to study the diffrac-
tion problem and the so-called Fresnel zones. Multidimensional parabolic
equation follow, and the general reflection method is explained. The chapter
concludes with a study of the moving boundary problem and the standard
physical problem of particle motion in a potential well.
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 59
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 2,
© Springer Science+Business Media New York 2013
60 Chapter 10. Diffusions and Parabolic Equations
equation
∂G D ∂ 2G
= + δ(x)δ(t). (2)
∂t 2 ∂x2
We will ask for only a forward-in-time, t ≥ 0, solution of (2), since in this
case, the problem is well posed. Recall that a problem is said to be well posed
if it has a unique solution that changes little under small perturbations of
the initial conditions. The fact that the initial value problem for (2) is well
posed only in the forward-in-time direction has profound physical signifi-
cance. Equations of this type describe irreversible physical processes such as
heat flow and molecular diffusion. For the Green’s function problem for the
diffusion equation to be well posed, the Green’s function itself has to satisfy
the causality condition
G(x, t < 0) = 0.
Hence, proceeding as in the case of the fundamental solution (equation (2.2.9)
in Volume 1) to the initial value problem for an ordinary differential equation,
we shall be seeking the Green’s function of the diffusion equation (2) in the
product form
G(x, t) = u(x, t)χ(t), (3)
where χ(t) is the familiar Heaviside function, and u(x, t) is an unknown
function. Substitute (3) into (2) and recall that χ (t) = δ(t). As a result, we
obtain that
∂u D ∂ 2 u
χ(t) − + u(x, 0)δ(t) = δ(x)δ(t),
∂t 2 ∂x2
from which it is evident that if we take
then u(x, t) satisfies the homogeneous diffusion equation (1) with the initial
condition (4).
the cumulative distribution of the particles. The function v(x, t) solves the
initial value problem
∂v D ∂ 2v
= , v(x, 0) = χ(x). (1)
∂t 2 ∂x2
Remarkably, problem (1) has a self-similar solution, by which we mean here
a solution determined by a function f of only one argument
ρ = tn x,
which is the product of power functions of the original variables t and x. The
self-similarity property of v can then be stated as follows: for every x and
t > 0,
v(x, t) = f (ρ) = f (tn x) = v(xtn , 1).
nx D
f = f .
tn+1 2
This equation is consistent only if the fraction on the left-hand side is a power
of the self-similar variable ρ, i.e., if
x
= ρα = tnα xα ,
tn+1
for some exponent α. A comparison of the powers of the variables x and t
on the left- and right-hand sides, respectively, yields the coupled equations
62 Chapter 10. Diffusions and Parabolic Equations
√
1 = α and −n − 1 = nα. Hence, α = 1, n = −1/2, ρ = x/ t, and the
equation for f (ρ) takes the form
ρ
f = − f.
D
This equation is easy to solve, since clearly, the function g(ρ) = f (ρ) satisfies
a first-order ordinary differential equation with the general solution
ρ2
g(ρ) = f (ρ) = C exp − ,
2D
f (−∞) = 0, f (+∞) = 1,
which appears in the solution (2), is called the error function, and its graph
is shown in Fig. 10.2.1.
FIGURE 10.2.1
The plot of the error function erf z.
10.2. Self-Similar Solutions 63
or equivalently,
u(x, t) = t−1/2 u(xt−1/2 , 1).
Once the fundamental solution is known, one can easily write the solution
of an arbitrary inhomogeneous diffusion equation
∂u D ∂ 2u
= + ϕ(x, t), u(x, t = 0) = u0 (x).
∂t 2 ∂x2
Indeed, it is of the form
t
u(x, t) = u0 (y)G(x − y, t)dy + ϕ(y, τ )G(x − y, t − τ ) dy dτ. (6)
0
The first convolution integral takes into account the initial condition, while
the presence of a source ϕ(x, t) is reflected in the second convolution integral.
64 Chapter 10. Diffusions and Parabolic Equations
x2 /Dt.
∂u D ∂ 2u
= , u(x, t = 0) = aeikx , (1)
∂t 2 ∂x2
with periodic initial data. Substituting the above initial condition into the
first integral in (10.2.6) and recalling that
2
π k
2
exp(−bx + ikx)dx = exp − , Re b > 0, (2)
b 4b
10.4. Complex Parabolic Equations 65
∂ψ 2 ∂ 2 ψ
i =− . (1)
∂t 2m ∂x2
Its more general version
∂ψ 2
i =− Δψ + U ψ
∂t 2m
is called the Schrödinger equation. It describes the quantum-mechanical laws
of particle motion in a force field with potential U = U (x). The causality
principle is assumed to be satisfied. The constant is called Planck’s con-
stant, and m denotes the particle’s mass. We shall discuss this more general
equation in greater detail in Sect. 10.9.
Here, we will seek a solution of (1) satisfying the initial condition
ψ(x, t = 0) = ψ0 (x).
66 Chapter 10. Diffusions and Parabolic Equations
By definition, the Green’s function G(x, t) of (1) solves the initial value prob-
lem
∂G 2 ∂ 2 G
i =− , G(x, t = 0) = δ(x),
∂t 2m ∂x2
and it can be obtained from the Green’s function (10.2.5) of the diffusion
equation by a formal substitution of the quantity i/m in place of the diffu-
sion coefficient D. Thus
m imx2
G(x, t) = exp . (2)
2πit 2t
which, for τ > t, means that the wave function ψ(x, t) at time t can be
reconstructed from its value ψ(x, τ ) at a later time. This can be accomplished
by solving (1) backward in time. Substituting t = 0 in (3) and expressing
ψ(y, τ ) through the initial state ψ(x, 0) by the same integral, we discover
that the Green’s function has to satisfy the equation
G∗ (x − y, τ )G(y − p, τ )dy = δ(x − p), (4)
where the asterisk denotes complex conjugation, and where we have taken
into account that
G(x, −τ ) = G∗ (x, τ ).
Let us prove the validity of (4) by considering an auxiliary “regularized”
integral
Jε (x, p) = G∗ (x − y, τ )G(y − p, τ ) exp(−ε2 y 2 ) dy. (5)
10.4. Complex Parabolic Equations 67
Substituting the explicit expression (2) for the Green’s function of (1), we
get that
m im 2 im
Jε (x, p) = exp (p − x )
2
exp (x − p)y − ε y dy.
2 2
2πτ 2τ τ
Evaluating this integral with the help of formula (10.3.2), we arrive at the
expression
1 im 2 (x − p)2
Jε (x, p) = √ exp (p − x ) −
2
,
πα 2τ α2
where
z
g(y, z) = (1 + ikr)e−ikr , r = |x| = |y|2 + z 2 , (7)
2πr3
represents the transfer function, or propagator, of the vacuum.
Let us check the behavior of this field far away from the radiation plane
z = 0. Assume that u0 (y) is equal to zero outside the disk of radius a with
center at y = 0, and that the observation point (y, z) is close to the z-axis
(|y|2 z 2 ) and is located in the wave zone (kz 1). In such a case, we are
justified in replacing the factor in front of the exponential function in (7) by
a term with the same principal asymptotics, and we can rewrite (7) in the
form
ik |y|2
g(y, z) = exp −ikz 1 + 2 . (8)
2πz z
If, additionally, the condition
3k|y|4 /8z 3 1
68 Chapter 10. Diffusions and Parabolic Equations
is satisfied, then we are justified (see Sect. 9.5.1) in retaining just the first
two terms in the Taylor expansion of the radical in the exponent of (7). As
a result, after dropping the nonessential factor exp(−ikz), we get that
ik ik|y|2
g(y, z) = exp − . (9)
2πz 2z
∂u
2ik = Δ⊥ u, u(y, z = 0) = u0 (y), (10)
∂z
which describes wave diffraction in the Fresnel approximation. Here, Δ⊥ de-
notes the Laplacian in transverse coordinates y.
ik
g(0, l) = . (2)
2πl
We will try to obtain this equality by other means, utilizing the Huygens–
Fresnel principle, which asserts that a wave in the plane z = l can be repre-
sented as a superposition of fields of secondary sources, which are determined
10.5. Fresnel Zones 69
by the wave incident on a certain auxiliary surface, say the plane z = l/2.
Hence,
g(0, l) = g 2 (y, l/2)d2 y.
Note that the above equality is a direct consequence of the parabolic equation
of quasioptics (10.4.10). Substituting expression (1) for the Green’s function,
we get 2
k i2k|y|2
g(0, l) = − exp − d2 y. (3)
πl l
Changing to polar coordinates ρ and ϕ and using the radial symmetry of the
integrand, we can reduce the above double integral to the single integral
∞
k
g(0, l) = − e−ix dx, (4)
2πl 0
where we have introduced the dimensionless variable of integration x =
2kρ2 /l. Observe that although each of the double integrals in (3) converges,
the integral (4) is a typical divergent integral. The situation is analogous to
a well-known phenomenon in the theory of infinite series: the product of two
convergent series may turn out to be divergent.1
In the case of integral (4), physicists save the day by employing a proce-
dure that essentially amounts to an application of the generalized summation
method satisfying the separation of scales condition. The Abel summation
method of Volume 1, in the continuous case, would require us to replace
(4) by
−k ∞ −ix −αx
lim e e dx.
α→0 2πl 0
After inclusion of the attenuation factor, the integral (4) can be rewritten in
the form ∞
k
g(0, l) = − f (x)e−ix dx. (5)
2πl 0
1
Of course, (3) contains just an ordinary Gaussian integral, which can be computed
by means of analytic continuation from the case of a real-valued exponent. However, our
reasoning emphasizes the usefulness of Abel summation in applied physics problems.
70 Chapter 10. Diffusions and Parabolic Equations
that includes the attenuation factor f (x). Its smooth decay to zero causes the
representing point If (t) to move along a spiral instead of a circle. As t → ∞,
the spiral converges to the center of the original circle (see Fig. 10.5.1b).
Hence, the vector diagram accounting for the smoothly decaying attenuation
factor again leads to the answer If (∞) = −i, which coincides with (6).
The attenuating factor is not necessary if we place a screen in the plane
z = l/2 that is transparent at the point of intersection with the z-axis but
becomes more and more opaque as we move away from the center. Then,
instead of the attenuating factor f (x), it suffices to insert into the integral
(6) a function P (x) that describes the screen’s transparency. This leads to a
convergent (if limx→∞ P (x) = 0 at a fast enough rate) integral
∞
I= P (x)e−ix dx. (7)
0
10.5. Fresnel Zones 71
FIGURE 10.5.1
Vector diagrams for the interference integral I(t).
Now suppose that the screen’s transparency varies in a discrete fashion ac-
cording to the function
∞ 2π(m+1)
−αm
I= e e−ix dx = 0, (9)
m=0 2πm
the screen placed in the plane z = l/2 when the light of a point source passes
through a practically transparent (small α) screen with the transparency
expressed by the formula (8). In the language of vector diagrams, this can
be explained by the fact that the auxiliary integral
t
J(t) = P (x)e−ix dx
0
travels, as t increases, along a family of circles with smaller and smaller radii,
with the common point of the circles located at the origin of the coordinate
system (see Fig. 10.5.1c).
1
N
∂G ∂ 2G
= Dn 2 , G(x, t = 0) = δ(x). (1)
∂t 2 n=1 ∂xn
Recall that the usefulness of the Green’s function arises from the fact that
one can recover with its help any solution of the equation by a straightforward
integration process. If at t = 0, the initial profile of the field is u0 (x), then for
t > 0, the corresponding solution of the 2-D parabolic equation (1) is given
by the convolution integral
1 (x − p)2
u(x, t) = u0 (p)G(x − p, t)d p =
2
u0 (p) exp − d2 p.
2πDt 2Dt
Let us evaluate this convolution for the Gaussian initial profile
μx2
u0 (x) = u0 exp − .
2
In this case, the convolution integral takes the form
1 μp2 (x − p)2
u(x, t) = u0 exp − − d2 p,
2πDt 2 2Dt
which can be evaluated by completing the square in the exponent, so that
2
μp2 (x − p)2 μx2 1+γ x
+ = + p− ,
2 2Dt 2(1 + γ) 2Dt 1+γ
where
γ = μDt
is a dimensionless parameter. Substituting this identity into the integral and
passing to new variables of integration
x
q =p− ,
1+γ
we get that
1 μx2 1+γ 2 2
u(x, t) = u0 exp − exp − q d q.
2πDt 2(1 + γ) 2Dt
The remaining integral splits into two Poisson integrals
2 π
exp(−bx )dx = .
b
The more general integral (10.3.2) reduces to the same Poisson integral for
k = 0. Note that in our case, b = (1 + γ)/2Dt, so that
u0 μx2
u(x, t) = exp − . (2)
1+γ 2(1 + γ)
74 Chapter 10. Diffusions and Parabolic Equations
The above solution of the 2-D diffusion equation indicates that if the
profile u0 (x) of the initial temperature field is Gaussian at t = 0, then it
remains Gaussian for all t > 0. Thus we have discovered a kind of “Gaus-
sian invariance principle” for solutions of parabolic equations in unbounded
homogeneous space. Also, note that the relation (2) contains a much richer
lode of information for the complex equation than was the case for (10.4.1),
or the parabolic equation of quasioptics. We shall exploit this fact to analyze
diffraction of a Gaussian beam.
From the complex wave amplitude u one can obtain a description of the
power characteristics of the wave in terms of the wave intensity
I(y, z) = |u(y, z)|2 .
In our case,
a2 y2
I(y, z) = I0 2 exp − 2 ,
a (z) a (z)
where I0 = |u0 |2 , and
2
z 2 δz F
a(z) = a|1 + γ| = a 1− + , δ= ,
F F ka2
10.7. The Reflection Method 75
FIGURE 10.6.1
Diffracting beam’s effective radius a(z)/a as a function of the dis-
tance from the radiation plane z = 0. For δ 1, its minimum
actually corresponds to the beam’s focusing plane z = F .
is the effective radius of the Gaussian beam at distance z from the radiation
plane. For z = 0, it is equal to the initial radius a, which reflects the weak
convergence, as z → 0, of the Green’s function of the parabolic equation of
quasioptics (10.4.9) to the Dirac delta. A typical graph of a(z) as a function
of z is shown in Fig. 10.6.1.
FIGURE 10.7.1
The domain of diffusion with a moving absorbing barrier.
f (x = h, t) = 0. (3)
2
For more information about Brownian motion, see Volume 3.
10.7. Reflection Method 77
a situation shown in Fig. 10.7.2, then one can construct a solution to the
mixed problem (1)–(3) using the already known Green’s function
1 x2
G(x, t) = √ exp − (5)
2πDt 2Dt
FIGURE 10.7.2
A uniformly moving absorbing barrier.
and for x ≥ αt, it simultaneously solves the boundary problem (1)–(3) with
a uniformly moving barrier. The last Dirac delta in (8), which is positioned
below the barrier x = αt, serves as a sort of “mirror image” of the “real”
Dirac delta in the original initial condition (2); hence the name “reflection
method” for the above procedure.
78 Chapter 10. Diffusions and Parabolic Equations
0.1
–0.1
f(x,t)
–0.2
–0.3
–0.4 x=h
–0.5
–8 –6 –4 –2 0 2 4 6 8
x
FIGURE 10.7.3
A typical solution of the diffusion equation with an absorbing bar-
rier. The parameter values in this figure are D = 2, α = 0.5, q = 2,
and t = 2 (h = 1).
Figure 10.7.3 shows a typical graph of the solution (6) of the diffusion
equation for t > 0. The solid line shows the part of the solution that is
physically meaningful; the dashed line shows its Alice-in-Wonderland-like
“behind-the-mirror” part.
∂g ∂g D ∂ 2 g
= α(t) + , y > 0, t > 0, (9)
∂t ∂y 2 ∂y 2
∂g ∂g D ∂ 2 g
=α + , y ∈ R, t > 0, (12)
∂t ∂y 2 ∂y 2
2qα
g(y, t = 0) = δ(y − q) − exp δ(y + q),
D
with the initial condition obtained by the reflecting condition (10) into the
“behind-the-mirror” region of the negative y’s, also solves the boundary prob-
lem (9)–(10). The solution can be written in the form
1 (y − q + αt)2 2qα (y + q + αt)2
g(y, t) = √ exp − − exp − .
2πDt 2Dt D 2Dt
(13)
Note its special feature that for every t > 0, it enjoys the spatial symmetry
property
2αy
g(y, t) = −g(−y, t) exp − , (14)
D
which is a result of the invariance property of (12). Here, invariance means
that the equation for the function
is identical to the equation for g. Hence, if the initial condition g0 (y) of (12)
enjoys the symmetry property
then the solution g(y, t) will enjoy the same property (14) at every time t > 0.
It is clear from (14) that if the function g is continuous in y, then g(0, t) ≡ 0,
and the boundary condition (11) is automatically satisfied.
The reflection method applied to the boundary problem
∂g ∂g D ∂2g
∂t
= α ∂y + 2 ∂y 2
, y > 0, t > 0, (16)
g(y, t = 0) = g0 (y), y > 0, (17)
g(0, t) = 0, t > 0, (18)
80 Chapter 10. Diffusions and Parabolic Equations
with
α, 0 < t < τ,
α(t) = (20’)
α1 , t > τ.
In this case, for 0 < t < τ , the solution of the boundary value problem (9)–
(11) is identical the the solution of the auxiliary initial value problem (12)
enjoying the symmetry property (14). At time t = τ , the coefficient of (12)
jumps from the value α to α1 , and the equation for g(y, t) assumes the form
∂g ∂g D ∂ 2 g
= α1 + , y > 0, t > τ.
∂t ∂y 2 ∂y 2
In order to apply the reflection method for t > τ , one has to augment this
equation by the initial condition
which, for y > 0, is equal to the solution g(y, τ ) of the initial value problem
(12), and for y < 0, is constructed in such a way that the new symmetry
10.8. Moving Boundary: The Detonating Fuse Problem 81
Suppose that a fuse of length l is stretched along the x-axis over the in-
terval [0, l]. Initially, its temperature is equal to zero: u(x, t = 0) = 0. At
time t = 0, the left endpoint of the fuse is ignited, and thereafter, the tem-
perature of the moving left endpoint is equal to the “burning” temperature
v. The temperature at the right endpoint is kept at zero at all times. The
side surface of the fuse is assumed to be insulated.
Under these circumstances, the temperature distribution u(x, t) in the
fuse is a solution of the following mixed moving boundary problem:
∂u D ∂ 2u
= , t > 0, h(t) < x < l, (1)
∂t 2 ∂ 2x
u(x, 0) = 0, 0 < x < l,
u(h(t), t) = v(t), u(l, t) = 0, t > 0.
Here h(t) represents the time-dependent coordinate of the fuse’s left endpoint,
which moves to the right as the fuse burns. In our concrete calculations we
shall put h = αt, where α is the constant speed of burning. However, for
now, let us keep h(t) to be a general smooth function equal to zero at t = 0.
Also, for the sake of generality, we shall allow the burning temperature v(t)
of the fuse’s left endpoint to vary in time.
Just as was the case in other linear problems considered thus far, the
solution of (1) is given by an integral of the corresponding Green’s function
G(x, q, t) solving the boundary value problem with homogeneous boundary
conditions
∂G D ∂ 2G
= , t > 0, h(t) < x < l, (2)
∂t 2 ∂ 2x
G(x, q, 0) = δ(x − q), 0 < q < l,
G(h(t), q, t) = G(l, q, t) = 0, t > 0.
We will derive a formula that expresses u(x, t) through G(x, q, t) by first
considering the expression
∂
u(q, τ )G(q, x, t − τ ) .
∂τ
Calculating the derivative with respect to τ , and taking into account equa-
tions satisfied by functions u(q, τ ) and G(q, x, t − τ ), we get that
∂ D ∂2 ∂2
(uG) = G(q, x, t − τ ) 2 u(q, τ ) − u(q, τ ) 2 G(q, x, t − τ ) .
∂τ 2 ∂q ∂q
10.8. Moving Boundary: The Detonating Fuse Problem 83
Now integrate both sides of the above equality over the entire length (h(τ ), l)
of the fuse. The integration by parts formula applied to the right-hand side
gives
l l
∂ D ∂u ∂G
(uG)dq = G −u .
h(τ ) ∂τ 2 ∂q ∂q q=h(τ )
Taking into account the boundary conditions in problems (1) and (2), the
equality reduces to the form
l
∂ D ∂
(uG)dq = v(τ ) G(q, x, t − τ ) .
h(τ ) ∂τ 2 ∂q q=h(τ )
Integrating with respect to τ over the interval [0, t], utilizing the initial con-
dition in problem (1) and the probing property of the Dirac delta appearing
in the initial condition in (2), we finally obtain
D t ∂
u(x, t) = v(τ ) G(q, x, t − τ ) dτ. (3)
2 0 ∂q q=h(τ )
Thus we have demonstrated that the integral of the Green’s function gives
the sought solution of the boundary value problem under consideration. Our
next job is to find the Green’s function itself using the reflection method. We
shall do so only in the case of constant burning speed h(t) = αt. As a matter
of fact, the constant α can be either positive or negative. In the latter case,
our moving boundary problem can be interpreted, e.g., as a description of
the temperature field of a crystallizing rod, its the length increasing as the
crystallization process progresses.
It turns out that here it is easier to first guess a solution of a more general
problem and then to obtain the desired solution as a special case. Consider
a function f (x, t) that solves the boundary value problem
∂f D ∂ 2f
= , t > 0, αt < x < l, (4)
∂t 2 ∂x2
84 Chapter 10. Diffusions and Parabolic Equations
∂f D ∂ 2f
= , t > 0, x ∈ R, (5)
∂t 2 ∂x2
f (x, t = 0) = f (x),
where inside the interval (0, l), the function f (x) coincides with the initial
condition
f (x) = f0 (x), 0 < x < l, (6)
of the original boundary value problem (4), and outside is extended so that
f (x, t) automatically satisfies the boundary conditions (4) at x = αt and
x = l. It follows from the discussion of the previous section that it is sufficient
for f (x) to satisfy the symmetry condition (10.6.14) relative to the point
x = 0, i.e.,
f (x) = −f (−x) exp(−2αx/D), (7)
and also the “antisymmetry” condition
relative to the point x = l. Combining (7) and (8), we arrive at the following
useful equality:
f (x) = f (x + 2l) exp(−2αx/D), (9)
which has to be satisfied by the function f (x). It allows us to extend the
definition of f (x) onto the whole x-axis once its values are known inside
any interval of length 2l. The function f (x) can be defined on an interval of
length 2l by either of the formulas (7) and (8). Before doing so, let us extend
the definition of the original initial condition function f0 (x) from (4) to the
whole x-axis by making it equal to zero outside the interval (0, l). Then it
follows from (8) that the values of f (x) on the interval (0, 2l) are given by
the formula
f 0 (x) = f0 (x) − f0 (2l − x). (10)
For x ∈ (0, 2l), the new function f 0 (x) coincides with f (x), and it is equal to
zero outside that interval. Denote by f m (x) a similarly defined function on
10.8. Moving Boundary: The Detonating Fuse Problem 85
!
m−1
2α
m 0
f (x) = f (x + 2ml) exp − (x + 2rl) . (11)
r=o
D
m−1
m(m − 1)
r= ,
r=1
2
we get that
2α
f (x) = f (x + 2ml) exp − m(x + ml − l) .
m 0
(12)
D
Hence, the final form of the initial condition for the auxiliary initial value
problem (5) is
∞
∞
2α
f (x) = m
f (x) = f0 (x + 2ml) exp − m(x + ml − l) (13)
m=−∞ m=−∞
D
∞
2α
− f0 (2l − x + 2nl) exp n(x − nl − l) .
n=−∞
D
In deriving this formula, we used equalities (10), (11), and (12), and replaced
the last sum’s index of summation m by −n.
The solution of the auxiliary initial value problem with the above initial
condition is of the form
f (x, t) = f (r)G0 (x − r, t)dr, (14)
where
1 (x − r)2
G0 (x − r, t) = √ exp − . (15)
2πDt 2Dt
Substituting in (14) the series (13) for f (r), and changing the variables of
integration to q = r + 2ml and q = 2l − r + 2nl, we obtain
l
f (x, t) = f0 (q) [G1 (x, q, t) − G2 (x, q, t)] dq, (16)
0
86 Chapter 10. Diffusions and Parabolic Equations
where
∞
2α
G1 (x, q, t) = G0 (x − q + 2ml, t) exp − m(q − ml − l) (17)
m=−∞
D
and
∞
2α
G2 (x, q, t) = G0 (x + q − 2l − 2nl, t) exp − n(q − nl − l) .
n=−∞
D
γ = αl/D, (20)
which characterizes the competition between two processes: the fuse burn-
ing and the heat diffusion. We shall also pass to the dimensionless Green’s
function
g(s, p, τ ) = lG(sl, pl, τ l2 /D), (21)
10.9. Particle Motion in a Potential Well 87
where
∞
1 (s − p + 2m)2
g1 (s, p, τ ) = √ exp − − 2γm(p − m − 1) (23)
2πτ m=−∞ 2τ
and
∞
1 (s + p − 2 − 2n)2
g2 (s, p, τ ) = √ exp − − 2γn(p − n − 1) .
2πτ n=−∞ 2τ
In the new dimensionless coordinates, the left endpoint of the fuse moves
according to the law
s = γτ. (24)
Substituting (24) in (23), it is not difficult to check that the terms of the
series with indices m and n = m − 1 coincide. So the function g from (22)
satisfies the required boundary condition g(γτ, p, τ ) = 0.
As a final step, let us investigate the convergence of the series (23). For
this purpose we shall group together, in the first series’ exponential, the terms
containing m2 to get
2 2 1
− m + 2γm = 2 γ −
2
m2 .
τ τ
∂G 2 ∂ 2 G
i =− , αt < x < l,
∂t 2m ∂x2
G(x, q, t = 0) = δ(x − q), 0 < x < l, (1)
G(αt, q, t) = 0, G(l, q, t) = 0, t > 0.
The above boundary value problem describes the wave function of a quantum-
mechanical particle located, at time t = 0, at the point x = q and positioned
between the ideal reflecting barriers at x = αt and x = l.
Having already found the Green’s function of the heat equation, we
obtain—without any extra effort—the solution of problem (1):
10.10 Exercises
1. Solve the initial value problem for the parabolic equation
∂f ∂f β(t) ∂ 2 f
+ α(t) = , t ≥ 0, x ∈ R, (1)
∂t ∂x 2 ∂x2
f (x, t = 0) = δ(x),
where α(t) and β(t) are arbitrary integrable functions on the semiaxis
t ≥ 0. Hint: Rewrite the initial value problem in the form of an inho-
mogeneous equation and with the help of the Fourier transform with
respect to the coordinate x, find its solution satisfying the causality
condition.
2. Suppose that the function f (x, t) satisfies the following initial value
problem:
∂f ∂ 1 ∂ ∂f
= [a(x)f ] + β(x) , t > 0, x ∈ R,
∂t ∂x 2 ∂x ∂x
f (x, t = 0) = δ(x), (2)
where a(x) and β(x) are everywhere differentiable functions. Addition-
ally, to ensure that the initial value problem under consideration is
correctly posed, assume that β(x) > 0. Check that
N (t) = f (x, t) dx ≡ 1, t > 0. (3)
90 Chapter 10. Diffusions and Parabolic Equations
3. Using the Fourier transform with respect to x, solve the parabolic equa-
tion
∂f ∂ β(t) ∂ 2 f
= α(t) (xf ) + , t ≥ 0, x ∈ R, (4)
∂t ∂x 2 ∂x2
with the initial condition
4. Find an explicit expression for the solution of the initial value problem
(4)–(5) with constant coefficients α(t) = α > 0 and β(t) = β > 0.
Explore the asymptotic behavior of these solutions as t → ∞.
Find the first two moments of the probability density function f (x, t)
that obeys (7) and the initial condition (5). Discuss their behavior in
time.
with the initial condition (5) (0 < y < ) and the periodic boundary
condition
f (x = 0, t) = f (x = , t), t > 0. (10)
f (x, t = 0) = δ(x),
plays an important role in probability theory. Using the Fourier trans-
form, solve the Kolmogorov–Feller equation and explore the asymptotic
properties (as t → ∞) of its solutions in the case that the kernel w(x)
is Gaussian, i.e., 2
1 x
w(x) = √ exp − . (12)
2π 2
11. Verify numerically the correctness of the asymptotics of the exact con-
tinuous part of the solution of the Kolmogorov–Feller equation with
Cauchy kernel. Utilize the solution to Problem 10 provided at the end
of this volume.
12. Using results of Chap. 4, Volume 1, examine the main asymptotics (for
an arbitrary t > 0 and x → ∞) of the continuous part of the solution
of the Kolmogorov–Feller equation with Cauchy kernel.
Chapter 11
Waves are everywhere, literally, and our senses are acutely attuned to waves
of various types: mechanical, water, acoustic, electromagnetic, optical, etc.
Moreover, quantum mechanics tells us that matter itself is, in a sense,
indistinguishable from waves. In this chapter we concentrate on properties
of linear waves in dispersive media; a discussion of nonlinear waves will
be postponed until Chaps. 12–14. Here the methods of choice are integral
transforms and asymptotic relations, especially the Fourier transform and
the stationary phase method. They will occupy a central role in what follows.
Their main advantage is that they reduce relevant partial differential, or even
integral, equations to algebraic, or in some cases transcendental, equations.
For example, the 1-D wave equation
∂ 2u 2
2∂ u
= c ,
∂t2 ∂x2
already familiar to the readers of this book, is reduced via the Fourier trans-
form to the simple algebraic equation
ω = ±ck.
As a result, in this chapter, partial differential and integral equations
describing wave propagation will be mostly hidden behind their corre-
sponding, and much more convenient and elementary, dispersion relations.
Recall that the above wave equation has the general solution
u(x, t) = f (x − ct) + g(x + ct),
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 93
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 3,
© Springer Science+Business Media New York 2013
94 Chapter 11. Waves and Hyperbolic Equations
where f (x) and g(x) are arbitrary functions. This formula hints at two
remarkable properties of such 1-D waves:
ω = W (κ). (8)
for a wave freely propagating in a medium. Formula (10) describes the so-
called wave packet, which contains a continuum of harmonic waves whose
contributions to the packet are measured by the complex amplitude f˜(κ).
Recall that if u(x, 0) = f (x) is a real function, then f˜(−κ) ≡ f˜∗ (κ).
Accordingly, to make the field u(x, t) in (10) real-valued for every t = 0,
we shall assume below that W (κ) is an odd functions of κ, i.e., W (−κ) =
−W (κ).
The sum
S = κx − W (κ)t + arg[f˜(κ)]
appearing in the exponent in (10) is called the phase of the harmonic wave.
Note that the phase is constant on straight lines x = ct + d. The constant
W (κ)
c= (11)
κ
is called the phase velocity of the harmonic wave.
11.2. Examples of Dispersive Media 97
and the wave propagates preserving its shape. If W depends not only on c
but also on the spatial scale l, then the medium is said to exhibit spatial
dispersion. Similarly, one can define media with temporal or spatiotemporal
dispersions. Even in the case of a single spatial scale, W (κ) can have a struc-
ture much more complex than (12). A good example is provided by the
function
W (κ) = cκ ϕ(lκ), (14)
where the function ϕ(z) is an arbitrary function of a dimensionless argument
defined by concrete properties of the medium.
Waves (1) cannot satisfy these boundary conditions, but their superpositions
FIGURE 11.2.1
Dispersion curves for the first four modes of a waveguide. The ver-
tical axis represents the dimensionless frequency Ω = ωl/c, and the
horizontal axis, the wavenumber κ = kl. The dashed line represents
the nondispersive asymptotics ω = ck.
11.2. Examples of Dispersive Media 99
FIGURE 11.2.2
Plot of the dispersion relation for gravitational water waves. The
axes
represent, respectively, the dimensionless frequency Ω =
ω h/g and the wavenumber κ = kh. The dashed lines represent
the curves corresponding to the deep- (top) and shallow-water (bot-
tom) approximations.
We shall call I(x, t) the wave intensity, leaving aside a discussion of the actual
physical energetic properties of waves. Consequently, in our context, a wave’s
energy is defined by the simplest quadratic integral functional,
E = |u(x, t)|2 dx. (2)
Substituting the right-hand side of (11.1.10) into (2), we arrive at the triple
integral
˜ ˜∗
E= f (κ)f (μ) exp i(κ − μ)x − it W (κ) − W (μ) dκ dμ dx.
11.3. Integral Laws of Motion for Wave Packets 101
and using the probing property of the Dirac delta, we arrive at Parseval’s
equality
E = 2π |f˜(κ)|2 dκ = const, (4)
which “averages” the function g(x) over the whole wave packet with the
weight proportional to the wave intensity, then arguments similar to those
used above to obtain the energy invariance principle lead to the equality
2π
g(x) = g̃(μ − κ)f˜(κ)f˜∗ (μ) exp it[W (μ) − W (κ)] dκdμ. (6)
E
Then, if one takes g(x) = x, one can think of x as the coordinate of the
wave packet’s “center of mass.” In view of (3.3.5) of Volume 1, the singular
Fourier image of this function is
∂
g̃(μ − κ) = i δ(μ − κ).
