You are on page 1of 340

Lecture Notes in

Applied Differential
Equations of
Mathematical Physics
This page intentionally left blank
Lecture Notes in
Applied Differential
Equations of
Mathematical Physics
Luiz C L Botelho
Federal University, Brazil

World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI
This page intentionally left blank
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

LECTURE NOTES IN APPLIED DIFFERENTIAL EQUATIONS OF


MATHEMATICAL PHYSICS
Copyright © 2008 by World Scientific Publishing Co. Pte. Ltd.
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.

ISBN-13 978-981-281-457-9
ISBN-10 981-281-457-4

Printed in Singapore.

ZhangJi - Lec Notes in Applied.pmd 1 12/10/2008, 7:41 PM


August 6, 2008 15:46 9in x 6in B-640 fm

For all those who do not think that doing research


and writing is just to sign and be “co-author”
in other’s papers.
August 6, 2008 15:46 9in x 6in B-640 fm

This page intentionally left blank


August 6, 2008 15:46 9in x 6in B-640 fm

Foreword

“. . . Both Mathematics and Physics are in a state of crisis in present days,


from intellectual and social sides simultaneously. Both have much to con-
tribute to each other’s disciplines, and I believe that a renewal of their
traditional liaison would go far to cure at least the intellectual side of the
malaise felt by workers in both disciplines . . . ” — Robert Hermann in Lec-
tures in Mathematical Physics — Vol. I & Vol. II.
My aim in this set of informal lecture notes on applied mathematical
analysis is to present mathematical author’s original research material
(Chaps. 3–9 and appendixes) revised and amplified of our two previous work
on path-integrals in statistical and quantum physics and strongly focused
at a pure and applied mathematical audience of readers.
We have a strong believe that mathematics students can find a source
of inspiration to stimulate mutual interaction of their discipline and physics
and thus, undertaking the reading of physics books, talking with physi-
cists and try to read our previous two monographs in classical and
quantum physics. The special emphasis in this supplementary volume 3
is on advanced mathematical analysis methods with its applications to
solve partial differential equations in finite and infinite dimensions arising
in applied settings. Another objective behind writing this set of author’s
mostly original results in the form of lectures is the hope it could serve
to show that the “aridity” and “sterility” that one finds in much modern
“estimate mathematical analysis” can be conterweighted side by side by
mathematical analysis formulae exactly used in the modern applications.
A word on the “methodology” in our set of lecture notes as a set of com-
plementary notes we feel that the students should read the chapter with a
pencil and paper at hand, after which they should make additional studies
in the specific topic in the literature and thus present their studies in sem-
inars to others students with all lectures mathematical details worked out.

Luiz C.L. Botelho


Professor
vii
August 6, 2008 15:46 9in x 6in B-640 fm

This page intentionally left blank


August 6, 2008 15:46 9in x 6in B-640 fm

Contents

Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Chapter 1. Elementary Aspects of Potential Theory


in Mathematical Physics . . . . . . . . . . . . . . . . . . . . . . . 1

1.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2. The Laplace Differential Operator
and the Poisson–Dirichlet Potential Problem . . . . . . . . 1
1.3. The Dirichlet Problem in Connected Planar Regions:
A Conformal Transformation Method for Green Functions
in String Theory . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4. Hilbert Spaces Methods in the Poisson Problem . . . . . . . 18
1.5. The Abstract Formulation of the Poisson Problem . . . . . 21
1.6. Potential Theory for the Wave Equation in R3 —
Kirchhoff Potentials (Spherical Means) . . . . . . . . . . . . 23
1.7. The Dirichlet Problem for the Diffusion Equation —
Seminar Exercises . . . . . . . . . . . . . . . . . . . . . . . 27
1.8. The Potential Theory in Distributional Spaces —
The Gelfand–Chilov Method . . . . . . . . . . . . . . . . . 28
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Appendix A. Light Deflection on de-Sitter Space . . . . . . . . . 31
A.1. The Light Deflection . . . . . . . . . . . . . . . . . . . . . . 31
A.2. The Trajectory Motion Equations . . . . . . . . . . . . . . 34
A.3. On the Topology of the Euclidean Space–Time . . . . . . . 36

Chapter 2. Scattering Theory in Non-Relativistic


One-Body Short-Range Quantum
Mechanics: Möller Wave Operators
and Asymptotic Completeness . . . . . . . . . . . . . . . . . 39

ix
August 6, 2008 15:46 9in x 6in B-640 fm

x Lecture Notes in Applied Differential Equations of Mathematical Physics

2.1. The Wave Operators in One-Body Quantum Mechanics . . 39


2.2. Asymptotic Properties of States in the Continuous Spectra
of the Enss Hamiltonian . . . . . . . . . . . . . . . . . . . . 44
2.3. The Enss Proof of the Non-Relativistic One-Body
Quantum Mechanical Scattering . . . . . . . . . . . . . . . 48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Appendix B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Appendix C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

Chapter 3. On the Hilbert Space Integration


Method for the Wave Equation
and Some Applications to Wave Physics . . . . . . . 58
3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.2. The Abstract Spectral Method — The
Nondissipative Case . . . . . . . . . . . . . . . . . . . . . . 59
3.3. The Abstract Spectral Method — The Dissipative Case . . 63
3.4. The Wave Equation “Path-Integral” Propagator . . . . . . 65
3.5. On The Existence of Wave-Scattering Operators . . . . . . 68
3.6. Exponential Stability in Two-Dimensional
Magneto-Elasticity: A Proof on a Dissipative
Medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.7. An Abstract Semilinear Klein Gordon Wave
Equation — Existence and Uniqueness . . . . . . . . . . . . 73
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Appendix A. Exponential Stability in Two-Dimensional
Magneto-Elastic: Another Proof . . . . . . . . . . 76
Appendix B. Probability Theory in Terms of Functional
Integrals and the Minlos Theorem . . . . . . . . . 80

Chapter 4. Nonlinear Diffusion and Wave-Damped


Propagation: Weak Solutions
and Statistical Turbulence Behavior . . . . . . . . . . . . 84
4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.2. The Theorem for Parabolic Nonlinear Diffusion . . . . . . . 85
4.3. The Hyperbolic Nonlinear Damping . . . . . . . . . . . . . 91
4.4. A Path-Integral Solution for the Parabolic
Nonlinear Diffusion . . . . . . . . . . . . . . . . . . . . . . . 94
August 6, 2008 15:46 9in x 6in B-640 fm

Contents xi

4.5. Random Anomalous Diffusion, A Semigroup Approach . . . 96


References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Appendix B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Appendix C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Appendix D. Probability Theory in Terms of Functional
Integrals and the Minlos Theorem —
An Overview . . . . . . . . . . . . . . . . . . . . . 107

Chapter 5. Domains of Bosonic Functional Integrals


and Some Applications to the
Mathematical Physics of Path-Integrals
and String Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.2. The Euclidean Schwinger Generating Functional as
a Functional Fourier Transform . . . . . . . . . . . . . . . . 111
5.3. The Support of Functional Measures — The
Minlos Theorem . . . . . . . . . . . . . . . . . . . . . . . . 114
5.4. Some Rigorous Quantum Field Path-Integral
in the Analytical Regularization Scheme . . . . . . . . . . . 124
5.5. Remarks on the Theory of Integration of Functionals on
Distributional Spaces and Hilbert–Banach Spaces . . . . . . 131
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Appendix B. On the Support Evaluations of Gaussian Measures 148
Appendix C. Some Calculations of the Q.C.D. Fermion
Functional Determinant in Two-Dimensions
and (Q.E.D.)2 Solubility . . . . . . . . . . . . . . . 150
Appendix D. Functional Determinants Evaluations
on the Seeley Approach . . . . . . . . . . . . . . . 160

Chapter 6. Basic Integral Representations


in Mathematical Analysis of Euclidean
Functional Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
6.1. On the Riesz–Markov Theorem . . . . . . . . . . . . . . . . 169
6.2. The L. Schwartz Representation Theorem on C ∞ (Ω)
(Distribution Theory) . . . . . . . . . . . . . . . . . . . . . 173
August 6, 2008 15:46 9in x 6in B-640 fm

xii Lecture Notes in Applied Differential Equations of Mathematical Physics

6.3. Equivalence of Gaussian Measures in Hilbert Spaces


and Functional Jacobians . . . . . . . . . . . . . . . . . . . 178
6.4. On the Weak Poisson Problem in Infinite Dimension . . . . 183
6.5. The Path-Integral Triviality Argument . . . . . . . . . . . . 186
6.6. The Loop Space Argument for the Thirring
Model Triviality . . . . . . . . . . . . . . . . . . . . . . . . 193
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
Appendix A. Path-Integral Solution for a Two-Dimensional
Model with Axial-Vector-Current-Pseudoscalar
Derivative Interaction . . . . . . . . . . . . . . . . 197
Appendix B. Path Integral Bosonization for an Abelian
Nonrenormalizable Axial Four-Dimensional
Fermion Model . . . . . . . . . . . . . . . . . . . . 202

Chapter 7. Nonlinear Diffusion in RD and Hilbert


Spaces: A Path-Integral Study . . . . . . . . . . . . . . . . . 209
7.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 209
7.2. The Nonlinear Diffusion . . . . . . . . . . . . . . . . . . . . 209
7.3. The Linear Diffusion in the Space L2 (Ω) . . . . . . . . . . . 216
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
Appendix A. The Aubin–Lion Theorem . . . . . . . . . . . . . . 223
Appendix B. The Linear Diffusion Equation . . . . . . . . . . . 225

Chapter 8. On the Ergodic Theorem . . . . . . . . . . . . . . . . . . . . . . 227


8.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 227
8.2. On the Detailed Mathematical Proof of the RAGE Theorem 228
8.3. On the Boltzmann Ergodic Theorem in Classical
Mechanics as a Result of the RAGE Theorem . . . . . . . . 231
8.4. On the Invariant Ergodic Functional Measure for Some
Nonlinear Wave Equations . . . . . . . . . . . . . . . . . . 234
8.5. An Ergodic Theorem in Banach Spaces
and Applications to Stochastic–Langevin
Dynamical Systems . . . . . . . . . . . . . . . . . . . . . . 241
8.6. The Existence and Uniqueness Results for Some
Nonlinear Wave Motions in 2D . . . . . . . . . . . . . . . . 245
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Appendix A. On Sequences of Random Measureson Functional
Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 249
August 6, 2008 15:46 9in x 6in B-640 fm

Contents xiii

Appendix B. On the Existence of Periodic Orbits in a Class of


Mechanical Hamiltonian Systems —An Elementary
Mathematical Analysis . . . . . . . . . . . . . . . . 251
B.1. Elementary May be Deep — T. Kato . . . . . . . . . . . . 251

Chapter 9. Some Comments on Sampling of Ergodic


Process: An Ergodic Theorem
and Turbulent Pressure Fluctuations . . . . . . . . . . . 254
9.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 254
9.2. A Rigorous Mathematical Proof of the Ergodic
Theorem for Wide-Sense Stationary Stochastic Process . . 255
9.3. A Sampling Theorem for Ergodic Process . . . . . . . . . . 257
9.4. A Model for the Turbulent Pressure Fluctuations
(Random Vibrations Transmission) . . . . . . . . . . . . . . 262
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
Appendix A. Chapters 1 and 9 — On the Uniform Convergence
of Orthogonal Series — Some Comments . . . . . 266
A.1. Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . 266
A.2. Regular Sturm–Liouville Problem . . . . . . . . . . . . . . . 270
Appendix B. On the Müntz–Szasz Theorem on Commutative
Banach Algebras . . . . . . . . . . . . . . . . . . . 272
B.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 272
Appendix C. Feynman Path-Integral Representations
for the Classical Harmonic Oscillator
with Stochastic Frequency . . . . . . . . . . . . . . 274
C.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 274
C.2. The Green Function for External Forcing . . . . . . . . . . 274
C.3. The Homogeneous Problem . . . . . . . . . . . . . . . . . . 278
Appendix D. An Elementary Comment on the Zeros of the Zeta
Function (on the Riemann’s Conjecture) . . . . . . 282
D.1. Introduction — “Elementary May be Deep” . . . . . . . . . 282
D.2. On the Equivalent Conjecture D.1 . . . . . . . . . . . . . . 283

Chapter 10. Some Studies on Functional Integrals


Representations for Fluid Motion with
Random Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
10.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 287
August 6, 2008 15:46 9in x 6in B-640 fm

xiv Lecture Notes in Applied Differential Equations of Mathematical Physics

10.2. The Functional Integral for Initial Fluid Velocity


Random Conditions . . . . . . . . . . . . . . . . . . . . . 287
10.3. An Exactly Soluble Path-Integral Model for Stochastic
Beltrami Fluxes and its String Properties . . . . . . . . . 292
10.4. A Complex Trajectory Path-Integral Representation
for the Burger–Beltrami Fluid Flux . . . . . . . . . . . . . 298
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
Appendix A. A Perturbative Solution of the Burgers Equation
Through the Banach Fixed Point Theorem . . . . 304
Appendix B. Some Comments on the Support of Functional
Measures in Hilbert Space . . . . . . . . . . . . . . 309

Chapter 11. The Atiyah–Singer Index Theorem:


A Heat Kernel (PDE’s) Proof . . . . . . . . . . . . . . . . . 312
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Appendix A. Normalized Ricci Fluxes in Closed Riemann
Surfaces and the Dirac Operator in the Presence
of an Abelian Gauge Connection . . . . . . . . . . 320

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
August 6, 2008 15:46 9in x 6in B-640 ch01

Chapter 1

Elementary Aspects of Potential Theory


in Mathematical Physics

1.1. Introduction

Many problems of elementary classical mathematical-physical analysis


are connected with the solution through integral representations of the
Dirichlet, Newmann, and Poisson problems for the Laplacian operator,
including its generalization to the Cauchy problem for wave and diffusion
equations. In this chapter, we present the basic and introductory mathe-
matical methods analysis used in obtaining such above-mentioned integral
representations.1

1.2. The Laplace Differential Operator


and the Poisson–Dirichlet Potential Problem

Let Ω be an open-bounded set in RN . We define C 2 (Ω) to be the vectorial


space composed by those functions f (x) ≡ f (x1 , . . . , xn ) ≡ f (r ). (Here we

adopt the physical notation r = N i
i=1 x ei with 
ei denoting the canonical
N
vectorial basis of R ), continuously differentiable until the second order in
Ω, namely,
∂ 2 f (x)
∈ C(Ω), for any (k, m) ∈ {1, . . . , n}.
∂xk ∂xm
Let us consider the following linear transformation with domain C 2 (Ω)
and range C(Ω)

∆(N ) : C 2 (Ω → C(Ω)
 
n
∂2
f (x) → (∆f )(x) = f (x1 , . . . , xn ). (1.1)
i=1
∂x2i

We now have the following simple result about the kernel of the above
linear transformation in R3 (holding true for any RN , with N ≥ 2).

1
August 6, 2008 15:46 9in x 6in B-640 ch01

2 Lecture Notes in Applied Differential Equations of Mathematical Physics

Lemma 1.1. The kernel of the Laplacean operator as defined by Eq. (1.1)
has infinite dimension in R3 .
2 ¯
Proof. Let f¯ ∈ C 2 (C × R) such that ∂∂2 w f
(w, t) exists and the function
1 2 3
defined below (x = x, x = y, x = z — the usual classical notation in
potential theory in the physical space R3 )

Ut (x, y, z) = f¯(ix cos t + iy sin t + z, t). (1.2)

We wish to show that


 
∆(3) Ut (x, y, z) = 0. (1.3)

This result is a simple consequence of the following elementary


identities:
∂2 ∂2 ¯
U t (x, y, z) = f (w, t)|w=z+ix cos t+iy sin t , (1.4a)
∂2z2 ∂2w
∂2 ∂2
2 2
Ut (x, y, z) = 2 f¯(w, t)|w=z+ix cos t+iy sin t ((i cos t)2 ), (1.4b)
∂ x ∂ w
∂2 ∂2
2 2
Ut (x, y, z) = 2 f¯(w, t)|w=z+ix cos t+iy sin t (i sin t)2 . (1.4c)
∂ y ∂ w
As a consequence
  ∂2 
∆(3) Ut (x, y, z) = 2 f (w, t)w=z+ix cos t+iy sin t (1 − cos2 t − sin2 t) ≡ 0.
∂ w
(1.5)

Note that a whole class of harmonic functions H̃(Ω) in the class of


C 2 (Ω)-functions in the kernel of the Laplacean operator Eq. (1.3) may be
obtained from Eq. (1.3):
 b
U (x, y, z) = dt g(t)(f¯(w, t)|w=ix cos t+i sin ty+z ), (1.6)
a

and for the special function below:

f¯(w, t) = wn eiαt , (α ∈ R, n ∈ Z),


August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 3

we have the harmonic legendre polynomials in the three-dimensional


spherical ball
 π
U (x, y, z) = eiαt (z + ix cos t + iy sin t)n dt
−π
 π
= dt eiαt (r cos θ + ir sin θ cos ϕ cos t + ir sin θ sin ϕ sin t)n
−π
 π
= 2rn eiαϕ dt cos(αt)(cos θ + (sin θ sin t)n
0

= rn eiαϕ Pn,α (cos θ). (1.7)

After these preliminary elementary remarks we pass on the discussion


of the famous classical Poisson–Newton problem: Let Ω ⊂ R3 be a bounded
open set of R3 with a boundary ∂Ω being an orientable simply con-
nected C 2 -surface. We search for the solution of the nonhomogeneous linear
problem in Ω through an integral representation (the Laplacean Green
function).

−∆3 U (x, y, z) = f (x, y, z). (1.8)

Here f (x, y, z) ∈ C 1 (Ω̄) (Ω̄ = Ω ∪ ∂Ω).


We have the basic result in answering the Poisson–Newton problem.


Theorem 1.1. The Laplacean operator

−∆3 : C 2 (Ω) H̃(Ω) → C 1 (Ω̄) (1.9)

is inversible, and its inverse has the following integral representation:


 
1    f (x , y  , z  )
U (x, y, z) = dx dy dz
4π Ω (x − x )2 + (y − y  )2 + (z − z  )2

def
≡ d3r  f (r  ) · G(0) (|r − r  |). (1.10)

August 6, 2008 15:46 9in x 6in B-640 ch01

4 Lecture Notes in Applied Differential Equations of Mathematical Physics

Proof. Let us re-introduce the “classic mathematical-physicist” vectorial


notation in our formulas

r = (x, y, z); r  = (x , y  , z  ). (1.11)

Let us consider the δ-approximant solution for the problem



1
d3r  f (r  ) G(0) (|r − r  |).
(δ)
U (δ) (r ) = (1.12)
4π Ω

Here the “regularized” Green function is given by


 1

 for |r − r  | > δ,
   |r − r  | ,
G(0) |r − r  | =
(δ)
(1.13)

 1 |r − r  |2
 3− , for |r − r  | ≤ δ.
2δ δ2
Note that the above “regularized” Green function is a continuous
function in both variables, i.e.

G(0) (|r − r  |) ∈ C(Ω × Ω)(since G(0) (δ + ) = G(0) (δ − )).


(δ) (δ) (δ)

We now have the estimate for δ → 0:


  12
 (δ)  1  2
sup |U − U |(r ) ≤ f L2(Ω) ×
2
d rG(δ) (|r − r  |)
3
r∈Ω
 Bδ (0)
1 √
≤ f L2(Ω) ( 2πδ),
2
(1.14)

which shows the uniform convergence of the sequence of approximate


solutions to the (well-defined) function U (r ) by the Weierstrass uniform
convergence criterium.
That the integral representation Eq. (1.10) defines a well-defined
function comes straightforwardly from the estimate around our infinitesimal
ball Bε (r ) = {r  | ||r  − r| ≤ ε}.
    ε 
 3  f ( r  )  1 1
 d 
r ≤ f  2
× 4π dv · v 2
·
 |r − r  | 
2
L (Ω) v2
Bε (r) 0
1 1
≤ (4πε) 2 f L2 2(Ω) < ∞. (1.15)

As a second step to the proof of Theorem 1.1, we point out that the
sequence of functions

1
∇ · U (δ) = d3r f (r  )∇ · G(δ) (|r − r  |) (1.16)
4π Ω
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 5

converges uniformly to the (well-defined) function in C(Ω̄)

 
3   1
I=+ d r f (r )∇r . (1.17)
Ω 4π|r − r  |

The proof of the above made claim is a consequence of the following


estimate:

    2 
 |r − r  |2 (x − x ) 
lim d3r   ∇rx − + 
 = 0. (1.18)
δ→0 Bδ (
r) δ 2 |r − r  |3

Let us now pass on the problem of evaluating second derivatives.


In this case we have the Gauss theorem (integration by parts in R3 !)