∂μ
A substitution of this formula into (6) permits evaluation of the integral with
respect to μ, which gives
2πi d ˜∗
x = − f˜(κ) exp (−itW (κ)) f (κ) exp (itW (κ)) dκ. (7)
E dκ
where
2πi d
a=− f˜(κ) f˜∗ (κ) dκ
E dκ
102 Chapter 11. Waves and Hyperbolic Equations
where
p(κ) = W (κ) − vκ.
It is the asymptotic behavior (t → ∞) of this integral that we shall study via
the stationary phase method. As usual, we begin by looking for the stationary
points, which are roots of the equation
p (κ) = W (κ) − v = 0. (1)
Suppose that the equation has only one root, equal to k. Then the sum
corresponding to (5.2.3) reduces to a single term, and
2π
U (t) ∼ ˜(k) exp it(vk − W (k)) ,
f (t → ∞), (2)
itW (k)
assuming that
p (k) = W (k) = 0.
In view of (2) and (1), the astonished experimenter “sees” a unique compo-
nent of the wave’s Fourier image f˜(k) running with the group velocity
v(k) = W (k). (3)
Example. Transverse vibration of a rod. For the transverse vibration
of a rod (Example 11.2.2), the dispersion relation is (11.2.6), and the group
velocity,
v(k) = 2c(k) = 2γk, (4)
is twice as large as the phase velocity. For deep-water waves, the dispersion
law (11.2.7) holds, and the group velocity
1 1 g
v(k) = c(k) = (5)
2 2 k
is only half as large as the phase velocity.
Substituting v = x/t in (2), we get the following asymptotic expression
for the wave at large x and t:
2π
U (x, t) ∼ ˜(k) exp i(kx − W (k)t) .
f (6)
itW (k)
The dependence of k on x and t can be found by solving the equation
x − W (k)t = 0 (7)
for the unknown k.
104 Chapter 11. Waves and Hyperbolic Equations
First of all, note that in the derivation of (11.4.6), we did not take into
account the fact that by (11.1.4), in addition to the stationary point κ =
k > 0, there always exists a conjugate point κ = −k. Both correspond to the
same group velocity. Hence, in the analysis of the energetic properties of the
wave, one has to replace (11.4.6) by
u(x, t) = U (x, t) + U ∗ (x, t) = 2 Re U (x, t).
The corresponding wave intensity is then equal to
I(x, t) = u2 (x, t) = 2|U (x, t)|2 + U 2 (x, t) + U ∗2 (x, t).
We shall split it into two components,
I(x, t) = I0 (x, t) + I1 (x, t), (2)
where
I0 (x, t) = 2|U (x, t)|2 , and I1 (x, t) = U 2 (x, t) + U ∗2 (x, t).
By (11.4.6), the first component satisfies
4π
I0 (x, t) = |f˜(k)|2 (k > 0). (3)
t|W (k)|
11.5. Energy Conservation Law in the Stationary Phase Method 105
This indicates that at large times, the Fourier components of the wave behave
as if they had separated spatially. Integrating (4) over the entire x-axis, we
obtain the conservation law
∞
I0 (x, t)dx = 4π |f˜(κ)|2 dκ = const.
0
Since the function f (x) is real-valued, the modulus squared of its Fourier
transform |f˜(κ)|2 is even. Hence, the last expression is equivalent to the
energy invariant (11.3.3), and we have discovered that the entire energy of
the wave is contained in the first component of the intensity, I0 (x, t).
So, what happened to the energy contained in the second component,
I1 (x, t)? The answer to this question will be provided by investigating the
integral
1 ∞ ˜2 dx
I1 (x, t)dx = 4πIm f (k) exp i(kx − W (k)t) .
t 0 v (k)
Introducing the new variable of integration k via the substitution x = v(k)t,
we get
∞
I1 (x, t)dx = 4πIm f˜2 (k) exp 2ik(v(k) − c(k))t dk.
0
The stationary √phase method predicts the vibration’s decay to zero with the
asymptotics 1/ t as t → ∞. It is also possible to obtain analogous results
about the asymptotic decay of similar integrals for other dispersion laws.
Hence, the conclusion is that the approximate intensity (2) of the wave
does not satisfy the energy conservation law (11.3.4). However, that law is
106 Chapter 11. Waves and Hyperbolic Equations
satisfied by the main component I0 (x, t) of the intensity, while the contri-
bution of the term I1 (x, t) decays to zero as t → ∞. One could eliminate
the contribution from I1 (x, t) and preserve a more accurate energy conserva-
tion law by going beyond the framework of the stationary phase method and
keeping further terms of the asymptotic expansion of the original integral
(11.1.10).
Multiplying (1) by the initial density f0 (y, r) of particles in the phase space
and integrating the result over all y’s and r’s gives the current density of the
gas of particles:
f (x, v, t) = f0 (y, r)δ(X(t; y, r) − x)δ(V (t; y, r) − v) dy dr. (2)
is singular. Here, ρ0 (x) is the initial density of particles, and v(x) is the
velocity of particles located originally at the point x. Substituting (4) into (3),
we obtain
f (x, v, t) = ρ0 (y)δ(v − v(y))δ(x − y − vt)dy. (5)
Integrating the particle density f (x, v, t) over all v’s, we obtain the time
evolution of the spatial density of the particles:
ρ(x, t) = f (x, v, t)dv. (6)
similar to the expression for the wave intensity in the framework of the sta-
tionary phase approximation (11.5.4).
Here, v1 and v2 stand for velocities of the extreme particles (v1 = v(y1 ),
v2 = v(y2 ) if v(y) increases monotonically, and v1 = v(y2 ), v2 = v(y1 ) if it
decreases). In addition, in view of the mass conservation law, the density of
a gas of diverging particles should decrease. Indeed, in view of the probing
property of the Dirac deltas of a composite argument, (8) implies that
ρ0 (y)
ρ(x, t) = , (9)
t|v (y)|
1 x
ρ1 (x, t) = ρ0 ,
t t
11.6. Wave as Quasiparticle 109
retains the shape of the initial matter density. In the second case, the nonuni-
formity of the flow of matter and a change in the order in which the particles
are arranged leads to a qualitatively different law,
2 t2 t
ρ (x, t) = 3 ρ0 .
x x
∂ ∂
I+ (v(x, t)I) = 0. (10)
∂t ∂x
Equations for k(x, t) and v(x, t) can be found by differentiating the equation
x = v(k)t
and where the support of f˜(κ) contains the origin κ = 0 and has width
Δ k. As a rule, W (κ) is a smooth function of κ, slowly varying on ab
interval of length Δ.
In such a case, W (κ) in (1) can be replaced, with a negligible error, by
the first three terms of its Taylor expansion,
1
W ≈ W (k) + v(k)(κ − k) + r(κ − k)2 , r = W (k).
2
This gives
˜ μ2
U (x, t) = exp[i(kx − ωt)] f (μ) exp iμ(x − vt) − irt dμ, (2)
2
11.8. Optical Wave Behind a Phase Screen 111
∂u
2ik = Δ⊥ u, u(x, z = 0) = u0 (x). (1)
∂z
Suppose that a phase screen is placed in the plane z = 0 and that it changes
the incident wave’s phase by kψ(x). If the incident wave is planar, propagat-
ing along the z-axis, then its complex amplitude just in front of the screen is
and the solution of the boundary value problem (1) is of the form
ik ik(y − x)2
u(x, z) = exp −ikψ(y) − d2 y. (3)
2πz 2z
is that the integral (3) is two-dimensional; its analysis requires a 2-D version
of the stationary phase method.
Let us begin by rewriting (3) in a form more convenient for analysis,
ik
u(x, z) = exp (−ik G(x, y, z)) d2 y, (4)
2πz
where
(x − y)2
G(x, y, z) = ψ(y) + (5)
2z
is a function independent of k. By analogy with the 1-D case, we need to find
stationary points of G in the y-plane, i.e., to solve the equation
∇y G = 0,
x = y + v(y)z, (6)
where
v(y) = ∇ψ(y) (7)
is a vector function.
For simplicity’s sake, assume that for a given x, there exists only one
stationary point y ∗ (x, z) in the y-plane. In a neighborhood of y ∗ (x, z), a
twice continuously differentiable function G has the asymptotic form
1
G ∼ G(x, y ∗ , t) + (yi − yi∗ )(yj − yj∗ )rij , (8)
2z
where
rij = δij + zτij . (9)
In the above formula, τij = τji ,
∂2
τij (x, z) = ψ(y) ,
∂yi ∂yj y =y ∗
and δij is the Kronecker symbol.
Let us express the variables of integration y in terms of the main axes
(eigenvectors) of the symmetric matrix rij (or equivalently, of τij ). In the new
coordinate system, the quadratic form (8) has a diagonal representation,
1
G ∼ G(x, y ∗ , t) + (yi − yi∗ )2 [1 + zτi ], (10)
2z
11.8. Optical Wave Behind a Phase Screen 113
The above Dirac delta probes values y = y(x, t) that are roots of the equation
x = y + v(y)z. (4)
FIGURE 11.9.1
Left: Rays behind the phase screen. Right: Intensity behind the phase
screen.
The plot clearly demonstrates that focusing of the rays by the screen
creates areas of increased intensity. The reader may have seen such patterns
created on a wall by sun rays passing through a window pane. A transparent,
seemingly smooth sheet of glass always has some slight thickness fluctuations
and acts as a 1-D phase screen, which explains the creation of luminous zones
on the wall facing the window.
11.10 Caustics
At larger distances from the screen (z > R), the multiray regime sets in,
whereby several rays radiating from different screen points can meet at the
same point (x, z). In this case, the intensity (11.9.3) becomes the sum
1
I(x, z) = , (1)
i
1 + v (y)z
y=yi (x,z)
where the summation extends over all the roots of (11.9.4). This sum
describes the intensity in the incoherent-superposition-of-waves approxima-
tion; all rays arriving at the point (x, z) are taken into account. The areas
of the multiray regimes are surrounded by caustic surfaces (caustic curves
in the (x, z)-plane in the case of a 1-D phase screen), or simply by caustics.
The equation
1 + v (y)z = 0 (2)
116 Chapter 11. Waves and Hyperbolic Equations
–2
–4
–6
0 0.5 1 1.5 2 2.5
z
FIGURE 11.10.1
A set of rays, caustics, and multiray regions.
ρ = (y1 + y2 )/2, s = y1 − y2 ,
and express the old coordinates through the new ones by the inverse relations
y1 = ρ + s/2, y2 = ρ − s/2,
11.10. Caustics 117
FIGURE 11.10.2
Intensity in the multiray area.
and we recover the familiar geometric optics approximation for the wave
intensity (11.9.3). Clearly, it is not good enough here, since it predicts infinite
values of the intensity on the caustics. Hence, the conclusion is that the
expansion should include the next nonvanishing term, which is of the third
degree in l:
lz lz ∼ z 3
k ψ ρ+ −ψ ρ− = zv(ρ)l + v (ρ)l3 .
2k 2k 24k 2
Inserting this approximation into (4), we obtain
1 1 33
L= exp −i(a − x)l − i b l dl, (7)
2π 3
where
z
a(ρ, z) = v(ρ)z + ρ, b(ρ, z) = 3
kv (ρ). (8)
2k
The integral (7) can be expressed via the special function
1 ∞ 1 3
Ai (x) = cos xt + t dt, (9)
π 0 3
called Airy’s function. Its graph is shown in Fig. 11.10.3.
Although as t → ∞, the integrand does not converge to zero, for large t,
the crests and troughs compensate each other sufficiently to guarantee the
integral’s convergence. As a result, the integral of Airy’s function over the
entire x-axis is well defined and equal to 1.
Clearly,
1 a−x
L(x, ρ, z) = Ai , (10)
b b
which weakly converges to (5) as b → 0. Thus in the present approximation,
the final expression for the wave intensity is
1 ρ + v(ρ)z − x
I(x, z) = Ai dρ. (11)
b(ρ, z) b(ρ, z)
A detailed analysis shows that for sufficiently large k, the formula (11)
actually coincides with the geometric optics approximations (11.9.5) and
(11.10.1) in the zones between caustics. Close to the caustics, Airy’s function
eliminates the singularities by limiting the maximal value of the intensity.
11.11. Telegrapher’s Equation 119
0.6
0.4
0.2
Ai(x)
0
–0.2
–0.4
–0.6
–30 –25 –20 –15 –10 –5 0 5
x
FIGURE 11.10.3
Graph of Airy’s function.
FIGURE 11.11.1
Schematic illustration of an infinitesimal section of a long trans-
mission line.
equality by dx, we obtain the first equation linking voltage and current in
the transmission line:
∂v ∂i
+L + Ri = 0. (1)
∂x ∂t
The second equation for the two unknown functions v and i will be obtained
by writing the balance of currents entering and leaving the line segment
[x, x + dx]:
∂i
−di = i(x, t) − i(x + dx, t) = − dx,
∂x
which is determined by the capacitative charging C dx vt of this line seg-
ment and the current drain G dx v due to the conductivity of the insulation.
Consequently,
∂i ∂v
− dx = C dx + G v dx ,
∂x ∂t
or
∂i ∂v
+C + Gv = 0. (2)
∂x ∂t
The system of two first-order equations (1) and (2) can now be reduced to
a single second-order equation for voltage v(x, t) or current i(x, t) as follows.
Differentiate (1) with respect to x and (2) with respect to t, and multiply
the latter by L to obtain
∂ 2v ∂ 2i ∂i
2
+ L +R = 0,
∂x ∂t∂x ∂x
∂ 2i ∂ 2v ∂v
L + LC 2 + LG = 0.
∂x∂t ∂t ∂t
Subtracting the second equation from the first one, we get
∂ 2v ∂ 2v ∂v ∂i
− LC − LG +R = 0.
∂x2 ∂t2 ∂t ∂x
Now we can eliminate the derivative ix using (2), arriving at the final form
of the telegrapher’s equation:
∂ 2v ∂ 2v ∂v
= LC + (CR + LG) + RG v . (3a)
∂x2 ∂t2 ∂t
11.11. Telegrapher’s Equation 121
Similarly, one can obtain the following equation for current i(x, t):
∂ 2i ∂ 2i ∂i
2
= LC 2
+ (CR + LG) + RG i . (3b)
∂x ∂t ∂t
Thus, voltage and current satisfy the same hyperbolic telegrapher’s equation.
The telegrapher’s equation may be simplified by the substitution
Indeed, substituting (4) into (3) and taking into account that
∂ 2v 2
−μt ∂ u ∂v ∂u
= e , = e−μt ( − μu)
∂x2 ∂x2 ∂t ∂t
and
∂ 2v 2
−μt ∂ u ∂u
2
= e ( 2
− 2μ + μ2 u) ,
∂t ∂t ∂t
we get an equation for the function u(x, t):
∂ 2u ∂ 2u ∂u
= LC + (CR + LG − 2μLC)
∂x2 ∂t2 ∂t
CR + LG
μ= , (5)
2LC
∂ 2u 2
2∂ u
= a + b2 u , (6)
∂t2 ∂x2
where
2
2 1 2 CR − LG
a = , b = . (7)
LC 2LC
122 Chapter 11. Waves and Hyperbolic Equations
and
C −Rt
i(x, t) = e L [ϕ(x − at) − ψ(x + at) + α] .
L
Finally, it is not difficult to see that the introduction of an arbitrary
integration constant α is not necessary, since a suitable redefinition of the
arbitrary functions ϕ and ψ produces the same effect. Indeed, one can take
α α
Φ(x) = ϕ(x) + , Ψ(x) = ψ(x) − ,
2 2
which gives the solutions in the final form
v(x, t) = e− L t Φ(x − at) + Ψ(x + at) ,
R
C −Rt (14)
i(x, t) = e L Φ(x − at) − Ψ(x + at) .
L
If, using equality (4), we transform (3) into (6), then it is necessary to
formulate initial conditions for (6) that are equivalent to the initial condi-
tions (15).
To accomplish this, note that
Combining the above equality with equalities (5), (15), and (16), we obtain
the following initial conditions for (6):
FIGURE 11.11.2
Voltage v(x, 0) = E(t) applied at the point x = 0 of a long transmis-
sion line.
FIGURE 11.11.3
A long transmission line with two endpoints grounded through
concentrated resistivities.
Now let us assume that one endpoint (at x = 0) of the line is grounded
through a concentrated resistance R0 (see Fig. 11.11.3). Then v(−0, t) = 0,
but the jump condition gives the boundary value
FIGURE 11.11.4
A long transmission line grounded at the left endpoint by a con-
centrated inductance and at the right endpoint by a concentrated
capacitance.
FIGURE 11.11.5
Schematic diagram of an infinite transmission line grounded at
x = 0 through a concentrated resistivity.
e− L t g(x − at) ,
R
v(x, t) =
C −Rt (26)
i(x, t) = e L g(x − at) .
L
128 Chapter 11. Waves and Hyperbolic Equations
The absence here of terms containing the expression x + at follows from the
obvious condition that in the right half of the transmission line there are no
waves incident on the grounded point (the so-called radiation condition).
Our problem will be solved if we find a function ϕ(x) for x > 0 and a
function g(x) for x < 0. Substituting (25) and (26) in (24), we arrive at the
following equation for ϕ and g:
where
C
ρ = R1
L
is the nondimensionalized grounding resistivity (see also Exercise 11.12.5).
The second equation,
results from the continuity condition v(−0, t) = v(+0, t) for the voltage func-
tion at the grounding point.
Substituting z = −at < 0 in the last two equations, we get
FIGURE 11.11.6
Dependence of the reflection coefficient K and the transmission co-
efficient T on the nondimensionalized grounding resistance ρ, for
an infinite transmission line grounded in the middle through a con-
centrated resistivity.
v(x, t) = e− L t f (x + at)
R
was present, and where f (x) = 0 for x ≤ 0. We shall determine the distribu-
tion of the voltage in this line for t > 0, given that from the moment t = 0
on, the line is short-circuited to the ground at a moving point of contact
x = ψ(t) (ψ(0) = 0).
In mathematical terms, the problem can be formulated as follows: We seek
a solution of (3a) (CR = LG) in the domain ψ(t) < x < +∞, 0 < t < +∞,
satisfying the moving boundary condition
v(ψ(t), t) = 0 (ψ(0) = 0)
To find the solution, we shall try to find the voltage function in the form
R
v(x, t) = e L t f (x + at) − ϕ(at − x) , t > 0,
where the unknown function ϕ(z) equals zero for z < 0, and for z > 0, it is
determined by the condition that voltage at the point of short circuit is zero:
Thus
ϕ(z) = f (y(z)), z > 0, (28)
where y is a function of z given parametrically:
Replacing z by at − x, we finally obtain that for t > 0 and x > ψ(t), the
voltage in a long transmission line evolves as follows:
v(x, t) = e− L t f (x + at) − f (y(x − at)) .
R
(30)
Physically, the above solution makes sense only when the velocity of mo-
tion of the contact point does not exceed the propagation velocity of the
wave in the transmission line, i.e., when |ψ̇(t)| < a. In this case,
ẏ a + ψ̇(t)
y (z) = = > 0,
ż a − ψ̇(t)
i.e., the function y(z) must be monotonically increasing, and the reflected
wave will satisfy the following causality principle: it should arrive at an ar-
bitrary point x > ψ(t) in the same time order in which the incident wave
propagated through x.
To better understand the phenomenon of reflection from a moving short-
circuited point, it is worthwhile to analyze it in a few typical cases of the
regime of motion.
Example 1. Boundary moving with constant velocity. Consider a contact
point sliding along the transmission line with constant velocity, that is, ψ(t) =
ct. In this case, it follows from (29) that
a+c
y= z. (31)
a−c
11.11. Telegrapher’s Equation 131
Now assume that the contact point moving with constant velocity is over-
taken by an incident sine wave,
where ω is the wave frequency, k = ω/a is its wave number, and χ(x) is the
Heaviside unit step function (equal to 1 for x > 0, and 0 for x ≤ 0).
FIGURE 11.11.7
Dependence of the frequency of the reflected wave on the nondi-
mensional parameter μ = c/a, describing the ratio of the velocity
of the boundary to the propagation velocity of the wave.
where
a+c a+c
ω = ω , k = k . (32)
a−c a−c
Equalities (32) represent the familiar Doppler effect: a wave reflected from
a moving boundary has frequency ω and wave number k different from the
frequency ω and the wave number k of the incident monochromatic wave.
Figure 11.11.7 shows the dependence of the frequency of the reflected wave
on the nondimensional parameter μ = c/a.
Observe that the Doppler effect results in compression of the reflected
wave profile (as compared to the incident wave) if the boundary is moving
toward the wave (c > 0), and in dilation of the reflected wave profile if the
boundary is moving away from the wave (c < 0).
132 Chapter 11. Waves and Hyperbolic Equations
FIGURE 11.11.8
The magnitude y (τ ) of the compression of the wave reflected from
a vibrating boundary as a function of the nondimensional variable
τ = Ωz/a. Graphs are shown for three different values of the nondi-
mensional parameter μ = c/a, representing the ratio of the maximal
velocity c = ΩH of the moving boundary point to the velocity of
the propagating wave.
Let us take a close look at this phenomenon by changing to a nondimen-
sional variable τ = Ωz/a. The above discussion gives
dy(τ ) 1 + μ cos ρ
= , τ = ρ − μ sin ρ .
dτ 1 − μ cos ρ
The only nondimensional parameter μ = c/a entering in the above formula
is the ratio of the maximal velocity c = ΩH of the reflection point to the
velocity of the propagating wave. Several graphs of the function y (τ ) for
different values of the parameter μ are shown in Fig. 11.11.8.
π R G
√ > − . (33)
l CL L C
Insulation of the right endpoint means that i(l, t) = 0, but also that it (l, t) =
0. Thus, taking (1) into account, the voltage at the right boundary satisfies
the condition
vx (l, t) = 0 (0 < t < +∞) .
Let us reformulate the initial conditions and consider a more general sit-
uation, which, as sometimes happens, leads to a simpler analysis. So as-
sume that the initial electric charge at time t = 0 has a linear charge
density ρ(x). Consider a small segment [x, x + dx] having length dx (see
Fig. 11.11.1). The charge of this segment is dQ = ρ(x)dx, and its capac-
itance is dC = Cdx. Accordingly, the potential of the above segment is
v(x, 0) = dQ/dC = (1/C) ρ(x). Now, if we substitute here the density of
the concentrated charge ρ(x) = Qδ(x − x0 ), we shall obtain the first initial
condition,
Q
v(x, 0) = δ(x − x0 ).
C
The second initial condition is derived with the help of (2). Since at the
initial time, i(x, 0) = 0 and ix (x, 0) = 0, we have vt (x, 0) = −(G/C) v(x, 0),
or taking into account the first initial condition,
QG
vt (x, 0) = − δ(x − x0 ).
C2
To solve this problem it is convenient to pass from the voltage function
v(x, t) to an auxiliary function u(x, t), connected with the voltage by the
equality
v(x, t) = e−μt u(x, t) , (34)
and satisfying the equation
From (34) and (17), it follows that the boundary and initial conditions for
u(x, t) corresponding to the above boundary and initial conditions for v(x, t)
are of the form
u(0, t) = 0 , ux (l, t) = 0 , (36a)
Q QG
v(x, 0) = δ(x − x0 ) , vt (x, 0) = − 2 δ(x − x0 ) . (36b)
C C
The tilde emphasizes the fact that the functions X̃n (x) are normalized on the
segment [0, l], i.e.,
l
X̃n (x), X̃n (x) = X̃n2 (x) dx = 1 .
0
11.11. Telegrapher’s Equation 135
As always, (f (x), g(x)) denotes the inner product of functions f (x) and g(x)
on the interval [0, l].
The family of functions {X̃n (x)}, n = 0, 1, 2, . . ., forms an orthonormal
system of the eigenfunctions of the boundary value problem (37). The corre-
sponding numbers λn are called the eigenvalues. The graphs of the first three
eigenfunctions (with l = 1) are shown in Fig. 11.11.9.
FIGURE 11.11.9
The first three eigenfunctions, X̃n (x), n = 0, 1, 2, . . . , of the boundary
value problem (37) (l = 1).
where
ωn = λ n a2 − b 2 .
∞
u(x, t) = (an cos ωn t + bn sin ωn t)X̃n (x) .
n=0
136 Chapter 11. Waves and Hyperbolic Equations
The coefficients an and bn can be found from the initial conditions (36b):
Q ∞
u(x, 0) = δ(x − x0 ) = an X̃n (x) ,
C n=0
Q ∞
ut (x, 0) = b δ(x − x0 ) = bn ωn X̃n (x) .
C n=0
The last two equalities give a series expansion of the functions u and ut
with respect to the orthonormal system of functions X̃n (x), with coefficients
Q Q
an = δ(x − x0 ), X̃n (x) = X̃n (x0 ) ,
C C
Q Q
bn ωn = b δ(x − x0 ), X̃n (x) = b X̃n (x0 ) .
C C
Consequently, the sought function u(x, t) is given by
∞
Q b
u(x, t) = cos ωn t + sin ωn t X̃n (x0 )X̃n (x) .
C n=0 ωn
Recalling that
b
cos ωn t + sin ωn t = An sin(ωn t − ϕn ) ,
ωn
where
b ωn2 + b2
tan ϕn = , An = ,
ωn ωn
we get another representation of the solution:
Q
∞
u(x, t) = An sin(ωn t − ϕn )X̃n (x0 )X̃n (x) .
C n=0
Finally, using formula (34), we can return from the function u(x, t) to the
voltage function v(x, t) to obtain
2Q
∞
−μt π(2n + 1) π(2n + 1)
v(x, t) = e An sin(ωn t − ϕn ) sin x0 sin x.
lC n=0 2l 2l
always make a tacit assumption that all quantities under consideration are
nondimensional, but in physics, the choice of which quantities to express as
nondimensional depends on the nature of the given problem and should be
carefully considered. In our case, first of all, notice that without any loss of
generality it is possible to take l = 1 and a = 1. This choice of spatial and
temporal scale is adequate and most suitable for our analysis. Thus b becomes
the only nondimensional parameter of the problem; any change in its value
will affect the properties of the solutions. Moreover, to simplify our analysis
we shall suppress the inessential factor in front of the sum. The damping
effect caused by the exponential factor e−μt will be easy to reestablish at the
very end. For the sake of concreteness, we shall also assume that x0 = 1/2;
in other words, the concentrated charge has been placed in the middle of the
transmission line.
So at this point, all that remains is a discussion of the behavior of the
auxiliary function
N
π π
U (x, t) = An sin(ωn t − ϕn ) sin (2n + 1) sin (2n + 1)x , (40)
n=0
4 2
FIGURE 11.11.10
The auxiliary function U (x, t) for N = 50, b = 1, and t = 0.3.
A further analysis shows that the small “splashes” have no physical mean-
ing and are only an artifact of the cutoff procedure for the infinite Fourier
series (i.e., its replacement by a finite sum). Indeed, the scale of the “splashes”
is close to the period of the first neglected eigenfunction, which is of magni-
tude 4/103 ≈ 0.039. Such undesirable artifacts can reduce the effectiveness of
the simple cutoff procedures for Fourier series, but can be repaired by the use
of different summation methods; some of these were discussed in Volume 1.
FIGURE 11.11.11
The result of Cesàro summation of the sum representing the aux-
iliary function U (x, t) for t = 0.3, b = 1, and N = 50.
As it turns out, our series is divergent, and here we shall use the Cesàro
method, which was introduced in Chap. 8, Volume 1. This method allows the
summation of some divergent (in the classical sense) series and the acceler-
ation of the convergence of some already convergent series; for example, the
11.11. Telegrapher’s Equation 139
FIGURE 11.11.12
The result of Cesàro summation of the sum representing the aux-
iliary function U (x, t) for t = 1.3, b = 1, and N = 50.
FIGURE 11.11.13
The time dependence of the auxiliary function U (0.8, t) obtained
via the Cesàro summation method applied to the sum containing
100 terms; here, the distortion parameter b is equal to zero.
reflects from the left (grounded) boundary, but the polarity does not change
when the signal is reflected from the right (insulated) boundary.
Finally, Fig. 11.11.14 shows the analogous picture in the realistic case of
a distortion parameter b = 1. In this case, dispersion is present but does not
destroy the pulses themselves; it just adds trailing “echoes” to them, such as
were also present in Figs. 11.11.11 and 11.11.12.
FIGURE 11.11.14
The time dependence of the auxiliary function U (0.8, t) obtained
via the Cesàro summation method applied to the sum containing
100 terms; here, the distortion parameter b is equal to 1.
11.12. Exercises 141
11.12 Exercises
1. Recall that the wave packets (11.1.10) possess (11.3.3) as an invariant.
Prove that they also have an infinite set of other invariants of the form
S = u∗ (x, t)v(x, t) dx = const, (1)
2. The equation
∂ 2u 2
2
2∂ u
+ ω0 u = a (2)
∂t2 ∂x2
arises in various physical and engineering applications. For large-scale
waves (ak ω0 ), it is close to the harmonic oscillator equation, while
for small-scale waves (ak ω0 ), it is close to the wave equation. Find
the phase and group velocities of the harmonic wave satisfying this
equation.
Linear partial differential equations discussed in Part III often offer only
a very simplified description of physical phenomena. To get a deeper under-
standing of some of them, it is necessary to move beyond the linear “universe”
and consider nonlinear models, which in the case of continuous media, means
nonlinear partial differential equations. Even today, their theory is far from
complete and is the subject of intense study. On closer inspection, almost all
physical phenomena in continuous media—from growing molecular interfaces
at atomic scales to the structure of the distribution of matter in the universe
at intergalactic scales—are nonlinear. The variety of nonlinear physical phe-
nomena necessitates the use of various mathematical models and techniques
to study them. In this part we shall restrict our attention to nonlinear waves
of hydrodynamic type in media with weak or no dispersion. Since weak dis-
persion has little influence on the development of many nonlinear effects, we
shall have a chance to observe typical behavior of these systems in strongly
nonlinear regimes. The basic features of strongly nonlinear fields and waves
are already evident in solutions of first-order nonlinear partial differential
equations, and we take them as our starting point.
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 145
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 4,
© Springer Science+Business Media New York 2013
146 Chapter 12. First-Order Nonlinear PDEs
Varying y, we shall obtain the laws of motion of all the particles in the
flow. Note that in addition to the time variable t, there appears here another
variable, y, which represents the particle coordinate at the initial time instant
t = 0. This coordinate, rigidly connected to each particle of the flow, is called
the Lagrangian coordinate of the particle.
On the other hand, in many physical applications, the observer is more
likely to describe the flow by measuring its velocity at a certain selected
12.1. Riemann Equation 147
point in space with a given coordinate, say x, which we shall call the Eule-
rian coordinate. The transformation of Lagrangian coordinates into Eulerian
coordinates is described by the equality
x = X(y, t) . (4)
In the case of uniform particle motion discussed here,
x = y + v0 (y)t . (5)
Suppose that the particle velocity field v(x, t) is known as a function of
the Eulerian coordinate x and the time t. If in addition, we know the rela-
tionship (4) which determines the transformation of Lagrangian coordinates
into Eulerian coordinates, then we can express the velocity field in terms of
the Lagrangian coordinates via the formula
V (y, t) = v(X(y, t), t) . (6)
In what follows, the fields describing the behavior of particles in the La-
grangian coordinate system will be called Lagrangian fields , and the fields
describing the motion of particles in the Eulerian coordinate system will be
called Eulerian fields. So, v(x, t) is the Eulerian field of the particles’ ve-
locity, but X(y, t) is the Lagrangian field of the Eulerian coordinates of the
particles. Since the motion of particles is assumed to be uniform, the velocity
V (y, t) of a particle with a given Lagrangian coordinate y does not depend
on the time, and is thus described by the simple differential equation
dV
= 0, (7)
dt
while its coordinate satisfies an equally simple equation
dX
=V . (8)
dt
Equations (7) and (8) are immediately recognizable as characteristic equa-
tions for the first-order partial differential equation (2), and solutions of the
Riemann equation can be obtained from solutions of the equations (7) and
(8) as soon as we find the inverse function to the function (4),
y = y(x, t) ,
displaying the Lagrangian coordinates in terms of the Eulerian coordinates.
In this case, the Lagrangian laws of motion (3) yield the following solution
of the Riemann equation:
v(x, t) = V (y(x, t), t) = v0 (y(x, t)) . (9)
148 Chapter 12. First-Order Nonlinear PDEs
and substituting the right-hand side of the resulting expression into (9), we
arrive at the solution of the Riemann equation in an implicit form,
On the other hand, the second equality in (10) gives the solution in the form
x − y(x, t)
v(x, t) = . (12)
t
Figure 12.1.1 shows a plot of the solution of the Riemann equation in the
case that the initial profile (dashed curve) of the velocity field is represented
by a Gaussian function
x2
v0 (x) = V0 exp − 2 . (14)
2
FIGURE 12.1.1
A solution of the Riemann equation for a Gaussian initial condition.
than the particles to the right, and the former catch up with the latter in the
course of time. By contrast, the decreasing slope of the left portion of the
velocity profile v(x, t) corresponds to the rarefaction in the particle density.