1
d3r  f (r  ) ∇r
Ω |r − r  |

1
=− d3r f (r  ) ∇r 
Ω |r − r  |

1  
= −O f (r  ) 
 ∠r  ) dS(r  )
cos(N
∂Ω |r − r |

  1
+ d3r  ∇r  · f (r  ) . (1.19)
Ω |r − r  |

Here cos(N  ∠r ) is the cosine between the normal field of the globally ori-
ented surface ∂Ω and the point r, argument of the function U (r ). At this
point we are using the fact that f ∈ C 1 (Ω̄). However, Eq. (1.19) has a rig-
orous meaning for f ∈ L2 (Ω̄) (left as an exercise to our diligent reader!).
It is important to remark that at this point of our exposition all these
topological–geometrical constraints imposed on the surface ∂Ω by means
of Gauss theorem must be imposed as complementary hypothesis on the
geometrical nature of the Poisson problem.
By deriving a second time the left-hand side of Eq. (1.19) and taking
into account that in the surface integral we always have r = r  (we do
not evaluate the solution U (r ) at the boundary points) and obviously
∇r  f (r ) ≡ 0, we arrive at the following result (without bothering ourselves
August 6, 2008 15:46 9in x 6in B-640 ch01

6 Lecture Notes in Applied Differential Equations of Mathematical Physics

with the 4π-overall factor in Eq. (1.10)):



1
∆r U (r ) = − O f (r  ) ∇r cos(N  ∠r  )dS(r  )
∂Ω |
r − r  |

  1
+ d3r  ∇r  f (r  ) − f (r ) ∇r
Ω |r − r  |

1
= − O f (r  ) ∇r cos(N  ∠r  )dS(r  )
∂Ω |
r − r  |

  1
+ O f (r  ) − f (r ) ∇r cos(N  ∠r  )dS(r  )
∂Ω |r − r  |

1
− (f (r  ) − f (r )) ∆r d3r  . (1.20)
Ω |
r − r  |
Now we have that
 
1  
f (r ) O ∇r cos(N ∠r )dS(r )

∂Ω |r − r  |
  · r  
N
= f (r ) O ds(r ) = 4πf (r ) (1.21)
∂Ω |r  |3
and (with xa = (x, y, z) and a = 1, 2, 3):
   
 (xa − xa )2 3  
  3
 (f (r ) − f (r ) − + d r 
 Bε (
r) |r − r  |3 |r − r  |5 

|r − r  |
≤ sup |Df |(r  ) × 3 d3r 
r  ∈Bε (
 r) Bε (
r) |r − r  |3
 
|xa − xa |2 |r − r  | 3  ε→0 .
+ d r →0 (1.22)
Bε (
r) |r − r  |5
Precisely, to obtain this estimate, we need the condition that the data
f (r ) must belong to the space C 1 (Ω̄) (including the continuity of the
derivative at the boundary ∂Ω!).
As a general exercise in this section we left to our reader to prove
Theorem 1.1 in the general case of the domain Ω ⊂ RN (N ≥ 3) with the
Green function in RN .
   
Γ N2
U (r ) = √ N × dN r  f (r  )|r − r  |2−N . (1.23)
(N − 2)2( π) Ω
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 7

Note that the regularized Green function G(δ) is now given by

G(δ) |r, r  )
 1

 if |r − r  | > δ,
 |r − r  |N −2
=  N

 δ 2−N 2 N −2 |r − r  |
 2− − if |r − r  | ≤ δ.
(N − 2)wN N 2 δ
(1.24)

In R2 , we can show that (left to our readers)


 
1 1
U (r ) = − d2r  g f (
r 
) . (1.25)
2π Ω |r − r  |

This case can be seen as coming from the general case Eq. (1.23) after
considering the (distributional) limit
  
Γ 22 1 r |
r− 1
lim √ e (2−N )g|
=− g|r − r  |. (1.26)
N →2 2( π)2 N − 2 2π

We consider now the well-known Dirichlet problem: Determine a U (r ) ∈


C 2 (Ω) ∩ C 1 (Ω̄) such that

∆ U (r ) = 0, for r ∈ Ω,
  (1.27)
U (r ) = g(r )
∂Ω ∂Ω
,

a very useful problem in string path integrals for Ω ⊂ R2 .


By a simple application of the first Green identity we obtain the
uniqueness property for the Dirichlet problem

 0
 
 
U · ∆U +|∇U | (r )d r = O (U · ∇N U )(r )dΓ(r ) = 0.
2 3
(1.28)
Ω ∂Ω

As a consequence,

|∇U |2 (r )d3r = 0 ⇔ U (r ) ≡ 0 in Ω. (1.29)

An explicit solution for Eq. (1.27) is easily obtained by considering the


Poincaré method of considering a Green function of the structural form
below.
August 6, 2008 15:46 9in x 6in B-640 ch01

8 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let
1
G(r, r ) = H(r, r  ) + , (1.30)
4π|r − r  |
where H(r, r  ) is a harmonic function in Ω in relation to the variable r  ,
satisfying the following condition:

1 

H(r, r )|r  ∈∂Ω = −  . (1.31)
4π|r − r  | r  ∈∂Ω
Then we have the following explicit representation: 
Theorem 1.2. We have the integral representation for the Dirichlet
problem
 
− 1   
 
U (r ) =    4π O g(
r )∇N G(
r , 
r ) dS(
r )
(−1/2π if r ∈ ∂Ω − exercise) ∂Ω

 
1 1
=− O g(r  )∇N H(r, r  ) + dS(
r 
) . (1.32)
4π ∂Ω 4π|r − r  |
Proof. Let us consider the regularized form for the Green function as
proposed by the Poincaré ansatz
G(δ) (r, r  ) = H (δ) (r, r  ) + U (δ) (r − r  ). (1.33)
Here

∆r H (δ) (r, r  ) = 0,
 (1.34)
H (δ) (r, r  ) = −U (δ) (r − r  ),
r  ∈∂Ω

with
 1

 , if |r − r  | > δ,

4π|
r − r  |

U (r − r ) =
(δ)
  (1.35)

 1 1 |r − r  |2
 3− , if |r − r  | ≤ δ.
4π 2δ δ2
By applying the second Green identity to the pair of functions U (r )
and G(δ) (r, r  ), we obtain the relation

  
U (r  ) ∆r  U (δ) (r, r  ) d3r 



= O U (r  ) ∇N G(δ) (r, r  )dS(r  ). (1.36)
∂Ω
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 9

By noting that we have explicitly the Laplacean action on the regu-


larized function U (δ) :

0, if |r  − r  | > δ,
(δ) 
∆r  U (r, r ) = 3 (1.37)
− , if |r − r  | ≤ δ,
4πδ 3

and by taking the limit of δ → 0, it yields the integral representation


Eq. (1.30).
Our problem now is to obtain an explicit procedure to determine the
Harmonic piece of the Green function, namely, H(r, r  ).
A solution was first given by H. Poincaré through the use of the so-called
double-layer potential Φ(r, r  ), satisfying the Poincaré integral equation in
L2 (∂Ω, dS):

1 Φ(r, r  ) 1 1
− = − O ∇N  × dS(w).
Φ(r, w) 
4π|r − r  | 2 ∂Ω 4π |
r − r  |
(1.38)
Note the somewhat geometrical complexity of the geometrical measure
dS on the surface ∂Ω, when the integral equation above is solved by the
standard Fixed Point Theorems.1,2
After solving Eq. (1.38), Poincaré showed that one easily has a formula
for the function H(r, r  ):

1
H(r, r  ) = O Φ(r, r  )∇N − |
dS(r  ). (1.39a)
∂Ω 4π|
r − 
r
Another point worth to call the reader’s attention is that it appears more
straightforward to consider the following integral equation for the surface
derivative of the harmonic function U (r ). Since
 
1 1   1
U (r ) = O ∇  U (r ) − U (r )∇N dS(r ),
4π ∂Ω |r − r  | N |r − r  |
(1.39b)
we have
 
∇  (r ) · N
 r U  (r ) = ∇  U (r ) = ρ(r )
N
r )
 
ρ(
  
1 (r − r ) 
= O − N (
r ) · ∇ r )
 U (
4π ∂Ω |r − r  |3 N


1
−g(r  )∇N ∇N dS(r  ). (1.39c)
|r − r  |
August 6, 2008 15:46 9in x 6in B-640 ch01

10 Lecture Notes in Applied Differential Equations of Mathematical Physics

One can in principle solve the above-written integral equation in


2
L (∂Ω, dS(r )) by the Banach fixed point theorem at least for small volumes
and for the normal derivative of the harmonic function U (r ) and then using
again the Liouville integral representation Eq. (1.39b) to determine U (r )
in Ω.
A more useful approach for the explicit determination of the harmonic
term H(r, r  ) in the Poincaré–Green function is by means of the eigenvalue
problem for our Laplacean operator with a Dirichlet condition
−∆ ϕn (r ) = λn ϕn (r ) for r ∈ Ω, (1.40)

ϕn (r ) = 0
∂Ω
for r ∈ ∂Ω,
or in the equivalent integral form (see Theorem 1.1)

ϕn (r  )
µn ϕn (r ) = d3r  (1.41)
Ω 4π|r − r  |
with
   
1
µn = = ϕ2n (r )d3r |∇ϕn |2 (r )d3r > 0.
λn Ω Ω

Now it is straightforward to see that


 
ϕn (r ) 1 (∇N ϕn (r  ))dS(r  )
d3r  G(r, r  )ϕn (r  ) = − O
Ω λn λn ∂Ω 4π|r − r  |

(−∆r  ϕn (r  )) 3 
+ H(r, r  ) d r .
Ω λn
(1.42)
At the point we left as an exercise to our reader to show that the surface
integral and the volume integral cancel out.
By the usual Mercer theorem (if supr∈Ω̄ |ϕn (r )| ≤ M , for ∀ n ∈ Z+ ), we
have the uniform convergence of the spectral  λ series defining the Dirichlet
Green Function as a result that limn→∞ |n| n
2 < ∞,
 |ϕn (r )ϕn (r  )|  1 

|G(r, r )| ≤ ≤M 2
< ∞. (1.43)
3
λn 3
|n|2
n∈Z n∈Z

The same result holds true for the case of the Newmann problem written
below:

∆ U (r ) = 0, for r ∈ Ω,

(1.44a)
∇N U (r ) = g(r ) O g(r  )dS(r  ) ≡ 0 ,
∂Ω
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 11

with

U (r ) = + O g(r )G(r, r  )dS(r  ), (1.44b)
∂Ω

where
 βn (r )βn (r  )
G(r, r  ) = , (1.44c)
n
λn

and with the eigenfunction/eigenvalue problem



−∆ βn (r ) = λn βn (r ), for r ∈ Ω,
(1.45)
∇N · βn (r ) = 0, for r ∈ ∂Ω.
Finally, it is left to our readers as a highly nontrivial exercise to gener-
alize all the above results (their proofs!) for the case of Ω possessing a Rie-
mannian structure: The famous Laplace–Beltrami operator acting on scalar
functions into the domain (Ω, gab (x)), Ω ⊂ RN , ∂a = ∂x∂ a (a = 1, . . . , n)

1 √
−∆g U = − √ ∂a ( g g ab ∂b ) U (xa ) (1.46)
g
mathematical object instrumental in quantum geometry of strings and
quantum gravity.4
For instance, for two-dimensional Riemann manifolds, one can always
locally parametrize the Riemann metric into the conformal form in the
trivial topological Riemann surface sector gab (x, y) = ρ(x, y)δab .
An important application in potential theory is the use of the famous
multipole expansion inside the Laplacean Green function (r = r|, r = |r  |):

1 1 r · r  3 r · r  (r )2
= + 3 + − + ··· . (1.47a)
|r − r  | r r 2π r2 2r3
A useful explicit expression for the Laplace Green function can be
obtained for the Ball Ba (0) = {r ∈ R3 | |r| = r ≤ a} in spherical
coordinates
 2π  π 
a(r2 − a2 )    f (θ , ϕ )
U (r, θ, ϕ) = dϕ sin θ dθ ,
4π 0 0 (a2 + r2 − 2ar cos Ω)
(1.47b)

where the composed angle in the above-written formula is defined as


 
Ω (θ, θ ); (ϕ, ϕ ) = cos θ cos θ + sin θ sin θ cos(ϕ − ϕ ). (1.47c)
August 6, 2008 15:46 9in x 6in B-640 ch01

12 Lecture Notes in Applied Differential Equations of Mathematical Physics

Another point worth remarking is that Eqs. (1.38), (1.39a) and (1.39c)
have potentialities for problems involving weak perturbations of the domain
boundary ∂Ω(0) → ∂Ω(0) + ε ∂ Ω(1) , since we have for instance the
explicit perturbation result for the integrand of Eq. (1.38) for a given
parametrization of the perturbed surface
  (0) 
  = W1 (u, v)ê1 + W2 (u, v)ê2 + W3 (u, v)ê3 = W
W  + εW  (1) (u, v)







 (u, v) = (W + εW ) × (W + εW ) = N
  (0)  (1)  (0)  (1)

 N  (0) + εN (1) + · · ·

 |(W
 (0) + εW (1) ) × (W
 (0) + εW
 (1) )| 3 R


  ·(0) εW

 1 1 (r − W  (1) )

 = + + ···

 |r − W
 (0) − εW
 (1) | |r − W
 (0) | |r − W  (0) |3




  ) = Φ(0) (r, W (0) ) + ε∇Φ(1) (r, W (0) ) · ∇W  (1) + · · · .
Φ(r, W
(1.47d)

Rigorous mathematical analysis of such ε-perturbative expansion is left


to our mathematically oriented readers, including the open problem in
advanced calculus to find explicitly the Green function for a toroidal surface
and the Möebius surface band and its weakly random perturbations. 
An important mathematical question is the one about the well-
posedness properties of the Dirichlet problem in relation to the uniform con-
vergence of the datum. We have the following result in this direction:

Lemma 1.2. If in Eq. (1.27) gn (r ) ∈ C 0 (∂Ω) converges uniformly for a


function g(r ) ∈ C0 (∂Ω), then we have the uniform convergence of the solu-
tions (1.32) associated with the sequence data {gn } to the solution associated
with g(r ).

Proof. Let Ḡ be a compact in Ω such that dist(Ḡ, ∂Ω) = δ > 0. Since we


have the following estimate for the Green function
 
 1 

sup |G(r, r )| ≤ sup   ≤ 1 (1.47e)
 r  ∈∂Ω
r ∈G;
 r∈G;
 r − r |  4πδ
 r  ∈∂Ω 4π|


and so on,
Area(∂Ω)
sup |Un (r  ) − U (r )| ≤ ||gn − g||C 0 (∂Ω) → 0. (1.47f)
r ∈Ḡ
 4πδ

August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 13

Finally, let us announce the combined problem of Poisson–Dirichlet in


3
R and its associated integral representation.

Theorem 1.3. Let Ω ⊂ R3 be a region satisfying the hypothesis of the


Gauss theorem. Let g(r ) and f (r ) be elements in C 1 (Ω̄); the solution of
this problem in C 2 (Ω) ∩ C 1 (Ω̄)

∆U (r ) = f (r ); r ∈ Ω,
  (1.48)
U (r )∂Ω = g(r )∂Ω ,

is (uniquely) given by the integral representation written below:



1 f (r  )
U (r ) = d3r  + v(r  ). (1.49)
4π Ω |r − r  |

Here the function v(r ) is the harmonic function related to the boundary
data

∆ v(r ) = 0,
 
 3  f ( r )  (1.50)
v(r )∂Ω = − d r + g(
r ) 
 .
Ω |r − r  | r ∈∂Ω

1.3. The Dirichlet Problem in Connected Planar Regions: A


Conformal Transformation Method for Green Functions
in String Theory

Let Ω be a simply connected region in R2 and W ⊂ R2 another region


with similar topological properties. A coordinate transformation (diffeo-
morphism) of W into Ω

F
W ⊂ R2 −→ Ω ⊂ R2
(1.51)
(ξ, η) (x(ξ, η), y(ξ, η))

is called conformal when one has the validity of the conditions in W :



 ∂x(ξ, η) ∂y(ξ, η)

 ∂ξ = ,
∂η
(1.52)


 ∂x(ξ, η) = − ∂y(ξ, η) .
∂η ∂ξ
August 6, 2008 15:46 9in x 6in B-640 ch01

14 Lecture Notes in Applied Differential Equations of Mathematical Physics

In this case one can see that z = x + iy = f −1 (ξ + iη) and f (z) is an


analytic function in Ω. We have the following relation among the functional
expressions of the operator Laplaceans among the two given domains W
and Ω:
2 
  ∂x
2
∂y 
∆ξ,η U (ξ, η) = ∆x,y Ū(x, y) +  , (1.53)
∂ξ ∂η 
ε+iη=f (x+iy)

where U (ξ(x, y), η(x, y)) ≡ Ū(x, y).


In the case of the given domain transformation F possesing a unique
extension as a diffeomorphism between the domain boundaries Γ1 = ∂W
and Γ2 = ∂Ω, we have the following relationships among the Dirichlet
problems in each region:


∆ξ,η U (ξ, η) = 0 (ξ, η) ∈ W,
(a) (1.54)
U (ξ, η) = f (ξ, η) ,

Γ1 Γ1



∆x,y Ū (x, y) = 0 (x, y) ∈ Ω,
(b) (1.55)
Ū (x, y)
 
= f¯(x, y)Γ .
Γ2 2

Then,

U (ξ, η) = Ū (x, y). (1.56)

It is a deep theorem of B. Riemann in the theory of one-complex


variables that given two simply connected bounded domains where the
boundaries are curved with continuous curvature there is a conformal trans-
formation of the given domains realizing a coordinate (sense preserving)
of the boundaries.3 A proof of such a result can be envisaged along the
following arguments: Firstly, one considers the universal domain B1 (0) =
{(x, y) | x2 + y 2 < 1} as the domain Ω. Secondly, one considers a triangu-
larization T(n) of the region W . It is possible to write explicitly a conformal
transformation of T(n) into B1 (0) by means of the well-known Schwartz–
Christoffell transformation of a polygon into the disc B1 (0) denoted by
fn (z). One can show now that the family of functions {fn (z)} is a uniformly
bounded (compact) set in the space H(Ω) (Holomorphic function in Ω).
By choosing the family of triangulations Tn in such a way that Tn ⊂ Tn+1 ,
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 15

one can expect that fn (z) converges into H(Ω) to a function f (z) ∈ H(Ω)
carrying the conformal transformation of W into B1 (0). (Details of our sug-
gested proof are left to our mathematically oriented readers again.)
At this point we intend to present a Hilbert space approach to construct
explicitly such canonical conformal mapping of the domain W into the disc
B1 (0).
Let H = {F (W, C), f : W → C, f is a holomorphic function in W }.
We introduce the following Hilbert space inner product in this space of
holomorphic (analytical) functions in W :

1
f, gc = dz ∧ dz̄ f (z)g(z). (1.57)
2i W
We have the following straightforward result for two elements of the
Hilbert space (H,  , c ) obtained by a simple application of the Green
theorem in the plane for any domain W̄ ⊂ W :
 
1
g(z)h̄(z) dz = (g(z) h (z))dz ∧ dz̄. (1.58)
∂ W̄ =Γ1 2i W̄
If one considers h(z) = z −z0 and W̄ = Br (z0 ) in Eq. (1.58), one obtains
that
  
g(z) r2
g(z)(z − z0 )dz = r 2
= g(z0 ). (1.59)
|z−z0 |=r |z−z0 |=r z − z0 2πi
In other words,
  
1 
2π  
|g(z0 )| ≤ 2  g(z) dxdy 
r 2 |z−z0 |≤r 

 1
2
 1
2
2π 1
≤ |g 2 (z)| dxdy dxdy
r2 2 |z−z0 |≤r |z−z0 |≤r

π √
≤ π · r ||g||2H(W )
r2

π 3/2
≤ ||g||2H(W ) ≤ c||g||2W (D) . (1.60)
r
By the Riesz theorem for representations of linear functionals in
Hilbert spaces, the bounded (so continuous) linear evaluation functional on
(H(D),  , c ) = HD has the following representation:

1
g(z0 ) = dz ∧ dz̄ R(z, z0 )g(z) (1.61)
2i D
August 6, 2008 15:46 9in x 6in B-640 ch01

16 Lecture Notes in Applied Differential Equations of Mathematical Physics

and


H
R(z, z0 ) = ek (z)ek (z0 ), (1.62)
k=1

where {ek (z)} denotes any orthonormal set of holomorphic functions in


HD .
Let f : W → B1 (0) be the aforementioned Riemann function mapping
holomorphically the given bounded simple connected-smooth boundary
domain in B1 (0). Since f (z) is univalent, we have that for f (z0 ) = 0 and
f  (z0 ) > 0 the following identity holds true:

1 g(z) g(z0 )
=  for any g ∈ HD . (1.63)
2πi ∂W =Γ1 f (z) f (z0 )

Let Cr = {(f (z))(f (z)) = r < 1} be the curve in the complex plane.
For this special contour, we have that (see Eq. (1.58))
 
1 g(z) 1
dz = g(z)f¯(z) dz
2πi Cr f (z) 2πir2 Cr

1  
= 2 g(z)f  (z) dxdy. (1.64)
πr (W \f −1 (Br (0))

Note that we have implicitly assumed that the domain W is a homotopical


deformation of its boundary Γ1 !
As a consequence (by considering the limit of r → 1)
 
1
g(z0 ) = f  (z0 ) g(z) f  (z) dxdy . (1.65)
π W

In view of the uniqueness of the Reproducing Kernel Eq. (1.57), we have


the explicit representation for the conformal transformation (f (z0 ) = 0 and
f  (z0 ) > 0)
1
2
 z 
π
f (z) = R(ζ, z0 )dζ . (1.66)
R(z0 , z0 ) z0

For instance, one could consider the Graham–Schmidt orthogonalization


process applied to the functions {z n , z̄ n }n∈Z+ and obtain Eq. (1.66) as an
infinite sum of holomorphic functions, or use triangulations of the domain D
in order to evaluate the functions en (z) in Eq. (1.62).
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 17

Finally let us devise some useful formulae for the Dirichlet problem in
planar “smooth” regions, useful in string path integral theory.
Let U0 f −1 : B1 (0) → R be the associated harmonic function on the
disc B1 (0) related to the Dirichlet problem in the region W . Note that by
construction f (z0 ) = 0 and f  (z0 ) > 0 for a given z0 ∈ W . As a consequence
of the mean value for harmonic function, we have that
 2π
1
U (z0 ) = (U0 f −1 )(0) = (U0 f −1 )(cos θ, sin θ)dθ
2π 0

1 f  (z)
= U (z) z0 dz. (1.67)
2πi ∂W fz0 (z)

From Eq. (1.67) we can write explicitly the Green functions G(z, z0 ) =
fz 0 (z) fz0 (z) for the given canonical regions. For instance, fz0 (z) = R(z −
z0 ) (R2 − z̄0 z) applies BR (0) conformally into B1 (0), leading to the Poisson
formulae in the circle of radius R
 2π
1 R2 − r 2
U (r, θ) = dφ U (Reiφ ) . (1.68)
2π 0 R2 + r2 − 2Rr cos(φ − θ)

The function fz0 (z) = (z−z0 ) (z+z0 ) applies conformally the right-half-
+
plane = {z = x + iy, x ≥ 0, −∞ < y < ∞} into B1 (0) and fz 0 (z0 ) = 0.
As a consequence, one has the integral representation for this region

x +∞ U (it) dt
U (x, y) = · (1.69)
π −∞ (t − y)2 + x2

The function fz0 (z) = (z 2 −z02 ) (z 2 −(z̄0 )2 ) solves the Dirichlet problem
in the first quadrant plane, leading to the integral representation below:
 ∞ 
4xy t U (t) dt
U (x, y) =
π 0 t4 − 2t2 (x2 − y 2 ) + (x2 − y  )2
 ∞ 
t U (it) dt
+ . (1.70a)
0 t4 + 2t2 (x2 − y 2 ) + (x2 − y 2 )
In general
fz0 (z  ) − fz0 (z)
G(z, z  ) = (1.70b)
1 − fz0 (z) fz0 (z  )

with f (z0 ) = 0 is the explicit Green function associated with the Dirichlet
problem in the planar region W .
August 6, 2008 15:46 9in x 6in B-640 ch01

18 Lecture Notes in Applied Differential Equations of Mathematical Physics

Just for completeness let us write the Green function of the Dirichlet
2
problem in the SR in polar coordinates
1 
G((ρ, θ), (ρ , θ )) = − log[ρ2 + ρ2 − 2ρρ cos(θ − θ )

 
ρ2 ρ2  
+ log R2 + − 2ρρ cos(θ − θ ) . (1.70c)
R2
Here

 

∆ρ,θ G((ρ, θ), (ρ , θ ) = 0, 
1 
 G((R, θ), (R, θ )) = + g|r − r  |  . (1.70d)
 2π  r =Reiθ


r  =Reiθ


1.4. Hilbert Spaces Methods in the Poisson Problem

Let us pass now to the formulation of the Poisson and Dirichlet problems in
the Hilbert spaces L2 (Ω), with Ω̄ a compact set of RN . This formulation is
known in the mathematical literature as the Hilbert space weak formulation
of the problem.
Let H01 (Ω) be the completion
 of the vector space Cc1 (Ω) in relation to the
inner product f, gH 1 = d4 x ∇f ∇g; the weak formulation of Poisson

problem reads as of

U, ϕH 1 = dN x ∇U (x) · (∇ϕ)(x)


= dN x(f · ϕ̄)(x) = f, ϕL2 ∀ ϕ ∈ H01 (Ω), (1.71)

where one must determine a U (x) ∈ H01 (Ω) for a given f ∈ L2 (Ω). That
such (unique) U (x) exists is readily seen from the fact that the functional
linear below is bounded in H01 (Ω):
Lf : H01 (Ω → C)

ϕ→ dN x(f · ϕ̄)(x) = Lf (ϕ) (1.72)

as a consequence of the Friedrichs inequality
 
 N −1
d x|U | (x) ≤ diam(Ω̄)
N 2
dN x|∇U |2 (x) . (1.73a)
 Ω
   Ω
 
||U ||2L2 (Ω) ≤ C(Ω) ||U ||2H 1 (Ω) (1.73b)
0
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 19

By the Riesz representation theorem, there is a unique element U (x) ∈


H01 (Ω),
such that it satisfies Eq. (1.72) and has the explicit expansion

H01 (Ω) 
U (x) = Lf (en )en (x)
n
 
= dN y f¯(y)en (y) ēn (x)
n Ω
   
= d y N ¯
ēn (x)en (y) f (y) = dN y G(x, y)f¯(y). (1.74)
Ω n Ω

Here, en (x) is an arbitrary orthonormal basis in H01 (Ω), which can be


constructed from the application of the Grahm–Schmidt orthogonalization
process to any linear independent set of Cc∞ (Ω), for instance, the set of the
polynomials {x|n| , x|m| } with x ∈ RN .
This method of determining weak solutions is easily generalized to the
Poisson problem with variable coefficients — the Lax–Milgram theorem, a
useful result in covariant euclidean path integrals in string theory.
Let aij (x) and U (x) be functions in C(Ω), with Ω̄ a compact set of RN .
Let us consider the following Hilbert space:

L2{a} (Ω) = topological closure of


 
∂ ∂ ¯
· f ∈ C0 (Ω) dN x aij (x) i f · j
f + dN x V (x)(f f¯)(x) < ∞
Ω ∂x ∂x Ω
with aij (x) = aji (x), aij (x)λi λj ≥ a0 |λ|2

and V (x) > V0 > 0 . (1.75)

We left as an exercise to our readers to mimic with some improvements


the previous study to prove the existence and uniqueness of the Poisson
problem in L2{a} (Ω) for f ∈ L2{a} (Ω) ⊇ H01 (Ω) and ϕ(x) ∈ H01 (Ω)
 
∂ ∂
dN x aij (x) i
U (x) j ϕ(x) + dN x V (x)U (x)ϕ̄(x)
Ω ∂x ∂x Ω

= dN x f (x)ϕ̄(x). (1.76)

It is another nontrivial problem to find conditions on the data function


f (x) (see Theorem 1.1) in order to have the weak solution U (x) to belong
August 6, 2008 15:46 9in x 6in B-640 ch01

20 Lecture Notes in Applied Differential Equations of Mathematical Physics

to the space C 2 (Ω) ∩ C 1 (Ω̄), which will lead to the strong solution of the
Poisson problem in Ω


N 
∂ ∂
− aij (x) U (x) + V (x)U (x) = f (x) x ∈ Ω. (1.77)
i,j=1
∂xj ∂xi

Another important generalization of the Poisson problem is the class


of semilinear Poisson problems defined by Lipschitzian nonlinearities as
written below for Ω ⊂ R3 :

(−∆U )(x) + F (U (x)) + V (x)U (x) = f (x). (1.78)

Here f (x) ∈ L2 (Ω) and F (z) are real Lipschitzian functions (i.e. ∀(z1 , z2 ) ∈
R × R ⇒ |F (z1 ) − F (z2 )| ≤ C|z1 − z2 | for a uniform bound c).
In order to show the existence and uniqueness through a construction
technique, we rewrite Eq. (1.78) into the form of an integral equation (see
Theorem 1.1)

1 F (U (x )) + V (x )U (x ) − f (x ) 3 
U (x) = − d x = (T U )(x).
4π |x − x |
(1.79)

One can see easily that T defines an application with domain L2 (Ω) and
range in L2 (Ω). Besides, T is a contraction application in L2 (Ω) for small
domains, as we can see from the following estimate:

||T u − T v||2L2 (Ω)


 
1 1
≤ ||u − v||2L2 (Ω) (c2 + ||V ||2L2 (Ω) ) × d3 x
16π 2 Ω |x − x |2

C 2 + ||V ||2L2 (Ω)


≤ (4π diam(Ω))||u − v||2L2 (Ω) . (1.80)
16π 2

As a consequence of the Banach Fixed Point Theorem, we have that the


sequence

Un = T (Un+1 ) (1.81)

converges strongly to the solution in L2 (Ω) of Eq. (1.79).


The reader should repeat the above-made analysis in the general case
Ω ⊂ RN (see Eq. (1.23)).
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 21

1.5. The Abstract Formulation of the Poisson Problem

Let A be a linear operator defined in a Hilbert space H = (H, ,) (A : H → H,


A(αU + βv) = αA(u) + βA(v)) such that its domain D(A) forms a (non-
closed) vectorial subspace dense in H. Note that D(A) has the explicit
analytical representation

D(A) = u ∈ H | Au ∈ H ⇔ Au, AuH < ∞ . (1.82)

Let λ ∈ C such that there is a vector Uλ ∈ H, such that AUλ = λUλ .


Such vector is called the eigenvector of A associated with the eigenvalue λ.
The set of all these eigenvectors {Uλ } makes an invariant subspace for the
operator A.
We have the following results in operator theory in separable Hilbert
spaces, useful to the Poisson problem solution.

Theorem 1.4. If the operator A satisfies the symmetric condition (so A


is called a symmetric operator) Au, vH = (u, Av)H (u, v ∈ D(A)), then all
eigenvalues λ are real numbers.

Proof. Let AUλ = λUλ . As a consequence of the symmetry property of


the operator A, we have that

AUλ , Uλ H = λUλ , Uλ  = λ||Uλ ||2 = Uλ , AUλ  = λ̄||Uλ ||2 . (1.83)




Theorem 1.5. If A is a symmetric operator, then different eigenvectors


corresponding to different eigenvalues are orthogonals. So A has at most a
countable number of eigenvalues in the separable H.

Proof. Let AUλ1 = λ1 Uλ1 and AUλ2 = λ2 Uλ2 (λ1 = λ2 ). Thus we have the
identity

0 = AUλ1 , Uλ2 H − Uλ1 , AUλ2 H = (λ1 − λ2 )Uλ1 , Uλ2 H , (1.84)

which means the result searched

Uλ1 , Uλ2 H = 0. (1.85)



We have our basic result in the theory of the Poisson problem in Hilbert
spaces.
August 6, 2008 15:46 9in x 6in B-640 ch01

22 Lecture Notes in Applied Differential Equations of Mathematical Physics

Theorem 1.6. (Dirichlet–Riemann Variational formulation of Poisson


problems.) Let A be a positive-definite symmetric operator in a given sep-
arable Hilbert space H = (H,  , ). (Exists c ∈ R+ such that AU, U H ≥
c(U, U )H for any U ∈ D(A).) Let the functional given below for any
F ∈ Range(A) be

F(A) (ϕ) = Aϕ, ϕH − f, ϕH . (1.86)

Then, the functional equation (1.86) has a minimum value at U ∈ D(A)


such that

A U = f, (1.87)

the converse still holds true.

Let us exemplify the above result for the Poisson problem in bounded
open sets in RN with a Riemannian structure {gab (xa ); a, b = 1, . . . , n}
(manifolds charts).
Thus we consider the following covariant Hilbert space associated with
the given metric structures
 
L2g (Ω) = closure of f ∈ Cc (Ω) |  ,  ≡ dN x g(x)f (x)f (x) ,

(1.88a)

 

1
H0,g (Ω) = closure of f ∈ C01 (Ω) |  ,  = dN x g g ab ∂a f ∂b f ,

(1.88b)

p
H0,g (Ω) = closure of f ∈ C0p (Ω) |  , H 0


= dN x g(∇a1 . . . ∇a(p/2) f )(x)

 
× {g a1 ,a(p/2+1) . . . g a(p/2) ap }(x)(∇a(p/2+1) . . . ∇ap f )(x). (1.88c)
1
In the covariant Hilbert space H0,g (Ω), we consider the functional
(positive-definite) associated to the Laplace–Beltrami operator ∆g =

− √1g ∂a ( g δ ab ∂b ) acting on real functions
 
√   √
F∆q (ϕ) = dN x g ϕ(−∆g )ϕ (x) − dN x g f (x)ϕ(x). (1.89)
W̄ W̄
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 23

The minimizing Ū ∈ H0,g 1


(Ω) of the above-written functional is the
solution of the covariant Poisson problem in the open domain W :

(−∆g Ū )(x) = f (x) for x ∈ W. (1.90)

Let us now analyze the eigenvalue problem in the Dirichlet problem in


the Hilbert space L2 (Ω).
We can accomplish such studies by considering the problem rewritten
into the integral operator form

1 Uλ (r  )
µ Uλ (r ) = d3r  = (T Uλ )(F ). (1.91)
4π Ω |r − r  |

First, let us see that the integral operator T has as its domain C 1 (Ω) ⊂
L (Ω) and as its range the functional space C 2 (Ω). However, it can be
2

extended to the whole space L2 (Ω) by the Hahn–Banach theorem

T : L2 (Ω) → L2 (Ω) (1.92)

since we have the estimate


 1
2
 1
2
1 1
|(T U )(r )| ≤ d3r  × d3r  U 2 (r  )
4π Ω |r − r  |2 Ω
1
2
1
≤ diam(Ω) ||U ||L2 (Ω) . (1.93)
3

We can see that T is a symmetric bounded compact operator with a set


of eigenvalues |µr | ≤ 4π
3 diam(Ω) and µn → 0 for n → ∞.

1.6. Potential Theory for the Wave Equation in R3 —


Kirchhoff Potentials (Spherical Means)

Let us start this section by defining the Poisson–Kirchhoff potential asso-


ciated with the given C 2 (Ω) functions. The first potential associated with a
certain f (r ) ∈ C 2 (Ω) is defined by the spherical means (r = (x, y, z) ∈ Ω)

t
U (1)
(r, t, [f ]) = f (x + αct, y  + βct, z  + γct)dΩ. (1.94)
4π 3 (
Sct r)
August 6, 2008 15:46 9in x 6in B-640 ch01

24 Lecture Notes in Applied Differential Equations of Mathematical Physics

3 3
Here, Sct (r ) = ∂{Bct (r )} is the spherical surface centered at the point
r = (x, y, z) with radius |r  − r| = ct. The spherical parameter α, β, and γ
are defined as usual as

α = cos ϕ sin θ,
β = sin ϕ sin θ,
(1.95)
γ = cos θ,
dΩ = sin θdθ · dϕ.

The second potential is defined in an analogous way as



1 ∂ 1 (1)
U (2) (r, t, [g]) = U (r, t, [g]) . (1.96)
4π ∂t 2

We have the following elementary initial-value properties for the above


wave potentials, which are the functions defined in the infinite domain
R3 × R+ :

U (1) (r, 0, [f ]) = 0,
 
∂ (1) t
U (r, 0, [f ]) = f (x, y, z) × lim dΩ = f (r ),
∂t t→0+ 4π 3 (
Sct r) (1.97)
U (2) (r, 0, [g]) = g(r ),

U2 (r, 0, [g]) = 0 (exercise).
∂t
The main differentiability property of the above-written wave potentials
is the following:
 2π  π 
(1) t
∆r U (r, t, [f ]) = dϕ sin θdθ∆r f (x + αct, y + βct, z + γct) .
4π 0 0
(1.98)

We should now evaluate the second-time derivative of the potential


(x̄ = x + αct, ȳ = y + βct, z̄ = z + γct)

∂U (1) (r, t, [f ])
∂t
 
U (1) (r, t, [f ]) t ∂ ∂ ∂
= + cα + cβ + cγ f (x̄, ȳ, z̄)d2 Ω
t 4π 3 (
Sct r) ∂ x̄ ∂ ȳ ∂ z̄
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 25

n·∇
 
    
U (1) (r, t, [f ]) 1 ∂ ∂ ∂
= + α +β +γ f (x̄, ȳ, z̄)dA
t 4πct 3 (
Sct r) ∂ x̄ ∂ ȳ ∂ z̄
(Gauss theorem)
 
U (1) (r, t, [f ]) 1
= + (∆ f )(x̄, ȳ, z̄)dx̄ dȳ dz̄
t 4πct 3 (
Bct r)

 ct  
U (1) (r, t, [f ]) 1 2
= + dρ d A(∆f )(x̄, ȳ, z̄) (1.99)
t 4πct 0 3 (
Sct r)

since

dx̄ dȳ dz̄ = dρ d2 A,
(1.100)
0 ≤ ρ ≤ ct.

Since we have the usual Leibnitz rule for derivatives inside integrals
 ct 
d
f (x) dx = f (ct) · c. (1.101)
dt 0

We have the result for evaluation of the second-time derivative of Eq. (1.94):

∂ 2 (1) ∂U (1) (r, t, [f ]) 1 1


2
U (
r , t, [f ]) = · − 2 U (1) (r, t, [f ])
∂t ∂t t t
 
1
+ ×c× 2
d A(∆f )(x̄, ȳ, z̄)
4πct Sct3 (
r)
 ct  
1
− dρ 2
d A(∆f )(x̄, ȳ, z̄) . (1.102)
4πct2 0 Sρ3 (
r)

Using Eq. (1.99) again to simplify Eq. (1.102), yields


 
∂ 2 U (1) (r, t, [f ]) 1 2
= d A(∆f )(x̄, ȳ, z̄)
∂2t 4πt Sρ3 (r )
 
2 1
=c t d Ω(∆f )(x̄, ȳ, z̄) = 2 ∆r U (1) (r, t, [f ]).
2
3 (
Sct r) c
(1.103)

As a result of the above differential calculus evaluations we have the


analogous of Theorem 1.1 (the Cauchy problem for the wave equation).
August 6, 2008 15:46 9in x 6in B-640 ch01

26 Lecture Notes in Applied Differential Equations of Mathematical Physics

Theorem 1.7. The solution for the Cauchy (initial values) problem for
the wave equation with “smooth” C 2 (Ω) data (including an external forcing
F (r, t), continuous in the t-variable), is given by the wave potentials
 2


∂ U (r, t) 1

 = 2 ∆r U (r, t) + F (r, t),
 ∂ t 2 c
U (
r , 0) = f (
r ), r ∈ Ω, (1.104)





Ut (r, 0) = g(r ), r ∈ Ω,

U (r, t) = U (1) (r, t, [f ]) + U (2) (r, t, [g])


 
1 F (r  , ct − |r − r  |) 3 
+ d 
r . (1.105)
4π |
r−r  |≤ct |r − r  |

In the case of Ω ⊂ R2 we have the Poisson wave potentials:


  
∂ 1 f (r  ) 2 
U (r, t) = d r
∂t 2π |r−r  |<ct c2 t2 − |r − r  |2
  
1 g(r  ) 2 
+ d r
2π |r−r  |<ct c2 t2 − |r − r  |2
  t  
1  F (r  , c(t − t )) 2 
+ dt d r . (1.106)
2π 0 | r  |<ct
r− (t )2 − |r − r  |2
Let us give a proof for the problem uniqueness for the Cauchy initial
value Eq. (1.104).
In order to show such a result, let us consider the energy wave (func-
tional) inside a spherical ball of radius R:

 2 
ER (t) = Ut + c2 |∇U |2 (r, t)d3r. (1.107)
BR (0)

Let us analyze its time derivative



dER (t)  
lim = 2Ut Utt + 2c2 ∇∂t U · ∇U (r, t)
R→∞ dt BR (0)
  
 
=2 Ut (Utt − c ∆U ) + O
2
Ut · ∇N U dS
3 (0)
BR 3 (0)
SR
 
 
= 2 lim O Ut · ∇N U dS = 0, (1.108)
R→∞ 3 (0)
SR
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 27

if one imposes a sort of Helmholtz–Sommerfeld radiation condition at


infinity (the vanishing of the surfaces term in Eq. (1.108). As a consequence,
if f (r) = g(r) ≡ 0 (vanishing of the initial conditions), we have that

E(t) = E(0)
  
   
= lim −c O 2
U · ∆U (r, 0) + lim O Ut · ∇N U dS = 0.
R→∞ BR (0) 3 (0)
SR
(1.109)

1.7. The Dirichlet Problem for the Diffusion Equation —


Seminar Exercises

Let us consider the following Dirichlet problem for the diffusion equation in
a given domain Ω ⊂ R3 with a nontrivial time-dependence on the boundary
(an interior problem):

 1 ∂

 ∆ U (r, t) = 2 U(r,t) ,

 a ∂t
 U (r, 0) = f (r ) ∈ C ∞ (Ω), (1.110)



  
U (r, t)∂Ω = g(r, t)∂Ω ∈ C ∞ (∂Ω).