Quantitatively, the magnitude of the compression and rarefaction effects is
measured by the Jacobian of the transformation that maps Lagrangian to
Eulerian coordinates,
∂X(y, t)
J(y, t) = . (16)
∂y
For a uniform flow of particles (5), the Jacobian is given by
J(y, t) = 1 + v0 (y)t . (17)
For a given Lagrangian coordinate y, the larger the Jacobian is, the larger
are the rarefaction effects seen in a neighborhood of y. For this reason, the
field J(y, t) is often called the divergence field of the particle flow. A graph of
the flow X(y, t) and its divergence field corresponding to the flow of particles
with Eulerian velocity field satisfying the Riemann equation (2), with initial
condition (14), is shown in Fig. 12.1.2.
The field J(y, t) described by (16) is a Lagrangian divergence field. The
corresponding Eulerian divergence field is obviously
j(x, t) = J(y(x, t), t) ⇐⇒ J(y, t) = j(X(y, t), t) . (18)
If the mapping y(x, t) of Eulerian coordinates into Lagrangian coordinates
is known, then the divergence field can be determined via the following,
geometrically obvious, relationship:
∂y(x, t) 1
= . (19)
∂x j(x, t)
FIGURE 12.1.2
Graph of the flow X(y, t) (top), and its divergence field (bottom),
corresponding to a flow of uniformly moving particles, with Eule-
rian velocity field satisfying the Riemann equation (2), with initial
condition (14). For values of J(y, t) greater (resp. smaller) than 1,
the particle density decreases (resp. increases).
FIGURE 12.2.1
The Eulerian field m(x∗ , t) representing the mass of particles of the
flow to the left of the point x∗ .
and it does not vary in time. In other words, the Lagrangian field of the mass
on the left satisfies the equation
dM
= 0.
dt
The equivalent equation for the corresponding Eulerian field m(x, t) is
∂m ∂m
+v = 0. (2)
∂t ∂x
Now we are ready to define the particle density. In the 1-D case under
consideration, the density field is simply the derivative of the mass on the
left with respect to x,
∂m(x, t)
ρ(x, t) = . (3)
∂x
12.2. Continuity Equation 153
∂ρ ∂
+ (vρ) = 0 . (4)
∂t ∂x
Remark 1. Note that in contrast to the Riemann equation for the velocity
field of uniformly moving particles, our derivation of the continuity equation
never made use of the fact that the motion was uniform. So, the continuity
equation expresses a universal law, valid for any particle motion.
∂y(x, t)
ρ(x, t) = ρ0 (y(x, t)) , (6)
∂x
or taking into account (12.1.19),
The last relationship has an obvious geometric meaning: the flow’s density
at a given point is equal to the initial density in the vicinity of the particle’s
initial location at t = 0, divided by the degree of compression (or rarefaction)
of the particle flow.
Separately, consider the density field for the flow of uniformly moving
particles in which the velocity field v(x, t) satisfies the Riemann equation,
and y(x, t) is expressed by the relation (12.1.10). In this case, in view of (6)
and (12.1.10), the flow’s density is expressed in terms of the solution of the
Riemann equation as follows:
∂v(x, t)
ρ(x, t) = ρ0 (x − v(x, t)t) 1 − t . (8)
∂x
154 Chapter 12. First-Order Nonlinear PDEs
which directly demonstrates the close relationship between the flow’s density
and the steepening of its velocity profile.
FIGURE 12.2.2
Graphs of the evolving density in the flow of uniformly moving
particles for a Gaussian initial velocity field (12.1.14) and constant
initial density field ρ(x, t = 0) = ρ0 = const.
only some of them, such as the energy and momentum conservation laws, have
a clearcut physical significance; in the present subsection we shall briefly dis-
cuss the latter. Recall that the cumulative momentum function at time t is
determined by the formula
x
p(x, t) = v(x, t)ρ(x, t) dx .
−∞
Substituting here the expressions (12.1.9) and (6) for the velocity and density
in the flow of uniformly moving particles and then passing to the Lagrangian
coordinates in the integral, we obtain that the cumulative momentum is given
by
x y(x,t)
∂y(x, t)
p(x, t) = v0 (y(x, t))ρ0 (x(y, t)) dx = v0 (y)ρ0 (y) dy .
−∞ ∂x −∞
∂p(x, t)
g(x, t) = ρ(x, t)v(x, t) =
∂x
satisfies the continuity equation
∂g ∂
+ (vg) = 0 .
∂t ∂x
of the density field. For this purpose, let us substitute in (12) the solution of
the continuity equation (6) to obtain
1
ρ̃(κ, t) = e−iκx ρ0 (y(x, t)) dy(x, t) .
2π
Switching to integration with respect to the Lagrangian coordinates, we fi-
nally arrive at the formula
1
ρ̃(κ, t) = e−iκX(y,t) ρ0 (y) dy . (13)
2π
Similar but slightly more complex calculations yield a formula for the Fourier
image of the velocity field,
i −ikX(y,t)
ṽ(κ, t) = e − e−iky dy . (14)
2πkt
Example 1. Harmonic initial velocity field. Formulas (13) and (14) are
remarkable because they express Fourier images of implicitly defined (for
example, by equality (12.1.11)) functions ρ(x, t) and v(x, t) through integrals
of explicitly given integrands. This fact gives us an opportunity to find an
explicit expression for the density field ρ(x, t) in the case that the initial
12.2. Continuity Equation 157
velocity field is given by a simple harmonic function and the initial density
field is constant,
In the case under discussion, the mapping of Lagrangian into Eulerian coor-
dinates is given by the formula
Substituting it into the formula (13) for the Fourier image of the density field
gives
ρ0
ρ̃(κ, t) = e−iμz−iμτ sin z dz ,
2πk
where we introduced the nondimensional variables of integration z = ky,
time τ = kat, and spatial frequency μ = κ/k. An application of the formula
(16) gives
ρ0
∞
ρ̃(κ, t) = Jn (−μτ ) e−i(μ−n)z dz .
2πk n=−∞
In view of the distributional formula (3), of Sect. 3.3, Volume 1, we have
1 1
e−i(μ−n)z dz = δ(μ − n) = δ(κ − kn) .
2πk k
So, the generalized Fourier image is given by
∞
ρ̃(κ, t) = ρ0 Jn (−nτ )δ(κ − kn) .
n=−∞
Jn (−w) = (−1)n Jn (w) , J−n (w) = (−1)n Jn (w) ⇒ J−n (−w) = Jn (w), (18)
158 Chapter 12. First-Order Nonlinear PDEs
FIGURE 12.2.3
A comparison of the sum of the first two and the first eleven terms
of the series (19) with the exact expression for the density field
obtained parametrically using the formula (10). Time τ = akt equals
0.7, the initial velocity is assumed to be harmonic (15), and the
initial density, constant.
we arrive at the formula for the density field expressed as a Fourier series:
∞
ρ(x, t) = ρ0 + 2ρ0 (−1)n Jn (nτ ) cos(kx) . (19)
n=1
A comparison of the sum of the first few terms of (19) with the exact form
of the density field obtained parametrically using formula (10) is given in
Fig. 12.2.3. It shows that even for moderate values of τ , the agreement is
quite good.
FIGURE 12.3.1
A schematic illustration of a moving forest fire front.
Furthermore, observe that the coordinate Z(t) can be expressed via the
previously introduced fire front function (1) as follows:
Z(t) = h(X(t), t) . (3)
160 Chapter 12. First-Order Nonlinear PDEs
FIGURE 12.3.2
A geometric illustration of the validity of equation (7).
12.3. Interface Growth Equation 161
For a planar wave propagating along the z-axis, this equation has a very
simple solution, h = ct. If we are interested only in the form of the wave
front but not in its absolute position, then we can introduce the “comoving”
coordinate system
w(x, t) = h(x, t) − ct . (8)
The new function w(x, t) satisfies a more elegant equation,
2
∂w c ∂w
= . (9)
∂t 2 ∂x
and the topography of the growing snow cover does not change in time; only
its elevation increases uniformly in time.
D(θ) = cδ(θ) .
Let us note that the limits in the integral (15) have to take into account
effects of shadowing of the incident stream of particles by the growing inter-
face h(x, t) and that they depend on the concrete topography of the interface
h(x, t). As a concrete example, consider a segment of the interface that forms
an angle θ > 0 with the z-axis. In this case, clearly (see Fig. 12.3.3),
π π
min θ− = θ − , max θ+ = .
2 2
The asymmetry of the minimal and maximal angles for the directional di-
agram is due to the fact that the local geometry (θ > 0) of the interface
164 Chapter 12. First-Order Nonlinear PDEs
FIGURE 12.3.3
A schematic illustration of the process of determining limits in the
integral (15).
Obviously, the function c(θ) is even. This means that the formula (16) re-
mains valid for both θ > 0 and θ < 0.
Now, for the system of equations (11) and (17), the corresponding charac-
teristic equations are
dX̃ dU dH
= C(U ) , = 0, = Λ(U ) , (19)
dt dt dt
where the fields along the characteristic curves are denoted by capital letters,
and
2 d Φ(u)
Λ(u) = Φ(u) − uC(u) = −u . (20)
du u
The auxiliary function X̃(y, t) introduced above should not be confused with
the function X(t), which must satisfy the first equation in (2),
dX u c(u)
= V (u) , V (u) = c sin θ = √ , (21)
dt 1 + u2
and which, in the case of optical wave fronts, has a clearcut geometric mean-
ing: it is the horizontal coordinate of the ray, always perpendicular to the
wave front propagating in an anisotropic (if c = c(u)) medium.
To distinguish X(t) from X̃(y, t), we shall call the latter function (to-
gether with Z̃ = h(X̃, t)) the isocline trajectory of the growing interface
h(x, t).
The solutions of (19) are of the form
U (y, t) = u0 (y) , X̃(y, t) = y + C(u0 (y))t ,
(22)
H(y, t) = h0 (y) + Λ(u0 (y))t .
In particular, it follows that the isoclines are straight lines, which is not
always true for rays, which in an anisotropic medium could be curved.
As in the case of the Riemann equation, the sought fields h(x, t) and
u(x, t) are obtained from (22) by substituting in H(y, t) and U (y, t) the func-
tion y = ỹ(x, t), which is the inverse function to x = X̃(y, t).
In the remainder of this section we shall provide two concrete examples
of the interface evolution discussed above.
This means in particular that rays and isoclines coincide and are represented
by straight lines perpendicular to the wave fronts. In this case, the interface
evolution can be described parametrically as follows:
u0 (y) ct
x=y+ ct , h = h0 (y) + . (23)
1 + u20 (y) 1 + u20 (y)
FIGURE 12.3.4
Time evolution of the interface described by the field h(x, t) (24)
in the case of positive growth rate (c > 0). Physically, it may be
interpreted as evolution of an optical wave front that started out
as a sine curve. The medium is assumed to be isotropic. In the
course of time, the crests become flatter, but the troughs become
sharper.
Now consider the initial sine profile
h0 (x) = h0 cos(kx) ⇒ u0 (x) = h0 k sin(kx).
In this case, we obtain the following parametric representation for h(x, t):
ετ sin μ τ
z =μ+ , η = ε cos μ + , (24)
1 + ε2 sin2 μ 1 + ε2 sin2 μ
where the nondimensional variables
kx = z , ky = μ , ckt = τ , kh = η , (25)
have been introduced, with the parameter
ε = kh0 . (26)
Graphs of the evolving interface h(x, t) are shown in Fig. 12.3.4 for different
τ and ε = 1/2.
12.4. Exercises 167
FIGURE 12.3.5
Time evolution of a melting interface h(x, t) from Example 4. In
contrast to the growing interface from Example 3 (Fig. 12.3.4), in
the course of time, the troughs of this interface become smoother
while the crests become sharper. The reader may have observed
this phenomenon while tasting the ice cubes melting in a glass of
Coke.
In this case, the profiles of the evolving interface are obtained by changing
in (12.3.3) the signs in front of τ from plus to minus. The corresponding
graphs are shown in Fig. 12.3.5.
12.4 Exercises
1. In the 1980s, Ya.B. Zeldovich, one of the fathers of the Russian hydro-
gen bomb, developed the theory of nonlinear gravitational instability,
which describes the evolution of the large-scale mass distribution in
the universe (at the scale of galaxy clusters). In its simplified form,
his theory reduces to an analysis of solutions of the Riemann equation
(12.1.2), and the continuity equation (12.2.4), with initial conditions
∂u ∂u ∂ ∂
+u = 0, + (u) = 0 ,
∂t ∂x ∂t ∂x
u(x, t = 0) = u0 (x) , (x, t = 0) = 0 (x) .
∂v ∂v 1
+v + v = 0, v(x, t = 0) = v0 (x) ,
∂t ∂x τ
where τ is a constant representing the characteristic velocity “dissipa-
tion” time. Find the evolution of the velocity field v(x, t) and determine
its asymptotic behavior as t → ∞.
3. Prove the formula (12.2.14) for the Fourier image of the solution of the
Riemann equation. Assume that v(x, t) and y(x, t) are smooth functions
of x such that v(x, t) decays to zero rapidly as x → ±∞, while y(x, t)
is strictly increasing and maps the x-axis onto the whole y-axis.
∂v ∂v e ∂ρ ∂
+v =− E, + (vρ) = 0 ,
∂t ∂x m ∂t ∂t (2)
v(x, t = 0) = v0 (x) , ρ(x, t = 0) = ρ0 (x) ,
describing the electron velocity v(x, t) and their density ρ(x, t), and the
equation
∂E
= −4πe(ρ − ρ0 )
∂x
12.4. Exercises 169
describing the longitudinal electric field E(x, t). Here e and m are re-
spectively the electron charge and mass, and ρ0 is the density of ions
that are assumed to be at rest. Solve the above equations by the method
of characteristics and discuss their solutions. Assume that as x → −∞,
the electron velocity field decays to zero, and so does the cumulative
mass function of the electrons,
x
m(x, t) = [ρ(x , t) − ρ0 ] dx .
−∞
D(θ) = c cos2 θ .
Study the evolution of the interface h(x, t). While constructing the
corresponding equation, use the small angle approximation and assume
that the inequalities |θ| 1 and |u| 1 are satisfied. Conduct a
detailed investigation of the case that the initial interface profile is given
by h0 (x) = h cos(kx) (kh 1).
170 Chapter 12. First-Order Nonlinear PDEs
9. Investigate the interfacial growth in the case that the growth velocity
depends on the angle of incidence θ as follows:
c(θ) = c cos2 θ .
10. Assuming the cosine initial condition (4), expand the solution of the
previous exercise in a Fourier series. For τ = 10, compare the graph
in Fig. 2 shown in the Appendix A: Answers and Solutions with the
graph of the partial sum of the first ten terms of the Fourier series.
Chapter 13
Generalized Solutions of
First-Order Nonlinear PDEs
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 171
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 5,
© Springer Science+Business Media New York 2013
172 Chapter 13. Generalized Solutions of Nonlinear PDEs
The lower limit in the integral has been deliberately omitted, indicating the
presence of an arbitrary constant; physical potentials are always defined only
up to a constant. The potential of the velocity field satisfies the following
nonlinear equation:
2
∂s 1 ∂s
+ = 0, s(x, t = 0) = s0 (x) . (4)
∂t 2 ∂x
The reader can verify the validity of the above equation by differentiating it
with respect to x to obtain again the Riemann equation (1) for the velocity
field.
solutions of the interface growth equation written in the small angle approx-
imation. Recall (see Chap. 12) that the forest fire boundary line h(x, t), or
the boundary of a 2-D wave front, propagating with the velocity c perpen-
dicularly to the interface h(x, t) satisfies, in the small angle approximation,
the equation
2
∂h c ∂h
= , h(x, t = 0) = h0 (x) , (5)
∂t 2 ∂x
∂u ∂u
+ cu = 0, u(x, t = 0) = u0 (x) . (7)
∂t ∂x
Moreover, if one introduces the change of variables
v = cu , s = −ch, (8)
∂P 1 ∂P ∂P n
− − βP + P =0 (9)
∂r c ∂t ∂t 2r
containing the parameter n. For n = 2, the equation (9) accurately describes
the propagation of spherical waves (cylindrical waves for n = 1 and flat waves
for n = 0). The constant c stands for the sound velocity in the medium, and
β is a parameter that quantifies the nonlinearity of the medium.
Usually, one reduces equation (9) to a more convenient form by introduc-
ing the local time coordinate
r
θ =t− ,
c
174 Chapter 13. Generalized Solutions of Nonlinear PDEs
which “delays” the wave by the time needed for the wave to propagate from
the origin r = 0 to the point r under consideration. This substitution elimi-
nates one term of the equation (9), so that we arrive at the equation
∂P ∂P n
− βP + P = 0.
∂r ∂θ 2r
The last term takes into account the wave attenuation due to its geometric
divergence, but it can also be eliminated via a change of variables. Indeed,
for example, in the case of a spherical wave (n = 2), one can introduce the
new variables
r r
p= βP , z = r0 ln ,
r0 r0
which transform the equation of nonlinear acoustics into the canonical Rie-
mann equation for the field p(z, θ),
∂p ∂p
+p = 0. (10)
∂z ∂θ
Usually, r0 is taken to be the radius of the spherical source radiating the
acoustic wave. Then the boundary condition in equation (10), which describes
vibrations of the surface of the radiator, is
Recall that this solution has a transparent geometric meaning: the flow
velocity at time t at the Eulerian coordinate x is equal to the velocity of a
uniformly moving particle that finds itself at x at time t. The time interval
(0, tn ) will be called the single stream motion interval.
FIGURE 13.2.1
Lagrangian-to-Eulerian coordinate mappings for the flow of uni-
formly moving particles at different instants of nondimensional
time τ = 1, 2, 4. The thick-line interval on the vertical x-axis in-
dicates the presence of a multistream motion there, at τ = 4.
176 Chapter 13. Generalized Solutions of Nonlinear PDEs
FIGURE 13.2.2
Eulerian-to-Lagrangian mapping at the nondimensional time τ = 4.
The thick-line interval on the x-axis indicates points where the mul-
tistream regime is present. Three branches of the function y(x, t)
are visible, and are numbered in the order of increase of the cor-
responding Lagrangian coordinates.
FIGURE 13.2.3
The velocity field v(x, t) at time τ = 4. The dashed line indicates the
initial Gaussian velocity profile (12.1.14) on which three particles
with different Lagrangian coordinates y1 , y2 , y3 are marked. They
all end up, at time τ = 4, at the same point in space. It is clear
that particles y1 , y2 , initially positioned to the left have a greater
velocity than y3 , and in the course of time, they catch up with the
particle with Lagrangian coordinate y3 . The points x1 and x2 bound
the multistream segment under consideration.
The expression contains the derivative of the mapping y = y(x, t), and is best
studied by investigating a “nicer,” everywhere differentiable, inverse function
x = X(y, t), see (1), and its derivative
13.2. Multistream Solutions 179
∂X(y, t)
J(y, t) = = 1 + v0 (y) t (11)
∂y
with respect to y. At the turnover time tn , the mapping X(y, t) ceases to be
strictly monotone, and the minimum value of its derivative (11), attained at
a certain point y∗ , becomes zero:
J(y∗ , tn )) = 1 + v0 (y∗ ) tn = 0 .
Since the point y∗ is also the point minimizing values of the function J(y, tn ),
not only does the function J(y, tn ) itself vanish there, but its derivative van-
ishes as well,
∂J(y, t)
= v0 (y∗ ) tn = 0 ⇒ v0 (y∗ ) = 0 .
∂y y=y∗
Consequently, the mapping X(y, t), see (1), has, in a vicinity of the point y∗ ,
cubic asymptotics,
v0 (y∗ )
x = X(y, t) ∼ x∗ + b(y − y∗ )3 , b=− , (12)
v0 (y∗ )
where the turnover point x∗ is the Eulerian coordinate of the singular point
y∗ discussed above.
In view of (12), the function y(x, t), inverse to X(y, t), and the velocity
field v(x, t), see (5), have, in the vicinity of the turnover point x∗ , the following
asymptotics:
3 |x − x∗ |
y(x, t) ∼ y∗ + sign (x − x∗ ) ⇒
b
3 |x − x∗ |
v(x, t) ∼ v∗ − sign (x − x∗ ) ,
d
x − x∗ v (y∗ )
v∗ = . d = 0 .
t [v0 (y∗ )]4
Consequently, at the turnover time, the derivative q(x, t), see (10), of the
field v(x, t) has an infinite singularity of the following type:
1 1
q(x, t) ∼ − √ .
3 d |x − x∗ |2/3
3
Recall that the density field ρ(x, t), see (8), is proportional to the deriva-
tive of the mapping y(x, t) with respect to x. It is this derivative that deter-
mines the character of the gradient catastrophe. In the language of particle
flows, the gradient catastrophe is caused by an infinite compression of the flow
at turnover time. The density at this point (for ρ0 (y∗ ) = 0) becomes infinite.
The nature of the density collapse as the time approaches the turnover time
tn is well illustrated in Fig. 12.2.2,√where the last time considered, τ = 1.5,
is close to the turnover time τ = e ≈ 1.65. The sharp peak in the density
field at that time does not fit in the figure.
The validity of the above formula can be verified using standard properties of
the Dirac delta distributions and in particular, the formula for a nonmonotone
change of variables; see Volume 1.
Fourier transform (4) is exactly the sum (1) of the multistream densities in
the flow of uniformly moving particles.
Although construction of the density field via the inverse Fourier trans-
form may seem somewhat artificial, the technique permits us to prove several
integral properties of the total density (1). For example, the theory of the
Fourier transform gives us the following equality:
ρ(x, t) dx = 2π ρ̃(κ = 0, t) .
For κ = 0, this fact and equality (3) imply that the total density (1) satisfies
the integral mass conservation law
ρ(x, t) dx = ρ0 (y) dy = const . (5)
n
n
i−1
v± (x, t) = (−1) vi (x, t) = (−1)i−1 v0 (yi (x, t)) . (6)
i=1 i=1
In contrast to the total density fields (1) and (4), the above function (6)
does not have any particular physical interpretation when applied to parti-
cle flows. This does not preclude the possibility that such an interpretation
might be found in the future for other physical, economic, etc., phenomena
described by the Riemann equation. Notice that the field (6) has an integral
invariant. Indeed,
v± (x, t) dx = v0 (y) dy = const . (7)
FIGURE 13.3.1
The field v± (x, t), found by summation of the first 50 terms of the
series (9), at time τ = kat = 2 (which is larger than the time
τn = 1 when the multistream regime appears for the first time).
The dashed line shows the multistream velocity field for the flow
of uniformly moving particles. The vertical lines mark one of the
multistream intervals. The pluses and minuses indicate which sign
was assigned for which stream. The sharp cusps reflect the lack of
differentiability for branches of the multistream velocity field at
the points corresponding to the onset of different regimes.
which was found in one of the exercises in Chap. 12 by taking the inverse
Fourier transform of the Fourier image (12.2.14) of the velocity field v± (x, t).
A comparison of the graph of the partial sum of the series with the corre-
sponding multistream solution shows the equivalence of the inverse Fourier
transform of the Fourier image (12.2.14) and the signed sum (6).
The field ρ(x − ut, t) is the original “cold” density field (2) shifted along the
x-axis by ut. The subscript ε in (14) points to the dependence of the density
field ρε (x, t) on the “temperature” ε.
13.3. Summing over Streams 185
The formula (14) means that the thermal scattering of particle velocities
leads to the spatial averaging of the hydrodynamic density field ρ(x, t). Such
averaging eliminates singularities of the original density field. We shall illus-
trate this fact with the example of the uniform original density ρ0 (x) = ρ0 =
const and the sinusoidal initial velocity field (8).
Substituting the Fourier expansion (12.2.19) for ρ(x, t) into (14), we ob-
tain
∞
ρ(x, t) = ρ0 + 2ρ0 (−1)n Jn (nτ )ϕn (x, t) , (15)
n=1
where
ϕn (x, t) = cos(knx − kunt) f (u) du .
∞
δ
ρ(x, t) = ρ0 + 2ρ0 (−1) Jn (nτ ) exp − τ 2 n2
n
. (16)
1
2
Here and in (15), we used the nondimensional time τ = kat as well as the
parameter δ = ε/a2 , which characterizes the relative contribution of thermal
velocity fluctuations to the flow’s behavior. For small δ, for a long stretch of
time, the field ρ(x, t) is formed by hydrodynamic compression and rarefaction
of the particle flow. For large δ, very quickly, namely at the time
2
τ∗ = ,
δ
we observe that the thermal effects gradually eliminate the lack of uniformity
of the density. The remaining small density fluctuations are asymptotically
described by the first two terms of the series (16):
2
ρ(x, t) ∼ ρ0 1 − 2 J1 (τ ) e−δτ /2 cos(kx) .
The graphs of the density (16), with δ = 0.002, are shown in Fig. 13.3.2 at
different moments of time.
186 Chapter 13. Generalized Solutions of Nonlinear PDEs
FIGURE 13.3.2
The density field (16) for δ = 0.002 and the time instants τ = 2
and 3. The dashed line depicts the multistream velocity field at
the time τ = 3. The value of the parameter δ is small, and as a
result, all the structural elements of the “hydrodynamic” density
field are clearly visible. In particular, the multistream intervals and
the characteristic density peaks near the interval boundaries can
be seen. The peaks’ heights diminish with time. The presence of
even weak thermal fluctuations of particle velocities eliminates the
infinite singularities.
∂s ∂s 1
+v = v2, s(x, t = 0) = s0 (x) ,
∂t ∂x 2 (1)
∂v ∂v
+v = 0, v(x, t = 0) = v0 (x) .
∂t ∂x
The corresponding characteristic equations are of the form
dX dV dS 1
=V , = 0, = V2.
dt dt dt 2
These equations have the familiar solutions
The corresponding Eulerian fields s(x, t) and v(x, t) are conveniently ex-
pressed in terms of the mapping y(x, t), inverse to the function x = X(y, t):
s0(x)
s(x,t)
Propagation Direction
FIGURE 13.4.1
The initial location of the fire front s0 (x) (the upper curve) and the
fire front function s(x, t) at a certain time t > 0 (the lower curve).
The two graphs are dilated along the z-axis to emphasize the di-
rection (opposite to the direction of the z-axis) of propagation of
the fire front. The shaded area indicates the part of the forest that
has burned down during the time interval (0, t). The generalized
weak solution of the fire front equation is obtained by rejecting the
dashed portions of the multivalued solutions.
A typical fire front function s(x, t) constructed parametrically with the help
of the corresponding Lagrangian fields
1 2
x = X(y, t) = y + v0 (y) t , s = S(y, t) = s0 (y) + v (y) t , (4)
2 0
FIGURE 13.4.2
Typical sawtooth weak solutions v(x, t), see (6), of the Riemann
equation, and the corresponding piecewise smooth weak solutions
s(x, t), see (5), of the equation (13.1.4).
This can be done, but there is a price to pay for the alternative approach.
To achieve our goal, it will be necessary to expand the number of variables of
the function under investigation. Namely, we shall take our familiar mapping
x = y + v0 (y) t of Lagrangian into Eulerian coordinates, which depends on
two variables y and t, and use it to construct a function of three variables,
R(y; x, t) = X(y, t) − x = v0 (y) t + (y − x) . (7)
Its graph, as a function of the variable y, for a fixed x and t, passes through
the zero level at certain points {yi (x, t)}, which in general, are values of the
multivalued mapping
y = y(x, t) (8)
of Eulerian into Lagrangian coordinates.
At this point, let us introduce another auxiliary function,
y y
G(y; x, t) = R(z, x, t) dz = (X(z, t) − x) dz , (9)
FIGURE 13.4.3
Search of the coordinate of the absolute minimum of the auxiliary
function G(y; x, t) via a geometric algorithm. Raising the parabola
P, centered at the point x, we find the coordinate yw (x, t) where the
parabola first touches the graph of the initial potential s0 (y).
the curve s0 (y) (hence the name osculating parabola), keeping track of the
coordinate yw (x, t) of the point of tangency.
FIGURE 13.4.4
Top: The initial potential s0 (y) and two critical parabolas. Bottom:
Their centers determine locations of the discontinuities (shocks) of
the weak solutions of the Riemann equation.
y2 x2
G(y; x, t) = s0 (y) + − xy + . (15)
2 2
As before, we are looking for the coordinate of the global minimum of
G(y; x, t) as a function of y. The first, immediate, observation is that the
last term in (15) plays no role in finding the desired location. Thus it can be
dropped. Moreover, if we introduce the new notation
y2
ϕ(y, t) = + s0 (y) t , (16)
2
then our problem is reduced to finding the coordinate of the global minimum
of the function
ϕ(y, t) − xy . (17)
FIGURE 13.4.5
The function ϕ(y, t) and its convex envelope ϕ̄(y, t). Note that the
convex envelope is always located above any of the tangent lines
to ϕ(y, t).
194 Chapter 13. Generalized Solutions of Nonlinear PDEs
xy + h (18)
against the graph of the function ϕ(y, t), see (16), thus selecting a particular
value of h. The coordinate of the tangency point of the line (18) and the
function ϕ(y, t) is the desired coordinate yw (x, t). Changing the slope x of
the line (18), we can find the values of yw (x, t) for all x.
The key observation is that the outcome of the above-described algorithm
will not change if the function ϕ(y, t) is replaced by its convex envelope ϕ̄(y, t),
and we search for the coordinate of the global minimum of the function
ϕ̄(y, t) − xy (19)
∂ ϕ̄(y, t)
Xw (y, t) = . (20)
∂y
Integrating this equation term by term over the whole x-axis and assuming
that v(x, t) decays rapidly to zero as x → ±∞, we arrive at the equality
d
v(x, t) dx = 0 .
dt
FIGURE 13.4.6
Geometric interpretation of the invariants (21) and (23) in the case
of a multistream field v(x, t). The sum of the areas of the shaded
domain is equal to the invariant I.
196 Chapter 13. Generalized Solutions of Nonlinear PDEs
Of course, there are infinitely many integral invariants for the Riemann
equation, but in nonlinear acoustic applications, the weak solutions vw (x, t)
are physically meaningful as long as they satisfy the relation (21). We shall
now construct solutions of this type (Fig. 13.4.6).
For that purpose, let us substitute in (21) the right-hand side of the
second equality in (3),
1
I= [x − y(x, t)] dx . (22)
t
A simple geometric argument shows that our invariant can be expressed via
the mapping x = X(y, t) inverse to y = y(x, t):
1
I= [X(y, t) − y] dy . (23)
t
The last integral makes sense even for the multistream velocity field v(x, t)
of uniformly moving particles. In this case, the right-hand side of (3) is equal
to the area contained between the graph of the multivalued function v(x, t)
and the x-axis. From (23) and from the explicit expression (13.2.1) for X(y, t),
equality (21) follows. This confirms that the invariant (21) is also present in
multistream flows.
Usually, construction of weak solutions of the Riemann equation that
possess the invariant (21) proceeds as follows: In the intervals where the
multistream field v(x, t) is multivalued, we draw vertical segments. Their
position is chosen so that the cut-off areas of the multivalued field v(x, t),
shown in Fig. 13.4.7 (left picture), are equal.
Representation of the invariant (21) in the integral form (23) implies that
the above rule of equal areas will be satisfied if one uses Maxwell’s rule,
which is well known in mechanics. According to this principle, one has to
replace the nonmonotone mapping X(y, t) = y + v0 (y) t by the piecewise
constant mapping Xw (y, t), placing horizontal segments so that the cut-off
pieces of nonmonotonicity shown in Fig. 13.4.7 (right picture) have equal
areas. Substituting then the inverse mapping yw (x, t) into (12), we will get
the required weak solution that satisfies the invariance condition (21).
FIGURE 13.4.7
Illustration of the rule of equal areas (left) and the equivalent
Maxwell’s rule (right). The former states that to obtain a weak
solution (thick line) of the Riemann equation, it is necessary to cut
off the multivalued branches of the multistream field while keeping
the left and right cut-off areas equal. The latter prescribes how
to remove the intervals of nonmonotonicity of the Lagrangian-to-
Eulerian mapping x = y + v0 (y) t. Again the upper and lower cut-off
areas must be equal. The dashed lines indicate the initial condi-
tions.
yk+
[X(z, t) − x∗k ] dy = 0 , (24)
yk−
where [yk− , yk+ ] is the interval of the y-axis that is a projection of the horizontal
segment described above.
In this subsection we shall establish the equivalence of Maxwell’s rule
with the Oleinik–Lax global minimum principle described in Sect. 13.4.2.
For this purpose, let us observe that if the mapping yw (x, t) jumps from
yk− to yk+ as we pass through the point x∗k from left to right, then for x = x∗k ,
the function G(y; x∗k , t) has two identical global minima attained at yk− and
yk+ . In other words, at every discontinuity point x∗k of the weak solution of
the Riemann equation, we have
G(yk− , x∗k , t) = G(yk+ , x∗k , t) . (25)
In view of the definition (9) of the function G(y; x∗k , t), the above condition
of equality of the two minima can be written in the form
y+
k
∗ − ∗
G(yk , xk , t) − G(yk , xk , t) =
+
[X(z, t) − x∗k ] dy = 0 ,
yk−
x∗ (t) = xc + v ∗ t , (1)
where
1
n
xc = ∗ m k xk (2)
m k=1
is the initial center of mass, and
p∗
v∗ = (3)
m∗
is the velocity calculated from the law of conservation of momentum with
n
n
m∗ = mk , p∗ = mk vk . (4)
k=1 k=1
Obviously, once all the particles stick together, the position of the thus
formed macroparticle coincides with its center of mass, and the macroparticle
itself moves according to (1). Observe that to arrive at the solution to the
above elementary problem it was not necessary to know the entire history
of individual particles. In particular, information about where and in what
order particles stuck together was not needed.