Let us introduce the double-layer like Poisson potential for a not-yet


determined density Φ(r, t)
  
1 t
Φ(r  , ζ)  − 2r  
U (1)
(r, t) = dζ O ∇ e 4a (t−ζ) dS(r )
4π 0 ∂Ω (t − ζ) N
  
1 t
Φ(r  , ζ) − 4a2 (t−ζ)
r
 (r )∠dS(r  ) .
= dζ O e · r · cos(N
8πa2 0 ∂Ω (t − ζ)2
(1.111)

We observe that U (1) (r, t) satisfies Eq. (1.110) and vanishes at t = 0


(exercise).
The solution for the Dirichlet-like problem Eq. (1.110) will be deter-
mined from the ansatz
U (2) (
r,t)
   
 2
(1) 1 3   − (
r −
r )
U (r, t) = U (r, t) + d r f (r )e 4a2 t . (1.112)
3
(2πa2 t) 2 Ω
August 6, 2008 15:46 9in x 6in B-640 ch01

28 Lecture Notes in Applied Differential Equations of Mathematical Physics

It is clear that the formal solution written above satisfies the boundary
condition in Eq. (1.110) if one can determine the density function Φ(r, t)
from the integral equation coming from the imposition of the boundary
condition as given by Eq. (1.110) (exercise):
  
U (r, t)∂Ω = g(r, t)∂Ω = U (2) (r, t)∂Ω
  
 1 t
Π(r  , t) − 4a2r(t−ζ)
2

+ −Φ(r, t)∂Ω + dζ O e · r
8πa2 0 ∂Ω (t − ζ)2

× cos(N ∠r  )dS(r  ) . (1.113)

 
One can show that for small densities Φ(r, t) → λΦ(r, t) with λ  1 ,
one can solve Eq. (1.113) by means of the fixed point theorem of Banach
(exercise) and yielding the solution as a power series in λ (the size of the
area of ∂Ω).

1.8. The Potential Theory in Distributional Spaces — The


Gelfand–Chilov Method

Seminar Exercises

In this brief section, we intend to show the general method of Kirchhoff–


Poisson potentials in the context of Schwartz distribution theory in S  (RN ).
Let us start this section by considering the problem of determining the
fundamental solution of an arbitrary elliptic differential operator of order
m in the whole space RN with constant coefficients


aα Dxα U (x) = δ (N ) (x). (1.114)
|α|≤m

Here U (x) shall belong to S  (RN ). In the distributional sense2


  t 
U (x), ϕ(x) = aα Dxα ϕ(x) . (1.115)
|α|≤m x=0

Let us consider the analytical regularization of the Dirac distribution in


the right-hand side of Eq. (1.114) (the well-known analytical regularization
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 29

scheme in quantum field theory) leading to a usual function-theoretic partial


differential equation expressed below:
 2|x|λ
aα Dxα U (λ) (x) =  
|α|≤m
ωn Γ λ+n
2
 
1  λ 
= n−1  λ+n  |x · r | dS(r ) .
(n)
ωn π 2 Γ 2 S1 (0)
(1.116)

At this point one can consider the following analytical regularized asso-
ciated problem:
 |r · r  |λ
(L U )(r ) = aα Drα v (λ) (r ) = n−1  · (1.117)
|α|≤m
ωn π 2 Γ λ+n 2

Due to the linearity of Eq. (1.116), we should search for a solution of this
analytical regularized equation in the linear-superposing form of a spherical
mean (Radon transforms)

(λ)
U (r ) = dΩ(r  ) v (λ) (r, r  ). (1.118)
(n)
S1 (0)

The solution of Eq. (1.117) reduces to the solution of an ordinary non-


homogeneous differential equation in R after introducing the variable
change ∂x∂ a = ωa dξ d
, where r = {xa }a=1,...,n and r  ∈ S1n (0) ⇔ r  =
{ω }a=1,...,n [(ω ) + · · · + (ω n )2 = 1)] with −∞ < ξ < +∞.
a 1 2

It yields the following simple result:

d |ξ|λ
L ωa v λ (ξ) = n−1  · (1.119)
dξ ωn π 2 Γ λ+n 2

The solution of Eq. (1.119) is easily written down in the distributional


sense in S  (R):
 +∞ 
1   λ 
(λ)
v (ξ) = n−1   G(ξ, ξ )|ξ | dξ . (1.120)
ωn π 2 Γ λ+n
2 −∞

Here, G(ξ, ξ  ) is the Green function of the ordinary differential operator in


Eq. (1.119), obtained through Fourier transforms in S  (R), namely,

d
L ωa G(ξ, ξ  ) = δ (1) (ξ − ξ  ). (1.121)

August 6, 2008 15:46 9in x 6in B-640 ch01

30 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now it can be shown case by case that the limit of λ → −n in our


already-obtained distribution-regularized solutions converges in the distri-
butional sense to the expected S  (RN )-distributional solution of Eq. (1.114).
As an exercise, the reader can find the distributional solution of the
α-power Laplacean through the analytical regularized nonhomogeneous
ordinary differential equation. (Here α ∈ C, and 0 < Real(α) < n).

 2α
d d2 |ξ|λ
L ωa = − ((ω1 )2 + (ω2 )2 + (ω3 )2 ) 2 v λ (ξ) =  ·
dω d ξ ω3 π Γ λ+n2

(1.122)

The full solution is given by the famous Riesz–Poisson potential in RN


(compare with Eq. (1.41):
 α

(−∆) 2 U (x) = f (x),
 
Γ (n−α) f (x ) (1.123)

U (x) =  α  α 2 (n/2) dN x .
Γ 2 2 ·π RN |x − x |n−α

References

1. C. Hilbert, Methods of Mathematics Physics, Vol. I, II (Interscience


Publishers, Inc., New York, 1978).
2. B. Simon, J. Math. Phys. 12, 140 (1971).
3. M. Lavrentiev and B. Chabat, Méthodes de la théorie des fonctions d’une
variable complexe, Édition MIR (Moscou, 1976).
4. L. C. L. Botelho, Methods of Bosonic and Fermionic Path Integrals Repre-
sentations: Continuum Random Geometry in Quantum Field Theory (Nova
Science Publisher, NY, USA, 2008).
5. C. Luiz and L. Botelho, Il Nuovo Cimento 112A, 1615 (1999).
6. S. Weinberg, Gravitation and Cosmology (John Wiley & Sons Inc., 1992).
7. P. G. Bergmann, Introduction to the Theory of Relativity (Prentice-Hall, Inc.,
England, Cliffs, NJ, USA).
8. V. Fock, The Theory on Space, Time and Gravitation, 2nd revised edn.
(Pergamon Press, 1964).
9. S. Sternberg, Lectures on Differential Geometry (Chelsea Publishing
Company, NY, 1983), Theorem 4.4, p. 63.
10. S. W. Hawking, General Relativity: An Einstein Centenary, Survey
(Cambridge University Press, 1979).
11. H. Flanders, Differential Forms (Academic Press, 1963).
12. J. Milnor, Topology from the Differentiable Viewpoint (University Press of
Virginia, 1965).
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 31

Appendix A. Light Deflection on de-Sitter Space

A.1. The Light Deflection

The most general covariant second-order equation for the gravitation field
generated by a given (covariant) enegy–matter distribution on the space–
time is given by the famous Einstein field equation with a cosmological
constant Λ with dimension (length)−2 (Ref. 5), namely,
1
Rµν (g) − gµν R + Λgµν (x) = 8πGTµν (x), (A.1)
2
where x belongs to a space–time local chart.
It is well known that studies on the light deflection by a gravitational
field generated by a massive point-particle with a pure time like geodesic
trajectory (a rest particle “sun” for a three-dimensional spatial space–time
section observer!) is always carried out by considering Λ ≡ 0.6,7
Our purpose in this appendix is to understand the light deflection
phenomena in the presence of a nonvanishing cosmological term in Einstein
equation (A.1), at least on a formal mathematical level of solving trajectory
motion equations.
Let us, thus, look for a static spherically symmetric solution of Eq. (A.1)
in the standard isotropic form6,7

(ds)2 = B(r)(dt)2 − A(r)(dr)2 − r2 [(dθ)2 + sin2 θ(dφ)2 ]. (A.2)

In the space–time region r = + | x |2 > 0, where the matter–energy


tensor vanishes identically, we have that the Einstein equation takes the
following form:
Rµν (g)(x) = −Λ(gµν (x)). (A.3)

In the above cited region, the Ricci tensor is given by


 
−ΛB(r) 0 0 0
 0 ΛA(r) 0 0 
−Λgµν =   2


0 0 Λr 0
2 2
0 0 0 Λr sin θ
 
Rtt 0 0 0
 0 Rrr 0 0 
=
 0
. (A.4)
0 Rθθ 0 
0 0 0 Rφφ
August 6, 2008 15:46 9in x 6in B-640 ch01

32 Lecture Notes in Applied Differential Equations of Mathematical Physics

We have, thus, the following set of ordinary differential equations in place


of Einstein partial differential Eq. (A.1)

B  1 B A B 1 B
Rtt = − + + − = −ΛB, (A.5)
2A 4 A A B r A

B  1 B A B 1 A
Rrr ≡ − + − = ΛA, (A.6)
2B 4 B A B r A

r A B 1
Rθθ = −1 + − + + = Λr2 , (A.7)
2A A B A

Rφφ = sin2 θRθθ = Λ(sin2 θ)r2 . (A.8)

At this point we note that


Rtt Rrr
+ = 0, (A.9)
B(r) A(r)

or equivalently
α
A(r) = , (A.10)
B(r)

where α is an integration constant.


Since Rθθ = −1 + αr B  + B 2
α = Λr , we get the following expression for
the B(r) function:

αΛr2 β
B(r) = +α+ , (A.11)
3 r
with β denoting another integration constant.
In the literature situation,6,7 one always considers the case Λ = 0 in
a pure classical mathematical vacuum situation context, the so-called de-
Sitter vacuum pure gravity. However, in our case it becomes physical to
consider that our solution depends analytically on the cosmological con-
stant. In other words, if the parameter Λ → 0 in our solution, it must
converge to the usual Schwarzschild solution with a mass singularity at the
origin r = 0, that is our boundary condition hypothesis imposed on our
solution.
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 33

As a consequence, one gets our proposed Schwarzschild–de-Sitter


solution
Λr2 2M G
(ds)2 = +1− (dt2 )
3 r

−1
Λr2 2M G
− +1− (dr)2 − r2 [(dθ)2 + (sin2 θ)(dφ)2 ].
3 r
(A.12)

At this point let us comment that for the space–time region exterior to
the spatial sphere r > ( 3mG
Λ )
(1/3)
, the field gravitation approximation leads
to the antigravity (a repulsion gravity force) if Λ < 0; So, explain from
this Einstein Gravitation theory of ours the famous “Huble accelerating
Universe expansion”.
In what follows we are going to consider a nonvanishing Λ < 0 and study
the path of a light ray on such negative cosmological constant Einstein
manifold Eq. (A.12).
We have the following null-geodesic equation for light propagating in
θ = π/2 plane (Einstein hypothesis) for light propagation in the presence
of the sun (Sec. A.2 for the related formulae):
2 2
1 dr dφ
0 = B(r) − − r2 . (A.13)
B(r) dt dt

At this point we note that

dφ B(r)J
= , (A.14)
dt r2
where J is an integration constant.
After substituting Eq. (A.14) into Eq. (A.13) we have the following
differential equation for the light trajectory as a function of the deflection
angle φ
2
dr dφ J 2 B 3 (r)
+ − B 2 (r) = 0, (A.15)
dφ dt r2

which is exactly integrable:


dr
dφ = ' . (A.16)
B(r)
r2 1
J2 − r2
August 6, 2008 15:46 9in x 6in B-640 ch01

34 Lecture Notes in Applied Differential Equations of Mathematical Physics

dr
By supposing a deflection point rm where dt = 0 and, thus, J =
rm / B(rm ), we get the deflection angle
 rm  1
dr rm dU
∆1 φ = B(rm ) B(r) (1/2)
= 2 − U 2 ) − 2M G(U 3 − U 3 )](1/2)
,
∞ r2 [ r 2 − r2 ] 0 [(Um m
m
(A.17)
which is exactly the one given in the pure (Λ = 0) Schwarzschild famous
case. However, if one supposes that there is no deflection (a continuous
monotone trajectory r = r(φ)!), the total deflection angle now depends on
the cosmological constant and is given formally by the expression below:
  

 

 rm  
  


1  1 
∆2 φ = dr ' +  = ∆1 φ. (A.18)
   3−ΛJ 2  
∞  r2 − 12 + 2mG
  
r 3 ( 3J 2 ) 

 r r3 1 + 2MG−r 

As a general conclusion of our note we claim that the usual light-


deflection experimented test does not make any difference between the usual
noncosmological Schwarzschild case and our case Eq. (A.12), and, thus,
it should not be considered as a definitive physical support for Einstein
General Relativity without cosmological constant.

A.2. The Trajectory Motion Equations

The body trajectory (t(p), r(p), θ(p), ϕ(p)) in the presence of the gravita-
tional field generated by the metric Eqs. (A.2)–(A.10) is described by the
following geodesic equations:

d2 t B dr dt
+ = 0, (A.19)
d2 p B dp dp

A B
2 2 2 2
d2 r dr r dθ r sin2 θ dφ dt
+ − − + = 0,
d2 p 2A dp A dp A dp 2A dp
(A.20)
2 2
d θ 2 dθ dr dφ
+ − sin θ · cos θ = 0, (A.21)
d2 p r dr dp dp

d2 φ 2 dφ dr dφ dθ
+ + 2cotg(θ) = 0. (A.22)
d2 p r dp dp dp dp
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 35

At this point we remark that by multiplying Eq. (A.19) by B(r(p)),


it reduces to the exact integral form relating the Einstein proper-time
(physical evolution parameter) p with the geometrical-dependent coor-
dinate Newtonian time t:
dt 1
= . (A.23)
dp B(r)

We remark either that Eq. (A.22) can be rewritten in the form

d dφ
n + nr2 + 2 n sin θ = 0, (A.24)
dp dp

which reduces to the following form:

dφ 2
r (p) sin2 (θ(p)) = J, (A.25)
dp

where J is an integration constant.


By substituting Eqs. (A.23) and (A.25) into Eqs. (A.20) and (A.21) we
obtain the full set of equations describing the body trajectory in relation
to the (r, θ) variables

B B
2 2
d2 r dr dθ J 2B
− − rB − + = 0, (A.26)
d2 p 2B dp dp r3 2B

d2 θ 2 dθ dr cos θ J 2
+ − = 0. (A.27)
d2 p r dr dp sin3 θ r4

For Einstein hypothesis of light propagation on the plane θ = π/2,


Eq. (A.27) vanishes and Eq. (A.26) takes the form

B B
2
d2 r dr J 2B
2
− − 3
+ =0 (A.28)
d p 2B dp r 2B

or in a more manageable alternative form after multiplying Eq. (A.28) by


2 dr
B ( dp ) and by using Eq. (A.23) for exchanging the geometrical parameter p
by the time manifold coordinate
2
dr 1 J2 1
+ − + E = 0, (A.29)
dt B3 r2 B

where E denotes another integration constant.


August 6, 2008 15:46 9in x 6in B-640 ch01

36 Lecture Notes in Applied Differential Equations of Mathematical Physics

2 3
dφ r 2
By writing r as a function of φ and using Eq. (A.25) dt B = J! , we
get our final trajectory equation
 (1/2)
dr 1 B BE
= ±r 2
− 2 − 2 , (A.30)
dφ J2 r J

which leads to the body trajectory geometric form



dr
φ=± . (A.31)
2
r B (1/2) [ J 2 B − JE2 − r12 ](1/2)
1

Note that for light propagation the integration constant E always


vanishes, a result used in the text by means of Eq. (A.16).

A.3. On the Topology of the Euclidean Space–Time

One of the most interesting aspects of Einstein gravitation theory is the


question of the nonexistence of “holes” in the space–time C 2 -manifold from
the view point of a mathematical observer situated on the Euclidean space
R9 associated with the “minimal” Whitney imbedding theorem of M on
Euclidean spaces.9
In order to conjecture the validity of such a topological space–time
property, let us suppose that M is a C 2 -manifold, and the analytically con-
tinued (Euclidean) matter distribution tensor generating the (Euclidean)
gravitation field on M allows a well-defined Euclidean metric tensor
(solution of Euclidean Einstein equation).10
At this point we note that M must be always orientable in order to
have a well-defined theory of integration on M and, thus, the validity of
the rule of integration by parts: Stokes’ theorem is always needed in order
to construct matter tensor energy momentum. Since Euclidean Einstein’s
equations say simply that the sum of sectional curvatures is a measure
of the (classical) matter energy density generating gravity, which must be
always considered positive, it will be natural to expect the positivity of the
Euclidean. Energy–Momentum of the matter content leads to the result
that the associated sectional curvatures are positive individually. Since M
is even-dimensional (four), the famous Synge’s theorem8 leads to the result
that M is simply connected (note that this topological property is obvi-
ously independent of the metric structure being Lorentzian or Euclidean!)
and as a direct consequence of this result, any physical geodesic (particles
August 6, 2008 15:46 9in x 6in B-640 ch01

Elementary Aspects of Potential Theory in Mathematical Physics 37

trajectory) on M can be topologically deformed to a point, and, thus, M


does not possess “holes” from the point of view of the Whitney imbedding
extrinsic minimal space R9 .
Finally, let us argue that the existence of a (symmetric) energy–
momentum tensor on M is associated with the “General Relativity”
description of the space–time manifold M by means of charts (the Physics
is invariant under the action of the diffeomorphism group of M ), which
by its turn leads to the existence of the matter energy–momentum tensor
by means of Noether theorem (a metric-independent result) applied to the
matter distribution Lagrangean (a scalar function defined on the tangent
bundle of M ).
As a consequence, let us conjecture again that the introduction of a
cosmological term on Einstein equation spoils the physical results presented
on the whole topology and the physical requirement of positivity of the
matter–energy universe moments tensor, given, thus, a plausible topological
argument for the vanishing of the cosmological constant at the level of the
global–topological aspects of the space–time manifold.
Finally, let us show the mathematical formulae associated with our ideas
and conjectures written above.
Let e0 , . . . , e3 be an orthonormal frame at a point of M (Euclidean). It
is well known that the Ricci quadratic form can be expressed in terms of
sectional curvatures

Ric (ei , ei ) = K(ei ∧ ej ), (A.32)
j=i

and the Einstein tensor is defined by


1
Gij = Rij − gij R. (A.33)
2
Since the Einstein equation reads in terms of quadratic forms associated
with the sectional curvatures as

G(ep , ep ) = + K(e⊥p ) = T (ep , ep ), (A.34)

with Tij being the matter energy tensor and e⊥ p the basis two-plane
orthogonal to ep , one can in principle write the sectional curvatures
K(ep ∧eq ) in terms of the quadratic energy–momentum sectional curvatures
Tij (ep , eq ) at least for “short-time” cylindrical geometrodynamical space–
time configurations as expected in a Quantum theory of gravitation (see
August 6, 2008 15:46 9in x 6in B-640 ch01

38 Lecture Notes in Applied Differential Equations of Mathematical Physics

Ref. 5). For the two-dimensional case this assertive is straightforward as


one can see from the relations below:

G(e0 , e0 ) = K(e1 ∧ e1 ), (A.35)


G(e1 , e1 ) = K(e0 ∧ e0 ). (A.36)

As a consequence, one should conjecture that the positivity of the


energy–momentum tensor T (ep , eq ) leads to the individual positivity of the
sectional curvature set K(er , es ) on the basis of Eq. (A.34), namely,
 

G(ep , eq ) = T (ep , eq ) = Ric (ep , eq ) − δpq  K(ei ∧ ej ) . (A.37)
i,j,i=j
August 6, 2008 15:46 9in x 6in B-640 ch02

Chapter 2

Scattering Theory in Non-Relativistic One-Body


Short-Range Quantum Mechanics: Möller Wave
Operators and Asymptotic Completeness

2.1. The Wave Operators in One-Body Quantum Mechanics

We know that the quantum dynamics in RN is postulated to be given (see


the previous chapter) by unitary group generated by a self-adjoint operator
of the form

H = H0 + V, (2.1)

where H0 = − 21 ∆ is the Laplacean operator defined in the Sobolev space


H 2 (RN ) and U (x) is a given multiplication operator in H 2 (RN ) (for
instance: V ∈ L2loc (RN ) and with a lower bound in RN ).
The problem in scattering theory is to prove that for a class of special
quantum initial states φ ∈ Hc (H) (the subspace of absolute continuity
of H), there are states φ± ∈ Dom(H0 ) such that the interacting dynamics
is asymptotically equivalent to the free dynamics. Namely,

lim e−iHt φ − e−itH0 φ± L2 (RN ) = 0 (2.2)


t→±∞

or equivalently

lim φ − eiHt e−iH0 t φ± L2 (RN ) = 0. (2.3)


t→±∞

Let us thus make a general formulation of the (formal) suggestion as


given in Eqs. (2.2) and (2.3).