13.5. E–Rykov–Sinai Principle 199
1
In fact, theoretical traffic studies often employ this model.
200 Chapter 13. Generalized Solutions of Nonlinear PDEs
FIGURE 13.5.1
The Lagrangian-to-Eulerian mapping X(y, t) of the original and cur-
rent density of the particles in a flow of inelastically colliding par-
ticles. The area of the shaded domain is equal to the mass m∗ of
the macroparticle located at the point x∗ . The discrete component
of the macroparticle density is schematically pictured as the thick
arrow on the left.
Note that the single but crucial difference between the function S(y; x, t)
and the function G(y; x, t), see (13.4.9), appearing in the absolute minimum
principle consists in the presence of the initial density ρ0 (x) under the integral
(9). The two functions have a common feature: the coordinates of the extrema
in both cases are equal to the Lagrangian coordinate yi (x, t) of noninteracting
particles that are located at x at time t.
The ERS principle can be formulated as follows: the value yw (x, t) of the
mapping appearing in the weak solution
x − yw (x, t)
vw (x, t) = (10)
t
of the Riemann equation is equal to the coordinate of the global minimum in
the variable y of the function S(y; x, t) (9).
is the cumulative mass of the particles located to the left of y. Formula (11)
also contains the cumulative momentum
y
P (y) = ρ0 (y) v0 (y) dy (13)
−∞
as well as the initial center of mass of all the particles located to the left of
y: y
N (y)
xc (y) = , N (y) = z ρ0 (z) dz .
M (y) −∞
Furthermore, observe that if for an arbitrary fixed time t and jump from
x∗ (t) − 0 to x∗ (t) + 0 of the parameter x, the coordinate y(x, t) of the global
minimum of S(y; x, t), see (9), jumps from y − (t) to y + (t) (y + > y − ), then
S(y; x∗ , t) has, because of its continuity with respect to x, two identical global
13.5. E–Rykov–Sinai Principle 203
minima, located at y − (t) and y + (t). This, in turn, implies the validity of the
following equality:
y + (t)
S(y + ; x∗ , t) − S(y − ; x∗ , t) = [x∗ (t) − y − v0 (y) t] ρ0 (y) dy = 0 . (14)
y − (t)
in (15) is related to the macroparticle’s position x∗ (t), see (15), through the
obvious kinematic relation (5).
Let us demonstrate this by differentiating the jump condition (14) with
respect to t:
y+ (t)
d
[x∗ (t) − y − v0 (y) t] ρ0 (y) dy = 0 . (18)
dt y− (t)
Note that the derivatives with respect to the integral limits vanish in view
of the obvious equalities
x∗ = X(y + , t) = y + + v0 (y + ) t and x∗ = X(y − , t) = y − + v0 (y − ) t , (19)
which mean that at the boundary of the interval of integration, the integrand
in (18) is equal to zero. Differentiation of the integrand leads to the required
relationship
d ∗ p∗ (t)
v ∗ (t) = x (t) = ∗ ,
dt m (t)
∗ ∗
where p (t) and m (t) are given by the formulas (16).
The second requirement, which we shall call the absorption rule, says that
in the course of time, macroparticles should absorb the surrounding particles
without releasing those that already had been absorbed previously.
Since requirements like the one above play a key role in the theory of weak
solutions of nonlinear partial differential equations, a comment about general
admissibility criteria is in order. They all reflect certain physical realities of
the system under consideration. For example, in studying the multivalued fire
front function h(x, t), it is clear that once the branches of the function fall
behind the front (see Fig. 13.4.1), they will never again influence its future
evolution.
In all of the above applications of weak solutions of the Riemann equation
(fire fronts, nonlinear acoustic waves, flows of inelastically sticking particles),
the admissibility criterion reduces to the inequalities
v − (t) ≥ v ∗ (t) ≥ v + (t) , (20)
where
v − (t) = v0 (y − ) = v(x∗ − 0, t) and v + (t) = v0 (y + ) = v(x∗ + 0, t)
are values of the field vw (x, t) immediately to the left and, respectively, right
of the jump point. In the mathematical literature, such criteria are also called
entropy conditions.2
2
See, for example, J. Smoller, Shock Waves and Reaction–Diffusion Equations,
Springer-Verlag, 1994, Chapter 16.
13.5. E–Rykov–Sinai Principle 205
2x∗ = y + + y − + (v + + v − ) t .
v∗ = U + W , (22)
where
v+ + v−
U= (23)
2
is the average of the velocities of microparticles arriving at the jump from
left and right, and
1 y+ + y−
W = − xc (24)
t 2
is the difference between U and the velocity of the moving jump. We shall
prove that v ∗ , see (22), satisfies the inequalities (20). For this purpose, we
shall evaluate the quantity xc that enters on the right-hand side of (24).
206 Chapter 13. Generalized Solutions of Nonlinear PDEs
FIGURE 13.5.2
The graph of a discontinuous field vw (x, t). Shown are all the veloc-
ities used in the proof of the fact that the ERS principle generates
weak solutions of the Riemann equation satisfying the admissibility
criterion.
Consequently, W in (24) is zero, and the shock moves with velocity U , see
(23), which is the arithmetic mean of the values of the velocity field vw (x, t)
immediately to the left and right of the discontinuity.
The mapping, yw (x, t) defining the weak solution (10) of the Riemann equa-
tion is a coordinate of the global minimum of the function (26). To find it,
it is necessary to lift the graph of the function
M (y)x + h (29)
from the minus infinity level until it touches the graph of the function φ(y, t)
(27).
Here, χ(z) is the Heaviside unit step function equal to 1 for z > 0 and 0 else-
where. Also, notice that the particle has no impact on the initial momentum
P (y), since its momentum is zero.
Before we proceed to explain the solution of our problem, let us carry out a
dimensional analysis of the parameters entering into the problem, something
that is routine in physics and engineering. These parameters can be used to
form the unique combinations that have the dimensionality of length and
time:
m m
= , θ= .
ρ vρ
Uniqueness of the above dimensional combinations means that and θ are
typical spatial and temporal intervals over which most of the events that are
important for the embedded particle are played out. For this reason, it makes
sense to pass to the nondimensional variables
x y t
η= , ζ= and τ = ,
θ
measuring the spatial coordinate and the time in scales that are natural for
the particle. Now multiply (27) and (29) by ρ/m2 , and replace φ(y, t) and
M (y)x by the nondimensional functions
ρ
ψ(ζ, τ ) = φ(y, t)
m2
and
ρ
μ(ζ, η) = M (y) x = η[ζ + χ(ζ)] .
m2
It is also convenient to take
(ζ + τ )2
ψ(ζ, τ ) = . (30)
2
Determination of the law of motion of the particle is equivalent to finding
the value η = η ∗ (τ ) for which the piecewise linear curve μ(ζ, η) is tangent to
the parabola (30) at two points. The graphics shown in Fig. 13.5.3 suggest
that the desired η ∗ satisfies the quadratic equation
(η ∗ − τ )2 = 2η ∗ ,
√
so that η ∗ = τ + 1 − 2τ + 1 . The resulting law of motion for the macropar-
ticle has the following asymptotics:
τ2
η∗ ∼ (τ → 0) ; η∗ ∼ τ (τ → ∞) .
2
13.5. E–Rykov–Sinai Principle 209
FIGURE 13.5.3
A particle placed in a flow of constant density. Two rectilinear
segments (ζ − τ for ζ < 0 and ζ − τ + η for ζ > 0) touch the parabola
(30) from below at two points.
If the initial flow density ρ0 (x) is positive everywhere, then there exists a
continuous inverse function y = Y(m). Substituting it in (26)–(29), we arrive
at a construction of the weak solution of the Riemann equation that relies
on the previously used formula (10), where now
yw (x, t) = Y(m(x, t)) , (32)
while m(x, t) is the “coordinate” of the lower point of tangency of the line
xm + h (33)
210 Chapter 13. Generalized Solutions of Nonlinear PDEs
For a time t ∈ (0, tn ), while the motion remains single-streamed, the corre-
sponding Lagrangian field M (y, t) is independent of time, and is described
by the expression (31). This means that the Eulerian field m(x, t) satisfies
the equation
x
∂m ∂m
+v = 0, m(x, t = 0) = ρ0 (x ) d x = M (x) . (37)
∂t ∂x −∞
13.5. E–Rykov–Sinai Principle 211
Substituting for y in (31) the mapping yw (x, t) obtained via the ERS
principle, we arrive at a generalized solution of equation (37),
This expression fully agrees with the laws of physics. In particular, cross-
ing through the discontinuity point x∗ of the mapping yw (x, t), the cumulative
mass function has a jump of size corresponding to the mass
y+
m∗ = ρ0 (z) dz (39)
y−
FIGURE 13.5.4
Schematic graphs of the generalized cumulative mass function
(top), and the density (bottom) of the 1-D flow of inelastically col-
liding particles. The thick arrows in the bottom graph indicate the
Dirac deltas describing the singular density of the macroparticles.
where the summation is taken over all the jump points {x∗k } of the mapping
yw (x, t). The constants {m∗k } in front of the Dirac deltas are equal to the
masses of the corresponding macroparticles, and the last term describes the
bounded density of uniformly moving microparticles located in the intervals
between the macroparticles. A schematic graph of the generalized density
field in the flow of inelastically colliding particles is shown in Fig. 13.5.4.
Another method of determination of the generalized density of the flow
of sticky particles is based on the representation of the density field with the
help of the Dirac delta:
ρw (x, t) = ρ0 (y) δ (Xw (y, t) − x) dy . (41)
In the case that X(y, t) is smooth and strictly monotone, the functional
on the right-hand side is defined via the usual rules of distribution theory.
However, if on some intervals of the y-axis the function X(y, t) is constant, as
is the case for the flow of sticky particles, then standard distribution theory
fails to apply. But even in this case, the expression (41) can be endowed with
a well-defined meaning relying on the concept of supersingular distributions
discussed in Sect. 2.9.5 of Volume 1.
FIGURE 13.5.5
The generalized mapping Xw (y, t) and the corresponding singular
density ρw (x, t). The initial density is Gaussian.
by a test function φ(x), integrate both sides with respect to x over the whole
real line, change the order of integration on the right-hand side, and finally,
use the probing property of the Dirac delta. As a result, we get the equality
φ(x)ρw (x, t) dx = ρ0 (y) φ(X(y, t)) dy , (42)
The braces on the right-hand side indicate that values of the derivative are
taken everywhere with the exception of the point x∗ , where the derivative
can be defined arbitrarily without affecting the value of the integral.
Equality (44) means that the distributional density of the flow of sticky
particles is equal to
"
∗ ∗ ∂yw (x, t)
ρw (x, t) = m δ(x − x ) + ρ0 (yw (x, t)) . (45)
∂x
It is in complete agreement with the equality (40), which was obtained via
different reasoning.
214 Chapter 13. Generalized Solutions of Nonlinear PDEs
∂v
+ (v · ∇)v = 0 , v(x, t = 0) = v 0 (x) . (1)
∂t
The nabla operator
∂ ∂ ∂
∇ = j1 + j2 + j3
∂x1 ∂x2 ∂x3
is equal to the sum of the partial derivatives with respect to the three Carte-
sian coordinates x1 , x2 , x3 , each multiplied by the corresponding unit coordi-
nate vector, j 1 , j 2 , j 3 .
If the velocity field is a potential field, that is, if there exists a scalar
function s(x, t) such that
∂s 1
+ (∇s)2 = 0 , s(x, t = 0) = s0 (x) . (4)
∂t 2
In what follows, we shall always assume that the velocity field is a potential
field and that the equality (3) is satisfied. In this case,
(∇s)2 = v 2 = (v · ∇)s,
13.6. Multidimensional Nonlinear PDEs 215
dX
=V , X(y, t = 0) = y , (8)
dt
for the Lagrangian velocity field V (y, t),
dV
= 0, V (y, t = 0) = v 0 (y) , (9)
dt
dS 1
= V2, S(y, t = 0) = s0 (y) . (10)
dt 2
and the solutions of the second and third equations provide, respectively, the
Lagrangian velocity and the Lagrangian velocity potential fields:
1
V (y, t) = v 0 (y) , S(y, t) = s0 (y) + v02 (y) t . (12)
2
y = y(x, t) . (13)
Then the velocity and the velocity potential can be expressed by the equalities
1
v(x, t) = v0 (y(x, t)) , and s(x, t) = s0 (y(x, t)) + v02 (y(x, t)) t . (14)
2
x − y(x, t)
v(x, t) = , (15)
t
(y(x, t) − x)2
s(x, t) = s0 (y(x, t)) + . (16)
2t
13.6. Multidimensional Nonlinear PDEs 217
where δij is the Kronecker symbol and sij are components of the symmetric
matrix ŝ, equal to the second-order partial derivatives of the initial velocity
field potential
∂ 2 s0 (y)
ŝ = [sij ] , sij (y) = . (19)
∂yi ∂yj
As is well known, by a rotation of the coordinate system {y1 , y2 , y3 }, different
for different points y, the symmetric matrix ŝ can be diagonalized, that is,
written in the form
ŝ = [λi δij ] ,
where {λ1 , λ2 , λ3 } are the eigenvalues of the matrix ŝ. In the local coordinate
system specified above, the Jacobian (18) turns out to be
!
3
J= (1 + λi t) . (20)
i=1
FIGURE 13.6.1
Transformation of an “elementary” 2-D ellipse in the general case
λ1 < λ2 < 0. Initially, in the interval 0 < t < −1/λ1 , the el-
lipse is being compressed in both directions. Then in the interval
−1/λ1 < t < −1/λ2 , it expands along the y1 -axis but continues to be
compressed along the y2 -axis. Finally, for t > −1/λ2 , it expands in
both directions. The “volume” of the shaded ellipse is negative.
is equal to the sum of the densities of all flows, and it preserves its physical
meaning. The above summation is over all n branches of the multivalued, in
general, mapping y = y(x, t).
The generalized density field in the 2-D plane and the case of a discon-
tinuous mapping y = y(x, t) will be discussed in the last section of this
chapter.
So, assume that a fire front surrounding a 3-D combustion region propa-
gates with unit velocity in a direction perpendicular to the front itself. It is
not hard to prove, arguing as in the case of equation (12.2.7), that the fire
front
z = h(x, t), (25)
which is a 2-D surface in a 3-D space, satisfies a nonlinear partial differential
equation of the form
∂h
= 1 + (∇h)2 , (26)
∂t
where x = {x1 , x2 } is a 2-D vector in the horizontal plane, and z = x3 is the
vertical coordinate.
If the combustion spreads predominantly upward along the z-axis, then
the inequality
(∇h)2 1
is satisfied, and equation (26) can be simplified. Indeed, if we expand the
right-hand side of (26) into a Taylor series with respect to (∇h)2 , retain only
the first two terms of the expansion,
1
1 + (∇h)2 ≈ 1 + (∇h)2 ,
2
and drop the first constant term because it represents a trivial constant-
velocity motion of the interface, then we arrive at the approximate equation
∂h 1
= (∇h)2 , h(x, t = 0) = h0 (x) , (27)
∂t 2
similar to equation (12.2.9).
Another field closely related to the geometry of the fire front is
Its geometric meaning is clear: its magnitude is equal to the tangent of the an-
gle between the normal to the front and the z-axis. Applying the ∇-operator
to the terms of (27), we arrive at the following equation for the vector field
u(x, t):
∂u
+ (u · ∇)u = 0, (29)
∂t
which is a 2-D analogue of equation (1) describing the velocity field of the
flow of uniformly moving particles.
13.6. Multidimensional Nonlinear PDEs 221
and
1
H(y, t) = h0 (y) − (∇h0 (y))2 t . (32)
2
Since the vector function
and its inverse (13) provide a one-to-one mapping of R2 onto R2 , the expres-
sions (31) and (32) allow us to determine the shape of the desired surface
(y(x, t) − x)2
h(x, t) = h0 ((y(x, t)) − (34)
2t
and the field
x − y(x, t)
u(x, t) = (35)
2t
at any point of the x-plane. But as soon as the mapping y = y(x, t) becomes
multivalued, the real physical surface hw (x, t) of the combustion region cor-
responding to the weak solution of equation (27) selects from all branches of
the multivalued field h(x, t) the one with the largest magnitude at a given
point x.
The above-mentioned weak solution of equation (27) can be found by
relying on the global maximum principle.
At this point, it is helpful to introduce an auxiliary function
(y − x)2
G(y; x, t) = h0 (y) − (36)
2t
of the vector argument y, with x and t considered parameters.
Vanishing of the gradient at x, that is, the condition
∇G(y; x, t) = 0 ,
222 Chapter 13. Generalized Solutions of Nonlinear PDEs
FIGURE 13.6.2
Top: The initial smooth shape of the growing interface. Bottom: The
same evolving interface after a certain time t > 0. The lines of
discontinuity of the gradient are clearly visible.
Comparing this equality with (32)–(34), we observe that all the extreme
points of y(x, t) turn the mapping (33) into an identity. Moreover, the value
of the function G(y; x, t) at these points is equal to the value of the Eulerian
field h(x, t) (34), which represents the elevation of the growing interface.
Consequently, substituting the coordinates y w (x, t) of the global maximum
of the function (36) into (34) and (35), we obtain the desired weak solutions
hw (x, t) and uw (x, t).
13.7. Exercises 223
13.7 Exercises
1. Prove that in a single-stream regime, the field q(x, t), see (13.2.9), of
the derivative with respect to x of the solution v(x, t) of the Riemann
equation (13.1.1) satisfies the inequality
1
q(x, t) < . (1)
t
Illustrate the above inequality by constructing a graph displaying the
dependence on x of the field tq(x, t) in the case of the initial condition
v0 (x) = a sin(kx) .
3. Suppose that the initial density and the initial velocity of a cloud of
uniformly moving particles are equal, respectively, to ρ0 (x) and v0 (x).
The initial mass of particles and their center of mass are finite:
1
M = ρ0 (y) dy < ∞ , xc = y ρ0 (y) dy < ∞ .
M
Determine the motion
1
x̄(t) = x ρ(x, t) dx
M
224 Chapter 13. Generalized Solutions of Nonlinear PDEs
of the center of mass of the particle cloud and the time dependence of
its dispersion
1
D(t) = (x − xc )2 ρ(x, t) dx .
M
Solve the problem by expressing x̄(t) and D(t) via the Fourier images
of the density field (13.3.3).
FIGURE 13.7.1
The initial shape h0 (x) of the growing interface and the osculating
disk of radius ct. Its center traces the shape of the interface at time
t > 0. For large enough t, the osculating disk will not able to touch
all the points of the initial interface shape, and the center will move
along the weak solution hw (x, t) of the interfacial growth equations
(13.1.5) and (12.2.7). A cusp, often encountered in weak solutions,
is clearly visible.
7. Visualize a 2-D ice floe bounded by the contour L0 and floating on the
surface of the water. The water surrounding the ice floe can be either
cold or warm. We shall assume that in cold water, the floe grows with
velocity c perpendicular to the floe’s boundary and that in warm water,
226 Chapter 13. Generalized Solutions of Nonlinear PDEs
it melts with the same velocity. Suppose that initially, the floe was
frozen by being surrounded by cold water for a period of time, say T .
As a result, the floe assumed a new shape with its boundary described
by the contour LT . Subsequently, the water was instantaneously heated,
and the floe started melting for another time period T . What conditions
on L0 guarantee that after a cycle of freezing and thawing, the floe
returns to its original shape?
8. Suppose that a cone with solid angle Ω is filled with an inert material
of 3-D density , except for the tip of the cone, which we define as the
subset of the cone where the distance d from its vertex O is small, say
d ≤ ε → 0. The tip is filled with an explosive of the same density as the
inert material. At t = 0, the explosive is ignited, acquiring momentum
p0 (momentum density is uniformly distributed throughout the explo-
sive and is equal to p0 /ε). Suppose that the solid angle is small enough
(Ω 1) that the material can be assumed to be moving practically
along the inner axis r of the cone (see Fig. 13.5.4). Furthermore, as-
sume that in the process of compression, the inert material forms an
infinitely thin pancake of sticky particles occupying the whole cross sec-
tion of the cone. The motion of this pancake can be called a detonation
wave. Using the ERS principle, find the law of motion of the detonation
wave and the rate of growth of the mass of the pancake.
9. Using the global maximum principle, find a weak solution of the Rie-
mann equation (13.1.10) in the case that the initial condition is pro-
portional to the Dirac delta:
v0 (x) = s δ(x) .
10. Assume that the initial velocity field of the flow of inelastically colliding
particles is
v0 (x) = V χ(−x) ,
and the initial density ρ0 (x) is an absolutely integrable function with
total mass
ρ0 (x) dx = M < ∞ .
11. Imagine a 1-D Universe with the very simple universal gravitation law
(UGL): The force of gravitational attraction of two bodies is propor-
tional to their masses and independent of their distance. Let us refor-
mulate this UGL in the language of mathematical formulas. Suppose
that two bodies, the “left,” one of mass Ml , and the “right” one, of mass
Mr , are located in the 1-D Universe (x-axis). Denote by Fl the force
acting on the “left” body, and by Fr the force acting on the “right”
body. According to the UGL, the force of interaction between the two
bodies is
Fl = −Fr = γMl Mr ,
where γ is a “gravitational constant.”
Now assume that the matter is continuously distributed in the above
universe with initial velocity v0 (x) and density ρ0 (x). Moreover, assume
that the total mass of the matter is given by
M = ρ0 (x) dx < ∞.
Find the Lagrangian velocity field and the density field of this flow of
interacting particles at times before their collisions begin.
12. Suppose that the initial density field in Exercise 11 is
2
ρ0 (x) = ρ0 , (27)
x2 + 2
and that the initial velocity field is identically zero. Construct graphs of
the Eulerian velocity field v(x, t) and density field ρ(x, t) for the flow of
gravitationally interacting particles at several time instants. Utilizing
these graphs, discuss the onset of gravitational instability.
13. Assume that the initial density of the flow of gravitationally interacting
particles in Exercise 11 is of the form
x
ρ0 (x) = ρ0 g , (28)
where g(z) is a continuous, nonnegative, and even function such that
g(0) = 1. Find an expression for the Lagrangian velocity and density
fields in the limit → ∞ in the case of uniform initial density ρ0 =
const. Assuming that the 1-D universe is initially expanding, that is,
v0 (y) = Hy, where H is the Hubble constant, find the time of the
collapse of the universe.
Chapter 14
The present chapter studies behavior of two standard 1-D nonlinear dynam-
ics models described by partial differential equations of order two and higher:
the Burgers equation and the related KPZ model. We shall concentrate our
attention on the theory of nonlinear fields of hydrodynamic type, where the
basic features of the temporal evolution of nonlinear waves can be studied in
the context of competition between the strengths of nonlinear and dissipative
and/or dispersive effects. Apart from being model equations for specific phys-
ical phenomena, Burgers–KPZ equations are generic nonlinear equations that
often serve as a testing ground for ideas for analysis of other nonlinear equa-
tions. They also produce a striking typically nonlinear phenomenon: shock
formation.
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 229
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 6,
© Springer Science+Business Media New York 2013
230 Chapter 14. 1-D Nonlinear PDEs of Higher Order
the z-axis). We shall take this additional motion into account by introducing
a new term, ∂g/∂x, into equation (1):
2
∂h ∂g 1 ∂h
+ = . (2)
∂t ∂x 2 ∂x
FIGURE 14.1.1
Schematic illustration of the process of isotropic deposition of par-
ticles on the interface z = h(x, t). The arrows above the interface
represent velocities of particles being deposited on the growing in-
terface. The arrows below the interface indicate the direction of
motion of particles sliding down the interface.
incident stream. This extension of the model requires an addition of the term
F (x, t) on the right-hand side of equation (2), resulting in the equation
2
∂h 1 ∂h ∂ 2h
= +μ + F (x, t) . (4)
∂t 2 ∂x ∂x2
∂v ∂v
+v = 0. (7)
∂t ∂x
In this context, v(x, t) represents the fluctuations of pressure of the medium
(for example, the atmosphere) induced by a passing acoustic wave. How-
ever, from the physical perspective, equation (7) does not take into account
important additional effects such as viscosity. The latter is the principal cause
of energy dissipation for an acoustic wave. A first principles derivation of the
equation of nonlinear acoustics in a viscous medium leads to the equation
∂v ∂v ∂ 2v
+v =μ 2, v(x, t = 0) = v0 (x) , (8)
∂t ∂x ∂x
14.1. Regularization of First-Order Nonlinear PDEs 233
∂u ∂u
+ Φ (u) = 0.
∂t ∂x
If the flux function is permitted to depend additionally on the gradient of
the density u, say
∂
φ(t, x) = Φ(u(t, x)) − ν u(t, x),
∂x
then the above conservation law leads to the equation
∂u ∂u ∂ 2u
+ Φ (u) = ν 2,
∂t ∂x ∂x
of which the Burgers equation is a special case.
where ũ0 (κ) is the Fourier image of the initial field u0 (x). Observe that the
right-hand side of the equality (9) can be treated as a solution of an inte-
grodifferential equation that can be written in the following symbolic form:
∂u ∂
+ iW u = 0. (10)
∂t i∂x
For the wave to propagate without energy dissipation, it is necessary and
sufficient that W (κ) be a real function of a real argument κ. We also require
that a real-valued initial condition u0 (x) result in a real-valued field u(x, t).
This forces W (κ) to be an odd function of κ.
Suppose that u(x, t) is a sufficiently smooth function of x with Fourier
image that essentially vanishes outside a small neighborhood of κ = 0, and
let W (k) be a smooth function of its argument. Then we can approximate
W (κ) in (9) and (10) by a few terms of its Taylor expansion around κ = 0:
W (κ) = c κ + μ κ3 + · · · .
The first term is responsible for a straightforward propagation of the wave
without any change of its shape. It can be easily eliminated by changing the
variables to a moving coordinate system. The remaining second term permits
replacement of the integrodifferential equation (10) by the purely differential
equation
∂u ∂ 3u
=μ 3.
∂t ∂x
Combining it with the Riemann equation gives the equation
∂u ∂u ∂ 3u
+u =μ 3, (11)
∂t ∂x ∂x
which is called the Korteweg–de Vries (KdV) equation. The KdV equation
takes into account the combined effects of nonlinearity and dispersion. The
coefficient μ reflects the magnitude of the dispersion effects and plays a com-
pletely different role from that of the dissipation coefficient μ in the Burgers
equation.
236 Chapter 14. 1-D Nonlinear PDEs of Higher Order
We shall begin with an analysis of the basic properties of the Burgers equa-
tion related to its fundamental symmetries. Later on, we shall produce a
general analytic formula for solutions of the Burgers equation, but we will
not hurry to present it. Instead, we shall start with a systematic investigation
of intrinsic structural properties of the equation.
Note that only rarely can one count on finding explicit general solutions
of a nonlinear equation. In their absence, the investigator tries to get some
idea about what they are like, studies typical special cases, and using this
information tries to assemble a coherent general picture applying qualitative
and asymptotic methods. We shall implement this approach in the case of the
KPZ and Burgers equations, where the qualitative and asymptotic methods
can be compared with explicit analytic solutions, thus reassuring us about
the effectiveness of the former.
On the other hand, the KPZ equation enjoys the standard invariance
under reflection property, that is,
This symmetry represents the fact that the interfacial growth is the same in
both directions of the base line.
In addition to translation and reflection symmetries, the Burgers equa-
tion (14.1.8) enjoys Galilean invariance, which is of fundamental physical
importance. We shall explain this concept assuming that u(x, t) represents
the velocity in a 1-D flow of a medium. Let ũ(x , t) be the velocity field
of a medium in the coordinate system x moving with constant velocity V .
Then the same velocity field in the stationary coordinate system x such that
x = x − V t will have the form
u(x, t) = V + ũ(x − V t, t) .
It is easy to verify that the KdV equation is also invariant under Galilean
transformations.
238 Chapter 14. 1-D Nonlinear PDEs of Higher Order
x v(s, t)
s= , u(s, t) = , (5)
U
14.2. Symmetries of Burgers and KPZ Equations 239
∂u U ∂u μ ∂ 2u
+ u = 2 2, u(s, t = 0) = u0 (s) . (6)
∂t ∂s ∂s
First, observe that in the special case of μ = 0, the above equation takes
the form
∂u U ∂u
+ u = 0, u(s, t = 0) = u0 (s).
∂t ∂s
The solution of the latter will not depend on the scales of the initial field if
the unit of time measurement is taken to be T = /U , or in other words, if
we pass to the dimensionless time
t
τ= = t. (7)
T U
In this case, the above equation is transformed into the canonical Riemann
equation
∂u ∂u
+u = 0, u(s, t = 0) = u0 (s) .
∂τ ∂s
Therefore, in the absence of viscosity (μ = 0), the identical in shape, but dif-
ferent in spatial scale and magnitude initial condition v0 (x) generate solutions
v(x, t) that vary in time in a similar fashion. The only feature differentiating
them is their rate of evolution. And it is not surprising that this scale invari-
ance of the Riemann equation extends to its weak, discontinuous solutions
as well.
In presence of viscosity (μ = 0), equation (6) is transformed into the
equation
∂u ∂u μ ∂ 2u
+u = , u(s, t = 0) = u0 (s) , (8)
∂τ ∂s U ∂s2
so that the behavior of the solutions of the Burgers equation becomes much
more complicated and depends qualitatively on the magnitude of the dimen-
sionless parameter
U
R= , (9)
μ
which appears on the right-hand side of equation (8). Traditionally, this
parameter is called the acoustic Reynolds number, or simply the Reynolds
number. A small Reynolds number (R 1) indicates that the influence of
the viscosity is large, the diffusive term on the right-hand side of the Burgers
240 Chapter 14. 1-D Nonlinear PDEs of Higher Order
equation (14.1.8) dominates the nonlinear term on the left-hand side, and
the equation can be well approximated by the linear diffusion equation
∂v ∂ 2v
=μ 2. (10)
∂t ∂x
For this reason, the Burgers equation is sometimes called the nonlinear diffu-
sion equation. For large Reynolds numbers, the evolution of the field v(x, t)
is dominated by propagating nonlinear effects. In other words, one can think
of the Reynolds number as a measure of the influence of the nonlinearity on
the behavior of the field v(x, t).
Let us take a look at the Reynolds number from yet another point of
view. The number can be represented as the ratio of two spatial scales,
R= ,
δ
where is an external, in relation to the Burgers equation, spatial scale
dictated by the initial conditions, while
μ
δ= (11)
U
is the internal characteristic spatial scale, intrinsic to the structure of the
Burgers equation itself. In what follows, we shall call and δ the internal
and external scales, respectively.
The concepts of the internal and external scales are difficult to express in
a rigorous mathematical language. But they are valuable in heuristic studies
of nonlinear fields and in particular, are widely used in the theory of strong
atmospheric turbulence. Also, if we select a time-dependent exterior scale
(t) and an interior scale δ(t) of the evolving nonlinear field, we can then
define the evolving, time-dependent Reynolds number
(t)
R(t) = . (12)
δ(t)
v = αy, x = y + α yt.
Eliminating y from these two equations, we obtain the relation (13), which
is often called the Hubble expansion solution.1
The characteristic property of the Hubble expansion is that over time,
the velocity field v(x, t) in (13) loses information about the initial velocity
profile, which in our case means a loss of information about the coefficient α.
Indeed, as t → ∞, the field v approaches the field
x
v(x, t) ∼ , (14)
t
which is independent of α.
1
Edwin Hubble was an astronomer who in the 1920s, discovered that the universe is
continuously expanding, with galaxies moving away from each other.
242 Chapter 14. 1-D Nonlinear PDEs of Higher Order
For α < 0, the solution (13) describes not an expansion but a compression
of the universe. This situation leads to the so-called gradient catastrophe, in
which the velocity field becomes infinite everywhere in finite time. As we will
see later on, similar, but local, gradient catastrophes are typical of solutions
of the Burgers equation with small values of the coefficient μ.
As a next step, consider a more general model of the 1-D Universe in
which the simple Hubble expansion considered above is augmented by other
perturbations. In other words, we shall try to solve the Burgers equation with
the initial condition
v(x, t = 0) = α x + v̂0 (x) ,
where the first term on the right-hand side takes into account the Hubble
expansion and the second term is a perturbation of the Hubble velocity field.
Not to belabor the point, we shall seek a solution of the Burgers equation in
the form of a sum
αx
v(x, t) = + v̂(x, t), (15)
1+αt
where the first summand corresponds to the Hubble expansion. Astrophysi-
cists traditionally call the second summand peculiar velocity. It reflects the
evolution of perturbations corresponding to the dilation of the initial veloc-
ity fluctuations. Let us explore the evolution of the peculiar velocity in a
coordinate system expanding with the fluctuations. Thus the new coordinate
system z is related to the old coordinate system x via the obvious equation
x
z= .