Definition 2.1. Let H and H0 be self-adjoint operators in a given Hilbert


space (H, , ) and PAC (H0 ) the projection onto the subspace of absolute
continuity of the “free” hamiltonian H0 . Let us consider the operator
family,1–4 where PAC (H0 ) denotes the operator projection on the absolute

39
August 6, 2008 15:46 9in x 6in B-640 ch02

40 Lecture Notes in Applied Differential Equations of Mathematical Physics

continuous spectral subspace of H0

W (t) = exp(iHt) exp(−iH0 t)PAC (H0 ). (2.4)

We define the wave operators W ± (H, H0 ) associated with the operator


family equation (2.4) by the strong limit (if it exists)

W ± (H, H0 ) = S − lim W (t). (2.5)


t→±∞

We now show the existence of the wave operators equation (2.5) for a
certain class of self-adjoint operators (H, H0 ).

First we need the following lemma1 :

Lemma 2.1. Let us suppose that W (t) is strongly continuous differentiable


and there are numbers δ± > 0, such that
    −δ −  
∞ d  d 
 W (t)f  < ∞,  W (t)f  < ∞ (2.6)
 dt   dt 
δ+ H −∞ H

for f ∈ H. Then W (t)f converges for t → ±∞.

Proof. It is a single consequence of the estimate


  
t1 d  (t1 ,t2 )→+∞
Wt1 f − Wt2 f H ≤ ds 
 ds W (t)f  −−−−−−−→ 0.
 (2.7)
t2 H

We now have the basic result on the existence of wave operators (Cook–
Kuroda–Jauch).1,3,4

Theorem 2.1. Let M be a dense subset of Dom(H) ∩ Dom(H0 ) with the


following properties:

(a) (H − H0 ) exp(−itH0 )f H is a continuous function in R.


(b) (H − H0 ) exp(−itH0 )f H is integrable in [δ, ∞) (or in (−∞, −δ]).

Then the wave operators are well defined.

Proof. By the Stone theorem

d
(W (t)f ) = (eitH (H − H0 )e−itH0 )f. (2.8)
dt
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 41

As a straightforward consequence of hypothesis (b), we have


 
d 
 W (t)f  = (H − H0 )e−itH0 f H ∈ L1 ([δ1 , ∞)). (2.9)
 dt 
H

Since M is dense in PAC (H)H and W (t) are uniformly bounded


(W (t)operator = 1) we have the validity of the lemma of Cock–Kuroda–
Jauch. 

Let us apply the above result to concrete examples.


First, let the quantum mechanical potential V (x) ∈ L2loc (RN ) be such
that there exists ε > 0 with
N
V (x)(1 + |x|)−( 2 −1)+ε L2 (RN ) = Nv < ∞. (2.10)

(The so-called short-range quantum mechanical scattering potential.)


Additionally, let us suppose that H = − 12 ∆ + V (x) is a self-adjoint
operator in L2 (RN ). Then there are wave operators associated with the
pair (H, H0 ) in L2 (RN ).
In order to show this result, let us remark that the set of functions


M = ψa (x) ∈ L2 (RN ) | ψa (x) = φ(x − a), for a ∈ RN

 

N
2

and φ(x) = 2 −N
2  xj e−|x| /4 
(2.11)

j=1

is dense in L2 (RN ) (Wiener theorem2 ) and that (exercise!)


 » –
N
22 
N
− (x−a)2

(exp(−it(−∆))ψa )(x) = 3n
 (xj − aj )e
4(1+it)
. (2.12)
(1 + it) 2
j=1

We have thus the estimate for any δ

|(exp(−it(−∆))ψa )(x)|
N
  3N −1−δ  
−N |x − a|−( 2 −1)+δ  x − a  2 |(x − a)|2
≤2 2
 1 + it  exp − .
|1 + it|1+δ 4(1 + t2 |
(2.13)
August 6, 2008 15:46 9in x 6in B-640 ch02

42 Lecture Notes in Applied Differential Equations of Mathematical Physics

1
As a consequence and by choosing 0 < δ < 2, we have the combined
estimate involving the scattering potential
N
N |V (x)|x − a|−( 2 −1)+δ |
|(V exp(−it(−∆))ψa )(x)| ≤ Ca 2− 2 , (2.14)
|1 + it|1+δ
3N
−1−δ −Ω2
where t > 0 and Ca = supΩ∈R (2Ω) 2 e L∞ (RN ) < ∞, here
1
Ω = |(x − a)|/(1 + t2 ) 2 .
If we choose δ < min( 12 , ), we have
NV
(V e−it(−∆) ψa )(x)L2 (RN ) ≤ , (2.15)
|1 + it|1+δ
which obviously leads to the searched integrability result
 ∞  ∞ 
dt
V e−it(−∆) ψa L2 (RN ) ≤ NV 1+δ
<∞ (2.16)
δ+ δ + |(1 + it)|

and to the existence of the wave operators.


Another important class of potentials well-behaved in quantum
mechanics is those which belongs to the Enss’ class formed by those func-
tions in L∞ (RN ) ∩ L2 (RN ) (in such a way that V ((−∆)+i1)−1 is a compact
operator) and satisfying the x-localizibility Enss property5
V (−∆ + z1)−M (1 − χBR (0) (x)operator = hz (R) ∈ L1 ([0, ∞)). (2.17)
Here, χBR (0) (x) ≡ F (|x| < R) is the characteristic function of the Ball of
radius R in RN and M a positive real number, M ≥ 1.
Let us consider f ∈ S(RN ), such that for a given a > 0
 
c 1
| exp(−it(−∆))(−∆ + 1) f |(x) ≤
N
F |x| < a|t| .
(1 + |x| + |t|)N +1 2
(2.18)
This set is dense in L2 (RN ) since supp f˜(x) ⊂ (Ba (0))c .
We have the identity
V exp(−it(−∆))f L2 (RN )
   
 1 

≤ V (H0 + i)−M
F |x| ≤ a|t| e −iH0 t M 
(H0 + i) f 
2 L2 (RN )
   
 1 

+ V (H0 + i)−M
F |x| ≥ a|t| e −iH0 t M 
(H0 + i) f  .
2 L2 (RN )
(2.19)
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 43

Since every compact operator is a bounded operator and the estimate


equation (2.18) added to the Enss localizibility property equation (2.17),
one can easily obtain the Cook–Kuroda–Jauch lemma validity for the
existence of wave operator for Enss potentials.
At this point of our exposition we call the readers’ attention to the fol-
lowing theorem which is easy to proof and is left to the reader’s completion.3
Theorem 2.2. Let us suppose that the wave operators W ± (H, H0 ) exist in
H, then we have
(a) W ± are partial isometries between HAC (H0 ) (the subspace of absolute
continuity of H0 ) and the final range spaces H + = Range(W + (H, H0 ))
(H − = Range(W − (H, H0 ))).
(b) HW ± = W ± H0 .
(c) H± ⊂ HAC (H).
(d) The S-matrix operator S = (W + )∗ (W − ) : H → H.
An important physical requirement on quantum mechanics is the
requirement that W ± (H, H 0 ) should be unitary applications between the
subspaces H(H0 ) and H(H), with the condition of H + = H − = HAC (H0 ),
and in such a way that the operator S̄ = (W + )∗ W − must be a unitary
operator between H− and H+ , and thus, insuring that the physical
S-matrix operator S̄ is a unitary operator, the backbone of the Copenhagen
School interpetration of the quantum phenomena.
We intend to show the confirmation of this result in two situations.
First, we show that it is possible to always have asymptotic completeness
by reducing the physical Hilbert space of the scattering states. Second, we
present mathematical details in next sections, the beautiful original Enss’s
geometrical work on the subject.5
Our first claim is a simple result of ours obtained from a result of Halmos
(exercise).
Halmos Lemma. Let T be an isometry in H. Let us consider the closed
subspace N = (Range T )⊥ . Then we have the following subspace decom-
position of the original Hilbert space in terms of invariant subspace of the
isometry T
∞ 

H=M⊕ k
T (N ) , (2.20)
k=0
k
with T denoting the kth power of T and most importantly, the composite
operator TM = T PM (when P : H → M is the orthogonal projection onto
the closed subspace M ) is a unitary operator TM : M → M .
August 6, 2008 15:46 9in x 6in B-640 ch02

44 Lecture Notes in Applied Differential Equations of Mathematical Physics

In our case M ⊂ H+ ∩ H− and SM = SPM is automatically a unitary


operator in M .

2.2. Asymptotic Properties of States in the Continuous


Spectra of the Enss Hamiltonian

H0
z }| {
Let H = (− 12 ∆) +V be a self-adjoint operator defined by an Enss potential
with M = 1 (see Eq. (2.13)). Then we have the result5 :
Enss Lemma. Let f ∈ Hc (H) be a state belonging to the continuity sub-
space of the Enss Hamiltonian. Then there is a sequence of times tn such
that limn→∞ tn = +∞ (tn ≤ tn+1 , ∀n ∈ Z) and

lim F (|x| < 13n)eiHtn f L2 (RN ) = 0, (2.21a)


n→∞

 n
lim dt|F (|x| < n)e−iH(t+tn ) (H + i1)f L2 = 0. (2.21b)
n→∞ −n

In order to show such result let us remark


 T that if a continuous function
has an ergodic mean zero, i.e. limT →∞ T1 ( 0 g(z)dz) = 0, then there is a
monotone sequence (tn ) such that limn→∞ g(tn ) = 0 with tn ≤ tn+1 and
tn → ∞. Now it is a deep theorem that (see Chap. 8)

1 T
lim dtf (|x| < n)e−iH(t) (H + i1)f L2 (RN ) = 0. (2.22)
T →∞ T 0

So, the result of lemma is a simple consequence of the above-stated


RAGE theorem applied to the z-functions below written together with
Fubbini theorem in order to interchange an integration order
 n
W2n (z) = F (|x| < n)e−iH(t+z) (H + i1)f L2 dt, (2.23)
−n

W2n+1 (z) = F (|x| < n)e−iHz f L2 . (2.24)

Let us now define the Enss’ scattering states for a given state f ∈ Hc (H)

fn = exp(−iHtn )f. (2.25)

We now have the following operatorial limit:

lim ((H + z1)−1 − (H0 + z1)−1 )F (|x| > (13n)operator = 0. (2.26)


n→∞
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 45

The validity of Eq. (2.26) is a simple consequence of the continuity and


integrability of the Enss function (see Eq. (2.13))

((H + z1)−1 − (H0 + z1)−1 )F (|x| > (13n)operator


≤ V (H0 + z1)−1 F (|x| ≥ (13n)operator = hz (R) (2.27)

since
 14n 
0 = lim dRhz (R) ≥ lim nhz (13n). (2.28)
n→∞ 13n n→∞

Similar result is obtained through derivatives in relation to the z-


parameter. As the polynomials in the variables (x + i)−1 and (x − i)−1 are
dense in the space of the continuous functions vanishing at ∞ (the Stone–
Weierstrass theorem — Chap. 1), by a simple density argument we arrive
at the following result for any function φ(x) ∈ C0 (R):

lim (φ(H) − φ(H0 ))fn L2 (RN ) = 0. (2.29)


n→∞

At this point we introduce the Enss scattering states fn ∈ Hc (H), the


physical requirement that they are really physical by possessing finite (inter-
acting) energy. This last condition is achieved by considering the energy
cut-off new-states (0 ≤ a < b < ∞)
 2

¯ 2 |b − a|
fn = EH 73a , fn , (2.30)
4

where EH ([a, b]) denotes the spectral projection of the full Enss
Hamiltonian H, associated with the spectral interval [a, b]. (Note that
Hc (H) ⊂ R+ since (H + z1)−1 is a compact operator — see Appendix C.)
Let us consider φ̃(x) ∈ C ∞ (R) such that φ(x) = 1 for x ∈ [73a2 ,
(b − a)2 /4] and φ(x) ≡ 0 for r ∈ [72a2 , (b − a)2 /4]. We now consider the
“physical free-energy packet” of scattering states
 
1
φn = ϕ̃ − ∆ f¯n . (2.31)
2

Let us point out that the Fourier transforms


 of the states φn (x) all have
√ (b−a) 2
support in the region 2 · 72a2 ≤ |P | ≤ 2 2 ⇔

supp(F [φn ](p)) ⊆ {p ∈ RN | 12a ≤ |p| ≤ b − a}. (2.32)


August 6, 2008 15:46 9in x 6in B-640 ch02

46 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point we remark that we need to further localize our scattering


states by taking into account fully its quantum nature as given by the
Heisenberg Principle.5 In order to achieve such a result, we introduce a
function s(x) ∈ S(RN ) such that its Fourier transform has support in the
ball Ba (0) in RN with S̃(0) = F [S](0) = 1.
Let us define the quantum mechanical classical localization operators
for a given Borelian A

F0 (A)(x) = dyS(x − y)χA (y). (2.33)
RN

At this point we can see that the family of functions {F0 (Aj )}j∈I is a
unity decomposition if {Aj }j∈J is a disjoint decomposition of RN
 
F0 (Aj ) = S(y)dN y = S̃(0) = 1. (2.34)
j∈I

Another important point to call the readers’ attention is the fact that
the quantum localized states F0 (A)φn all have their momenta in the Ball
with radius b. This result is easily obtained through the result that for
m(A) < ∞, we have

supp F [F0 (A)](p) ⊆ supp S̃(p) ⊆ Ba (0). (2.35)

In the case of m(A) = +∞ where χA (x) ∈ / L1 (RN ), we left it for our


readers by just considering the equivalent expression

χA (x − y)
F0 (A) = dN y · S(y)(|(x − y)2N + 1|) . (2.36)
(|x − y|2N + 1)

Let us now consider the best quantum mechanical spatial clas-


sical localization decomposition of RN in Balls and truncated cones
(j) (j)
{B12n (0), C12n }j∈I ,6 where the cones C12n = {x ∈ RN | |x| ≥ 12n, xej ≥
|x|; 2} have axis along the vector {êj }, set of unity vector to be suitably
choosen.
Then we have the following results:

lim F0 (B12n )φn L2 = 0, (2.37a)


n→∞
  
  
 (j) 
lim fn − F0 (C12n )φn  = 0. (2.37b)
n→∞   2
j∈I L
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 47

The proof is achieved by using the estimates

F0 (B12n )φn − F (|x| ≤ 13n)f¯n L2


≤S̃(0)=1)
  
≤ ϕ̃(H0 )f¯n − ϕ̃(H)f¯n L2 F0 (B12n )L2
+ [(F0 (B12n ) − f (|x| ≤ 13n)]ϕ̃(H)f¯n L2 (2.38a)

and
 
  
 
lim  F0 (B12n (fn ))  = 0. (2.38b)
n→∞  
j∈I L2

Let us finally consider the physical Enss localized scattering states by its
classical motion direction. Let hj (p) ∈ C0∞ (RN ) be the functions satisfying
the conditions
hj (p) = 0 for p · êj ≤ −2a,
hj (p) + hj (−p) = 1 for |p| < b (2.39)
(j)
with êj denoting the vector in the direction of the cone’s axis C12n as
previously stated.
The Enss OUT and IN states are explicitly given by5
(j)
φn (j)OUT = F0 (C12n )hj (p)φn , (2.40a)
(j)
φn (j)IN = F0 (C12n )hj (−p)φn . (2.40b)

We now have the important technical result of free evolution of the


above considered directional scattering states (for a detailed proof see
Appendix A).
Enss Theorem (Technical). We have the following result:
 ∞
lim dtF (|x| ≤ n + at)e−iH0 t (H0 + i1)φn (j)OUT  = 0, (2.41a)
n→∞ 0
 0
lim dtF (|x| ≤ n − at)e−iH0 t (H0 + i1)φn (j)IN  = 0. (2.41b)
n→∞ −∞

After displaying their geometrical properties of the Enss scattering


states, we now pass to the proof of the asymptotic completeness of the Enss
Hamiltonian in Sec. 2.3.
August 6, 2008 15:46 9in x 6in B-640 ch02

48 Lecture Notes in Applied Differential Equations of Mathematical Physics

2.3. The Enss Proof of the Non-Relativistic One-Body


Quantum Mechanical Scattering

Enss Theorem (Asymptotic completeness).5 Let H be the Enss


Hamiltonian and its associated wave operators W ± (H, H0 ). Then, the sub-
spaces H± ≡ Range(W ± ) coincide with the subspace of continuity Hc (H)
of the Enss Hamiltonian.
Proof. Let ψ ∈ M be one of the quantum mechanical localized state of
the previous section. Note that M is dense in Hc (H). We now observe that
(Eq. (2.37b))
  
   
 IN OUT 
lim ψn − φn (j) + φn (j) 
n→∞   2
j∈I j∈I L
   
  
 (j) 
lim ψn − F0 (C12n φn ) = 0. (2.42)
n→∞   2
j∈I L

Let us suppose that there is a state ψ ∈ Hc (H) such that it is orthogonal


to H+ = Range(W + ).
We observe now that the estimate below holds true

lim |ψn , ψn |
n→∞
  
  
 IN OUT
= lim  ψn , ψn − φn (j) + φn (j)
n→∞ 
j∈I j∈I
 
  
IN OUT 
− φn (j) + φn (j) 

j∈I j∈I
  
   
 IN OUT 
≤ lim  ψn , ψn − φn (j) + φn (j) 
n→∞  
j∈I j∈I
   
     
 IN   OUT 
+ lim  ψn , φn (j)  + lim  ψn , φn (j) 
n→∞   n→∞  
j∈I j∈I
(2.43a)
   
   
 
≤ lim ψn ψn − φn (j)IN + φn (j)OUT 
n→∞  
j∈I j∈I
(2.43b)
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 49

 
  
 IN 
+ lim ψn , φn (j)  (2.43c)
n→∞  
j∈I
 
  
 + OUT 
+ lim  ψn , W φn (j)  . (2.43d)
n→∞  
j∈I

Here, we have used the technical lemma (see Appendix B)

lim (W + − 1)φn (j)OUT  = 0, (2.44a)


n→∞

lim |φn , φn (j)IN  = 0. (2.44b)


n→∞

As a consequence that ψ, W + w = 0, ∀w ∈ Dom(W + ), we have that


the only remaining term in the estimate considered above is the following:
  
  
 IN OUT 
lim  ψn , ψn − φn (j) + φn (j)  = 0. (2.45)
n→∞  
j∈E

In view of Eq. (2.42), we have the contradiction

0 = ψ2 = lim |ψn , ψn |


n→∞
  
  
 IN OUT 
≤ lim  ψn , ψn − φn (j) + φn (j) =0 (2.46)
n→∞  
j∈I

The proof of H− = Hc (H) is similar and only need to consider φn (j)IN


in Eq. (2.44a). Details are left as an exercise. Note that this result implies
that the singularly continuous spectrum of the Enss Hamiltonian is empty
and all bound states are orthogonal to HAC (H). 

References

1. S. T. Kuroda, On the existence and the unitary property of the scattering


operator, Nuovo Cimento 12, 431–454 (1959).
2. N. Wiener, The Fourier Integral and Certain of its Applications (Cambridge
University Press, NY, 1933).
3. W. O. Amrein, M. M. Jauch and B. L. Sinha, Scattering Theory in Quantum
Mechanics (W.A. Benjamin, Ins., Massachussetts, 1977).
4. J. B. Dollard, Adiabatic switching in the Schrödinger theory of scattering,
J. Math. Phys. 802–810 (1966).
5. V. Enss, Asymptotic completeness for quantum mechanical potential
scattering, Commun. Math. Phys. 61, 285–291 (1979).
August 6, 2008 15:46 9in x 6in B-640 ch02

50 Lecture Notes in Applied Differential Equations of Mathematical Physics

6. E. B. Davies, Quantum Theorem of Open Systems (Academic Press, New York,


1976).
7. S. Steinberg, Meromorphic families of compact operators, Arch. Rational
Mech. Anal. 31 (1969).

Appendix A
IN
Theorem A.1. Let us consider the Enss scatterbug states φn (j)OUT . Then
we have the result
 ∞
lim dtF (|x| ≤ n + at)e−iH0 t (H0 + i1)φn (j)OUT  = 0, (A.1)
n→∞ 0
 ∞
lim dtF (|x| ≤ n − at)e−iH0 t (H0 + i1)φn (j)IN  = 0. (A.2)
n→∞ 0

Proof. Our aim is to reduce the above estimates to the one-dimensional


case. Thus, let us analyze Eq. (A.1) by considering a new decomposition of
the Enss scattering states φn (j)IN in the x-space RN . First, let us observe
that the half-spaces x · F ≤ n + at with f = {fn }1≤k≤N ∈ RN and f  = 1
contains the Ball Bn+at (0) = {x ∈ RN | |x| ≤ n + at}. We let the function
hj (p) with {p ∈ RN | 12a ≤ |p| ≤ b−a} be decomposed into a finite number
(j) (j)
of summands ξk (p) ∈ C0∞ (RN ) in such a way that support {ξk (p)} ⊆
{p ∈ R | P · fn ≥ 2a}.
N
(j)
Let us choose fk unity vectors with an angle of π/9 rads with the cones
(j)
vectors axis êj of C12n .
(j) (j)
As a consequence x · fk > 2n for x ∈ C12n . We remark that Eq. (A.1)
will be a result of the validity of the following limit:
 ∞
(j) (j) (j)
lim dtF (x · fk ≤ n + at)e−iH0 t (H0 + i1)F0 (C12n )ξk (p)φn  = 0
n→∞ 0
(A.3)
since any state φn has its “momenta” support in the region 12a ≤ |p| ≤ b−a.
Within this region one has decomposition of the Enss scattering state
(j)
 (j) (j)
φn (j)OUT = F0 (C12n )h0 (p)φn = F0 (C12n )ξk (p)φn . (A.4)
k

Just in order to simplify the estimates which follows and choose the axis
(k)
x1 parallel to fj under consideration. Let us thus consider
(j)
Fm = (H0 + i1)F0 (C12n ∩ {2n + m ≤ x1 ≤ 2n + m + 1}). (A.5)
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 51

Obviously


(j)
(H0 + i1)F0 (C12n ) = Fm . (A.6)
m=0

It will be necessary to implement the action of (H0 + i1) by means of


a multiplication operator in p-space by means of a function with compact
support. This can be implemented by noting that all the states F0 (A)φn ,
with A denoting a Borelian of RN , have p-momentum support in the region
|p| ≤ b. Another point worth remarking is that the momentum component
p1 (related to the x1 -axis) of the states
(j) (j)
(H0 + i1)F0 (C12n ∩ {2n + m ≤ x1 ≤ 2n + m + 1})ξk (p)φn (A.7)

is in the interval a ≤ p1 ≤ b.
Let us now pass to the estimates. First, the estimate equation (A.3)
reduces to the one-dimensional case as a consequence of the dependence of
solely the coordinate x1 into our sliced up states equation (A.7).
We have thus reduced our analysis to the one-dimensional case
 ∞  2
! 
 −i ∂ ∂ t 
dtF (x1 ≤ n + at)e 2x
1 φ(x1 , x̂) 2 (A.8)
0 L

with φ(x, x̂) denoting our new sliced up states.


Let us define the following operator family of operators in L2 (RN ):
2 " +∞ #  2 2$ % &
eix1 /2ζ iy 2 iy1 x1
(Wζ φ) = dy1 1 + 1 + 2 e−i ζ y1 φ(y1 , x̂) .
(2πiζ)1/2 −∞ 2ζ 2ζ
(A.9)

It is a consequence of the support of the state φ(y1 , x̂) that

(Wt φ)(x) = 0 for x1 < at (A.10)

by just considering Eq. (A.9) by means of the Fourier transform


2 #  
eix1 /2t i ∂2
(Wt φ)(x) = 1 +
(2πit)1/2 2t ∂ 2 ( xt1 )
 2   %
i ∂3 x1
+ F y1 [φ] , x̂ (A.11)
2t ∂ 3 ( xt1 ) t

and Fy1 [φ](p1 , x̂) ≡ 0 for p ≤ x1


ζ < a.
August 6, 2008 15:46 9in x 6in B-640 ch02

52 Lecture Notes in Applied Differential Equations of Mathematical Physics

We now note the validity of the result written as


 #   %
 −i ∂∂
2 t 
(1 + t)(−x1 + 2at)2 e 2x 2
− Wt φ (x) 

 1

L2
  2 # iy12 % 
 ∂ iy12 
 2
= (1 + t)t − i + 2a e ib21 /2t
−1+ + 2t
φ(y1 , x̂)
∂y1 2t 2  2
L
  
 (1 + t)t 2  ∂ 
≤ |P1∗ (|y1 |)|(φ(y1 , x̂)) + |P2∗ (|y1 |)| φ(y1 , x̂)
t3 ∂y1
 2 
∂ 
+ |P3 (|y1 |)|  2 φ(y1 , x̂) 
∗ 
 2, (A.12)
∂y1 L

where P∗ (|y1 |), = 1, 2, 3 are polynomials satisfying the relations (x1 ≥ ã)
'
P3∗ (|y1 |) = |y1 |6 48,
4a|y1 |6 2|y1 |5
P2∗ (|y1 |) = + ,
48 8
     
1  1 

P1 (|y1 |) ≥ P |y1 |, ≥ P |y1 |,   .
 (A.13)
ã t
We have this for x1 ≤ at the estimate
  2 #  % 
 iy12 (iy12 ) 1 
(1 + t)t2 − i ∂ + 2a e iy12 /2t
− 1 + + · φ(y , x̂)
 ∂y1 2t (2t) 2
1 
 
a  ∗ ∂
≤ +1  ∗
P1 (|y1 |)φ(y1 , x̂) + P2 (|y1 |) ∂y1 φ(y1 , x̂)

  "    
∂ 2 φ(y1 , x̂) 
 = a  P1∗ ∂
 ∂
+ P3∗ (|y1 |) + P ∗
ip1
∂ 2 y1  ã  ∂p1 2
∂p1
  & 
∂ 
+ P3∗ (−p1 )2 F [φ](p1 , p̂)  2 ≤A (A.14)
∂p1 L

since supp φ̃(p1 , p̂) ⊂ {p ∈ RN | |p| ≤ b}.


As a consequence (see Eqs. (A.10) and (A.14))
 2
! 
 2 − i ∂ 2 2t 
(1 + t)(−x1 + 2at) F (x1 ≤ at)e ∂x1
φ(x1 , x̂) ≤ A. (A.15)

The result of Eq. (A.15) is the following:


 −i ∂ 2 2t !
2 
  A
F (x1 ≤ −v + at) e ∂x1 φ (x) ≤ (A.16)
(1 + t)(u + at)2
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 53

since for v > 0, (x1 ≤ −v + at)

1 1
≤ · (A.17)
(−x1 + 2at)2 (v + at)2

As a consequence we have the following estimate for v = m + n and


(j)
φ(x) = Fm ζkj (p)φn (x1 + 2m + n, x̂) and x1 ∈ C12n ∩ {2n + m ≤ x1 ≤
2n + m + 1} (Ref. 5) by using translational invariance x1 → x1 + v:
(j)
F (x1 − 2n − m < −(n + m) + at)e−iH0 t Fm ζk (p)φn (x1 , x̂)L2
A
≤ . (A.18)
|(1 + t)|m + n + at|2 |

By noting again that



 (j)
F (x1 − 2n − m < −n − m + at)e−iH0 t Fn ξk (p)φn (x)
m=0

 (j)
= F (x1 < n + at)e−iH0 t Fn ξk (p)φn (x).
m=0

We have
(j)
F (x1 < n + at)e−iH0 t (H0 + i1)ξk (p)φn (x)
  ∞ 
  
 −iH0 t (j) 
= F (x1 < n + at)e Fn ξk (p)φn (x) 
 
m=0

 (j)
≤ F (x1 + n + at)e−iH0 t Fn ξk (p)φn (x)
m=0
 ∞

 1
≤A .
m=0
|(1 + t)(m + n + at)2 |

We now left as an exercise to our diligent reader to establish the result5


 ∞
(j) (j) (j)
lim dtF (x · fk ≤ n + at)e−iH0 t (H0 + i1)F0 (C12n )ξk (p)φn 
n→∞ 0
 ∞   
 ∞ 1
≤ A lim dt = 0.
n→∞
m=0 0
|(1 + t)(m + n + at)2 |

August 6, 2008 15:46 9in x 6in B-640 ch02

54 Lecture Notes in Applied Differential Equations of Mathematical Physics

Appendix B

Let us show the following technical result used in the proof of the Enss
theorem.
IN
Lemma B.1. Let W± be the Enss wave operators and φn (j)OUT considered
in Sec. 2.2, then we have the result (  =  L2 ).

lim (W + − 1)φn (j)OUT  = 0, (B.1)


n→∞

lim (W − − 1)φn (j)IN  = 0. (B.2)


n→∞

Proof. Since
 t 
 
(W (t) − W (0)f  ≤ 
 dζ(V e −iH0 ζ
f )
 (B.3)
0

we have that
 ∞ 
 
+
(W − 1)φn (j) ≤  OUT
dζV e −iH0 ζ
φn (j)OUT 

0
 ∞ 
−1 −iHo ζ OUT
≤ V (H0 + i1) op dζF (|x| ≤ n + aζ)e (H0 + i1)φn (j) 
0
 ∞ 
−1
+ dζV (H0 + i1) F (|x| ≥ n + aζ)op
0

× e−iH0 ζ (H0 + i1)φn (j)OUT L2 . (B.4)

The first term in Eq. (B.4) goes to zero by the long estimate given in
Appendix A.
The second term can be estimated as follows (e−iH0 ζ (· · · ) ≤ (· · · )!)

 ∞
(H0 + i1)φn (j)OUT  h(n + aζ)dζ
0
 ∞ 
≤ (H0 + i1)φn (j)OUT  h(R)dR . (B.5)
n
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 55

Since

(H0 + i1)φn (j)OUT L2 = (p2 + i1)F (φn (j)OUT )(p)L2


≤ (b2 + 1)F (φn (j)OUT )L2 < ∞ (B.6)

since F [φn (j)OUT ](p) has support in |p| ≤ b and obviously due to the fact
that h(R) ∈ L1 (R), we have
 ∞ 
lim dR h(R) = 0. (B.7)
n→∞ n

We now give a proof of the following additional result in the Enss proof of
asymptotic completeness presented in Sec. 2.3

lim |φn (j)IN , ψn | = 0. (B.8)


n→∞

First, we remark that

|φn (j)IN , ψn | ≤ |(W − φn (j)IN , e−iHtn ψ + (I − W − )φn (j)IN , e−iHtn ψ|
(B.9)
IN − IN
≤ |W− e iH0 tn
φn (j) , ψ| + (I − W )φn (j) ψ. (B.10)

Here we have used the result (exercise)

eiHt W− = W− eiH0 t . (B.11)

As a consequence

|φn (j)IN , ψn | ≤ (I − W − )φn (j)IN  + |eiH0 tn φn (j)IN , (W − )∗ ψ| (B.12)

since

lim (1 − W − )φn (j)IN  = 0 (B.13)


n→∞

we only need to prove that

lim |eiH0 tn φn (j)IN , (W − )∗ ψ| = 0 (B.14)


n→∞

in order to show the validity of Eq. (B.8).


August 6, 2008 15:46 9in x 6in B-640 ch02

56 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now, we have the set of estimates

|eiH0 tn φn (j)IN , (W − )∗ ψ|


= |(F (|x| ≤ n + atn ) + F (|x| ≥ n + atn ))eiH0 tn φn (j)IN , (W − )∗ ψ|
≤ {F (|x| ≤ n + atn )(W − )∗ ψF (|x| ≤ n + atn )eiH0 tn φn (j)IN 
+F (|x| ≥ n + atn )(W − )∗ ψF (|x| ≥ n + atn )eiH0 tn φn (j)IN 
≤ (W − )∗ ψF (|x| ≤ n + atn )eiH0 tn φn (j)IN 
+ F (|x| ≥ n + atn )(W − )∗ ψφn (j)IN . (B.15)

At this point we observe that

lim F (|x| ≤ n + atn )eiH0 tn φn (j)IN  = 0 (B.16)


n→∞

by a similar analysis made in Appendix A.5 

Appendix C

In the appendix, we intend to show that compact resolvent perturbations


of the free quantum dynamics H0 = − 12 ∆ does not modify the essential
operator spectrum: defined as the union of the continuous spectrum and the
accumulation points of the bound-states (point spectrum). Let us announce
this deep result in its general terms.
Theorem C.1. Let H = H0 + U and H0 self-adjoint operators acting on a
separable Hilbert Space H. Let us suppose that all H0 spectrum is contained
in positive real axis and such that (H0 +U +λ1)−1 −(H0 +λ1)−1 is a compact
operator for some complex λ0 , with Im(λ0 ) = 0. Then the spectral part
of the operator H contained in Real(λ) ≤ 0 is formed entirely by discrete
points possesing zero as the only possible accumulation point for them.

Proof. This result is obtained through an application of the Holomorphic


Fredholm theorem to the following operatorial-valued holomorphic
function:

f (λ) = 1 − (λ − λ0 )[1 − (λ − λ0 )(H0 + λ1)−1 ]−1


× [(H0 + U + λ1)−1 − (H0 + λ1)−1 ] (C.1)
  
C(λ)-compact operator
August 6, 2008 15:46 9in x 6in B-640 ch02

Möller Wave Operators and Asymptotic Completeness 57

Since there is λ0 ∈ C\R+ such that f (x) is inversible (f (λ0 ) = 1!), we


have that C(λ) has inverse in all C\R+ , with the exception of a discret
set S ⊂ C\R+ . Since for λ with non-zero imaginary part C(λ) is always
inversible, we have that S ⊂ R− . It is straightforward to see that S coincides
with the negtive part of the spectrum of H. 
Just for completeness let us announce the holomorphic Fredholm
theorem.7
Holomorphic Fredholm Theorem. Let Ω be a connected open set of C
(a domain). Let

f : Ω → (L(H, H), uniform ),

a holomorphic bounded-operator valued function such that for each z ∈ Ω,


f (z) is a compact operator. Then we have the Fredholm alternatives:
(a) (1 − f (z))−1 does not exists for any z ∈ Ω
or
(b) (1 − f (z))−1 exists for z ∈ Ω\S, where S is a discret subset of Ω. In
this case (1 − f (z))−1 is a meromorphic function in Ω, holomorphic in
Ω\S and in z ∈ S, the equation f (z)ψ = ψ, has solely a finite number
of linearly independent solutions in H.
August 6, 2008 15:46 9in x 6in B-640 ch03

Chapter 3

On the Hilbert Space Integration Method for the Wave


Equation and Some Applications to Wave Physics∗

3.1. Introduction

One important topic of mathematical methods for wave propagation physics


is to devise functional analytic techniques useful to implement approximate-
numerical schemes.1 One important framework to implement such time-
evolution studies is the famous Trotter–Kato–Feynman short-time method
to approximate rigorously the full-wave evolution by means of a finite
number of suitable short-time propagations — the well-known Feynman
idea of a path-integral representation for the Cauchy problem.2
In this chapter we propose a pure Hilbert space analytical framework —
the content of Secs. 3.2 and 3.3 — and suitable to generalize rigorously
to the hyperbolic case, the powerful Trotter–Kato–Feynman formulae in
Sec. 3.4. In Sec. 3.5, we present as a further example of usefulness of
our abstract Hilbert space Techniques, a simple proof (by means of the
use of the integral Cook–Kuroda criterium) of the existence of reduced
(finite-time) scattering wave operators W± for elastic wave scattering by
a compact supported inhomogenous medium (a “small” space-dependent
Young medium elasticity modulus E(x) ∈ Cc∞ (R3 ), i.e. sup | E(x)
E0 |  1 and
∇E(x) 3–7
sup | E0 |  1).
In Sec. 3.6 and Appendix A, we address the very important problem of
giving a complete proof for the asymptotic stability for systems in thermoe-
lasticity in the relevant linearizing approximation for the nonlinear magne-
toelasticity wave equations,8,9 i.e. a coupling between the divergences Lamé
system of elastic vibrations and the convective parabolic equation for a
constant external magnetic field.
We give a (new) proof of stability for this system by producing an
exponential uniform bound in time for the total energy of the linearized

∗ Author’s original results.

58
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 59

magnetoelastic model through the systematic use of simple estimates on the


Trotter formula representations for our magnetoelastic wave propagators in
our proposed Hilbert space abstract scheme.
The resulting exponential bound solve an important long-standing
problem on the structural form of the decay behavior of magnetoelastic
systems in the presence of external constant magnetic fields.
Finally, in Sec. 3.7, we use again Hilbert space techniques to produce a
proof of existence and uniqueness of a semilinear problem of Abstract Klein
Gordon propagation11 as another example of usefulness of our proposed
abstract approach.

3.2. The Abstract Spectral Method — The


Nondissipative Case

Let A be a positive (essential) self-adjoint operator with domain D(A)


acting on a separable Hilbert space H (0) with the inner product denoted
by (, )(0) .
Associated to the above abstract operator A, one can consider the wave
equation (Cauchy problem) on H (0) × [0, T ]
∂2U
= −AU + f, (3.1)
∂2t
U (0) = U0 ; Ut (0) = v0 (3.2)

with initial dates given in suitable subspaces of H (0) to be specified later


and f ∈ L2 ([0, T ], H (0) ).
It is an important problem in the mathematical physics associated with
Eqs. (3.1) and (3.2) to determine exactly the subspaces where the initial
date should belong in order to insure the existence of a global solution of
Eqs. (3.1) and (3.2) in C ∞ ((0, ∞), H (0) ) when the time-interval [0, T ] of
the external source f (x, t) is infinite (T = +∞) and besides, when there
exist a natural extension of this global solution to the whole time interval
(−∞, ∞), namely,

U (x, t) ∈ C ∞ ((−∞, ∞); H (0) ).

Let us present in this section a solution for this problem by means of a


convenient generalization method exposed in Ref. 1 and based uniquely on
the Stone spectral theorem for one-parameter unitary groups in Hilbert
space.
August 6, 2008 15:46 9in x 6in B-640 ch03

60 Lecture Notes in Applied Differential Equations of Mathematical Physics

In order to implement this Hilbert space abstract method,1 we introduce


the following Hilbert space:

H (1) = {(ϕ, π) | ϕ ∈ D(A); π ∈ H (0) ,  ; }, (3.3a)

where the closure is taken in relation to the inner product below:

(ϕ1 , π1 ); (ϕ2 , π2 ) = (Aϕ1 , ϕ2 )(0) + (π1 , π2 )(0) . (3.3b)

Let us introduce the following operator LA on H (1) and defined by


the rule

LA (ϕ, π) = (π, Aϕ). (3.4)

Note that Eq. (3.4) is well defined since we assume that the range of A is
dense on H (0) . LA is a symmetric operator on H (1) since

(ϕ1 , π1 ); LA (ϕ2 , π2 ) = (Aφ1 , π2 )(0) + (π1 , Aφ2 )(0)


= LA (ϕ1 , π1 ), (ϕ2 , π2 ). (3.5)

Let us consider the smallest closed extension of LA , denoted by L̄A . We


claim that its deficiency indices n+ (L̄A ) and n− (L̄A ) are identically zero and
as a consequence leading to its self-adjointness on its domain D(L̄A ) ⊂ H (1)
which is explicitly given by the vectors (ϕ, π) such that (Aϕ, ϕ)(0) < ∞ and
π ∈ H (0) .
Let (ϕ1 , π1 ) be a nonzero vector such that (ϕ1 , π1 ) ∈ (Range(i + LA))⊥ ,
or equivalently for any ϕ2 ∈ D(A) and π2 ∈ H (0) :

(ϕ1 , π1 ); (i + LA )(ϕ2 , π2 )(1) = 0, (3.6)

which is the same content of the equation below:

−i(Aϕ1 , ϕ2 )(0) + (π1 , Aϕ2 )(0) + i(π1 , π2 )(0) + (Aϕ1 , π2 ) = 0, (3.7)

and due to the fact that Range(A ± i) is dense on H (0) and since A is self-
adjoint it leads to ϕ1 = 0 and π1 = 0. In the same way one can prove that
n− (L̄A ) = 0.
Collecting all the above exposed results, one has the following theorem:

Theorem 3.1. Let the spectral theorem be applied to the one-parameter


group U (t) = exp(itL̄A ) acting on D(L̄A ); one has that it satisfies the
abstract wave equation (3.1) on the strong sense for dates U0 ∈ D(A)
and v0 ∈ H (0) and leading to a global (time-reversing) solution on
C ∞ ((−∞, ∞), H (0) ).
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 61

Proof. It is straightforward that in components the wave equation asso-


ciated with the one-parameter semigroup U (t)(ϕ0 , π0 ) = (ϕ(t), π(t)) is
given by

ϕ(t) = iπ(t), (3.8)
∂t

π(t) = i(Aφ)(t), (3.9)
∂t
which is the same as Eqs. (3.1) and (3.2) without the external term f (x, t):

∂ 2 ϕ(t)
= −(Aϕ)(t) (3.10)
∂ 2t
with the initial date

ϕ(0) = U0 , (3.11)
ϕt (0) = v0 . (3.12)