1+αt
In the new coordinate system, the solution velocity field takes the form
where
u(z, t) = v̂ (z(1 + αt), t) .
Substituting (16) into (14.1.8), we arrive at the following equation for the
peculiar velocity:
∂u 1 ∂u αu μ ∂ 2u
+ u =− + .
∂t 1 + αt ∂z 1 + αt (1 + αt)2 ∂z 2
The above equation makes it clear that the expanding fluctuations dimin-
ish the influence of the nonlinearity, weaken the field’s dissipation, but also
14.2. Symmetries of Burgers and KPZ Equations 243
attenuate (for α > 0) the field itself. The last effect can be intuitively
explained by the fact that “particles” moving with high velocity overtake the
Hubble expansion, so that the difference between the peculiar velocity and
Hubble expansion becomes smaller.
We will take attenuation of the peculiar velocity into account by intro-
ducing the substitution
w(z, t)
u(z, t) = ,
1 + αt
thus arriving, finally, at the equation
∂w 1 ∂w 1 ∂ 2w
+ w = μ .
∂t (1 + αt)2 ∂z (1 + αt)2 ∂z 2
Note the similarity of coefficients for the nonlinear and dissipative terms. It
reflects the fact that after transition to the new time
t
τ= , (17)
1+αt
which takes into account the simultaneous attenuation of the nonlinear effects
and viscosity in the expanding background, we “miraculously” return to the
original Burgers equation
∂w ∂w ∂ 2w
+w =μ 2 , w(z, τ = 0) = v̂0 (z) . (18)
∂τ ∂z ∂z
Implementation of all of the above transformations yields the relation
1 x t
v̂(x, t) = w , . (19)
1 + αt 1 + αt 1 + αt
Now it is clear that the interaction of the perturbation field v̂(x, t) with the
dilating (for α > 0) background leads to attenuation of the field’s magnitude
and growth of its spatial scale. For this reason, Hubble expansion slows down
the evolution of the peculiar field, which, in the absence of expansion, is equal
to w(x, t). As a result, during the infinite time interval t ∈ (0, ∞), one realizes
only a portion of the evolution of the field w(x, t), namely that corresponding
to the finite time interval (0, t∗ ), where t∗ = 1/α.
In the opposite case, α < 0, a global gradient catastrophe takes place: an
avalanche-type growth of the magnitude of the peculiar field occurs, squeezing
the scales of the initial perturbation. Thus the course of events is accelerated
dramatically. More precisely, all stages of the evolution of the perturbation
w(x, t) on an infinite time interval are played out in the finite time interval
t ∈ (0, t∗ ), where t∗ = 1/|α| .
244 Chapter 14. 1-D Nonlinear PDEs of Higher Order
v(x, t) = υ(x − V t) ,
where υ = υ(z) depends on only one variable, z = x − V t. Stationary waves
move with constant velocity V without changing their shape.
To find stationary solutions of the Burgers equation, it suffices, in view of
its Galilean invariance, to find a static, time-independent solution and then
“move it” using the relationship (3).
A static solution υ(x) must satisfy the equation
d 1 2 dυ(x)
υ (x) − μ = 0, (20)
dx 2 dx
which has been written in a form immediately providing us with the first
integral,
dυ(x) 1 2
μ + U − υ 2 (x) ≡ 0 , (21)
dx 2
where U is an arbitrary constant of integration, which coincides, obviously,
with the extremal value of υ(x). Because of the translational invariance of the
Burgers equation, it is easy to see that as the second constant of integration
we can use an arbitrary shift a of the solution along the x-axis: υ = υ(x − a).
Thus we arrive at a final expression for the static field:
Ux
υ(x) = −U tanh . (22)
2μ
Finally, an arbitrary stationary wave solution of the Burgers equation can
be constructed by applying to (22) the Galilean transformation (3) and an
arbitrary shift:
U
v(x, t) = V − U tanh (x − V t − a) . (23)
2μ
Dimensionless graphs of u(x) = υ(x)/U are shown in Fig 14.2.1 for several
values of the viscosity coefficient μ.
It is evident that the stationary wave solution describes a transition from
the maximum value v− (on the left) of the field to its minimum value v+ (on
the right), where
It follows from (23) that the effective width of the transition zone is equal to
the internal scale δ, see (11), and that the velocity V of the stationary wave
and its amplitude U are related to the above maximal and minimal values
by the familiar relationships
v+ + v− v+ − v−
V = , U= . (24)
2 2
Recall that we already encountered these relationships in the context of
weak solutions of the Riemann equation obtained via the global minimum
principle. There, similar equalities expressed the velocity and amplitude of
the discontinuity of the weak solution vw (x, t) in terms of the field’s values
immediately to the left, and right, of the jump.
FIGURE 14.2.1
Static dimensionless solutions υ(x)/U of the Burgers equation are
shown here in a fixed scale on the x-axis for three values of the
viscosity coefficient: μ1 , μ2 = μ1 /3, μ3 = μ2 /3. The spatial scale on
the x-axis was selected so that the transition from the maximum
v − = 1 to the minimum v + = −1 was best illustrated for the middle
value μ2 of the viscosity coefficient.
A direct relationship between the stationary wave solution and the local
behavior of the weak solutions at a jump point is completely understandable
if one observes that for μ → 0+ , the stationary solution (23) weakly converges
to the shock wave
For this reason, one sometimes says that the introduction of the viscosity
(μ > 0) leads to a smoothing out of jumps in a weak solution vw (x, t) of the
Riemann equation. If the internal scale of a jump of the weak solution vw (x, t)
is much smaller than the external scale (δ(t) (t)), then the fine structure
of the jump can be reconstructed by means of the stationary solution (23)
and relations (24), substituting into them the current values v ± (t) to the
right and to the left of the jump of the weak solution vw (x, t).
FIGURE 14.2.2
The dimensionless special Khokhlov’s solution (33) of the Burgers
equations shown for different values R = 100, 10, 1 of the Reynolds
number (which is inversely proportional to time). The larger the
Reynolds number, the closer the graph becomes to the limiting
shock wave solution u(z) = z − sign(z).
2μg = g 2 − ρg , g = −f . (35)
2μp = ρp − 1
for which the integrating factor is exp(− ρ dρ/2μ) = exp(−ρ2 /4μ). Thus,
its solution is
2
π ρ ρ
p(ρ) = exp C −Φ √ ,
μ 4μ 2 μ
where the auxiliary function is
z
1 2 1
Φ(z) = √ e−w dw = [1 + erf (z)] . (36)
π −∞ 2
14.2. Symmetries of Burgers and KPZ Equations 249
As a result,
2
μ exp − 4μ
ρ
1
g(ρ) = = . (37)
p(ρ) π C − Φ √ρ
2 μ
where
S
R= . (39)
2μ
Hence, the self-similar solutions of the homogeneous KPZ equation are of
the form
x
hR (x, t) = f √ . (40)
2 μt
Differentiating (40) with respect to x, we obtain the self-similar solutions
of the Burgers equation:
∂h μ x
v=− ⇒ v(x, t) = VR √ , (41)
∂x πt 2 μt
The graphs of VR (z), as functions of the variable z, are shown in Fig. 14.2.3
for several values of the parameter R.
To understand the deeper nature of the above self-similar solutions, we
shall now consider mechanisms that lead to their formation. First of all, let
us check what kind of initial conditions are associated with such solutions.
Note that the weak limit of (40), as t → 0+, is proportional to the Heaviside
unit step function
h(x, t = 0+) = −S χ(x) ,
250 Chapter 14. 1-D Nonlinear PDEs of Higher Order
and that the self-similar solution (41) of the Burgers equation weakly con-
verges, as t → 0+, to the Dirac delta,
Next, let us elucidate the role of the parameter R. First, observe that by
applying the scaling (5) and (7) to the Burgers equation, we get the equa-
tion (8). With new scaling, the Dirac delta initial condition is transformed
into the initial condition
S
u0 (s) = δ(s) .
U
This solution is independent of the scales of the initial condition only if
U ∼ S .
Comparing this equality with (39) and (9), we arrive at the conclusion that
R in (39) can be treated as a Reynolds number of the singular initial condi-
tions (43).
FIGURE 14.2.3
Graphs of the function VR (z) for R = 0.5, 5, 10, 15, 20. It is clear that
the larger R is, the more asymmetric (triangular) the shape of the
self-similar solution of the Burgers equation becomes.
coincides with the main asymptotics of the self-similar solution of the lin-
ear diffusion equation (10). This should come as no surprise, since at small
Reynolds numbers, the nonlinear effects are negligible. Let us also note that
a similar “linearized” asymptotics,
μ −R x2
v(x, t) ∼ (1 − e ) exp − (z → −∞) , (45)
πt 4μt
is encountered for every R, as long as the values of the function Φ(z) are
small enough. Indeed, for all z < 0, they are.
Next, let us discuss the behavior of the self-similar solution (41-42) for
large values of the Reynolds number (R 1) and for z 1. For this purpose,
we need to know the asymptotics of Φ(z) as z → ∞. It is not hard to show
that
1 2
Φ(z) ∼ 1 − √ e−z (z → ∞) . (46)
2z π
Substituting the above expression into (41)–(42), we obtain
x 1 x
v(x, t) ≈ √ , z= √ . (47)
t 2z π exp(z 2 − R) + 1 2 μt
It follows from (47) that as long as the first term in the denominator remains
small, that is, √
2z π exp(z 2 − R) 1 , (48)
the asymptotics are the familiar, linear in x, asymptotics of the solutions of
the Burgers equation,
x
v(x, t) ≈ . (49)
t
The critical value z∗ at which the linear growth (49) is replaced by an expo-
nential decay is determined by the transcendental equation
√ 2
2z∗ πez∗ = eR .
A rough estimate of its solution gives
√ S
z∗ ≈ R = . (50)
2μ
2 (t) 2S
R= ≈ = const . (53)
2μt μ
This bring us to an important conclusion: The Reynolds number of a
self-similar solution of the Burgers equation is independent of time.
For large Reynolds numbers (R 1), we can also clarify the shape of
the self-similar solution of the Burgers equation in a transition zone. There,
expression (47) can be rewritten in the form
(t) 1
v(x, t) ≈ z 2 −z 2 . (55)
t e ∗ + 1
motion for a stationary wave solution), the above expression coincides with
expression (23).
In summary, one can think of the self-similar solutions of the Burgers
equation as being assembled, like a Lego blocks toy, from known particular
solutions (45), (49), and (56):
⎧
⎪
⎪ μ x2
⎪
⎪ exp − , x << 0 ,
⎪
⎪ πt 4μt
⎨
x
v(x, t) ≈ , 0 < x < , (57)
⎪
⎪ t
⎪
⎪
⎪
⎪ U
⎩ U − U tanh (x − ) , x .
2μ
The first piece is a solution of the linear diffusion equation (10) with initial
condition
2
v0 (x) = 2μδ(x) = S δ(x) .
R
The second piece coincides with the characteristic incline (49) and reflects
the influence of the nonlinearity. Finally, the third piece describes a sharp
decay of a self-similar solution, mimicking the shape of a stationary wave
solution.
FIGURE 14.2.4
A self-similar solution (41) of the Burgers equation, at a fixed time
t > 0, is shown here for R = 50. Three characteristic pieces of the
solution reflecting, respectively, the linear diffusion, the shock wave
structure, and the stationary wave behavior are indicated. Also, the
triangular “skeleton” of the self-similar solution, reflecting the case
of R = ∞, is shown.
254 Chapter 14. 1-D Nonlinear PDEs of Higher Order
∂h(x, t)
v(x, t) = − (2)
∂x
14.3. General Solutions of Burgers Equation 255
∂v ∂v ∂ 2v
+v = μ 2 + f (x, t) , v(x, t = 0) = v0 (x) , (3)
∂t ∂x ∂x
where
∂F (x, t) ∂h0 (x)
f (x, t) = − , v0 (x) = − . (4)
∂x ∂x
Moreover, both equations are very close relatives of the linear diffusion equa-
tion
∂ϕ ∂ 2ϕ 1
=μ 2 + F (x, t) ϕ . (5)
∂t ∂x 2μ
Indeed, write ϕ(x, t) in the form
h(x, t)
ϕ(x, t) = exp . (6)
2μ
Substituting (6) and the above equalities into (5) and canceling the common
factors, we discover that if ϕ(x, t) is a solution of the diffusion equation (5),
then h(x, t) appearing in the exponent in (6) obeys the KPZ equation (1).
Thus, every solution of the KPZ equation (1) can be expressed in terms
of a solution of the linear diffusion equation (5) via the logarithmic transfor-
mation
h(x, t) = 2μ ln ϕ(x, t) . (7)
To ensure that this solution satisfies the initial condition indicated in (1),
it is necessary that equation (5) be solved taking into account the initial
condition
h0 (x)
ϕ0 (x) = exp . (8)
2μ
Formula (7), reducing the nonlinear KPZ equation (1) to a linear diffusion
equation (5), is called the Hopf–Cole substitution.
Differentiating (7) with respect to x and taking (2) into account, we ar-
rive at the conclusion that the general solution of the inhomogeneous Burgers
256 Chapter 14. 1-D Nonlinear PDEs of Higher Order
equation (3) is expressed in terms of the solution of the initial value prob-
lem (5) and (8) as follows:
∂
v(x, t) = −2μ ln ϕ(x, t), (9)
∂x
or equivalently,
2μ ∂ϕ
v(x, t) = − . (9a)
ϕ ∂x
In particular, the Hopf–Cole substitution (9) reduces the homogeneous Burg-
ers equation
∂v ∂v ∂ 2v
+v =μ 2, v(x, t = 0) = v0 (x), (10)
∂t ∂x ∂x
to the linear diffusion equation
∂ϕ ∂ 2ϕ s0 (x)
=μ 2, ϕ0 (x) = exp − , (11)
∂t ∂x 2μ
where x
s0 (x) = v0 (x ) dx (12)
is the initial potential of the solution of the Burgers equation. We have en-
countered it in the previous chapters. The solution of the initial value problem
for the linear diffusion equation (11) is a convolution of the initial data with
the Gaussian kernel,
1 (y − x)2
ϕ(x, t) = √ ϕ0 (y) exp − dy, (13)
2 πμt 4μ t
and has been discussed in Chap. 10. In our case, it takes the form
1 1 (y − x)2
ϕ(x, t) = √ exp − s0 (y)t + dy . (13a)
2 πμt 2μt 2
Note that the expression in the exponent contains the function (13.4.10),
1
G(y; x, t) = s0 (y)t + (y − x)2 , (14)
2
which we have already encountered in the global minimum principle. Utilizing
this notation, the solution ϕ(x, t) can be written in a more compact form,
1 G(y; x, t)
ϕ(x, t) = √ exp − dy , (15)
2 πμt 2μt
which is known as the Hopf–Cole formula.
14.3. General Solutions of Burgers Equation 257
In other words, the brace operation applied to the function g(y) is defined
by the equality
{g(y)}(x, t) := g(y)f (y; x, t) dy . (18)
which proves our assertion. The monotonicity property of the average La-
grangian coordinate, together with (16), implies that every solution of the
Burgers equation satisfies the inequality
∂v(x, t) 1
≤ . (20)
∂x t
(2) Let {y}(x, t) and {y}(x, t|V ) be the average Lagrangian coordinates cor-
responding to the initial fields v0 (x) and v0 (x) + V , respectively. Then
x−Vt x
V + ≡ .
t t
Finally, let us observe that the weak limit limμ→0+ f (y; x, t) is given by
where yw (x, t) is the coordinate of the global minimum of the function (14).
Consequently, as μ → 0+ , the solution of the Burgers equation converges to
the weak solution (13.4.12) of the Riemann equation found via the Oleinik–
Lax global minimum principle.
where R is the initial Reynolds number, familiar from the self-similar solution
(14.2.39), and Φ(z) is the auxiliary function (14.2.36).
Now we can find the desired solution of the Burgers equation with the
singular initial condition by substituting (3) into (14.3.9a). Naturally, it co-
incides with the self-similar solution (14.2.41), but the time spent on finding
it via a different route has not been wasted. Indeed, the present approach will
now permit us to apply our knowledge of properties of solutions of the linear
diffusion equation to obtain a better understanding of the universal impor-
tance of self-similar solutions of the Burgers equation; up to this point, they
were considered to be just another type of special solution. More specifically,
relying on known properties of solutions of the linear diffusion equation, we
arrive at the following result: If the initial field v0 (x) is integrable, with
v0 (x) dx = S ,
3
See also E. Zuazua, Weakly nonlinear large time behavior in scalar convection–
diffusion equation, Differential and Integral Equations 6 (1993), 1481–1491.
14.4. Evolution and Characteristic Regimes for Burgers Equation 261
then for every arbitrarily large initial Reynolds number R, the solution of the
Burgers equation is asymptotically described by the linear diffusion equation.
In other words, the nonlinear stage of the evolution of the field v(x, t) is
replaced by the linear stage.
Let us begin by finding the asymptotics of solutions of the Burgers equa-
tion in the linear stage. Observe that inequality (4), applied to the potential
s0 (x) of the initial field v0 (x), implies that the limiting values of s0 (x) as
x → ±∞ are identical:
lim s0 (x) = lim s0 (x) .
x→−∞ x→∞
where
Q0
Q(t) = √ , Q0 = q0 (y) dy . (9)
2 πμt
Substituting (8) into the right-hand side of (14.3.9a), we obtain the following
asymptotics of the solution of the Burgers equation:
x Q(t)
v(x, t) ∼ . (10)
t exp x2
+ Q(t)
4μt
FIGURE 14.4.1
A singular initial field v0 (x), see (13), (top), and the corresponding
fields s0 (x) and q0 (x) (bottom). Bold vertical arrows in the top graph
symbolize the Dirac delta components of the initial condition (13).
So in this case, the transition time to the linear regime is close to the
characteristic time td of the linear diffusion; see (7).
Now let us trace the evolution of the field v(x, t) corresponding to the
initial condition (13) employing the exact solution of the Burgers equation
2μ R
v(x, t) = − × (15)
πτ
2
1 − e−R exp −R z τ+1 sinh 2Rz τ
,
1 − (1 − e−R ) Φ R
τ
(1 − z) − Φ − R
τ
(1 + z)
where the dimensionless time and space coordinates are
x 2St
z= , τ= ,
2
264 Chapter 14. 1-D Nonlinear PDEs of Higher Order
FIGURE 14.4.2
The N -wave solution of the Burgers equation at various times τ .
The initial condition is that of (13), and the initial Reynolds num-
ber is R=25. The scales of the horizontal axes in all four pictures
are identical, but the scales on the vertical axes are different. The
vertical scale in the last picture corresponding to τ = 30 is much
greater than that on the first three to make it possible to see the
shape of the solution, drastically attenuated by dissipation, of the
Burgers equation at the linear stage.
At time τ = 0.5, the field consists of two triangular waves moving to-
ward each other and practically noninteracting. At τ = 1, the fronts of the
triangular waves meet and begin to annihilate each other. By τ = 2, their
mutual absorption has substantially reduced their amplitudes. Nevertheless,
the absolute value of the parameter (9),
4R
Q(τ ) − ,
πτ
is still large (|Q(τ = 1)| 4), and the nonlinear stage of the evolution
continues. For τ = 1 and τ = 2, the characteristic profiles of the field v(x, t)
resemble the letter N . Hence the name N -waves. Finally, for τ = 30, when
|Q(τ = 30)| 1, the process of mutual destruction of colliding triangular
14.4. Evolution and Characteristic Regimes for Burgers Equation 265
At first glance, this solution looks similar to the N -wave solution (15).
However, on closer inspection, it turns out that its behavior is drastically
different. In the present case, the triangular waves generated by the Dirac-
delta-shaped components of the initial conditions (16) run away from each
other (see Fig. 14.4.3). Thus, unlike the N -waves, the triangular components
of the U -waves interact only through their tails. For large initial Reynolds
numbers, this interaction is significantly weaker than the mutually destruc-
tive collision of the fronts of N -waves. As a result, the nonlinear stage of
U -waves lasts much longer than that of N -waves.
Let us confirm the above conclusions by a quantitative evaluation of the
transition time (11) to the linear regime. Indeed, in this case,
4 R
tl = e − 1 td .
π
Thus, for the same (large) initial Reynolds number, the nonlinear stage of a
U -wave lasts incomparably longer—approximately 1010 times as long—than
the similar stage of an N -wave.
FIGURE 14.4.3
U -wave solution of the Burgers equation at various times τ . The
initial condition is that of (16), and the initial Reynolds number is
R = 25.
of the initial condition v0 (x), but only on the waves’ integral characteristic
Q0 ; see (9). If the latter is positive, then the field eventually assumes the
N -wave shape. If Q0 < 0, then the field converges to a U -wave. Relying on
the universal asymptotic formula (10), Fig. 14.4.4 shows a graph of an exact
solution of the Burgers equation with initial condition (16) for R = 25 and
τ = 100 and compares it with a graph of the corresponding U -wave. Their
similarity is obvious.
FIGURE 14.4.4
An exact solution of the Burgers equation with initial condition (16)
for R = 25 and τ = 100, compared with the graph of the U -wave (10)
for the corresponding value Q 8 · 109 .
2
x Q exp
x
2μt
− exp − 2μt
x
v(x, t) = − .
t t Q2 exp x
+ exp − 2μt
x
2μt
This is the familiar solution (14.2.31) of the Burgers equation moving with
velocity V .
268 Chapter 14. 1-D Nonlinear PDEs of Higher Order
The resulting solution field v(x, t) will be called a sawtooth wave, and we
shall write it in the form (14.3.16),
x − {y}(x, t)
v(x, t) = , (21)
t
where {y}(x, t) is the averaged Lagrangian coordinate, which in this partic-
ular case is
1 (x − yk )2
yk exp − sk t +
2μt 2
{y}(x, t) = k
. (22)
1 (x − yk )2
exp − sk t +
k
2μt 2
(x − yk )2
Πk (x) = sk t + (23)
2
and selection of the one that has the minimal value for x. Accordingly,
{y}(x, t) turns out to be a step function,
{y}(x, t) = yk ,
At the points where the minimal critical parabolas intersect, the skeleton
of the function {y}(x, t) has a jump. The coordinates of the points xk,m (t) of
the intersection of the kth and mth parabolas can be easily found. Indeed,
yk + ym sm − sk
x0k,m = , Vk,m = , Mk,m = ym − yk (k < m) . (24)
2 Mkm
If the parabolas Πk and Πm immediately to the left and to the right of
the point xk,m (t) are below other parabolas, then the function {y}(x, t) has
a jump at xk,m (t), and
The fact that the velocities of jumps are constant between collisions and
that they change at the point of impact according to the law of inelastically
colliding particles allows us to describe the behavior of the sawtooth wave in
the language of a 1-D stream of inelastically colliding particles. The skeletons
of the sawtooth field v(x, t), constructed via the formula (14.3.16), are shown
270 Chapter 14. 1-D Nonlinear PDEs of Higher Order
FIGURE 14.4.5
Construction of osculating parabolas of the skeleton of the function
{y}(x, t). The parabolas in the upper part of the picture are drawn
for t = 0, and those in the middle for t > 0. One can see that
the latter emerged from a competitive struggle for the right to
achieve the least value. As a result, instead of two jumps of the
function {y}(x, 0) at the points x0k−1,k and x0k,k+1 , there is only one
jump of{y}(x, t) at the point xk−1,k+1 (t).
FIGURE 14.4.6
The skeleton of a sawtooth wave created as a result of a nonlinear
transformation of the initial field. The zeros of the linear portions
of the graph, as well as the jump coordinates, are indicated.
where
∂ 2 s0 (x) ∂v0 (x)
σk = σ(yk ) , σ(x) = 2
= . (27)
∂x ∂x
In what follows, we shall assume that there exists an ε such that for all k,
σk > ε > 0 .
εt 1 , (29)
the function ϕk (x, t) (28) will not change much if the right-hand side of (26)
is replaced by the Dirac delta:
4πμ sk
ϕ0 (x) ∼
k
exp − δ(x − yk ) .
σk 2μ
272 Chapter 14. 1-D Nonlinear PDEs of Higher Order
FIGURE 14.4.7
Pattern of trajectories of merging jumps. The solid lines mark tra-
jectories of jump points before and after they underwent collisions
with other jump points.
Combining the contributions of all the minima of the initial potential and
taking into account the fact that in the case of a solution of the Burgers
equation, the initial potential is defined up to an arbitrary constant factor,
we arrive at the following assertion: If the condition (29) is satisfied, then the
solution of the Burgers equation can be expressed via the formula (14.3.9a),
where ϕ(x, t) is the solution of the linear diffusion equation with initial con-
dition (19), and
1 sk
Qk = √ exp − .
σk 2μ
The factor appearing in front of the exponential function can be treated
as a small correction to sk :
1 sk s μ
√ exp − = exp − k , sk = Sk + ln σk ,
σk 2μ 2μ 2
which can be neglected for small μ. As a result, we arrive at the initial
conditions (19) and (20).
To begin with, let us take a look at the case in which the initial field is a
simple harmonic function,
a
v0 (x) = a sin(kx) ⇒ s0 (x) = − cos(κx) . (30)
k
The question of applicability of the sawtooth approximation needs to be
clarified first. In this case, the characteristic curvature is the same at all the
minimum points of the initial potential and equals
1 1
σk = ε = aκ = , tn = ,
tn aκ
where tn is the time of the jump creation. Accordingly, one can rewrite the
condition (29) in the form
t tn . (31)
Let us find a sawtooth wave corresponding to the initial harmonic
field (30). Since the minima s0 (x) of the potential are identical, the val-
ues sk in (22) can be assumed to be equal to zero. Moreover, yk = 2πk/κ.
As a result, the expression for the sawtooth wave approximating the original
harmonic wave is of the form
⎛ ⎞
R
k exp − (z − 2πk) 2
a⎜⎜ 8τ ⎟
v(x, y) = ⎜z − 2π k
⎟⎟, (32)
τ⎝ R ⎠
exp − (z − 2πk) 2
k
8τ
where
2a
z = κx , τ = aκt , R= . (33)
μκ
Plots of the field v(x, t) in (32), for different values of τ , are shown in
Fig. 14.4.8.
The Reynolds number of the initial harmonic wave varies in time. If R1,
then there are three, clearly separated, stages of evolution of the field v(x, t).
For times
t < tn ,
there are no jumps, and the current Reynolds number is close to the initial
one:
a πa π
R(t) = R ( = ) .
μ μκ κ
274 Chapter 14. 1-D Nonlinear PDEs of Higher Order
FIGURE 14.4.8
The sawtooth wave in (32) at R = 100. The vertical scales of the
plots are selected so that the shape of the field v(x, t) is visible.
As τ becomes larger, the jumps are smoothed out, and the field
assumes a sine-like shape, characteristic of the linear stage of the
evolution.
For t tn , the size of the jump U (t) and its width Δ(t) change according
to the formulas
2μ
U (t) , and Δ(t) ,
t U (t)
so that the Reynolds number decays like 1/t,
U (t) 2
R(t) .
μ μt
Finally, for t ∼ tl , where tl is the time when the jump’s width becomes
comparable to the half-period of the field v(x, t),
π2
Δ(tl ) ⇒ tl R tn ,
2μκ2
we see the onset of the linear stage, and the Reynolds number tends to zero
quickly. A semiqualitative plot of the current Reynolds number R(t) of the
initial harmonic field v(x, t) is shown in Fig. 14.4.9.
14.4. Evolution and Characteristic Regimes for Burgers Equation 275
FIGURE 14.4.9
Semiqualitative plot of the dependence of the current Reynolds
number of a periodic field on the dimensionless time τ = aκt. It is
assumed that the initial Reynolds number is large, R 1.
FIGURE 14.4.10
Evolution of the solution of the Burgers equation with the initial
condition (34) and R = 50. Merging of jumps leads to creation of a
sawtooth wave with period equal to the period of the large-scale
component of the initial field.
14.5 Exercises
1. The functional
u(x, t) dx = const
5. Find an exact solution of the Burgers equation with the initial condition
v0 (x) = S [δ(x) + δ(x − )] ( > 0 , S > 0) . (3)
Construct plots of v(x, t) for R = S/2μ 1 and several values of the
dimensionless time τ = 2St/2 .
6. Find the skeleton of the solution of the previous problem. Using the
fact that a jump’s velocity is the average of the values of the field v(x, t)
to the left and to the right of the jump, write equations of motion of
the jumps and estimate the time of merger of two triangular waves into
a single triangular wave.
7. Find the form of jumps of the averaged Lagrangian coordinate {y}(x, t),
see (14.4.19), taking into account the competition of two dominating
summands in the vicinity of the jump.
8 Using the well-known formula
∞
R cos(κx)
e = I0 (R) + 2 In (R) cos(nκx) , (4)
n=1
where In (z) is the modified Bessel function of order n, find the asymp-
totic behavior (at the linear stage) of the solution of the Burgers equa-
tion with a harmonic initial condition
v0 (x) = a sin(κx) . (5)
Explore the dependence of the solution field’s amplitude on the initial
amplitude a at the linear stage.
9. Solve the Burgers equation in the case of
ϕ0 (x) = x2 .
This chapter builds on the material of Chap. 14 and reviews other stan-
dard nonlinear models that can be described by partial differential equations.
We begin with the model equations of gas dynamics, expand the Burgers–
KPZ model to the multidimensional case and the related concentration fields,
study in detail the Korteweg–de Vries (KdV) equations, in which one can ob-
serve the creation of solitary waves (solitons), and finally, discuss nonlinear
flows in porous media.
There are new ingredients here. The KdV equation includes partial deriva-
tives of order three, and the porous medium equation contains nonlinearity
intertwined with the highest (second) derivative. The latter complicates the
game considerably. All the other nonlinear equations we have considered thus
far were linear in the highest derivative; in the mathematical literature, such
equations are called quasilinear. The porous medium equation contains the
Laplacian superimposed on top of a nonlinearity and thus is classified as a
strongly nonlinear equation. Nevertheless, the reader will notice that the non-
linearities we consider are always quadratic or of power type. The quadratic
ones are the simplest possible, obtained by retaining the first nonlinear term
in the Taylor expansions of more complex, and perhaps more realistic, non-
linearities that describe various physical phenomena.
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 281
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3 7,
© Springer Science+Business Media New York 2013
282 Chapter 15. Other Nonlinear Models of Higher Order
a role as model equations of gas dynamics. For now, we will stay in the
one-dimensional universe.
∂v ∂v ∂ 2v
+v =μ 2 (1)
∂t ∂x ∂x
was proposed by Johannes Burgers as a model of strong hydrodynamic tur-
bulence that takes into account competing influences of inertial nonlinearity
and viscosity. However, from the physics perspective, the Burgers equation
does not provide an accurate description of the gas dynamics phenomena.
Most importantly, it does not take into account pressure forces that prevent
excessive compression of the particles. To model these forces, let us modify
equation (1) by an additional term to obtain the equation
∂v ∂v 1 ∂P ∂ 2v
+v + =μ 2. (2)
∂t ∂x ρ ∂x ∂x
The quantity P (x, t) introduced above represents the pressure of the model
gas, and ρ(x, t) its density.
In the remainder of this section, we shall restrict our attention to the
case of so-called polytropic gas, whereby gas pressure and gas density are
tied together by the relation
κ2 γ
P (x, t) = ρ (x, t) , κ > 0, γ > 0. (3)
γ
Equations (2)–(3) describe the behavior of just the velocity field and
are not closed, since they contain three unknown fields, v(x, t), P (x, t) and
ρ(x, t). Fortunately, we also know that the velocity and density fields are tied
together by the continuity equation
∂ρ ∂ ∂ 2ρ
+ (ρv) = ν 2 , (4)
∂t ∂x ∂x
where ν is the coefficient of molecular diffusion, which takes into account the
influence of the Brownian motion of gas molecules on the evolution of the
gas density.