It is worth noting that one has always a weak solution of Eqs. (3.1)
and (3.2) with a nonzero external term
     t   
ϕ itL̄A U0 i(t−s)L̄A 0
(t) = e + ds e , (3.13)
π −iϕ0 0 −if (s)

namely, ϕ(·, t) ∈ C((−∞, ∞); H (0) ), unless f (·, t) belongs to


D(A)(f : (−∞, ∞) → D(A)), in which case it produces a global C ∞ time-
differentiable solution.
At this point, we make one of the most important remark that comes
from such a theorem applied to the subject of random wave propagation.2
It is important to know the correct mathematical meaning of considering
the Cauchy problem for the wave equation with both initial conditions now
being stochastic process with realizations on L2 (Ω) space of functions (see
Appendix B). In this case, Eq. (3.13) in its weak-integral form should be
considered the correct mathematical solution for this problem since one can
take safely the initial dates (U0 , v0 ) both in the Hilbert space H (0) , which
in the usual case of wave propagation is the L2 (Ω) functional Hilbert space
(H (0) = L2 (Ω)).
Let us point out an abstract version of the wave total energy conser-
vation during the time evolution.
First, we observe that for any −∞ < t < ∞ we have the identity

Im{(π(t), Aϕ(t))(0) } ≡ 0, (3.14)


August 6, 2008 15:46 9in x 6in B-640 ch03

62 Lecture Notes in Applied Differential Equations of Mathematical Physics

which is obtained as an immediate consequence of the unitary property


of the semigroup U (t) on the two-component Hilbert space H (1) given by
Eqs. (3.3a) and (3.3b):
exp(itL̄A )(ϕ0 , π0 ); exp(itL̄A )(ϕ0 , π0 )
= (ϕ0 , π0 ); (ϕ0 , π0 ), (3.15)
and by deriving the above equation in relation to time, it yields
iL̄A (ϕ(t), π(t)), (ϕ(t), π(t))
+ (ϕ(t), π(t)); iL̄A (ϕ(t), π(t)) = 0, (3.16)
which is Eq. (3.14) when it is written in terms of components.
At this point one can easily see that the energy functional given by
1 1
E(t) = (π(t), π(t))(0) + (φ(t), (Aφ)(t)) (3.17)
2 2
is such that

dE(t) 1 1
= (i(Aφ)(t), π(t)) + (π(t), i(Aφ)(t)) 0
dt 2 2

1 1
+ (iπ(t), (Aφ)(t)) + ((Aφ)(t), iπ(t)) 0
2 2
= 0. (3.18)
As a consequence, we have the energy abstract conservation law
1 1
E(t) = (v0 , v0 )(0) + (U0 , AU0 )(0) . (3.19)
2 2
Finally, we remark that by a straightforward application of the Banach
fixed-point theorem within the context of our operatorial scheme, one has
the existence and uniqueness of the nonlinear abstract wave equation initial
value problem on H (1) :
∂ 2U
= −AU + F (U ), (3.20)
∂2t
U (0) = U0 ; Ut (0) = v0 ,
for F (x) denoting a Lipschitzian function on the real line (|F (x) − F (y)| ≤
C|x − y|). This result is the consequence of the fact that the application
T : C([0, T ], H (1) ) → C([0, T ]; H (1) )
 t
T (ϕ, π) = ds{ei(t−s)L̄A (0, F (ϕ))} (3.21)
0
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 63

is a contraction on H (1) , i.e. ((Aϕ, ϕ)H ≥ C0 (ϕ, ϕ)H ):

sup ||(0, F (ϕ1 (s)) − F (ϕ2 (s))||H (1)


0≤s≤T

≤ sup {(A0, 0)(0) + (F (ϕ1 (s)) − F (ϕ2 (s)), F (ϕ1 (s) − F (ϕ2 (s)))(0) }
0≤s≤T

≤ C sup (ϕ1 (s) − ϕ2 (s), ϕ1 (s) − ϕ2 (s))(0)


0≤s≤T
≤ CC0 sup {||(ϕ1 , π1 ) − (ϕ2 , π2 )||H (1) (s)}, (3.22)
0≤s≤T

which after successive iterations produces bounds leading to the global


weak solutions of Eq. (3.20) in C((−∞, ∞), H (1) ), a nontrivial result easily
obtained in this abstract spectral framework (see Sec. 3.7 for further studies
on this problem).

3.3. The Abstract Spectral Method — The Dissipative Case

In this section we generalize the Hilbert space method of Sec. 3.2 to the
case of a wave equation in the presence of damping terms, a very important
problem on statistical continuum mechanics.4
We, thus, consider the abstract problem in the notation of Sec. 3.2:
 
∂2U ∂U
+ B + (AU )(t) = 0, (3.23)
∂2t ∂t
U (0) = U0 ∈ D(A), (3.24)
Ut (0) = v0 ∈ H (0) , (3.25)

where A and B are essential positive self-adjoint operators such that


D(B) ⊃ D(A).
Note that we can write Eqs. (3.23)–(3.25) in the two-component Hilbert
space H (1) of Sec. 3.2 as
    
d U 0 −1 U
=− . (3.26)
dt π A B π

0 −1
Let us observe that the operator  = A B
is positive on H (1) as a
consequence of the operator B positive definiteness, namely,

Â(U1 , π1 ), (U1 , π1 ) = (Bπ1 , π1 )(0) > 0. (3.27)


August 6, 2008 15:46 9in x 6in B-640 ch03

64 Lecture Notes in Applied Differential Equations of Mathematical Physics

Since ∃ λ ∈ R ≥ 1 such that λ1 + Â is onto and bijective on H (1)


(a topological isomorphism), one can apply, thus, the well-known Hille–
Yoshida theorem1 to the operator  in order to consider a contractive C0
semigroup on H (1) in the interval [0, ∞) possessing  as its generator, and
naturally producing the unique global weak solution of the abstract wave
equation with damping on C ∞ ([0, ∞), H (1) ) if the initial conditions are
chosen to be in H (1) .
      
U (x, t) 0 −1 U0
dU = exp −t (3.28)
dt (x, t)
A B v0

as much as in our study presented in Sec. 3.2.


For example, let us consider A as a strongly positive elliptic operator
of order 2m on a suitable domain Ω ⊂ RN and satisfying the Gärding
condition there.3

A= (−1)|α| Dxα (aαβ (x)Dxβ , (3.29)


|a|≤m
|β|≤m

D(A) = H 2m (Ω) ∩ H0m (Ω), (3.30)


Real(Af, f )L2 (Ω) ≥ C0 ||f ||H 2m (Ω) . (3.31)

The damping operator B is taken to be explicitly a strictly positive


function ν(x) of C ∞ (Ω).
As a result of this section, we have weak global solution on
C ([0, ∞), H 2m (Ω) ∩ H0m (Ω)) of the damped wave equation problem in Ω

with f (x, t) ∈ L∞ ([0, ∞), L2 (Ω)).4

d2 U dU
= −(AU )(x, t) − ν(x) (x, t) + f (x, t) (3.32)
d2 t dt
U (x, 0) = f (x) ∈ H 2m (Ω) ∩ H0m (Ω) (3.33)
2
Ut (x, 0) = g(x) ∈ L (Ω). (3.34)

Finally, let us show that the operator on D(A) ⊕ H (0) = H (1) :

 
0 −I
 = (3.35)
A B
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 65

is a topological isomorphism from its (dense) domain to the complete


Hilbert space H (1) as a consequence of the fact that the self-adjoint operator
A + B on D(A) satisfies the Hille–Yoshida criterion: For any real number
λ > 0, the equation (A + B + λ1)y = f has a unique solution on D(A) for
each f ∈ H (0) . The same Hille–Yoshida criterion applied to the full two-
component operator Eq. (3.35), shows us that for each (f 1 , f 2 ) ∈ H (1) , one
can straightforwardly find (U 1 , U 2 ) ∈ D(A) × D(A) with λ = 1, namely,
U 1 = (f 1 + U2 ) and (A + B + 1)U 2 = f 2 − Af 1 ∈ H (0) (Af 1 ∈ H (0) , since
f 1 ∈ D(A)!). The uniqueness of (U 1 , U 2 ) is a direct consequence of the
strict positivity of the operator A + B on D(A).

3.4. The Wave Equation “Path-Integral” Propagator

In this section of a more physicist-oriented nature, we consider the classical


unsolved problem of writing a rigorous Trotter like path-integral formulae
for the Cauchy problem on R3 associated with the elastic vibrations of an
infinite length string with a specific mass (ρ(x)−ρ0 ) ∈ Cc∞ (R3 ) and a Young
elasticity modulus (E(x) − E0 ) ∈ Cc∞ (R3 ). The general wave evolution
equation governing such vibrations is given by3,4
 
∂ ∂U (x, t)
ρ(x) = ∇(E(x)∇ · U (x, t)), (3.36)
∂t ∂t
U (x, 0) = f (x) ∈ H 2 (Ω), (3.37)
2
Ut (x, 0) = g(x) ∈ L (Ω). (3.38)

Just for simplicity of our mathematical arguments, we consider the case


of a string with a constant specific mass ρ(x) = ρ0 in our next discussions.
Firstly, let us show that the operator
 
E0 (E(x) − E0 )
A= ∆+∇ ∇ = A(0) + B (3.39)
ρ0 ρ0

is a self-adjoint operator on the domain H 2 (Ω). This result is an imme-


diate consequence of the fact that the “perturbation” operator B =
∇ E(x)−E
ρ0
0
∇ is A(0) -bounded with A(0) -bound less than 1. As a conse-
quence, one can apply straightforwardly the Kato–Rellich theorem to guar-
antee the self-adjointness of A on H 2 (Ω). The A(0) -boundedness of B is a
August 6, 2008 15:46 9in x 6in B-640 ch03

66 Lecture Notes in Applied Differential Equations of Mathematical Physics

result of the following estimates for ϕ ∈ H 2 (Ω) (where E(x)  E0 and


|∇E(x)|  E0 ):
 
E(x) − E0
∇ ∇ ϕ
ρ0 L2 (Ω)
 
|E(x) − E0 | ∇E(x) E0
≤ sup , γ(Ω) ∆ϕ , (3.40)
x∈Ω E0 E0 ρ0 L2 (Ω)

where γ(Ω) is the universal constant on the Poincaré relation for function
ϕ(x) on H 2 (Ω)

||∇ϕ||L2 (Ω) ≤ γ(Ω)||∆ϕ||L2 (Ω) . (3.41)

By the direct application of Theorem 3.1, we have the rigorous solution


for the Cauchy problem Eqs. (3.36)–(3.38) if f (x) ∈ H 2 (Ω) and g(x) ∈
L2 (Ω) for any −∞ ≤ t ≤ +∞, namely,
  
   0 1  f (x)
U (x, t)
= exp it   . (3.42)
∂U
∂t (x, t)
 ∇ E(x) ∇ 0  g(x)
ρ 0

At this point, we apply the Trotter–Kato theorem to rewrite Eq. (3.42)


in the short-time propagation form
     n  
U (x, t) t 0 1 f (x)
∂U (x,t) = Strong limit exp i .
∂t (H 2 (Ω)⊕L2 (Ω)) n→∞ n ∇ E(x)
ρ0 ∇ 0 g(x)
(3.43)

However, for short-time propagation, we have the usual asymptotic


result
    
 0 1  
lim exp iε  ∂ E(x) ∂  f (x)
ε→0+  0 g 
∂x ρ ∂x 0

 0 2 31 

 0 17C 

 
 B 6
5A   

+∞  @iε4
− E(x) 2 E  (x)
p + ρ0 (ip) 0 f˜ 
= lim dpeipx e ρ0
(p)
ε→0+ −∞ 
 g̃ 


 

 

(3.44)
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 67

with f˜(p) and g̃(p) denoting the Fourier transforms of the functions f (x)
and g(x), respectively, and E  (x)p means the scalar product ∇E(x)p for
short.
It is worth calling to the reader’s attention that
!  "  
0 1 f˜(p)
exp iε 
− E(x)
ρ0 p2 + Eρ(x)
0
ip 0 g̃(p)
   12   ˜ 
E(x) 2 E  (x) f (p)
= cos ε − p + ip
ρ0 ρ0 g̃(p)
 1 
2 E  (x) 2  
−i sin ε − E(x)
ρ0 p + ρ0 ip iεg̃(p)
#   (1/2) .
2 E  (x)
− E(x) 2 E  (x) ε − E(x)
ρ0 p + ρ0 ip f˜(p)
ρ0 p + ρ0 ip
(3.45)
As a consequence of the above formulae we have our discretized path-
integral representation for the initial-value Green function of Eqs. (3.36)–
(3.38), i.e. by writing the evolution operator as an integral operator-
Feynman Propagator
 2 3

 0 17

 7
6

 6
5
 
 it4 ∂ E(x) ∂ 0  f
∂x ρ0 ∂x
e

 
 g

 


 

 +∞   

G11 (x, x , t) G12 (x, x , t) f (x )
= dx (3.46)
−∞ G21 (x, x , t) G22 (x, x , t) g(x )

we have (for A = 1, 2; B = 1, 2) the “path-integral” product representation


 n  +∞  +∞
 1
GAB (x, x , t) = strong − lim dp1 · · · dpn dxn−1 · · ·
n→∞ 2π −∞ −∞
 
 +∞ n 
$   
t
× dx1 ×  GAj Aj+1 xj , pj ; 
−∞ n 
j=1

× exp(ipj (xj − xj−1 ), (3.47)

where GAj Bj (xj , pj , nt ) is the 2 × 2 matrix on the integrand of Eq. (3.31)


for x = xj and p = pj . Note that xn = x and x0 = x .
August 6, 2008 15:46 9in x 6in B-640 ch03

68 Lecture Notes in Applied Differential Equations of Mathematical Physics

3.5. On The Existence of Wave-Scattering Operators

In this section we give a proof for the existence of somewhat reduced wave-
scattering operators associated to the following wave dynamics pair (H̃, H̃0 )
on the H (1) = H 2 (Ω) ⊕ L2 (Ω) (with Ω = RN , N ≥ 3)7 :
 
0 1
H̃0 = E0
, (3.48)
ρ0 ∆ 0
 
0 1
H̃ = E(x) . (3.49)
∇· ρ0 ∇ 0

We need, thus, to show the strong convergence on H (1) of the following


family of one-parameter evolution wave operators (s = strong) for T arbi-
trary (but finite):

W ± (H̃, H̃0 ) = S − lim (exp(itH̃) exp(−itH̃0 ). (3.50)


t→T

Let us remark that the existence of the strong limit of Eq. (3.50) is
a direct consequence of the validity of the famous Cook–Kuroda–Jauch
integral criterion6 applied to the pair (H̃, H̃0 ) in a dense subset of the
domain of H̃. This integral criterion means that there exists a δ > 0, such
that the following integral is finite for any (f, g) belonging to a dense set of
H (1) , i.e.
 T  
f
dt (H̃ − H̃0 )e iH̃0 t
< ∞. (3.51)
δ g H (1)

Let us denote Ẽ(x) = (E(x) − E0 ) ∈ Cc∞ (RN ), the scattering wave


potential in what follows.
We can, thus, rewrite Eq. (3.51) in the following explicit form:
 T
dt((−∇(Ẽ(x)∇)g(x, t), g(x, t)))L2 (RN ) < ∞ (3.52)
δ

or
 T
dt((+∇(Ẽ(x)∇)f (x, t), +∇(Ẽ(x)∇)f (x, t)))L2 (RN ) < ∞, (3.53)
δ
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 69

where the pair of functions f (x, t) and g(x, t) satisfy the free wave equation
 
d2 f (x, t) E0
= ∆ f (x, t), (3.54)
d2 t ρ
f (x, 0) = f ∧ (x), (3.55)

ft (x, 0) = g (x), (3.56)

and
d
g(x, t) = f (x, t). (3.57)
dt

Let us choose the initial dates f ∧ (x) and g ∧ (x) as the Fourier trans-
forms of L2 (RN ) functions polynomial-exponentially bounded ( ∈ Z+ and
Λ ∈ R+ ):


f (x) = dN keikx f˜Λ (k); |f˜Λ (k)| ≤ M e−Λ|k| |k| , (3.58)
RN


g (x) = dn keikx g̃Λ (k); |g̃Λ (k)| ≤ M e−Λ|k| |k| . (3.59)
RN

Note that we have an explicit representation in momentum space for


the free waves f (x, t) and g(x, t), namely,

g̃Λ (|k|)
f˜(k, t) = f˜Λ (|k|) cos(|k|t) + sin(|k|t), (3.60)
|k|

g̃(k, t) = (f˜(k, t)). (3.61)
∂t
At this point we observe the estimates for any δ > 0:
 T
dt|(Ẽ∆f + ∇Ẽ · ∇f, Ẽ∆f + ∇Ẽ · ∇f )L2 (RN ) (t)|
δ
 T 
≤ dt sup |Ẽ(x)|2 (k 4 f˜(k, t), f˜(k, t))L2 (RN )
δ x∈RN

+ sup (max(∇i Ẽ(x), Ẽ(x)))(k 2 ki f˜(k, t), f˜(k, t))L2 (RN )


x∈RN

˜ ˜
+ sup max(∇i Ẽ(x), ∇j Ẽ(x))(ki kj f (k, t), f (k, t))L2 (RN ) . (3.62)
x∈RN
August 6, 2008 15:46 9in x 6in B-640 ch03

70 Lecture Notes in Applied Differential Equations of Mathematical Physics

By using the explicit representations Eqs. (3.60) and (3.61) and the
simple estimate below for α > 0 with t ∈ [δ, T ) and δ > 0 (with T < ∞)
 ∞
I(t) = d|k||k|α+ cos2 (|k|t)e−Λ|k|
0
 ∞ 
1 ΛU M̄
≤ × dU U α+ e− T ≤ ∈ L1 ([δ, ∞)), (3.63)
tα+1+ 0 t1+α

one can see that by just choosing δ > 0, the existence of the reduced
(T < ∞) wave operators of our pair Eqs. (3.48)–(3.49) is a direct con-
sequence of the application of the Cook–Kuroda–Jauch criterion,6 since
the set of functions polynomial-exponentially bounded is dense on C0 (RN )
(continuous functions vanishing at the infinite), and so, dense in L2 (RN ).7
The complete wave operator (with T → +∞) can be obtained by means
of Zorn theorem applied to the chain of reduced wave operators {W± (T )}
ordered by inclusion, i.e. T1 < T2 ⇒ W± (T1 ) ⊂ W± (T2 ).

3.6. Exponential Stability in Two-Dimensional


Magneto-Elasticity: A Proof on a Dissipative
Medium

One interesting question in the magneto-elasticity property of an isotropic


(uncompressible) medium vibration interacting with an external magnetic
is whether the energy of the total system should decay with an exponential
uniform bound as time reaches infinity.8
We aim in this section to give a rigorous proof of this exponential bound
for a model of imaginary medium electric conductivity as an interesting
example of abstract techniques developed in the previous sections. (See
Appendix A for a complementary analysis on this problem.)
The governing differential equations for the electric–magnetic medium
displacement vector U = (U 1 , U 2 , 0) ≡ U (x, t) depending on the time
variable t ∈ [0, ∞) with x ∈ Ω and the intrinsic two-dimensional magnetic
field h(x, t) = (h1 , h2 , 0) are given by

∂ 2 U (x, t)
= ∆U(x, t) + ([∇ × h](x, t) × B), (3.64)
∂2t
  
∂ h(x, t) ∂U
β = ∆h(x, t) + β ∇ × (x, t) × B . (3.65)
∂t ∂t
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 71

Here, B = (B, 0, 0) is a constant vector external magnetic field, and β


is the (constant) medium electric conductivity.
Additionally, one has the initial conditions

U (0, x) = U0 (x); Ut (0, x) = U1 (x); h(0, x) = h0 (x) (3.66)

and the (physical) Dirichlet-type boundary conditions with n(x) being the
normal of the boundary of the medium Ω9 :

U (x, t)|Ω = 0; (n × (∇ × h))(x, t)|Ω = 0. (3.67)

Let us follow the previous studies of Sec. 3.1 to consider the contraction
semigroup defined by the following essentially self-adjoint operators in the
H-energy space:

ĤΩ = {(U , π, h) ∈ (L2 (Ω))3 | (∇U , ∇U)L2 (Ω)

+ (π, π)L2 (Ω) + (h, h)L2 (Ω) < ∞},

namely,
 
0 i 0
L0 = i∆ 0 0  (3.68)
0 0 −∆

and
 
0 0 0
 
V = 0 0 [∇ × ( )] × B  . (3.69)
0 −(∇ × [( ) × B]) 0

We, thus, have the contractive C0 -semigroup acting on ĤΩ (see


Appendix A for technical proof details):

Tt (U , π, h) = exp(t(+iL0 − V ))(U , π, h)(0). (3.70)

It is worth to call the reader’s attention that we have proved the essential
self-adjointness of the operator V by using explicitly the boundary condi-
tions on the relationships below:

([∇ × (h)] × B) = (B, B(∂1 h2 − ∂2 h1 )), (3.71)


2 2
([∇ × (h)] × B) = (−B∂2 π , B∂1 π , 0). (3.72)
August 6, 2008 15:46 9in x 6in B-640 ch03

72 Lecture Notes in Applied Differential Equations of Mathematical Physics

By standard theorems on contraction semigroup theory,8 one has that


the left-hand side of Eq. (3.70) satisfies the magneto-elasticity equations in

the strong sense with β = i = −1, a pure imaginary electric conductivity
medium, physically meaning that the medium has “electromagnetic dissi-
pation” and in the context that the initial values belong to the subspace
(H 2 (Ω) ∩ H01 (Ω))2 ⊕ (L2 (Ω))2 ⊕ (L2 (Ω))2 ⊂ ĤΩ .
Let us note that Eq. (3.70) still produces a weak integral solution on
the full Hilbert space ĤΩ = (H01 (Ω))2 ⊕ (L2 (Ω))2 ⊕ (L2 (Ω))2 , the result
suitable when one has initial random conditions sampled on L2 (Ω) by the
Minlos theorem (see Appendix B).
Thus, for any (U0 , π0 , h0 ) ∈ ĤΩ we have the following energy estimate:

||(U , ∂t U , h)||H̃Ω

≡ dx(|∇U |2 + |∂t U|2 + |h|2 )(x, t)

≤ ||(exp(t(iL0 − V )))(U (0), ∂t U(0), h(0))||H̃Ω


    n 
it t
≤ S − lim exp L0 exp − V (U0 , U1 , h0 )
n→∞ n n H̃Ω

≤ exp(−tω(V ))||U0 , ∂t U (0), h(0)||H̃Ω , (3.73)

where ω(V ) is the infimum of the spectrum of the self-adjoint operator V


on the space Ĥ.
In order to determine the spectrum of the self-adjoint operator V , which
is a discrete set since Ω is a compact region of R2 , we consider the associated
V -eigenfunctions problem

[∇ × (hn )] × B = λn πn , (3.74)

−∇ × [(πn ) × B] = λn hn , (3.75)

which, by its turn, leads to the usual spectral problem for the Laplacean
with the usual Dirichlet boundary conditions on Ω:
 
λ2n
∆π2n = − π2n , (3.76)
B

π2n (x)|∂Ω = 0. (3.77)


August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 73

As a result of Eqs. (3.75) and (3.76), one finally gets the exponential
bound for the total magneto-elastic energy equation (3.73):

||(U , ∂t U , h)||H̃Ω (t)


' ( )
≤ exp − t Bλ0 (Ω) ||(U0 , ∂t U (0), h(0)||H̃Ω , (3.78)

where λ0 (Ω) = inf{spec(−∆)} em H 2 (Ω) ∩ H01 (Ω).