15.1. Model Equations of Gas Dynamics 283
γ = 3, and ν = μ, (5)
equations (2)–(4) have an explicit analytic solution for a broad class of initial
conditions
v(x, t = 0) = v0 (x) , ρ(x, t = 0) = ρ0 (x) . (6)
This fact becomes clear once we rewrite (2) and (4), taking (5) into account,
as a set of equations
∂v ∂v ∂c ∂ 2v ∂c ∂c ∂v ∂ 2c
+v +c =μ 2, +v +c =μ 2. (7)
∂t ∂x ∂x ∂x ∂t ∂x ∂x ∂x
The quantity
c(x, t) = κρ(x, t) (8)
introduced above is called the local sound velocity. For convenience, we shall
also introduce two auxiliary fields
so that the original velocity and density fields can be expressed by the for-
mulas
u+ (x, t) + u− (x, t) u+ (x, t) − u− (x, t)
v(x, t) = , ρ(x, t) = . (10)
2 2κ
Adding and subtracting equations (7), we discover that each of the auxiliary
fields (9) obeys its own Burgers equation
∂u± ∂u± ∂ 2 u±
+ u± =μ , (11)
∂t ∂x ∂x2
with the initial conditions
From the perspective of physics, the most fundamental case is that of the
constant initial density
The crux of the matter is that, in this particular case, the solutions obey the
momentum conservation law
ρ(x, t)v(x, t) dx = const. (14)
For this reason, we shall analyze this case in greater detail and observe first
that the initial conditions (12) of the Burgers equations (11) assume the form
and due to the Galilean invariance (14.2.3), the desired solutions of equa-
tions (11) have the form
The remaining auxiliary field u(x, t) satisfies the Burgers equation with the
standard initial condition
∂u ∂u ∂ 2u
+u =μ 2, u(x, t = 0) = v0 (x) . (16)
∂t ∂x ∂x
Substituting (15) into (10), we can now express solutions of the model gas
dynamics equations (7) through a solution of the Burgers equation (16):
1
v(x, t) = [u(x − c0 t, t) + u(x + c0 t, t)] (17)
2
and
u(x + c0 t, t) − u(x − c0 t, t)
ρ(x, t) = ρ0 1− . (18)
2c0
Let us verify that the above solutions obey the momentum conservation
law (14). Indeed, multiplying the right-hand sides of (17) and (18), we obtain
ρ0
ρ(x, t) v(x, t) = [u(x − c0 t, t) + u(x + c0 t, t)]
2
1 2
+ u (x − c0 t, t) − u2 (x + c0 t, t) . (19)
2κ
Integrating both sides of the above equality with respect to x in infinite limits,
we find that the integral of the last term on the right-hand side vanishes in
view of its symmetry. The integral of the first term is independent of time,
because the quantity
u(x, t) dx = v0 (x) dx = const
15.1. Model Equations of Gas Dynamics 285
is an invariant of the solution of the Burgers equation (16). Thus, the total
momentum of the model gas does not depend on time and equals
ρ(x, t)v(x, t) dx = ρ0 v0 (x) dx .
1
v(x, t) = [v0 (x − c0 t) + v0 (x + c0 t)] .
2
1
vw (x, t) = [uw (x − c0 t, t) + uw (x + c0 t, t)] ,
2
uw (x + c0 t, t) − uw (x − c0 t, t)
ρ(x, t) = ρ0 1− , (24)
2c0
can be viewed as a weak solution of the system of 1-D polytropic gas equations
∂v ∂v ∂ρ ∂ρ ∂
+v + κ2 ρ = 0, + (ρv) = 0 , (25)
∂t ∂x ∂x ∂t ∂x
with the polytropic exponent γ = 3. Observe that they satisfy the momentum
conservation law. In this case, for large times (t ∼ tn ), the fields v(x, t) and
ρ(x, t) become discontinuous, but they remain bounded as a result of pressure
forces that prevent creation of domains of high density.
x Ut
z= , τ= ,
and consider the dimensionless velocity and density fields
v ρ
V (z, τ ) = , R(z, τ ) = .
U ρ0
15.1. Model Equations of Gas Dynamics 287
From (24) and (14.2.27) we see that the above fields are as follows:
1 τ τ
V (z, τ ) = g z+ +g z− ,
2(1 + τ ) M M
M τ τ
R(z, τ ) = 1 − g z+ −g z− , (26)
2(1 + τ ) M M
g(z) = z − sign(z) .
FIGURE 15.1.1
Graphs of the generalized velocity and density fields (26) of a poly-
tropic gas for Mach number M = 2. They are shown here for times
τ = 0.1 and τ = 0.3.
Their plots are shown in Fig. 15.1.1. It is clear that in the vicinity of the
point z = 0, where the jump of the initial velocity field was located, an area
of increased density is created. For larger times, the pressure forces lead to
an expansion of this area and to a reduction of the density within it.
Finally, let us discuss the behavior of a model gas in the limiting case of
Mach number M → ∞. Physically, this limit corresponds to a pressureless
gas, whose evolution is determined solely by a competition between the iner-
tial nonlinearity and viscosity. In this case, equations of the model gas take
the form
288 Chapter 15. Other Nonlinear Models of Higher Order
∂v ∂v ∂ 2v ∂ρ ∂ ∂ 2ρ
+v =μ 2, + (ρv) = μ 2 , (27)
∂t ∂x ∂x ∂t ∂x ∂x
and the density field is described by the formula
∂v(x, t)
ρ(x, t) = ρ0 1 − t , (28)
∂x
which follows from the formulas (17) and (18) as c0 → 0. For μ > 0, the
density field (28) is uniformly bounded. However, in contrast to the previous
case, in which an increase of the density was restricted by pressure forces, in
the present case, the boundedness of the density is caused by viscosity and
molecular diffusion.
FIGURE 15.1.2
Density plots for a pressureless gas with initial momentum applied
at the origin. The resulting motion leads to creation of a high-
density area trailed to the left by a low density tail. As the Reynolds
number increases (i.e., μ decreases), the maximum of the density
increases as well.
Sample plots of this density, for different values of the Reynolds number
R = S/μ, are shown in Fig. 15.1.2.
z = ct + h(x, t) . (1)
The first term on the right-hand side represents the growth of the flat surface
perpendicular to the z-axis, and the second takes into account the surface’s
fluctuations and other effects influencing the surface growth rate. For sim-
plicity’s sake let us assume that the constant growth rate is c = 1. Assuming
that the angles between the surface gradient vectors and the z-axis are small,
the elevation function h(x, t) obeys, in this small angle approximation, a mul-
tidimensional analogue of the equation (14.1.4):
∂h 1
= (∇h)2 + μΔh + F (x, t) . (2)
∂t 2
1
For more and in-depth information, see, S. Gurbatov, A. Malakhov and A. Saichev,
Nonlinear Random Waves and Turbulence in Nondispersive Media: Waves, Rays and Par-
ticles, Manchester University Press, Manchester–New York, 1991, and W.A. Woyczyński,
Burgers–KPZ Turbulence: Göttingen Lectures, Springer-Verlag, Berlin–Heidelberg, 1998.
290 Chapter 15. Other Nonlinear Models of Higher Order
Here F (x, t) takes into account the nonuniformity of the stream of particles
deposited on the surface, and the term containing μ reflects the diffusive
mobility of the particles along the surface.
The logarithmic substitution
h(x, t) = 2μ ln ϕ(x, t) , (3)
similar to that for equation (14.1.7), reduces equation (2) to the linear diffu-
sion equation
∂ϕ 1
= μΔϕ + F (x, t)ϕ , (4)
∂t 2μ
which needs to be solved with the initial conditions
h0 (x)
ϕ(x, t = 0) = exp . (5)
2μ
A complete analysis of the properties of solutions of equation (4) and
relevant solutions of the KPZ equation is outside the scope of this book. In
this section, we shall provide an explicit solution of equations (2) and (4)
only in the homogeneous case, that is, when F (x, t) ≡ 0. In this case,
n
1 1 (y − x)2
ϕ(x, t) = √ . . . exp h0 (y)t − dn y . (6)
2 πμt 2μt 2
Here n is the dimension of the x-space; in our context, n = 2.
The potentiality of this field means that there exists a scalar potential s0 (x)
such that
v 0 (x) = ∇s0 (x) . (10)
In this case, the solution v(x, t) of the Burgers equation remains potential
for every t > 0, and is described by the expression
x − {y}(x, t)
v(x, t) = , (11)
t
which follows from (6) and (7). By analogy with (14.3.16), the braces {. . . }
denote here the spatial average with respect to the weight function
exp − 2μt
1
G(y; x, t)
f (y; x, t) = , (12)
. . . exp − 2μt1
G(y; x, t) dn y
where
(y − x)2
G(y; x, t) = s0 (y) t + . (13)
2
Recall more generally that in our present context, the space average {g(y)}
is defined as the n-tuple integral
{g(y)}(x, t) := . . . g(y) f (y; x, t) dn y . (14)
equation (2) to the linear parabolic equation (4), and we shall hope that
similar substitutions applied to a system of linear parabolic equations will
produce nonlinear equations with a promising interpretation in the theory of
interfacial growth or gas dynamics.
Let us begin by considering a pair of interdependent linear parabolic
equations
∂ϕ ∂ψ
= μΔϕ + a ϕ + b ψ , = μΔψ + d ϕ + e ψ . (15)
∂t ∂t
We shall apply the substitution (3) to the first equation. For this purpose,
multiply the first equation (15) by 2μ/ϕ. As a result, we arrive at the equation
∂h 1
= (∇h)2 + μΔh + 2μ a + 2μ b C , (16)
∂t 2
where
ψ(x, t)
C(x, t) = . (17)
ϕ(x, t)
Now let us find an equation satisfied by the field C(x, t) (23). First, write
out the time derivative of this field:
∂C 1 ∂ψ ψ ∂ϕ
= − 2 . (18)
∂t ϕ ∂t ϕ ∂t
Substituting for ϕ and ψ the right-hand sides of the equations (15), one
obtains
∂C Δψ ψΔϕ
=μ − + d + (e − a) C − b C 2 . (19)
∂t ϕ ϕ2
To complete the operation, it is necessary to express, through C and h, the
first term of the right-hand side. This can be accomplished by noticing that
a formula analogous to (18) holds for the gradient of the field C:
∇ ψ ψ∇ ϕ
∇C = − . (20)
ϕ ϕ2
In turn, taking the scalar product of both sides of this equality with the
vector ∇, we find that
Δψ ψΔϕ 2 (∇ ϕ · ∇ ψ) ψ(∇ϕ · ∇ϕ)
ΔC = − − − . (21)
ϕ ϕ2 ϕ ϕ ϕ2
15.2. Multidimensional Nonlinear Equations 293
Taking into account the relation (20) and the expression for the gradient of
field h,
∇ϕ
∇ h = 2μ ,
ϕ
we can rewrite formula (21) in the convenient form
Δψ ψΔϕ 1
− 2
= Δ C + (∇ h · ∇)C . (22)
ϕ ϕ μ
In view of this relation, equation (19) now assumes the form
∂C
= μΔ C + (∇ h · ∇)C + d + (e − a) C − b C 2 , (23)
∂t
containing only the fields h and C.
In this fashion, we have arrived at a closed system of nonlinear equations
(16) and (23), whose solutions can be obtained explicitly from the solution
of the linear parabolic equations (15). The arguments in favor of the physical
relevance of the above system, for arbitrary parameters a, b, d, e, are outside
the scope of this book. However, we will provide an example of such an
interpretation of equations (16) and (23) in the special case that
1
a=e= F (x, t) , b ≡ 0, d = d(x, t) . (24)
2μ
In addition, as we have done before, we shall pass from the field h to the
potential field v(x, t) (7), interpreting it as a gas velocity field. In view of (16)
and (23), the relevant equations for the fields v(x, t) and C(x, t) have the
form
∂v ∂C
+ (v · ∇) v = μΔv + f (x, t) , + (v · ∇) C = μΔC + d(x, t) . (25)
∂t ∂t
We are already familiar with the first one, so at this point, we shall show only
that the second equation also has a physical meaning. Indeed, it describes
the evolution of the concentration C(x, t) of a passive tracer suspended in a
gas moving with the velocity field v(x, t). Recall that the passive tracer con-
centration is equal to the ratio of the passive tracer density and the medium
density and that it does not vary with compression or rarefaction of the
medium. Therefore, disregarding the influence of the molecular diffusion, the
passive tracer concentration satisfies the following equation:
DC ∂C
=0 ⇒ + (v · ∇) C = 0.
Dt ∂t
294 Chapter 15. Other Nonlinear Models of Higher Order
Equation (25) for the field C takes into account, in addition to the hydrody-
namic transport of passive tracer, molecular diffusion with diffusivity μ and
the presence of the passive tracer’s sources described by the function d(x, t).
Also, note that the first equation in (25) includes a potential force field
f (x, t) = −∇ F (x, t) ,
which influences the motion of the medium.
The above calculations show that the solution of the coupled equa-
tions (26) with the initial conditions
v(x, t = 0) = v 0 (x) , C(x, t = 0) = C0 (x) , (26)
can be reduced by substitutions
∇ϕ ψ
v = −2μ , C= , (27)
ϕ ϕ
to a solution of the linear equations of parabolic type
∂ϕ 1 ∂ψ 1
= μΔϕ + F (x, t)ϕ , = μΔψ + F (x, t)ψ + d(x, t) ϕ , (28)
∂t 2μ ∂t 2μ
with the initial conditions
s0 (x) s0 (x)
ϕ0 (x, t = 0) = exp − , ψ0 (x) = C0 (x) exp − . (29)
2μ 2μ
Here s0 (x) is the potential of the initial velocity field, connected with the
latter via the equality (10).
In particular, in the absence of external forces (f ≡ 0) and sources of
passive tracer (d ≡ 0), solutions of equations (26) take the form
x − {y}(x, t)
v(x, t) = , C(x, t) = {C0 (y)}(x, t) , (30)
t
where the braces signify, as before, spatial averaging defined by equalities (12)
and (14).
for the field v(x, t), which we interpret as the velocity of a hydrodynamic flow
of particles, each of them moving uniformly along the x-axis with constant
velocity. Accordingly, v0 (x) describes the dependence of the particles’ veloci-
ties on their spatial coordinate x at the initial time t = 0. As an illustration,
let us consider first the Gaussian initial velocity
x2
v0 (x) = V0 exp − 2 . (2)
2
In Fig. 12.1.1, we have already seen that due to different velocities of dif-
ferent particles, the right-hand side of the velocity field profile v(x, t) be-
comes, in the course of time t > 0, progressively steeper. Since the Riemann
equation (1) describes the motion of noninteracting particles, the steepening
fronts lead, in finite time, to the so-called gradient catastrophe described in
Chap. 13 and subsequently to appearance of the multistream regime shown
in Fig. 13.2.3 for the same Gaussian initial profile (2). In reality, in most ap-
plications of the Riemann equation, the arrival of the gradient catastrophe
and the multistream regime means that the Riemann equation no longer ad-
equately describes the particular physical phenomena under consideration;
gradient catastrophes and multistream regimes are obviously prohibited for
them. In these cases one has to replace the Riemann equation by an equa-
tion that can take into account more subtle physical effects that prevent the
infinite steepening of the field v(x, t) and its multistream behavior.
Recall that in Chap. 14, we already discussed a regularization of the
Riemann equation by taking into account the viscosity effects and replacing
equation (1) by the Burgers equation
∂v ∂v ∂ 2v
+v = μ 2, v(x, 0) = v0 (x) . (3)
∂t ∂x ∂x
The main characteristic peculiarity of the field v(x, t) satisfying the Burg-
ers equation is dissipation of energy of the field v(x, t) due to the viscous
term appearing at the end of equation (3). Nevertheless (see Sect. 14.1.2),
many nonlinear fields satisfy the energy conservation law while being sub-
ject to dispersive effects. The simplest equation taking into account both
the nonlinearity and the dispersion effects that at the same time is suitable
for diverse physical applications is the Korteweg–de Vries (KdV) equation
(14.1.11), which we rewrite here in the form
∂v ∂v ∂ 3v
+v + γ 3 = 0, v(x, 0) = v0 (x) . (4)
∂t ∂x ∂x
296 Chapter 15. Other Nonlinear Models of Higher Order
The relation ⇐⇒ means that the left-hand side in (5) satisfies the KdV
equation if and only if its right-hand side does. Note that the dependence of
the solution v(x, t, γ) on the parameter γ was explicitly indicated because its
role in the symmetry property (5) is essential.
One can say that relation (5) (excluding the changing sign of the pa-
rameter γ) resembles the symmetry property (14.2.2) of the solutions to the
Burgers equation. On the other hand, reflection in time produces the follow-
ing symmetry property for the KdV equation:
Another obvious symmetry property of both the Burgers and KdV equations
is their translational invariance (14.2.1). In particular,
x v(x, t)
τ= t, s= , u(s, τ ) = .
U U
Passing in the KdV equation (4) to the dimensionless field u(s, τ ), we can
now rewrite (4) in the nondimensional form
∂u ∂u ∂ 3u
+u + γ 3 = 0, u(s, 0) = u0 (s) . (11)
∂τ ∂s ∂s
Here, the nondimensional factor
1 U 2
γ= , where D= , (12)
D γ
u = αw , τ = βθ , α, β = 0 .
As a result, the initial value problem for the new renormalized field w(s, θ)
takes the form
∂w ∂w ∂ 3w 1
+ αβw + βγ 3 = 0, w(s, 0) = u0 (s) .
∂θ ∂s ∂s α
3
For clarity’s sake, note that the first equality in (12) defines a dimensionless dis-
persive factor, while the right-hand side of the second equation contains its dimensional
counterpart.
15.3. KdV Equation and Solitons 299
∂w ∂ 3 w ∂w D
+ 3 = 6w , w(s, 0) = −C u0 (s) , C= > 0. (14)
∂θ ∂s ∂s 6
Obviously, once we find a solution w(s, θ) of the initial value problem (13),
then the solution u(s, τ ) of the initial value problem (11) can be obtained
from the relation
u(s, τ ) = σγw(s, γτ ) . (15)
In the commonly considered case σ = −6, we have u(s, τ ) = −6γw(s, γτ ) .
τ =50
τ =40
0
1
τ =30
u(s,τ)
0
1
τ =20
0
1
τ =10
0
1
τ =0
0
−20 −15 −10 −5 0 5 10 15 20 25 30
s
FIGURE 15.3.1
Plots of the solution u(s, τ ) of the initial value problem (11), (10),
for τ = 0, 10, 20, 30, 40, 50, and for D = 10 (γ = 0.1). The main part
of the solution is a pulse (soliton) moving with uniform velocity
without changing its shape.
between the inertial nonlinearity and dispersive effects that is inherent for
waves obeying the KdV equation.
The existence of solitons is a remarkable feature of diverse nonlinear waves
in dispersive media. Below, we shall discuss in some detail properties of soli-
tons in the context of KdV equations. In later subsections we shall find the
exact analytic shape of the KdV solitons, but for now, let us continue the
numerical investigation of the field u(s, τ ) and its characteristic properties.
First, let us elucidate the influence of the nonlinearity on the behavior of so-
lutions of the KdV equation. We already observed that the larger the value
of the number D (12) (the smaller the factor γ in equation (11)), the stronger
becomes the influence of nonlinearity on the field u(s, τ ). We shall study this
phenomenon quantitatively.
Figure 15.3.2 shows plots of the field u(s, τ ) for different values of τ and
for D = 25 (γ = 0.04). In this, more nonlinear, case, not one but two solitons
arise. They have different amplitudes, and they move uniformly with different
velocities. Additionally, notice that the soliton with the higher amplitude
moves faster. Numerical calculations for even larger values of the constant
D show that the more nonlinear field u(s, τ ) is (in our case, the larger the
15.3. KdV Equation and Solitons 301
τ =50
0
1
τ =40
0
1
τ =30
u(s,τ)
0
1
τ =20
0
1
τ =10
0
1
τ =0
0
−20 −15 −10 −5 0 5 10 15 20 25 30
s
FIGURE 15.3.2
Plots of the solution u(s, τ ) of the initial value problem (11), (10), in
the case D = 25, and for τ = 0, 10, 20, 30, 40, 50. The initial unimodal
Gaussian field splits into two solitons, both moving uniformly but
with different velocities.
number D), the larger is the number of solitons generated by the initial
unimodal field w(s, 0) (in our case, the unimodal Gaussian field (10)).
So, what about interactions of different solitons? The above exploration
indicates that as long as the larger (and faster) soliton is located to the right
of the smaller (slower) soliton, the two move happily, independently of each
other, preserving their shapes; the distance between them increases. However,
if the larger soliton is located to the left of the smaller soliton, the former will
start approaching the latter, and eventually the two will collide. To see what
happens at the collision time and thereafter, we have calculated numerically
a solution of the KdV equation (11), taking as the initial field the sum of
two separated Gaussian peaks,
2
s (s − )2
u(s, 0) = 2 exp − + exp − , = 10 . (16)
2 2
As it turns out, in this case there are three qualitatively different stages of
evolution of the solution of the KdV equation. We shall review them in the
case D = 10.
302 Chapter 15. Other Nonlinear Models of Higher Order
2
3
1
τ=5
0
2
τ=4
0
2
τ=3
u(s,τ)
0
2
τ=2
0
2
τ=1
0
2
τ=0
0
−5 0 5 10 15
s
FIGURE 15.3.3a
Plots of a numerically calculated solution u(s, τ ) of the initial value
problem (11), (16), for D = 10, and τ = 0, 1, 2, 3, 4, 5. The first Gaus-
sian peak (on the left) splits into two solitons, while the second
peak (on the right) generates a single soliton.
Figure 15.3.3a shows the plots of the field u(s, τ ) in the first stage τ ∈
(0, 5) when the original Gaussian peak on the left splits into two solitons,
while the smaller Gaussian peak on the right generates a single soliton. As a
result, at τ 5, we have three solitons present, which are marked 1, 2, and
3 (from left to right).
In the second stage, for τ ∈ (5, 30) (depicted in Fig. 15.3.3b), the largest
soliton (number 2), which moves with velocity higher than that of soliton
number 3, catches up with the third soliton and then overtakes it. Due to
nonlinearity, the resulting shape of the colliding solitons is not equal to the
shape of the solitons’ sum. After the collision, at times τ 25, the original
shapes and velocities of the solitons return. However, a more accurate track-
ing of the colliding solitons’ positions reveals a subtle shift disturbing their
uniform motion as if they had become tangled up during their collision. We
shall discuss this phenomenon in more detail in a later subsection.
Figure 15.3.3c shows the field u(s, τ ) in the last, third, stage, when solitons
are arranged, from left to right, by their increasing amplitudes (and velocities)
and move uniformly without collisions.
15.3. KdV Equation and Solitons 303
2
2 3
1
τ =30
0
2
τ =25
0
2
τ =20
u(s,τ)
0
2
τ =15
0
2
τ =10
0
2
2 3
1
τ =5
0
−10 −5 0 5 10 15 20 25 30 35
s
FIGURE 15.3.3b
Plots of a numerically calculated solution u(s, τ ) of the initial value
problem (11), (16), for D = 10, and τ ∈ (5, 30). The second and third
solitons are colliding nonlinearly, but after the collision, they return
to their original shapes.
2
2
3
1
τ = 80
0
2
τ = 70
0
2
τ = 60
0
u(s,τ)
τ = 50
0
2
τ = 40
0
2
2
3
1
τ = 30
0
−10 0 10 20 30 40 50 60 70 80
s
FIGURE 15.3.3c
Plots of a numerically calculated solution u(s, τ ) of the initial value
problem (11), (16), for D = 10, and τ > 30. After their nonlinear col-
lision, solitons 2 and 3 reverse their order but regain their original
shapes.
2
s
u0 (s) = − exp − . (17)
2
0.5 τ =5
0
−0.5
0.5 τ =4
0
−0.5
0.5 τ =3
0
−0.5
u(s,τ)
0.5 τ =2
0
−0.5
0.5 τ =1
0
−0.5
0.5 τ =0
0
−0.5
FIGURE 15.3.4a
Plots of the KdV field u(s, τ ) at time instants τ = 0, 1, 2, 3, 4, 5 in the
case D = 10 (γ = 0.1) and the negative initial condition (17).
0.5
τ =25
0
−0.5
0.5
τ =20
0
−0.5
0.5
τ =15
0
−0.5
u(s,τ)
0.5
τ =10
0
−0.5
0.5 τ =5
0
−0.5
τ =0
0,5
0
−0,5
−100 −90 −80 −70 −60 −50 −40 −30 −20 −10 0 10
s
FIGURE 15.3.4b
Plots of the KdV field u(s, τ ) at time instants τ = 0, 5, 10, 15, 20, 25 in
the case D = 10 (γ = 0.1) and the negative initial condition (17).
306 Chapter 15. Other Nonlinear Models of Higher Order
where
W (k) = −γκ3 ,
and ũ0 is the Fourier image of the initial field u0 (s). In the case of initial
condition (17), it is equal to
2
1 1 κ
ũ0 (κ) = u0 (s)e−iκτ ds = − √ exp − . (18)
2π 2π 2
−0.5
ulin(s,τ), u(s,τ)
0.5
linear dispersion
0
inertial nonlinearity
−0.5
−0,5
FIGURE 15.3.5
From bottom to top: Plots of (a) the linearized field ulin (s, τ ) (19),
(b) a solution of the KdV equation (11) satisfying the negative ini-
tial condition (17), and (c) solution u(s, τ ) of the KdV equation (11)
satisfying the positive initial condition (10).
Figure 15.3.5 shows (bottom to top): (a) linearized field ulin (s, τ ) (19),
(b) a solution of the KdV equation (11) satisfying the negative initial con-
dition (17), and (c) solution u(s, τ ) of the KdV equation (11) satisfying the
positive initial condition (10). One can see that in the first two cases, the
solutions are moving to the left. For the linearized field, this phenomenon
may be explained by the fact that the group velocity v(κ) (14.4.3) governing
the movement of wave packets is negative. Indeed,
The characteristic length of the linearized wave packet ulin (s, τ ) can be esti-
mated using the relation
In the case of the negative initial condition (17), the inertial nonlinearity in
the KdV equation tends to force the field u(s, τ ) further to the left. As a
result, the oscillating wave packet of the nonlinear field u(s, τ ) to the left of
the initial Gaussian field u0 (s) becomes even longer than in the linearized
case.
On the other hand, in the case of the positive initial condition (10), the
linear dispersion and inertial nonlinearity in (11) act on the field u(s, τ ) in
the opposite directions. This competition compresses the waves to produce
well-localized solitons.
In the following subsections we shall discuss analytic representations of
(multi)soliton solutions of the KdV equation (11) that form the main part
of solutions in the case of positive initial conditions u0 (s) and small values
of the constant γ.
12k 2 ∂ 2 −z 12k 2 ez
w= ln(1 + e ) = .
σ ∂z 2 σ (1 + ez )2
Using the definition of the hyperbolic secant,
1 2
sech(z) = = z ,
cosh(z) e + e−z
the previous relation can be rewritten in a form more convenient for analysis:
3k 2 z
w= sech2 .
σ 2
Substituting the explicit expression (30) for the auxiliary argument z gives
3k 2 2 k
w(s, θ) = sech (s − k θ) .
2
σ 2
The translational invariance property (8) of the KdV equation now permits
us to write a more general version of the above solution,
3k 2 2 k
w(s, θ) = sech (s − d − k θ) ,
2
(31)
σ 2
where d is an arbitrary shift of the previous special solution along the s-axis.
Also, notice that in view of the evenness of the function sech(z), one may,
without loss of generality, select k to be positive (k > 0). A 3-D plot of the
solution (31) corresponding to the constants σ, k = 1 and d = 0 is shown in
Fig. 15.3.6.
We shall now provide a physical interpretation of the solution (31) using
first, the relation (20) to convert it into the solution u(s, τ ) of the more
physically transparent KdV equation (11),
c
u(s, τ ) = σγw(s, γτ ) = 3c sech 2
(s − cτ − d) , (32)
4γ
15.3. KdV Equation and Solitons 311
FIGURE 15.3.6
A 3-D plot of the exact soliton solution (31) of the KdV equation
corresponding to constants σ, k = 1 and d = 0. It is shown in the
time interval 0 < θ < 4.
The solution (32) obtained above is obviously a wave, moving from left
to right with velocity c without changing its shape. So it is plausible that
it is the soliton that we discovered earlier in numerical experiments with
solutions of the KdV equation. Equations (32) and (33) indicate that our
soliton possesses the following
characteristic properties: its amplitude 3c and
its effective width (∼ c/4γ) are tied uniquely to its velocity c. Qualitatively,
the higher the soliton’s velocity, the larger the amplitude and the narrower
the width of the soliton (for a given γ).
In some physical applications, solitons can be interpreted as quasiparti-
cles, whose momentum is equal to
M = u(s, τ )ds . (34)
312 Chapter 15. Other Nonlinear Models of Higher Order
Employing the above operator notation, one may rewrite (37) in the form
P[ϕ, ϕ] = 0 . (39)
At this point, let us state a few useful properties of the bilinear operator
P[f · g].
Property 1. For a fixed f (or g), P[f · g] is a linear operator in g (or f ).
Consequently, it possesses the following bilinear property:
# $
P c i fi , d j gj = ci dj P[fi , gj ] , (40)
i j i j
where
1
P (k, ω) = k k 3 − ω (42)
2
is a symmetric function, that is,
P[1, 1] = P (0, 0) = 0 ,
(46)
P[eωθ−ks , eωθ−ks ] = P (0, 0) e2ωθ−2ks = 0 .
and calculate the left-hand side of (39). Using the bilinear property (40) of
the operator P, and relations (46) and (41), we obtain
3 3
P[ϕ, ϕ] = c1 P (k1 , k13 )ek1 θ−k1 s + c2 P (k2 , k23 )ek2 θ−k2 s
3 3
+2c1 c2 P (k1 − k2 , k13 − k23 )e(k1 +k2 )−(k1 +k2 )sθ ,
15.3. KdV Equation and Solitons 315
P[ϕ · ϕ] =
3 3
2 c1 c2 P (k1 − k2 , k13 − k23 ) + ρP (k1 + k2 , k13 + k23 e(k1 +k2 )θ−(k1 +k2 )s .
The right-hand side is equal to zero if
2
P (k1 − k2 , k13 − k23 ) k1 − k2
ρ = −c1 c2 = c1 c2 .
P (k1 + k2 , k13 + k23 ) k1 + k2
Replacing the constant ρ in (50) by the right-hand side of the above equality,
we obtain the desired new solution of equation (39):
2
k13 θ−k1 s k23 θ−k2 s k1 − k2 3 3
ϕ = 1 + c1 e + c2 e + c1 c2 e(k1 +k2 )θ−(k1 +k2 )s .
k1 + k2
Finally, proceeding as in the transition from (47) to (48), we replace constants
c1 and c2 by arbitrary space shifts d1 , d2 ∈ (−∞, ∞) to produce a more
geometrically transparent expression for the function ϕ(s, θ):
where 2
k1 − k 2
zi = ki (s − di − k θ) ,
2
R= . (52)
k1 + k2
316 Chapter 15. Other Nonlinear Models of Higher Order
In the next subsection we shall demonstrate that the field w(s, θ) (29),
where ϕ(s, θ) is given by the equality (51), describes a two-soliton collision
phenomenon, which we have already observed in Fig. 15.3.3b.
Remark 2. Extension to n-soliton solutions. By similar but more sophis-
ticated arguments, Hirota4 obtained multisoliton solutions for the KdV equa-
tion. To better comprehend their structure, let us rewrite the formula (51)
in a form involving a determinant:
√
−z1 2 k1 k2 − z1 +z2
1+e e 2
ϕ= √ k1 + k2 . (53)
2 k1 k2 − z1 +z2 −z2
e 2 1+e
k1 + k2
Hirota proved that the elegant generalization
2 ki kj zi + zj
ϕ = ψi,j (s, θ) , ψi,j = δi,j + exp − ,
ki + kj 2 (54)
i, j = 1, 2, . . . , n,
of the relation (53) gives, in combination with the equality (29), a solution of
the KdV equation (13) describing n-soliton collisions. Here δi,j is the usual
Kronecker delta, equal to one if i = j and zero otherwise.
It has a transparent geometric sense. Namely, h̄(y, θ) describes the time evo-
lution of the KdV equation’s solution in the coordinate system moving uni-
formly with the first soliton, corresponding to the parameter k1 . In other
words, the new comoving coordinate y and the old fixed coordinate s are
joined by the relation
y = s − d1 − k12 θ ⇐⇒ s = y + d1 + k12 θ .
where
χ = d2 − d1 + (k22 − k12 )θ .
Intuitively, χ can be thought of as a measure of the “distance” between the
first and the second solitons. If |k1 | = |k2 |, then χ changes in the course of
time θ. Suppose for the sake of definiteness, that
k1 , k2 > 0 , k2 > k 1 .
(i) In the first stage, when χ −1, the first soliton is immovable (in the
coordinate system y) and the second approaches it from the left with
uniform speed.
(ii) In the second stage, for |χ| 1, a collision of the two solitons occurs.
(iii) Finally, in the third stage, for χ 1, the first soliton becomes im-
movable again while the second moves away from it to the right with
uniform speed.
Let us consider the above three stages in more detail. At the very begin-
ning of the first stage, when
χ → −∞ ⇒ eχ → 0,
318 Chapter 15. Other Nonlinear Models of Higher Order
χ→∞ ⇒ eχ → ∞ ,
It differs from the original (before the collision) soliton (57) only by the
space shift ξ. Figure 15.3.7 shows a contour plot of the field w̃(y, θ) obtained
via (57) from h̄(y, θ) (55). The corresponding 3-D plot is shown in Fig. 15.3.8.
15.3. KdV Equation and Solitons 319
FIGURE 15.3.7
A contour plot of the two-soliton collision. Here σ = 1. The time
frame is −15 ≤ t ≤ 15. The phase shifts caused by the interaction
of the two solitons in a neighborhood of x = 0, t = 0, are clearly
visible.
The picture in Fig. 15.3.7 can be interpreted in two different ways. The
first is that in the course of time, the second soliton overtakes the first one,
so that the ultimate result of the collision is only its finite space shift is com-
parison with the original motion of the solitons. However, some physicists
see the collision interaction of the solitons as representing the real physical
process of an inelastic collision of two quasiparticles. Each of them has a
complex structure involving waves compressed into the well-localized wave
packet(quasiparticle) and subject to two competing influences: linear disper-
sion and inertial nonlinearity (see Fig. 15.3.5). During the collision, quasi-
particles do not swap places but repulse each other, exchanging their masses
and, due to the energy conservation law, velocities as well. This phenomenon
can be observed in Fig. 15.3.7.5
5
Fascinating videos of KdV solitons, both simulated and photographed at ocean
beaches, can be found at www.youtube.com/watch?v=ZsTe2N5_eZE.
320 Chapter 15. Other Nonlinear Models of Higher Order
FIGURE 15.3.8
A 3-D illustration of a two-soliton collision shown as a contour plot
in Fig. 15.3.7.
6
See, e.g., F. Otto, The geometry of dissipative evolution equations: the porous medium
equation, Communications in Partial Differential Equations 26 (2001), 101–174.
15.4. Flows in Porous Media 321
P = κρα , (3)
∂
u(x, t) = div (1 + α)uα (x, t)∇u(x, t) , (5)
∂t
7
A comprehensive calculation showing how Darcy’s law can be obtained from the
general equations of flow in porous media can be found, e.g., in W.G. Grey and K. O’Neil,
On the general equations of flow in porous media and their reduction to Darcy’s law, Water
Resources Research 12 (1976), 148–154, or U. Hornung, Homogenization and Porous Media,
Springer-Verlag, New York 1997, pp. 16ff.
322 Chapter 15. Other Nonlinear Models of Higher Order
with the factor in front of the gradient an increasing function of u. This fact
plays an important role in the determination of the behavior of solutions of
the Cauchy problem for (4).
The case α = 0, not considered here, obviously corresponds to the stan-
dard diffusion equation discussed in detail in Chap. 10. So from now on, we
will always assume
α > 0.
Several special cases describe specific physical phenomena: α = 1 often is
used to model thin saturated regions in porous media, α ≥ 2 is a general
model of percolation of gas through porous media, α = 3 has been used
in the study of thin liquid films spreading under gravity, and α = 6 has
appeared in the investigation of radiative heat transfer by so-called Marshak
waves.8
The fundamental property of the porous medium equation is that it pos-
sesses solutions that are compactly supported for all t > 0, a behavior dra-
matically different from the behavior of solutions of other parabolic equations
we have considered thus far, in which even the solution produced by the Dirac
delta initial data spread over an infinite spatial interval for every t > 0.
α |x|2 1/α
uB (x, t) = max 0, Bt−dα/(dα+2) − · (8)
2(α + 1)(dα + 2) t
8
See, e.g., G.I. Barenblatt, On some unsteady motions of a liquid and a gas in a
porous medium, Prikladnaya Matematika i Mekhanika 16 (1952), 67–78; D.G. Aronson,
The Porous Medium Equation, Lecture Notes in Mathematics 1224, Springer-Verlag 1224,
1986; M.E. Gurtin and R.C. MacCamy, On the diffusions of biological populations, Math-
ematical Biosciences 33 (1977), 35–49.
15.4. Flows in Porous Media 323
1 x2
uB (x, t) = max 0, Bt−1/3 − · . (10)
12 t
FIGURE 15.4.1
Evolution of Barenblatt’s self-similar solution (8) of the porous
medium equation for times t ∈ (0, 1]. We have here the constant
B = 1.
for every solution u of (4) with nonnegative and integrable initial value
u(x, 0) = u0 (x) with the same mass as uB .9
15.5 Exercises
1. In the case in which the field v(x, t) models a 1-D gas flow, the kinetic
energy of the flow is given by the expression
1
T = ρ(x, t) v 2 (x, t) dx , (1)
2
which is different from the formula (2) for E in the answers and solu-
tions to Chap. 14. Show that in the case of a pressureless gas whose
velocity and density are described by equations (15.1.27), there is a
close relationship between the integrals E and T .
2. It is well known that the KdV equation has infinitely many invariants.
Let us derive the two “most physical” of them. More precisely, show
that if the solution u(s, τ ) of the KdV equation (15.3.11) converges to
zero fast enough as x → ±∞, then the “momentum” M (15.3.34) and
“energy” K (15.3.36) of the field u(s, τ ) are constant, i.e., they do not
depend on τ .
9
See, e.g., J.L. Vazquez, Asymptotic behavior for the porous medium equation posed
in the whole space, Journal of Evolution Equations 2 (2002), 1–52.
15.5. Exercises 325
we have
12
lim w(s, θ) ds = π n. (2)
θ→±∞ σ
Obtain an analogous relation for the solution u(s, τ ) of the KdV equa-
tion (15.3.11).
5. Neglecting the impact of the oscillating tail and assuming that the
number of solitons in the solution u(s, τ ) of the KdV equation (15.3.11)
is equal to 1 (n = 1), estimate the dependence of the soliton’s velocity
c on its momentum M and parameter γ. Produce the same estimate
by relying on the value of the kinetic energy K.
Appendix A
2.
1 −ik(x+y) e−iky sin kx, for x < y;
G(x, y) = [e − e−ik|x−y| ] =
2ik e−ikx sin ky, for x > y.
Finally, substituting the explicit expression for the normal derivative of the
Green’s function, we arrive at the celebrated Poisson integral formula:
π
1 1 − ρ2
u(ρ, ϕ) = f (ψ) dψ.
2π −π 1 + ρ2 − 2ρ cos(ϕ − ψ)
Chapter 9. Potential Theory and Elliptic Equations 329
which is one of the possible forms of the solution to the original problem.
Physicists often interpret it as follows: The first summand represents a super-
position of the planar wave that freely propagate in space. The quantity f˜(p)
is the complex amplitude of the planar waves propagating in the direction
of the wavevector k. Its projections on the x- and y-axes are, respectively, q
and p. The norm |k| = k of the wavevector is equal to the wavenumber. The
second component describes a localized field that decays exponentially with
x and that is concentrated in a thin layer adjacent to the boundary plane
x = 0.
330 Appendix A Answers and Solutions
of both sides with respect to the spatial variable x. This operation transforms
the partial differential equation (2) into an ordinary differential equation,
df˜ 1
− iκα(t)f˜ + β(t)κ2 f˜ = δ(t) . (3)
dt 2
If one takes into account the causality condition, then the solution of the
above equation has the form
1
f˜(κ; t) = exp iκh(t) − b(t)κ ,
2
t > 0, (4)
2
where t t
h(t) = α(τ ) dτ , b(t) = β(τ ) dτ . (5)
0 0
Applying the inverse Fourier transform
∞
1
f (x, t) = f˜(κ; t)e−iκx dκ (6)
2π −∞
to the right-hand side of (4) and taking into account the formula (10.3.2),
one finally obtains
1 (x − h(t))2
f (x, t) = exp − . (7)
2πb(t) 2b(t)
10
Note that this version of the Fourier transform differs, by the minus sign in the
exponent, from the Fourier transform introduced in Volume 1. This version of the Fourier
transform is more convenient in the applications to random fields discussed in Volume 3.
Chapter 10. Diffusions and Parabolic Evolution Equations 331
∂ f˜ ∂ f˜ β(t) 2 ˜
+ α(t)κ + κ f = 0, f˜(κ; t = 0) = eiκy (11)
∂t ∂κ 2
for the Fourier image f˜(κ; t) of the solution f (x, t) of the parabolic equation
(10.10.4).
The latter equation can be solved by the method of characteristics de-
scribed in Volume 1. Indeed, suppose that κ is a function of t, that is,
κ = Ω(κ0 ; t), where κ0 = Ω(κ0 ; t = 0) is an arbitrary initial value. Then
equation (11) splits into two characteristic equations,
describes the behavior of the desired Fourier image f˜(κ; t) along a chosen
characteristic line.
Solutions of the equation (12) have the form
where h(t) has been defined by the first equality in (5), and
κ20
F (κ0 , t) = exp iκ0 y − E(t) , (15)
2
with t
E(t) = β(τ )e2h(τ ) dτ. (16)
0
Now to get the Fourier image f˜(κ; t) of the solution of the initial value prob-
lem (10.10.4–5), it suffices to insert the inverse function κ0 = Ω−1 (κ; t) =
κe−h(t) into the right-hand side of (15). As a result, we obtain
f˜(κ; t) = F Ω−1 (κ; t); t = exp iκyeh(t) − κ2 E(t)e−2h(t) , (17)
and an application of the inverse Fourier transform yields the final result:
eh(t) (xeh(t) − y)2
f (x, t) = exp − . (18)
2πE(t) 2E(t)
4. In this case,
β 2αt
h(t) = αt, E(t) = e −1 .
2α
Substituting these expressions into the right-hand side of (18), we get
α α (x − ye−αt )2
f (x, t) = exp − . (19)
πβ(1 − e−2αt ) β 1 − e−2αt
As t → ∞, the above solution converges to the stationary solution
α α 2
lim f (x, t) = fst (x) = exp − x . (20)
t→∞ πβ β
5. By definition, the desired stationary solution fst (x) does not depend
on time. So, in equation (10.10.4), we can neglect the term containing the
Chapter 10. Diffusions and Parabolic Evolution Equations 333
derivative with respect to the time variable, which gives an equation of the
form
dGst (x)
= 0, (21)
dx
where
β dfst (x)
Gst (x) = −αxfst (x) − (22)
2 dx
is the stationary flow. Equality (21) means that in the stationary state, the
flow is constant everywhere, that is,
β dfst (x)
− αxfst (x) − = C, (23)
2 dx
where C is the magnitude of the flow. This magnitude has to be found from
the normalization condition
fst (x) dx = 1. (24)
However, the formula (34) gives limt→∞ m1 (t) = y. This seemingly paradox-
ical situation shows the perils of recklessly interchanging limit operations
without a proper justification.
A similar calculation for the second moment leads us to the initial value
problem
dm2 (t)
= (3γ − 2α)m2 (t) + β, m2 (0) = y 2 .
dt
Its solution is
β
m2 (t) = y02 e(3γ−2α)t + e(3γ−2α)t − 1 . (39)
3γ − 2α
8. The stationary solution satisfies the following ordinary differential
equation:
β dfst
+ αfst = C, (40)
2 dx
where C is the magnitude of the stationary flow. The general solution of the
above equation has the form
C 2α
fst (x) = + A exp − x . (41)
α β
The boundary condition (10.10.10) forces A = 0, and the initial condition
(10.10.5) implies that the stationary solution has to be normalized on the
interval x ∈ [0, ]. This means that
1
fst = , (42)
336 Appendix A Answers and Solutions
df˜(κ; t)
= ν[w̃(κ) − 1]f˜(κ; t), f˜κ; t = 0) = 1. (43)
dt
Its solution is
f˜(κ; t) = exp −ν 1 − w̃(κ) t , t > 0. (44)
Taking the inverse Fourier transform, we obtain the desired solution of the
Kolmogorov–Feller equation:
1
f (x, t) = exp −iκx − ν 1 − w̃(κ)t t dκ. (45)
2π
Let us discuss in more detail a particular case of the Gaussian kernel,
2
κ
w̃(κ) = exp − , (46)
2
which converges to 0 as κ → ∞. In this case, it makes sense to split the
Fourier image of the solution of the Kolmogorov–Feller equation into a con-
stant part (with respect to κ) and an absolutely integrable part
where
f˜c (κ; t) = eν[w̃(κ)−1]t − e−νt . (48)
Thus, the solution of the Kolmogorov–Feller equation can be split into the
sum of singular and continuous parts,
where 2
1
fc (x, t) = exp νt e−κ /2 − 1 − e−νt e−iκx dκ. (50)
2π
Chapter 10. Diffusions and Parabolic Evolution Equations 337
For large times (when νt 1), the asymptotic behavior of the continuous
part of the solution is determined by the behavior of its Fourier image in a
small vicinity of κ = 0. There, the Gaussian function (46) can be replaced
by the first two terms of its power expansion,
2
−κ κ2
exp =1− + ..., (51)
2 2
and the small term e−νt can be neglected. As a result, we obtain the fol-
lowing asymptotic formula for the continuous part of the solution of the
Kolmogorov–Feller equation:
1 νtκ2
fc (x, t) ∼ exp − − iκx dκ, (νt 1), (52)
2π 2
and it tends to zero as |κ| → ∞. So the solution has a structure similar to that
of (49) in the preceding problem. Moreover, the singular part of the solution
is the same as in the case of the Gaussian kernel, while the continuous part
is given by
1
fc (x, t) = exp νt e−|κ| − 1 − e−νt e−iκx dκ. (55)
2π
0.8
0.7
0.6
ϕ(z,μ)
0.5
0.4
0.3
0.2
μ=1
0.1
0
0 1 2 3 4 5 6
z
FIGURE 1
The plots of graphs of functions ϕ(z, μ) (59), for μ = 1, 100, and of
the function ϕ(z, ∞) (60) (from bottom to top). Note that ϕ(z, 100)
almost coincides with the Cauchy curve ϕ(z, ∞). The conclusion is
that for, say, μ = νt > 10, the continuous part of the solution of the
Kolmogorov–Feller equation with Cauchy kernel is well approxi-
mated by the Cauchy density itself.
11. First of all, notice that the asymptotic Cauchy density (57) is self-
similar. So changing variables
x
z= , μ = νt, (58)
νt
and introducing an auxiliary function,
μ
ϕ(z, μ) = πμfc μz, , (59)
z
Chapter 10. Diffusions and Parabolic Evolution Equations 339
we have
1
ϕ(z, ∞) = . (60)
1 + z2
Thus our numerics will have to demonstrate that ϕ(z, μ) → ϕ(z, ∞) as
μ → ∞.
Let us rewrite ϕ(z, μ) in a form more convenient for numerical calcula-
tions, ∞
ϕ(z, μ) = ψ(r, μ) cos(rz) dr, (61)
0
where
r
ψ(r, μ) = exp μ exp − − 1 − exp(−μ). (62)
μ
To calculate the improper integral (61) numerically, we have to replace
its upper limit by a large number, say R, and in order to avoid the Gibbs
phenomenon (see Volume 1), multiply the integrand by the Cesàro factor
(1 − r/R). As a result, it remains to evaluate numerically the integral
R
r
ϕ(z, μ) ≈ ψ(r, μ) 1 − cos(rz) dr. (63)
0 R
12. To find the required main asymptotics, let us rewrite the asymptotic
series (4.3.8) from Volume 1 in the form needed for our task,
∞ m+1
e−ixτ 1
fc (x, t) ∼ f˜c(m) , (64)
2π m=0 ix
(m)
or its derivatives have discontinuities. Recall that f˜c is the size of the
jump of the mth derivative (with respect to κ) of f˜c (κ; t) at the discontinuity
point τ ,
f˜c(m) = f˜c(m) (κ = τ + 0; t) − f˜c(m) (κ = τ − 0; t). (66)
340 Appendix A Answers and Solutions
100 z−2
μ = 10
μ=1
10–1 μ = 0.5
ϕ(z,μ)
10–2
10–3
FIGURE 2
Log-log plots of functions ϕ(z, μ) (61), for different values of μ. It is
evident that for every μ, the continuous part of the solution of the
Kolmogorov–Feller equation with the Cauchy kernel has a slowly
decaying power tail ∼ x−2 .
In our case, the function (65) is continuous, but its derivative has a jump
at κ = 0 (τ = 0),
f˜c (+0; t) − f˜c (−0, t) = −2νt. (67)
Thus it follows from (64) that the main asymptotics of the continuous part
of the Kolmogorov–Feller equation are as follows:
νt
fc (x, t) ∼
, (x → ∞). (68)
πx2
Figure 2 depicts the log-log plots of the numerically calculated functions
ϕ(z, μ) for different values of the dimensionless time parameter μ, and com-
pares them with the adjusted asymptotics ϕ(z, μ) ∼ 1/z 2 (68).
so that
v(x, t) = ṽ(ω, κ)e−iωt+iκx dω dκ.
Using the relation (11.1.9) and the probing property of the Dirac delta
with respect to ω, we obtain
v(x, t) = 4π 2
f˜(k)h̃(W (k), k)ei(kx−W (k)t) dk.
Substituting u(x, t) from (11.1.10) and the above expression for v(x, t) into
(11.12.1), utilizing the relation (11.3.3),
eix(κ1 +κ2 ) dx = 2πδ(κ1 + κ2 ),
the probing property of the above Dirac delta, and the fact that W (κ) is an
odd function, we finally obtain
S = (2π)3 |f (k)|2 h̃(W (k), k) dk = const. (1)
ω 2 = a2 k 2 + ω02 .
So
ω = ±W (k) = ± a2 k 2 + ω02 .
Correspondingly,
W (k)
c= = a 1 + ω02 /(a2 k 2 ) > a
k
and
v(k) = W (k) = a/ 1 + ω02 /(a2 k 2 ) < a.
Note that
c(k)v(k) = a2 = const.
342 Appendix A Answers and Solutions
where
˜ ˜
f (k, t) = f (k) exp −ikat 1 + ω0 /(a k ) .
2 2 2
From the formula (4.3.3) of Volume 1, we know that if the function f (x) =
u(x, t = 0) has a jump, then its Fourier image has the asymptotics
f −ikx0
f˜(k) ∼ e , (k → ∞).
2πik
Consequently, the spatial Fourier image of the wave packet u(x, t) has the
asymptotics
f −ik(x0 +at)
f˜(k, t) ∼ e , (k → ∞). (2)
2πik
This means that the original jump in the wave packet does not disappear but
preserves its size f , and moves with the velocity a, as if it were evolving in
a nondispersive medium.
Consequently,
v(x, t) =
1 −Rt 1 L −Rt
e L [v0 (x − at) + v0 (x + at)] + e L [i0 (x − at) − i0 (x + at)] ,
2 2 C
and
i(x, t) =
1 −Rt 1 C −Rt
e L [i0 (x − at) + i0 (x + at)] + e L [v0 (x − at) − v0 (x + at)] .
2 2 L
Physical effects: In practice, one often tries to generate waves that prop-
agate along the transmission line in a given direction. The above solution
implies, for example, that if the initial voltage and current are related by the
condition
L
v0 (x) = i0 (x) ,
C
then the wave will move only to the right, and in this case, the voltage is
described by the following simple expression:
v(x, t) = e− L t v0 (x − at) .
R
Substituting functions (6) into the boundary condition (4), after omitting
e− L t , we obtain
R
C
Φ(−at) + Ψ(at) = −R0 [Φ(−at) − Ψ(at)] , 0 < t < +∞ .
L
√ √
R0 C − L
Φ(x) = f (−x) √ √ , x < 0.
R0 C + L
Consequently, we find that the solution of the problem is of the form (6),
where
⎧ √ √
⎨ R0 C − L
f (−x) √ √ , for − ∞ < x < 0;
Φ(x) = R C + L
⎩ 0
0, for 0 < x < +∞.
Ψ(x) = f (x), for 0 < x < +∞ .
where
ρ−1
K = K(ρ) = .
ρ+1
Chapter 11. Waves and Hyperbolic Equations 345
FIGURE 1
A semi-infinite transmission line: Dependence on the grounding
resistivity of the reflection-from-the-grounded-end-point
coefficient.
Note some peculiarities of the reflection coefficient: For small values of the
grounding resistivity, the coefficient becomes negative, and in the presence of
a short circuit (R0 = 0), when the voltage at the left endpoint remains zero at
all times, the coefficient K is equal to −1. This means that the reflected wave
in the vicinity of the grounded point (i.e., for x = +0) is equal to the negative
of the incident wave. For R0 → ∞, the reflection coefficient converges to 1,
while if one chooses the grounding resistivity to be R0 = L/C (the so-called
matched load), the reflection coefficient is K(1) = 0, and there is no reflected
wave. This effect is widely used to achieve damping of reflected waves in long
transmission lines.
Solution: As before, the functions Φ(x) and Ψ(x) in (6), for x > 0, are
given by the equalities (7), and for x < 0, they have to be defined so that the
346 Appendix A Answers and Solutions
condition (8) is satisfied. Substituting in (8) the equalities (6) and canceling
the factor e−(R/L)t , we have
C R C
Φ(−at) + Ψ(at) = −L0 [Φt (−at) − Ψt (at)] + +L0 [Φ(−at) − Ψ(at)] .
L L L
Introducing an auxiliary variable z = −at < 0, we obtain, for z < 0, an
ordinary differential equation for Φ(z):
Φ − k Φ = p f (−z) − f (−z) .
the first corresponding to the wave reflected from the short-circuited point
of the line, and the second corresponding to the inertial process of reflection.
For L0 → 0, as k → ∞, the second summand disappears.
Chapter 11. Waves and Hyperbolic Equations 347
From the radiation condition, which in our case means that the voltage source
applied at the left endpoint of the line can only generate a wave propagating
to the right, we obtain that Ψ ≡ 0. The function Φ(x) in (6) can be found
by checking the boundary condition
E(t) = e− L t Φ(−at) .
R
The functions v0 (x) and i0 (x) describe a stationary state in the line at time
t = 0, which can be found from the telegrapher’s equation (9) with the
proviso that v0 (x) does not depend on t:
d 2 v0
− GR v0 = 0 , v(0) = E (0 < x < +∞) .
dx2
The general solution of this equation is
√ √
v0 (x) = A e− RGx
+Be RGx
.
It follows that Φ(−at) = 0 for t > 0. In other words, we have defined the
function Φ(x) for negative values of x. For other values of x,
√
Φ(x) = E χ(x) e− GRx
,
Similarly,
C √
i(x, t) = E χ(x − at)e− GRx .
L
Physical effects: As we could have guessed, the radiation condition implies
that a sudden change of the regime at x = 0 creates a wave propagating to
the right. One could think about it as a rigid stationary profile v0 (x), i0 (x),
preserved for x > at, when the information about the changed conditions at
the line’s left endpoint has not yet been received. The profile is “evaporating”
inside the segment 0 < x < at attached to the endpoint.
As in Section 11.11, instead of voltage v(x, t), we will work with a more suit-
able auxiliary function u(x, t) satisfying equation (11.11.6) and the boundary
and initial conditions obtained via (11)–(12), (11.11.4), and (11.11.17):
∂u(0, t)
= 0, u(l, t) = E0 eμt sin ωt ,
∂x
∂u(x, 0)
u(x, 0) = 0 , = 0.
∂t
350 Appendix A Answers and Solutions
Additionally, if conditions
where
1
α + iβ = (μ + iω)2 − b2 . (21)
a
Substituting (21) into (19), we obtain
and
Φ(l) = c1 e(α+iβ)l + c2 e−(α+iβ)l = E0 .
Therefore
E0
c1 = c2 = .
2 cosh(α + iβ)l
Consequently, according to equality (20), we have
cosh(α + iβ)x iωt
μt
w(x, t) = E0 e Im e . (22)
cosh(α + iβ)l
∞
ũ(x, t) = (an cos ωn t + bn sin ωn t) X̃n (x) , (23)
n=0
where 2
2 π(2n + 1)
X̃n (x) = cos λn x , λn = ,
l 2l
l √
2 cosh(α + iβ)x cos λn x
an = (−w(x, 0), X̃n (x)) = E0 Im dx ,
l cosh(α + iβ)l
0
1
bn =(−wt (x, 0), X̃n (x))
ωn
l √
μ ω 2 ch(α + iβ)x cos λn x
= an − E0 Im dx .
ωn ωn l ch(α + iβ)l
0
Substituting (22) and (23) into (13) and returning, via equalities (4), from
the auxiliary function u(x, t) to the voltage function v(x, t), we finally obtain
352 Appendix A Answers and Solutions
cosh(α + iβ)x iωt
v(x, t) = E0 Im e +
cosh(α + iβ)l
∞
+e−μt (an cos ωn t + bn sin ωn t) X̃n (x) .
n=0
The Eulerian coordinates are connected with the Lagrangian ones by the
formula
x = y(1 + Ht) + u0 (y)t . (2)
To find the Eulerian velocity field, we have to substitute the inverse function
y(x, t) in the right-hand side of (1). The Eulerian density field is described
by the formula (12.2.6),
∂y(x, t)
ρ(x, t) = 0 (y(x, t)) . (3)
∂x
Furthermore, observe that the relationship (2) between Lagrangian and Eu-
lerian coordinates can be written in a more familiar form if we introduce new
Eulerian coordinates and time,
x t
x = , t = .
1 + Ht 1 + Ht
Then the equality (2) can be rewritten in the form
x = y + u0 (y) t . (4)
θ = τ (1 − e−t/τ ) .
Comparing this solution with the solution of the standard Riemann equation,
it is easy to see that
v(x, t) = e−t/τ u(x, θ) ,
where u(x, θ) satisfies the Riemann equation
∂u ∂u
+u = 0, u(x, θ = 0) = v0 (x) . (8)
∂θ ∂x
Physical effects: For t → ∞, the auxiliary time is θ → τ . Physically,
this means that because of the velocity “dissipation,” the nonlinear effects
become weaker with the passage of time, and the shape of the velocity field
“freezes,” that is, it remains the same as if the time elapsed never exceeded τ .
μ = κ/k , z = ky , τ = kat .
Using the probing property of the Dirac delta, the fact that J0 (0) = 1, as
well as the symmetry properties of the Bessel functions, we obtain
∞
Jn (nτ )
ṽ(κ, t) = ia (−1)n [δ(κ − kn) − δ(κ + kn)] .
n=1
nτ
Substituting this expression into the inverse Fourier integral, we obtain the
following expansion of the solution of the Riemann equation:
∞
Jn (nτ )
v(x, t) = 2a (−1)n+1 sin(nkx) , (9)
n=1
nτ
and ρ(x, t) are related by the equality (12.2.9), which can be interpreted as
an ordinary differential equation for v(x, t),
dv(x, t) 1 ρ(x, t)
= 1− ,
dx t ρ0
and which contains the time t as a parameter. Substituting here the series
(12.2.19) obtained earlier for the density field and taking into account the
obvious boundary condition v(x = 0, t) = 0, we obtain again the desired
solution.
where
sin ωt
θ= ,
ω
and u(x, θ) is the solution of the Riemann equation (8).
D(θ ) = cδ(θ − θ0 ) .
The condition (12.3.3) guarantees that for every point of the initial profile
h0 (x), the Dirac delta is not concentrated outside the interval of integration
in (12.2.16). Consequently, the velocity of the snow accretion on the interface
segment that has the normal inclined at the angle θ to the z-axis is
c⊥ = c cos θ0 , c = c sin θ0 ,
7. It is not difficult to show that growth of the interface h(x, t) also has
to satisfy the linear equation developed in the solution for Exercise 6, where
in the present case,
π/2 π/2
c⊥ = D(θ ) cos θ dθ , c = D(θ ) sin θ dθ .
−π/2 −π/2
FIGURE 1
Time evolution of the interface h(x, t) from Exercise 8, for ε = 0.4
and τ = 0, 10, and 25.
In the special case of the sine initial profile, these equations take the form
τ τ
kx = μ + ε3 sin3 μ , kh = ε cos μ − ε4 sin4 μ ,
3 4
where the nondimensional variables
μ = ky , ε = kh , τ = ckt
have been introduced.
Graphs illustrating the evolution of the interface h(x, t) for various τ are
shown in Fig. 1 above.
9. In this case, the evolution equation (12.2.11) is of the form
∂h c
=√ .
∂t 1 + u2
The corresponding small angle approximation equation (after changing to a
coordinate system comoving with the average speed ct) is
2
∂h c ∂h
+ = 0.
∂t 2 ∂x
360 Appendix A Answers and Solutions
FIGURE 2
Time evolution of the interface h(x, t) from Exercise 9, for ε = 0.4
and times τ = 0, 5, and 10.
The corresponding graphs are shown in Fig. 2. To better show the evolution
of the profile, the graphs are shifted slightly upward along the vertical axis.
If we think of the scenario of the wave front evolution in an isotropic
medium shown in Fig. 12.3.4 as a standard deposition, then in the current
problem, the interface evolves in the reverse direction, following a “corrosion”
scenario as shown in Fig. 12.3.5. Its characteristic features are sharpening of
the crests and smoothing out of the troughs.
10. Note that the gradient function u(x, t) (12.2.10) of the interface h(x, t)
is described by the parametric equations
FIGURE 3
The interface h(x, t) from Exercise 10, for τ = 10 and ε = 0.1. The
lower curve was constructed from the parametric equations, and
the upper curve by taking the first ten terms of the Fourier series.
which coincide with the solution of the Riemann equation with the initial
condition v0 (x) = −khc sin(kx) and time running backward. We have already
obtained the Fourier series for the solution of the Riemann equation with the
sine initial condition. Hence, we can write immediately
dh(x, t) ∞
Jn (nετ )
=2 (−1)n sin(nkx) .
dx 1
nτ
∞
Jn (nετ )
k h(x, t) = 2 (−1)n cos(nkx) + C(t) .
1
n2 τ
Here C(t) is constant. If we are just interested in the shape of the interface
h(x, t) rather than its actual position, we can disregard C(t) altogether. Fig-
ure 3 shows a graph of the interface h(x, t) for τ = 10 and ε = 0.1. The
lower curve was constructed from the parametric equations, and the upper
curve by taking the first ten terms of the Fourier series. One can see that the
shapes are almost identical.
362 Appendix A Answers and Solutions
Here we employed our standard convention and denoted by Q(y, t) the La-
grangian field corresponding to the Eulerian field q(x, t). Let us find Q(y, t)
by noticing that the following relationship holds:
−1
∂y(x, t) ∂X(y, t) 1
= = .
∂x x=X(y,t) ∂y J(y, t)
v0 (y)
x = y + v0 (y) t , q= . (1)
1 + v0 (y) t
τ cos ζ
η = ζ + τ sin ζ , θ= ,
1 + τ cos ζ
where
η = kx , ζ = ky , τ = kat and θ = tq
Chapter 13. Generalized Solutions of Nonlinear PDEs 363
FIGURE 1
The field tq(x, t), where q(x, t) is the spatial derivative of the solu-
tion of the Riemann equation with sinusoidal initial condition. All
the variables have been nondimensionalized. The graph has been
constructed for the value τ = 0.7 of the nondimensional time. In-
equality (13.7.1) is obviously satisfied. Additionally, one can see
the typical deep troughs of the field q(x, t) corresponding to the
steepening fronts of the velocity field v(x, t) (the areas of the pro-
file v(x, t), where ∂v/∂x is negative).
Hence, it suffices to study the Lagrangian field q = Q(y, t), which is more
convenient for analysis. It follows from (1) that
z
t Q < c(z) , c(z) = , z = tv0 . (2)
1+z
364 Appendix A Answers and Solutions
νt<1 (4)
3. First, let us find the center of mass of the particle cloud. For this
purpose, let us multiply both sides of the equality (13.3.4) by x and integrate
them over the whole x-axis. As a result, after changing the order of integration
in the iterated integral on the right-hand side, we get
x ρ(x, t) dx = dκρ̃(κ, t) dx x eikx .
Chapter 13. Generalized Solutions of Nonlinear PDEs 365
The last integral, see Volume 1, is the derivative of the Dirac delta,
∂
x eikx dx = −2πi δ(κ) .
∂κ
Thus we have
∂
x ρ(x, t) dx = 2πi ρ̃(κ, t)|κ=0 . (6)
∂κ
Substituting here the derivative with respect to κ of the integral xρ(x, t) dx,
see (13.3.3), we obtain
xρ(x, t) dx = X(y, t) ρ0 (y) dy .
x(t) = xc + v 0 t , (7)
where
1
v0 = v0 (y) ρ0 (y) dx .
M
Now let us calculate the particle cloud’s dispersion. Some simple algebra
yields
D(t) = x2 (t) − x2 (t) , (8)
where the density’s second moment is given by
1
2
x (t) = x2 ρ(x, t) dx .
M
By analogy with (7),
∂2
x2 ρ(x, t) dx = −2π ρ̃(κ, t)|κ=0 ,
∂κ2
so that in view of (13.3.3),
1
x2 (t) = X 2 (y, t) ρ0 (y) dy .
M
Substituting now the explicit expression (13.2.1) for X(y, t) and utilizing
equalities (7) and (8), we obtain
D(t) = Dy + t2 Dv + 2t y v0 (y) − y v0 (y) ,
366 Appendix A Answers and Solutions
where
2
Dy = y 2 − y 2 , Dv = v02 (y) − v0 (y) ,
and where we have also introduced the operator
1
f (y) = f (y) ρ0 (y) dy
M
of spatial averaging with respect to the initial density ρ0 (y) of the particle
cloud.
4. Let us write the equation of the original interface shape in the form
r = r 0 (s),
where
r 0 (s) = l ζ(s) + m η(s) ,
and l and m are unit vectors on the coordinate axes x and z, respectively.
The normal vector to the initial interface can be written in the form
l η̇ − m ζ̇
n= ,
ζ̇ 2 + η̇ 2
where the dot denotes differentiation with respect to the parameter s. Obvi-
ously, the equation of the growing interface is
r = r 0 (s) + ct n . (9)
Substituting the above formula for the normal vector and passing to the coor-
dinate notation, we obtain the desired parametric description of the growing
interface, ⎧
⎪
⎪ ctη̇(s)
⎪
⎪ x = ζ(s) + ,
⎪
⎪
⎨ ζ̇ 2 + η̇ 2
⎪
⎪ ctζ̇(s)
⎪
⎪ z = η(s) −
⎪
⎪
.
⎩ ζ̇ 2 + η̇ 2
Now suppose that the initial condition is given explicitly in the form
z = h0 (x). Then the initial parametric conditions are
x = s, z = h0 (s) .
Chapter 13. Generalized Solutions of Nonlinear PDEs 367
ctḣ0 (s) ct
x=s+ , z = h0 (s) − . (10)
1 + ḣ20 (s) 1 + ḣ20 (s)
Here ḣ0 (x) has a clear geometric meaning: it is the tangent of the angle
between the normal to h0 (x) and the z-axis. If it is small, that is, if
|h0 (x)| 1 ,
then with good accuracy, we can expand the right-hand side of the equalities
(10) in a Taylor series in powers of ḣ0 and retain only the first nonzero term
of the expansion. This gives
ct 2
x = s + ct ḣ0 (s) , z = h0 (s) + ḣ (s) .
2 0
It is not difficult to verify that the above formulas give a parametric repre-
sentation of the exact solutions of the approximate equation (13.1.5).
5. Let us introduce the local unit vectors (er , eϕ ) of the polar coordinate
system and write the equation of the initial contour in the vector form
r = 0 (ϕ) er . (11)
r = (ϕ, t) er
The relationship between the vectors discussed above is shown in Fig. 2 (left).
It is clear from Fig. 2 that mapping of the points of the original contour into
points of the evolving contour changes the angular coordinate; the points of
the evolving contour not only escape from the origin of the coordinate system
but are subject to a rotating motion as well. For that reason, in analogy
with the Lagrangian and Eulerian coordinates of particles in a hydrodynamic
flow, it is convenient to introduce the “Lagrangian” angle ψ, and the current
“Eulerian” angle ϕ; see, Fig. 2 (right). The picture makes it clear that these
two angles are related by the equation
ϕ = ψ + β(ψ) ,
where β is the angle between the angular coordinates of the point on the
original contour and the corresponding point on the evolving contour. It can
be expressed via the angle α between the normal vector n and the unit vector
er of the polar coordinate system. The mutual position of the vectors n and
er indicates that
ct ct
sin β = sin α ⇒ β = arcsin sin α ,
where
= (ϕ, t) = (r · r) = c2 t2 + 2ct0 cos α + 20
is the distance from the origin to the point of the evolving contour.
Since
0 0
tan α = − , sin α = − 0 , and cos α = ,
0 2 2 2 2
0 + 0 0 + 0
we finally get
⎧
⎪
⎪
⎪
⎪ = (ψ, t) = c2 t2 + 2ct0 (ψ) cos α(ψ) + 20 (ψ) ,
⎪
⎪
⎪
⎪
⎨ ct sin α(ψ)
ϕ = ψ(ψ, t) + arcsin , (12)
⎪ (ψ, t)
⎪
⎪
⎪
⎪
⎪
⎪ α(ψ) = − arctan 0 (ψ) .
⎪
⎩
0 (ψ)
FIGURE 2
A geometric construction of the evolving contour in polar coor-
dinates. The picture on the left shows a fragment of the initial
contour, its radius vector, the local unit vectors of the polar co-
ordinate system, and the normal vector at a selected point M0 of
the contour L0 . The picture on the right shows fragments of the
contour at the initial time t = 0 and some t > 0. Shown are the
radius vector of the selected point M0 on the original contour and
its image M on the evolved contour Lt at time t > 0. The vector
ctn connecting the above two points is also pictured.
= 0.5
= 0.1
= 0.2
0 2
FIGURE 3
Asymptotic shapes of the evolving contour as described by their
dependence on the angle ϕ in the polar coordinate system. The case
of three values of ε are explored. The vertical scales on different
pictures are different and adjusted to the amplitude of the original
condition (dashed lines). The two bottom graphs display multi-
valued fragments. In these cases, the actual shape of the growing
contour corresponds to the upper branch of the function ∞ (ϕ). In
the two bottom graphs, the actual growing contour is marked by
a thick line.
Notice that if the first term on the right-hand side of the first equation
in (13), which does not affect the shape of the contour, is dropped, then we
obtain a parametric equation of the contour that is time-independent:
⎛ ⎞
2
0 (ψ) 0 (ψ)
= ∞ () = , ϕ = ψ − arcsin ⎝ ⎠.
2 2 2 2
0 (ψ) + 0 (ψ) 0 (ψ) + 0 (ψ)
(14)
In other words, in the course of time, the shape of the growing contour does
not change. Figure 3 shows the “frozen” shape of the evolving contour in the
case that the original shape is given by the equality (13.7.2).
Solution 7. Let us recall how to determine the shape of the ice floe at
t = T . The procedure is to roll a disk CR of radius R = cT around the exterior
Chapter 13. Generalized Solutions of Nonlinear PDEs 371
of the original contour L0 . The locus of the centers of the rolling disk is the
desired boundary LT of the freezing ice floe at t = T ; see Fig. 4.
FIGURE 4
Determination of the shape of an ice floe subject to a freezing and
thawing cycle. The original shape L0 and its image given by the
equations (9) and (12) are shown on the left. The picture on the
right shows only the segments of the mapping that correspond to
the actual shape of the contour LT . The thick line represents the
envelope HcT of the original contour L0 . All the disks of radius
R = cT that touch the original contour at exactly two points are
shown. Their arcs between the points of tangency form pieces of
the envelope HcT .
Let us take a closer look at what happens when the disk CR is rolled
around the contour L0 . Sometimes, it will not be able to touch all the points
of the original contour L0 , just as happened in the case pictured in Fig. 4. In
this case, the form of the contour LT will not change if the original contour
L0 is replaced by its (nonconvex) envelope HcT . The envelope HR (see Fig. 4)
of the contour L0 is here defined as a contour consisting of all the points of
tangency of the external circles of radius R with L0 , complemented, in the
case of separated adjacent points of tangency, by a circular arc of radius R
connecting those two points.
Let us now take a look at the process of thawing. To find the floe’s shape
after the thawing process is complete, one has to roll the same disk of radius
cT , but this time around the interior side of the contour LT . The locus L2T
of the centers of such internally osculating disks will provide the shape of the
boundary of the thawed floe after time 2T . It is clear that the final shape of
the floe at t = 2T will coincide with the shape of the envelope HcT shown in
Fig. 4.
372 Appendix A Answers and Solutions
Remark 2. Physicists often like to see what happens if one reverses the
course of time and to check whether the events retrace themselves. In this
spirit, we can think about the process of thawing as a process of freezing
in reverse time. The above analysis shows that the above process of the
contour growth is time-reversible only if the original contour coincides with its
envelope. In the general case, the process of contour growth is not reversible.
Indeed, in the process of contour growth, the information contained between
the original contour and its envelope is irretrievably lost. On the other hand,
the growth of a convex contour is completely reversible, since it coincides
with its convex envelope.
ρ0 (r) = Ω r2 .
According to the ERS principle, to solve the problem we need to find the
coordinate q(r, t) of the global minimum of the function
where
q
1
φ(q, t) = P (q)t + N (q), and M (q) = ρ0 (r) dr = Ω q3 , (16)
0 3
Chapter 13. Generalized Solutions of Nonlinear PDEs 373
γ t + 3q 4 = 4q 3 r and q = r,
9. Recall that the weak solution of the Riemann equation is of the form
x − yw (x, t)
vw (x, t) = , (17)
t
where yw (x, t) is the coordinate of the point of tangency of the initial potential
s0 (y) = s χ(y)
and the parabola (13.4.13). Here, as before, χ(y) is the Heaviside unit step
function equal 0 for y < 0 and 1 otherwise. The graph of the mapping
y = yw (x, t) and the accompanying geometric construction are shown in
374 Appendix A Answers and Solutions
FIGURE 5
Left: The initial potential function and the osculating (from below)
parabolas. Three typical positions of the parabolas are shown that
correspond to different portions of the mapping y = yw (x, t). Right
(top to bottom): Graphs of the Eulerian-to-Lagrangian mapping, the
solution velocity field, and the density field of the flow of sticky
particles. The thick arrow in the bottom picture indicates the sin-
gularity of the generalized density field.
Fig. 5. The graph and equation (17) demonstrate that in our case, the weak
solution of the Riemann equation is a sawtooth-shaped function of the form
⎧
⎪
⎨ 0x, x < 0 ,
v(x, t) = , 0 < x < x∗ (t) , (18)
⎪
⎩ t ∗
0 , x > x (t) ,
where √
x∗ (t) = 2st (19)
is the coordinate of the discontinuity (shock front) of the weak solution.
A physical explanation of the shape of the weak solution (18) is as follows:
A comparison of the ERS and the global minimum principles shows that the
latter describes the motion of inelastically colliding particles given uniform
initial density ρ0 . The assumed Dirac delta initial condition means that for
t = 0, the matter does not move, with the exception of the particle located
at x = 0, which is given momentum ρ0 s. Its running coordinate is given by
Chapter 13. Generalized Solutions of Nonlinear PDEs 375
where ⎧
⎨ ρ0 , x < 0,
ρc (x, t) = 0 , 0 < x < x∗ (t) ,
⎩
ρ0 , x∗ (t) < x .
Remark 1. Observe that for x contained in the interval [0, x∗ ], the flow
density is zero (see also Fig. 5). The reason for this is obvious: the macroparti-
cle sweeps all the matter located in its path, compressing it into an infinitely
dense cluster with a singular density. Moreover, the sawtooth field vw (x, t)
(18), if interpreted as a velocity field of the flow of sticky particles, turns out
to be vacuous—it describes the velocity of matter of zero density and mass.
376 Appendix A Answers and Solutions
Because of this phenomenon, the above weak solution of the Riemann equa-
tion may seem to be devoid of physical meaning. But this is not the case. In
nonlinear acoustics, where vw (x, t) describes the pressure field, such a solu-
tion is fully realizable, and the corresponding pressure field can be measured
experimentally.
10. According to the ERS principle, the coordinate x∗ (t) of the macropar-
ticle can be determined from the fact that the function (13.5.26) has two iden-
tical global minima at two different points with coordinates y − and y + . In the
context of our problem, it is easy to establish the relationship between those
two coordinates. Indeed, since the particles to the right of the macroparticle
do not move, their Lagrangian coordinates coincide with their Eulerian co-
ordinates. This observation also applies to the particle immediately adjacent
to the macroparticle. This means that
Similar considerations apply to the particles on the left that move with the
identical velocity V . Thus we get another useful formula,
To obtain an equation for x∗ (t), it remains to equate the values of the function
(13.5.26) at specified points, taking advantage of the equality x = x∗ . Hence,
the desired equation is of the form
where we used the relationships (13.5.28) and took into account the fact that
in our case, the momentum field is
Let us take a look at the limit cases of equation (20). If, for example, the
initial density is the same everywhere, then
y2
M (y) = y and N (y) = .
2
We assumed here, without loss of generality, that the initial uniform density
was equal to 1. Substituting the above expression into (20), we arrive at the
formula
V
x∗ (t) = t .
2
Chapter 13. Generalized Solutions of Nonlinear PDEs 377
This formula, together with the initial condition, implies that the weak so-
lution of the Riemann equation found via the global minimum principle is
V
vw (x, t) = V χ t−x .
2
In other words, in view of the global minimum principle, the shock front of
the field v(x, t) propagates with velocity V /2, equal to one-half the magnitude
of the discontinuity.
Now let us find the asymptotic solution of the equation (20) for t → ∞,
assuming, as demanded by the problem, that the total mass of the flow is
finite. Physically, it is evident that as t → ∞, all the material in the flow
becomes glued into one macroparticle moving with a constant speed. In this
case, physical intuition is in agreement with the mathematical derivations.
Namely, one can show rigorously that the following asymptotics are valid:
M (x∗ ) − M (x∗ − V t) → M,
N (x∗ ) − N (x∗ − V t) → xc M, (t → ∞) ,
∗
M (x − V t) → −M− ,
where M is the total mass in the flow, xc is its initial center of mass, and
0
M− = ρ0 (y) dy
−∞
11. Before we attempt to find the velocity and density fields, it is useful
to study the motion of individual particles. Consider a particle in the flow
with an arbitrary Lagrangian coordinate y. As always, denote its Eulerian
coordinate by x = X(y, t). Motion of the particles is subject to Newton’s
second law, which in the case of a one-dimensional UGL takes the form
d2 X
= γ [Mr (y, t) − Ml (y, t)] ,
dt2
where Mr (y, t) is the cumulative mass of the flow to the right of the selected
particle, and Ml (y, t) is the cumulative mass to its left. As long as the particles
378 Appendix A Answers and Solutions
maintain their original order, the above cumulative masses are independent
of time and equal to
y ∞
Ml (y) = ρ0 (y ) dy and Mr (y) = ρ0 (y ) dy ,
−∞ y
Ml (y) + Mr (y) ≡ M .
As a result, we have
On the other hand, we know that the Lagrangian density field is always
described by the expression
ρ0 (y)
R(y, t) = , (24)
J(y, t)
where in our case, the Lagrangian-to-Eulerian Jacobian is given by
∂X γ
J(y, t) = = 1 + v0 (y)t − ρ0 (y)t2 . (25)
∂y 2
Chapter 13. Generalized Solutions of Nonlinear PDEs 379
12. First of all, let us calculate the cumulative mass located to the left of
a point with Lagrangian coordinate y:
y
2 dz M y π
M (y) = ρ0 2 2
= arctan + ,
−∞ z + π 2
where
M = πρ0
is the total mass of the matter in the flow (Fig. 6).
Separately, let us write the combination of the cumulative masses on
the left and on the right that enter into the expressions for the Lagrangian
velocity field (22) and the Lagrangian-to-Eulerian mapping (23):
y
M − 2M (y) = −2ρ0 arctan .
Substituting this expression into (22) and (23), we obtain a parametric de-
scription of the Eulerian velocity and density fields:
1
η = ζ − τ 2 arctan (ζ) , u = −2τ arctan (ζ) , r= ,
1 + ζ2 − τ 2
where we have introduced the nondimensional time, velocity, and density
x y √ v ρ
η= , ζ= , τ= γρ0 , u= √ , r= .
γρ0 ρ0
13. In this case, utilizing the fact that g(z) is even, it is possible to write
the combination of cumulative masses entering into equations (22) and (23)
in the form
y y/
M − 2M (y) = ρ0 (y ) dy = ρ0 g(z) dz ,
−y −y/
which demonstrates that the behavior of the velocity field depends essentially
only on the density of particles in the interval [−y, y], and is independent of
380 Appendix A Answers and Solutions
FIGURE 6
Graphs of the nondimensional Eulerian velocity and density fields
for the gravitationally interacting particles shown at a progression
of time instants. It is clear that the gravitational interaction leads
to an ever accelerating convergence of the particles to the origin,
and to the growth of the density curve in the neighborhood of the
center of the 1-D universe.
the behavior of the initial density field outside this interval. This opens up
a possibility of easy transition to the limit of uniform initial density. Indeed,
letting to infinity and observing that in view of the continuity of the function
g(z), the last integral converges to 2y/, we get
lim [M − 2M (y)] = 2ρ0 y .
→∞
Substituting the limit value into the right-hand sides of (23) and (24), we
can find the velocity field and the Lagrangian-to-Eulerian mapping for the
initially uniformly distributed flow of gravitationally interacting particles:
x = y 1 − γρ0 t2 + v0 (y) t , v = v0 (y) + 2γρ0 y t .
Consequently, in view of the expressions (24) and (25), the Lagrangian den-
sity field is given by
ρ0
R(y, t) =
.
1 + v0 (y) t − γρ0 t2
For a uniformly expanding universe, when v0 (y) = Hy, the density remains
uniform but diverges to infinity as τ → τn , where the nondimensional collapse
time is given by
√ √ H
τn = δ + 1 + δ 2 , τ = γρ0 t , δ= .
4γρ0
Chapter 14. Nonlinear Waves and Growing Interfaces 381
The second integral on the left-hand side is obviously equal to zero. Thus, to
prove the invariance of the above functional, it is sufficient to show that the
integral on the right-hand side is zero. Integration by parts shows that this
indeed is the case:
2 2 ∞
∂ 3u 1 ∂ ∂u ∂u
u 3 dx = − dx = = 0.
∂x 2 ∂x ∂x ∂x
−∞
which reflects the fact that the viscosity (μ > 0) dissipates the “energy”
1
E= v 2 (x, t) dx (2)
2
of the wave.
of the stationary solution appearing under the integral sign in (14.5.1). Sub-
stituting this expression into the integral (14.5.1), we obtain
U4 dx U3 dz
Γ= = .
4μ cosh 4 Ux 2 cosh4 (z)
2μ
2 3
Γ= U . (3)
3
FIGURE 1
Plots of the dissipation of the density function γ(x) of a station-
ary wave, shown here for identical U ’s and different μ: μ1 = μ,
μ2 = μ/2, μ3 = μ/4. It is clear that as μ decreases, the densities
concentrate around their center point while keeping the area un-
derneath constant.
Chapter 14. Nonlinear Waves and Growing Interfaces 383
whose the integral over x gives the dissipation rate of the field (14.5.1). The
graphs of dissipation density for the stationary wave
U4
γ(x) = (5)
4μ cosh4 Ux
2μ
3. Let us first recover the initial conditions of the linear diffusion equa-
tion (14.3.11) corresponding to the given initial conditions of the Burgers
equation. The initial potential is
x
U1 U2
s0 (x) = v0 (y) dy = −2μ ln cosh (x − 1 ) cosh (x − 2 ) .
0 2μ 2μ
Here we have taken into account the fact that the initial potential s0 (x) is
defined only up to an additive constant, and the initial field ϕ0 (x) is defined
only up to a multiplicative constant. The notation is as follows:
U2 2 ± U1 1
U± = U2 ± U1 , ± = .
U2 ± U1
The solution ϕ(x, t) of the linear diffusion equation is obtained by noticing
that the diffusion equation
∂ϕ ∂ 2ϕ
= μ 2,
∂t ∂x
with the initial condition cosh(ax), has a solution with separable variables,
2
eμa t cosh(ax) .
Thus
U+2 U+
ϕ(x, t) = exp t cosh (x − + )
4μ 2μ
FIGURE 2
Merging stationary wave solutions, shown as functions of x, for
several values of the dimensionless time τ = U t/ and the dimen-
sionless parameter U /μ = 25. Observe that the effective width of
the transition area of the merged wave is roughly only one-half the
analogous areas for the initial stationary wave solutions.
Chapter 14. Nonlinear Waves and Growing Interfaces 385
U−2 U−
+ exp t cosh (x − − ) .
4μ 2μ
Substituting this expression in (14.3.9a), we finally get
U+ sinh U+
2μ
(x − + ) + exp − U1μU2 t U− sinh U−
2μ
(x − − )
v(x, t) = − .
cosh U+
2μ
(x − + ) + exp − U1μU2 t cosh U−
2μ
(x − − )
(6)
For concreteness, assume that U1 U2 > 0. Then, as t grows, the first terms
in the numerator and denominator of (8) dominate the second terms, and
the solution of the Burgers equation tends asymptotically to the stationary
solution
U+
lim v(x, t) = −U+ tanh (x − + ) . (7)
t→∞ 2μ
In other words, for large times, two initially separate stationary waves
form a single stationary wave with amplitude U+ = U1 + U2 . The wave is
centered on the point
U2 2 + U1 1
+ = . (8)
U2 + U1
In the special case U1 = U2 = U , 2 = −1 = , the solutions (6) of the
Burgers equation are shown in Fig. 2. The plots illustrate the process of jump
merging.
4. The skeleton (in the inviscid limit μ → 0+ ) of the initial field v0 (x),
see (14.5.2), is shown in Fig. 3.
The theory of weak solutions of the Riemann equation, see Sect. 13.5.5,
tells us that the jump’s (shock) velocity is the average of the velocities of the
jumps immediately to the left and to the right. In the case under considera-
tion, this means that the left jump moves to the right with velocity U2 , and
the right jump moves to the left with velocity U1 .
Thus, the equations of jumps’ motion before their collision are
FIGURE 3
The skeleton of the initial field v0 (x), see (14.5.2). The jumps move
in the directions indicated by the arrows.
the jumps merge, creating a standing stationary wave, whose jump coordi-
nate is
1 (t∗ ) = 2 (t∗ ) = + .
5. In this case, the initial potential and the initial condition of the corre-
sponding linear diffusion equation (14.3.11) are
and
ϕ0 (x) = 1 + e−R − 1 χ(x) + e−2R − e−R χ(x − ) .
Substituting the above expression in (14.3.13), we obtain
−R x −2R −R
x−
ϕ(x, t) = 1 + e − 1 Φ √ + e −e Φ √ .
2 μt 2 μt
−R x2
−2R 2
FIGURE 4
Plots of the solution (12) of the Burgers equation illustrating
merging of triangular waves.
The plots of this solutions, for different values of τ , are shown in Fig. 4. They
illustrate the process of merging of two triangular waves corresponding to
the initial condition in the form of a sum of two Dirac deltas (14.5.3).
6. At the initial stage, the skeleton of the field consists of two separate
triangles with the same area S. The triangles are bounded on the right by
the jumps. Let us denote the coordinates of the jumps by x1 (t) and x2 (t),
respectively. As long as the triangular waves do not overlap (see the upper
graph in Fig. 5), the jumps’ coordinates are easy to find using the fact that
the areas of the triangles are preserved:
√ √ t 2St
x1 (t) = τ , x2 (t) = (1 + τ) , τ= ∗
= 2 . (10)
t
At the time
2
t∗ = (τ = 1),
2S
the left jump begins to overlap the right triangle (see the lower graph in
Fig. 5), which changes the law of motion of the left jump. Let us find that
law by solving the left jump’s equation of motion,
dx1 (t) 1
= x1 (t) − , x1 (t∗ ) = .
dt t 2
388 Appendix A Answers and Solutions
This equation can be derived by recalling that the jump’s velocity is equal to
the average of the jumps’ velocities to the left (v− (t) = x1 (t)/t) and to the
right (v+ (t) = (x1 (t) − )/t).
FIGURE 5
The skeleton of the solution (9) of the Burgers equation.
As a result, we get
x1 (t) = (τ + 1) . (11)
2
Comparing the coordinates of the right (10) and left (11) jumps, we obtain
the desired merger time:
√
τm = (1 + 2)2 ≈ 5.83
Remark 1. Observe that at the beginning (for τ < 1), the field v(x, t)
consists of two nonoverlapping triangular waves, for each of which the
Reynolds number is R = S/2μ. In the course of time, they merge into a
single triangular wave with doubled area and doubled Reynolds number.
Thus the merger of jumps leads to growth in the current Reynolds number
and a loss of information about the fine structure of the initial field v0 (x).
7. Suppose that at the jump point of the function {y}(x, t), the parabolas
Π1 (x) and Π2 (x) intersect. Retaining the two corresponding summands in
(14.4.19), we shall rewrite {y}(x, t) as
Chapter 14. Nonlinear Waves and Growing Interfaces 389
(x−y1 )2 2
y1 exp − 2μt
1
s1 t + 2
+ y2 exp − 2μt
1
s2 t + (x−y2 2 )
{y}(x, t) ≈ .
(x−y1 )2 (x−y2 )2
exp − 2μt
1
s1 t + 2
+ exp − 1
2μt
s 2 t + 2
(12)
Define
y1 + y2 y2 − y1
x1 = , = , z = x − x1 .
2 2
Now, we can rewrite the expression (12) as
s2 t − z s1 t + z
exp − − exp −
2μt 2μt
{y} ≈ x1 + .
s2 t − z s1 t + z
exp − + exp −
2μt 2μt
Thus, finally,
y1 + y2 y2 − y1 s2 − s1
{y} ≈ + tanh (x − x1 − U t) , U= .
2 2 2μt y2 − y1
∞
n2 τ
ϕ(x, t) = I0 (R) + 2 In (R) exp − cos(nz) , (13)
n=1
2R
where
a
z = κx , τ = aκt , and R = .
2μκ
Substituting (13) in (14.3.9a) and retaining only the first summands in the
numerator and denominator, we get
2a I1 (R) − τ
v≈ e 2R sin(z) (τ R) . (14)
R I0 (R)
FIGURE 6
A plot of the dimensionless amplitude for an initially harmonic field
at the linear evolution stage. As long as it is well approximated by
the graph of the function g(R) = R, the nonlinear effects in the
evolution of the field v(x, t) can be neglected.
Plots of the equation at three consecutive time instants are shown in Fig. 7.
At first sight, the above solutions are counterintuitive. Indeed, our mental
picture would indicate that for x < 0, where v(x, t) > 0, the field should
move to the right, and the negative field, for x > 0, should be displaced to
the left. However, a more careful analysis of the solution and the graphics
below shows that just the opposite is true. This phenomenon is due to the
fact that the effective Reynolds number of the field (15) is close to 1, so that
the anticipated inertial effects are overwhelmed by a diffusive washout.
10. The function f (y; x, t), see (14.3.17), is uniquely determined by the
form of the function G(y; x, t). In the presence of V , it is
1
G(y; x, t|V ) = s0 (y)t + V yt + (y − x)2 .
2
Multiply the numerator and denominator of the right side of the equality
(14.3.17) by
V x V 2t
exp − .
2μ 4μ
FIGURE 7
Plots of the solution (15) of the Burgers equation at three successive
times, t1 = 1, t2 = 2, t1 , t3 = 3t1 . The “anomalous” behavior of the
solution is explained by the dominance of the diffusive effects over
the nonlinear inertial effects.
T = ρ0 E .
Integrating this equality term by term over all of s values and assuming that
the integral of u is finite while its first derivatives tend to zero as s → ±∞,
we obtain
∂
u(s, τ )ds = 0 ⇒ M = u(s, τ )ds = const . (2)
∂τ
Analogously, one can obtain the integral equality
# 2 $
∂ u2 ∂ u3 ∂2 1 ∂u
+ +γ u 2 − = 0.
∂τ 2 ∂s 3 ∂s 2 ∂s
also work satisfactorily for the intermediate values of the parameters. Let us
check whether this observation applies to the relation (4). Substituting the
initial condition (15.3.10) into (4), we obtain
1
n √ .
3πγ
5. To begin, assume, for the sake of generality, that the initial field u0 (s)
triggers an arbitrary number n of solitons. Neglecting the impact of oscillating
tails and taking into account that for τ large enough, solitons don’t overlap
and are described by the relation (15.3.32), we obtain
√ √ √ √
n n
u(s, τ ) 6 γ ci sech2 (z)dz = 12 γ ci
n=1 i=1
and
√ √ √ √
n n
u2 (s, τ ) 18 γ ci ci sech4 (z)dz = 24 γ ci ci .
n=1 i=1
Here, we have used the relation (15.3.32) and the integral formula (15.3.34).
Assuming that the soliton is unique and employing the momentum and
energy invariants (2) and (3), we obtain two approximate relations for the
soliton’s velocity,
√ M √ K
c √ , c c √ .
12 γ 12 γ
Let us check the accuracy of the above approximate relations in the case
of the initial field (15.3.10) and γ = 0.1. The corresponding numerically
calculated plots of the field u(s, τ ) are pictured in Fig. 15.3.1. It is easy to
calculate that in the case of the initial field (15.3.10), the momentum and
energy are equal to √
√ π
M = 2π , K= .
2
Accordingly, the values of a soliton’s velocity, depending on the momentum
and energy, are
c 0.437 and c 0.379.
Chapter 15. Other Standard Nonlinear Models 395
Let us check the above estimates, relying on the results of numerical calcula-
tions for the solution (15.3.11) of the KdV equation depicted in Fig. 15.3.1.
It follows from results of numerical calculations that the maximal value of
the field u(s, τ ), for τ = 50, is equal to
Assuming that this maximal value is equal to the soliton’s maximal value,
which is equal to three times the soliton’s velocity, we obtain
3c = 1.1289 ⇒ c = 0.376.
This last result agrees well with the above velocity estimate, which relied on
the energy, and is in satisfactory agreement with the estimate relying on the
momentum.
Appendix B
Bibliographical Notes
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 397
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3,
© Springer Science+Business Media New York 2013
398 Appendix B Bibliographical Notes
other areas of mathematics can be studied from many sources, starting with
the massive multivolume works
[7] I.M. Gelfand et al. Generalized Functions, six volumes,
Moscow, Nauka 1959–1966 (English translation: Academic Press,
New York and London, 1967), and
[8] L. Hörmander, The Analysis of Linear Partial Differential
Operators, four volumes, Springer, 1983–1985,
to smaller, one-volume research-oriented monographs
[9] E.M. Stein, G. Weiss, Introduction to Fourier Analysis on
Euclidean Spaces, Princeton University Press 1971,
[10] L.R. Volevich, S.G. Gindikin, Generalized Functions and
Convolution Equations, Moscow, Nauka 1994, (English transla-
tion: Gordon and Breach, 1990) and
[11] A. Friedman, Generalized Functions and Partial Differen-
tial Equations, Prentice Hall, Englewood Cliffs, N.J., 1963.
to textbook-style volumes less dependent on the locally convex topological
vectors space technology,
[12] R. Strichartz, A Guide to Distribution Theory and Fourier
Transforms, CRC Press, Boca Raton 1994,
[13] V.S. Vladimirov, Equations of Mathematical Physics,
Moscow, Nauka 1981 (English translation available on: http://
www.sps.org.sa/BooksandMagazinesLibraryFiles/Book_19_
152.pdf),
[14] G. Grubb, Distributions and Operators, Springer-Verlag
2009,
[16] J.J. Duistermaat, J.A.C. Kolk, Distributions: Theory
and Applications, Birkhäuser-Boston, 2010.
An elementary, but rigorous, construction of distributions based on
the notion of equivalent sequences was developed by
[14] J. Mikusiński, R. Sikorski, The Elementary Theory of
Distributions, I (1957), II (1961), PWN, Warsaw, and
[17] P. Antosik, J. Mikusiński, R. Sikorski, Generalized
Functions, the Sequential Approach, Elsevier Scientific, Amster-
dam 1973.
Appendix B Bibliographical Notes 399
which however, have a different spirit and do not cover some of the modern
areas covered by our book.
The classics on Fourier integrals are
The linear problems considered in Part III are classical, and there is an
enormous literature on the subject going back to
[32] J.-B. D’Alembert, Recherches sur la courbe que forme une
corde tendu mise en vibration, Histoire de l’Académie royale des
sciences et belles lettres de Berlin, 3 (1747), 214–219,
for the hyperbolic wave equations, and
[33] J.-B. Fourier, Théorie analytique de la chaleur, Paris 1822,
for the parabolic diffusion (heat) equations. A modern mathematical
treatment can be found in the above-mentioned four-volume treatise by Lars
Hörmander. The monumental, multivolume classic
[34] L. Landau and E. Lifschitz, Course of Theoretical
Physics, Moscow 1948–1957 (English translation: The first eight
volumes were translated into English by the late 1950s. The last
two volumes were written in the early 1980s. Vladimir Berestet-
skii and Lev Pitaevskii also contributed to the series. They were
published by various publishers: Pergamon press, Addison-Wesley
and Butterworth–Heinemann),
Appendix B Bibliographical Notes 401
The latter is a popular text for PDE courses for American graduate students
in pure mathematics. On the other hand,
[44] L. Garding, Linear hyperbolic partial differential equations
with constant coefficients, Acta Math., 85 (1950), 1–62.
provides an example of an in-depth modern research article.
Nonlinear problems discussed in Part IV have a vast literature both
mathematical and physical, and we just provide a small sample of some our
favorites:
[45] L. Hörmander, Lectures on Nonlinear Hyperbolic Differen-
tial Equations, Springer-Verlag, New York 1997,
A.I. Saichev and W.A. Woyczyński, Distributions in the Physical and Engineering Sciences, 405
Volume 2, Applied and Numerical Harmonic Analysis, DOI 10.1007/978-0-8176-4652-3,
© Springer Science+Business Media New York 2013
406 Index
T. Qian, V. Mang I, and Y. Xu: Wavelet Analysis and Applications (ISBN 978-3-7643-7777-9)
G.T. Herman and A. Kuba: Advances in Discrete Tomography and Its Applications
(ISBN 978-0-8176-3614-2)
M.C. Fu, R.A. Jarrow, J.-Y. J. Yen, and R.J. Elliott: Advances in Mathematical Finance
(ISBN 978-0-8176-4544-1)
P.E.T. Jorgensen, K.D. Merrill, and J.A. Packer: Representations, Wavelets, and Frames
(ISBN 978-0-8176-4682-0)
M. An, A.K. Brodzik, and R. Tolimieri: Ideal Sequence Design in Time-Frequency Space
(ISBN 978-0-8176-4737-7)
G.S. Chirikjian: Stochastic Models, Information Theory, and Lie Groups, Volume 1
(ISBN 978-0-8176-4802-2)
C. Cabrelli and J.L. Torrea: Recent Developments in Real and Harmonic Analysis
(ISBN 978-0-8176-4531-1)
P. Massopust and B. Forster: Four Short Courses on Harmonic Analysis (ISBN 978-0-8176-4890-9)
O. Calin, D. Chang, K. Furutani, and C. Iwasaki: Heat Kernels for Elliptic and Sub-elliptic Operators
(ISBN 978-0-8176-4994-4)
D. Joyner and J.-L. Kim: Selected Unsolved Problems in Coding Theory (ISBN 978-0-8176-8255-2)
J.A. Hogan and J.D. Lakey: Duration and Bandwidth Limiting (ISBN 978-0-8176-8306-1)