3.7. An Abstract Semilinear Klein Gordon Wave


Equation — Existence and Uniqueness

Let A be a positive self-adjoint operator as in Sec. 3.2 and (B1 , B2 ) a pair of


bounded operators with domain containing D(A) (the A-operator domain).
We can, thus, associate to these operatorial objects the following semilinear
Klein–Gordon equation with initial dates (zero time field Hilbert-valued
configurations):
∂2U ∂U
= −AU − B1 + B2 U + F (U, t)
∂2t ∂t
(3.79)
U (0) = U0 ∈ D(A)
Ut (0) = V0 ∈ H 0 .

Here, F (U, t) is a Lipschitzian function on the Hilbert space H (0) with


a Lipschitz-bounded being a function L(t) on Lq (R+ ) = {set of mensurable
'* ∞ )1/q
functions f (t) on R+ such that ||f ||Lq = 0 ds|f (s)|q < ∞ and q ≥ 1},
namely,

|F (U1 , t) − F (U2 , t)|H (0) ≤ L(t)|U1 − U2 |H (0) , (3.80a)


F (0, t) ≡ 0. (3.80b)

It is an important problem in the mathematical physics associated with


Eq. (3.79) to find a functional space, where one can show the existence and
uniqueness of the Cauchy problem for the global time interval [0, ∞).11
Let us address this problem through the abstract techniques developed
in the previous section of this work and the Banach fixed-point theorem.
First, let us write Eq. (3.79) in the integral operatorial form
     t  
U t(iLA +V̂ ) U0 (t−s)(iLA +V̂ ) 0
(t) = e + dse 1 , (3.81)
π V0 0 i F (U, s)
August 6, 2008 15:46 9in x 6in B-640 ch03

74 Lecture Notes in Applied Differential Equations of Mathematical Physics

where V̂ is the bounded operator with operatorial norm µ acting on the


two-component space H (1) by means of the rule:

V̂ (U, π) = (V1 U, V2 π), (3.82)

where the (bounded) component operators are written explicitly in terms


of the “damping” operator B1 and “mass” operator B2 , as given below:

V1 = B1 − V2 , (3.83)
2V2 = −B1 ± (B12 − 4B2 ) 1/2
. (3.84)

Let us introduce the following exponential weighted space of H (1) -valued


continuous functions:
+
Ck ([0, ∞), H (1) ) = g : R+ → H (1) , with K > 0
,
(k)
and ||g||H (1) = sup (e−kt ||g(t)||H (1) ) < ∞ . (3.85)
0≤t<∞

We are going to show next that there exists a value of the exponential
weight parameter k defining the above-considered space such that the appli-
cation T̂ written below is a contraction between the space of Hilbert-valued
continuous functions (Eq. (3.85)), i.e.

T̂ : Ck ([0, ∞), H (1) ) → Ck ([0, ∞), H (1) )


 t  
0
Φ = (ϕ1 , ϕ2 ) → dse(t−s)(iLA +V̂ ) 1 . (3.86)
0 i F (ϕ1 (s), s)

In order to show this result, let us point out the usual exponential
bound for the free Klein–Gordon abstract propagator (operator norm) (see
Eq. (A.23)):

|| exp[t(iLA + V̂ )]||op ≤ || exp tV̂ ||op


≤ exp(t||V̂ ||op ) ≡ exp(tµ). (3.87)

As a consequence of this bound Eq. (3.87), we have the contraction


estimate for Φ, Φ elements of Ck ([0, ∞), H (1) ):
 t 
||T̂ Φ − T̂ Φ ||Ck ([0,∞),H (1) ) ≤ sup dse−k(t−s) eµ(t−s) L(s)
0≤t<∞ 0

× ||Φ − Φ ||Ck ([0,∞),H (1) ) . (3.88)


August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 75

At this point, it is a straightforward step to obtain the general expression


for the contraction coefficient on Eq. (3.88) (where p1 + 1q = 1):
 t 
sup dse−K(t−s) eµ(t−s) L(s)
0≤t<∞ 0
 1/p 
t
−p(k−µ)(t−s)
≤ sup dse ||L||Lq (R+ )
0≤t<∞ 0
 (1/p)
1
= · ||L||Lq (R+ ) ≡ C(k, µ). (3.89)
p(k − µ)

So, T has a fixed point on Ck ([0, ∞), H (1) ) if one makes the choice of
the value k such that (C(k, µ)) < 1, namely,
 
1
K> (||L||Lq (R+ ) )p + µ . (3.90)
p
A very important result obtained as a consequence of the above “fixed-
point theorem” analysis is that the sequence of iteratives
 
(n) (n) U0
U (t) = (T̂ ) (3.91)
V0

converges in Ck ([0, ∞), H (1) ) to the solution of Eq. (3.81) with the error
estimate

sup (e−kt ||U (n) − U ||H (1) (t))


0≤t<∞
   
(C(k, µ))n−1 U0
≤ sup exp t(iLA + V̂ ) . (3.92)
1 − C(k, µ) 0≤t<∞ V0 H (1)

References

1. M. Reed and B. Simon, Methods of Modern Mathematical Physics, Vol. 2,


(Academic Press, 1975); Khandekar and Lawande, Path Integral Methods and
their Applications (World Scientific, Singapore, 1993).
2. L. C. L. Botelho and R. Vilhena, Phys. Rev. 49E(2), R1003–R1004 (1994); L.
C. L. Botelho, Mod. Phys. Lett. 13B, 363–370 (1999); L. C. L. Botelho, Mod.
Phys. Lett. 14B, 73–78 (2000); L. C. L. Botelho, J. Phys. A: Math. Gen. 34,
L131–L137 (2001); L. C. L. Botelho, Mod. Phys. Lett. 16B, 793–806 (2002).
3. S. G. Mikhlin, Mathematical Physics, An Advanced Course, Series in Applied
Mathematics and Mechanics (North-Holland, 1970).
August 6, 2008 15:46 9in x 6in B-640 ch03

76 Lecture Notes in Applied Differential Equations of Mathematical Physics

4. A. P. S. Selvadurai, Partial Differential Equations in Mechanics 2 (Springer-


Verlag, 2000).
5. R. Dautray and J.-L. Lions, Analyse Mathématique et calcul numérique,
Vol. 7 (Massun, Paris, 1988).
6. S. T. Kuroda, Il Nuovo Cimento 12, 431–454 (1959).
7. V. Enss, Commun. Math. Phys. 61, 285–291 (1978).
8. G. P. Menzala and E. Zuazua, Asymptotic Anal. 18, 349–367 (1998).
9. J. E. M. Riviera and R. Racue, IMA J. Appl. Math. 66, 269–283 (2001).
10. L. Schwartz, Random Measures on Arbitrary Topological Spaces and Cylin-
drical Measures (Tata Institute, Oxford University Press, 1973).
11. T. Cazennave and A. Haraux, An Introduction to Semi-Linear Evolution
Equations, Vol. 13 (Clarendon Press, Oxford, 1998).

Appendix A. Exponential Stability in Two-Dimensional


Magneto-Elastic: Another Proof

In this complementary technical appendix to Sec. 3.6 on the exponential


decay of the magneto-elastic energy associated with the (imaginary electric
medium conductivity) magneto-elastic wave
   
U U
 π  (t) = exp{it(+L0 + iV )}  π  (0) (A.1)
h h
with the essential self-adjointness operators on the Hilbert energy space
H̃ 1 (Ω)
 
0 +i 0
L0 = +i∆ 0 0 (A.2)
0 0 ∆
and
 
0 0 0
 
iV = 0 0 i[∇ × ( ) × B  , (A.3)
0 i∇ × [( ) × B] 0

we have used the fact that the operator −V + iL0 is the generator of a
contraction semigroup on this energy Hilbert space H̃ 1 (Ω) = (H 1 (Ω))3 ⊕
(L2 (Ω))3 ⊕ (L2 (Ω))3 in order to write Eq. (A.1) in a mathematically
rigorous way.
Let us give a proof of this mathematical result by means of a direct
application of the Hille–Yoshida theorem9 to the operator (−V + iL0 ).
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 77

First, the domain of (−V + iL0 ) is everywhere dense on H̃ 1 (Ω) as a


consequence of the self-adjointness of the operators V and L0 on H̃ 1 (Ω).
Second, the existence of a solution for the elliptic problem

(−V + iL0 )x = αy (A.4)

for x ∈ D(L0 ) ⊂ D(V ) and every y ∈ H̃ 2 (Ω) with α > 0 is a standard


result even if Ω has a nontrivial topology (holes inside!), i.e. Ω is a multiple-
connected planar region.9
The unicity of the solution x is a straightforward consequence of the
fact that the spectrum {λn } of the self-adjoint operator V coincides with
the positive Laplacean ωn (−∆), i.e. λ2n /B = ωn (−∆). As a consequence,
the solution of the equation

(−V + iL0 )x = 0 (A.5)

leads to the relation below due to the self-adjointness of the operators V


and L0 on H̃ 1 (Ω):

V x, xH̃ 1 (Ω) = V x, xH̃ 1 (Ω) = iL0 x, xH̃ 1 (Ω) = iL0 x, xH̃ 1 (Ω) , (A.6)

from which we conclude that

V x, xH̃ 1 (Ω) = 0, (A.7)


L0 x, xH̃ 1 (Ω) = 0, (A.8)

or equivalently

x ∈ Ker{V } ∩ Ker{L0 }, (A.9)

which is zero if Ω is a simply connected planar domain as supposed in


the text.
Finally, for any x > 0, we have that
1
||(α1 − (−V + iL0 ))−1 ||H̃ 1 (Ω) ≤ , (A.10)
α
since for every z ∈ Dom(L0 ) ⊂ Dom(V ),

α2 ||z||2H̃ 1 (Ω) ≤ ||((α1 − (−V + iL0 )))z||2H̃ 1 (Ω)

= α2 ||z||2H̃ 1 (Ω) + ||L0 z||2H̃ 1 (Ω)

+ ||V z||2H̃ 1 (Ω) + 2α(z, V z)H̃ 1 (Ω) . (A.11)


August 6, 2008 15:46 9in x 6in B-640 ch03

78 Lecture Notes in Applied Differential Equations of Mathematical Physics

Note that we have used the fact that V is a positive operator on H̃ 1 (Ω).
As a consequence of the above-exposed results we have that −V + iL0
is a generator of a contractive semigroup on the space (1 − PKer(V )∩Ker(L0 ) )
H̃ 1 (Ω), where PKer(V )∩Ker(L0 ) is the orthogonal projection on the Kernel
subspaces of the self-adjoint operators V and L0 .
As a final remark to be made in this Appendix A, let us call the physicist-
oriented reader’s attention to the following (somewhat formal) abstract
Lemmas, alternatives to the Banach space methods used in Ref. 9.

Lemma A.1. Suppose that the matrix-valued operator with self-adjoint


operators entries on suitable Hilbert spaces H1 , H2
 
0 A 0

L0 = B 0 0. (A.12)
0 0 C

Then, it will be a self-adjoint on the Hilbert space H̃ = H1 ⊕ H1 ⊕ H2 with


an inner product given by (D is a positive self-adjoint operator on H1 )

(U, π, h); (U  , π  , h )H̃ = (D∗ DU, U  )H1 + (π, π)H1 + (h, h)H2 , (A.13)

if we have the constraints below for the operators on the Hilbert component
spaces:

C = C∗ on H2
∗ ∗
A D D=B on H1 (A.14)
Range(A) ⊂ Dom(D).

Proof. Let us consider the inner product on H̃


-   .
U
    
L0 π ; (U , π , h ) = (DAπ, DU  )H1 + (BU, π  )H1 + (Ch, h )H2
h H̃
(A.15)

and

(U, π, h); L0 (U  , π  , h )H̃ = (DU, DAπ  )H1 + (π, BU  )H1 + (h, Ch )H2 .
(A.16)
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 79

On the basis of Eqs. (A.15) and (A.16) one can see that L0 is a closed
symmetric operator. Besides, if there is a (U0 , π0 , h0 ) ∈ H̃ such that for
every (U, π, h) we have the orthogonality relation

(U0 , π0 , h0 ); (i + L0 )(U, π, h)H̃ = 0, (A.17)

then

(U0 , π0 , h0 ); (Aπ, +iU, BU + iπ, Ch + ih)H̃ = 0. (A.18)

As a result

(D∗ DU0 , (Aπ + iU ))H1 = 0 (A.19)


(π0 , BU + iπ)H1 = (h0 , Ch + ih)H2 = 0. (A.20)

We thus have by self-adjointness of the operators A, B, and C that U0 =


π0 = h0 = 0. As a consequence, the deficiency indices of L0 vanish what
formally concludes the self-adjointness of L0 on vectors of H̃1 such that
(the domain of L0 since L0 is a closed symmetric operator):

L0 (U, π, h), (U, π, h)H̃ < ∞. (A.21)

Lemma A.2. Let U be the matrix-valued operator acting on H̃ given by


 
0 0 0
V = 0 0 V1  . (A.22)
0 V2 0

Then V is self-adjoint on H̃ if and only if on H2 we have the constraint


relationship below between the adjoints of the closable symmetric operators
V1 and V2

V1∗ = V2 . (A.23)

Proof. One can use similar arguments of the proof of Lemma A.1 to arrive
at such a result.
Let us now take into account the case of the spectrum of L0 contained
on the negative real line (−∞, 0] and that one associated with V is by
its turn, contained only on the upper-bounded negative infinite interval
(−∞, −c] (with c > 0); then we have by a straightforward application of the
August 6, 2008 15:46 9in x 6in B-640 ch03

80 Lecture Notes in Applied Differential Equations of Mathematical Physics

Trotter–Kato formulae for t ∈ [0, ∞) the following estimate on the norm of


semigroup evolution, operator-generated by L0 + V0 :
t t
||et(L0 +V ) || ≤ lim ||e n L0 e n V ||n
n
t
≤ lim ||e n V ||n ≤ ||etV || ≤ e−tc . (A.24)
n

In the following magneto-elastic wave problem with real conductivity


β > 0,
    
U 0 −1 0 U
∂t  π  (t) = −∆ 0 0   π  (t)
1
h 0 0 β∆ h
  
0 0 0 U
 
× 0 0 −B × [∇ × ( )]  π  (t).
0 −∇ × [( ) × B] 0 h
(A.25)

One can see that it satisfies the conditions of the above lemmas with the
operator D = ∇ and with the Hilbert space

H̃ = (L2 (Ω))3 ⊕ (L2 (Ω))3 ⊕ (L2 (Ω))3

and c = inf spec(−∆) and H 2 (Ω) ∩ H01 (Ω) (see Sec. 3.5), and thus, leading
to the expected total energy exponential decay showed in Sec. 3.5. 

Appendix B. Probability Theory in Terms of Functional


Integrals and the Minlos Theorem

In this complementary appendix, we discuss briefly for the mathematics-


oriented reader the mathematical basis and concepts of functional integrals
formulation for stochastic processes and the important Minlos theorem on
Hilbert space support of probabilistic measures.
The first basic notion of Kolmogorov’s probability theory framework is
to postulate a convenient topological space (Polish spaces) Ω formed by
the phenomenal random events. The chosen topology of Ω should possess a
rich set of nontrivial compact subsets. After that, we consider the σ-algebra
generated by all open sets of this — the so-called topological sample space,
denoted by FΩ . Thirdly, one introduces a regular measure dµ on this set
algebra FΩ , assigning values in [0, 1] to each member of FΩ .
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 81

The (abstract) triplet (Ω, FΩ , dµ) is thus called a probability space in


the Kolmogorov–Schwartz scheme.10
Let {Xα }α∈Λ be a set of measurable real functions on Ω, which may be
taken as continuous injective functions of compact support on Ω (without
loss of generality).
It is a standard result that one can “immerse” (represent) the
abstract space Ω on the “concrete” infinite product compact space R∞ =
/
α∈Λ (Ṙα ), where Ṙ is a compact copy of the real line. In order to achieve
such a result we consider the following injection of Ω in R∞ defined by the
family {Xα }α∈Λ :

I : Ω → R∞
(B.1)
ω → {Xα (ω)}α∈Λ .

A new σ-algebra of events on Ω can be induced on Ω and affiliated to


the family {Xα }α∈Λ .
It is the σ-algebra generated by all the “finite-dimensional cylinder”
subsets, explicitly given in its generic formulas by

CΛfin = ω ∈ Ω | Λfin is a finite subset of Λ

$
and [(Xα )α∈Λfin (Ω)] ⊂ [aα , bα ] .
α∈Λfin

The measure restriction of the initial measure µ on this cylinder set of


Ω will be still denoted by µ on what follows. Now it is a basic theorem of
/
Probability theory that µ induces a measure ν (∞) on α∈Λ (Ṙ)α , which can
be identified with the space of all functions of the index set Λ to the real
compactified line Ṙ; F (Λ, R) with the topology of Pontual convergence.
At this point, it is straightforward to see that the average of any Borelian
(mensurable) function G(ω) (a random variable) is given by the following
functional integral on the functional space F (Λ, R):
 
dµ(ω)G(ω) = dν (∞) (f ) · G(I −1 (f )). (B.2)
Ω F (Λ,R)

It is worth to recall that the real support of the measure dν (∞) (f ) in


most practical cases is not the whole space F (Λ, R) but only a functional
subspace of F (Λ, R). For instance, in most practical cases, Λ is the index set
/
of an algebraic vectorial base of a vector space E and α∈Λ (R)α (without
August 6, 2008 15:46 9in x 6in B-640 ch03

82 Lecture Notes in Applied Differential Equations of Mathematical Physics

the compactification) is the space of all sequences with only a finite number
of nonzero entries as it is necessary to consider in the case of the famous
extension theorem of Kolmogorov. It turns out that one can consider the
support of the probability measure as the set of all linear functionals on
E, the so-called algebraic dual of E. This result is a direct consequence
of considering the set of random variables given by the family {e∗α }α∈Λ .
For applications one is naturally lead to consider the probabilistic object
called characteristic functional associated with the measure dν (∞) (f ) when
F (Λ, R) is identified with the vector space of all algebraic linear functionals
E alg of a given vector space E as pointed out above, namely,

Z[j] = dν (∞) (f ) exp(if (j)), (B.3)
E alg
0
where j are elements of E written as j = α∈Λfin xα eα with {eα }α∈Λ
denoting a given Hammel (vectorial) basis of E.
One can show that if there is a subspace H of E with a norm || ||H
coming from an inner product (, )H such that

dν (∞) (f ) · ||f ||2H < ∞ (B.4)
E∩H

one can show the support of the measure of exactly the Hilbert space
(H, (, )H ) — the famous Minlos theorem.
Another important theorem (the famous Wienner theorem) is that when
we have the following condition for the family of random variables with the
index set Λ being a real interval of the form Λ = [a, b]:

dν (∞) (f )|Xt+h (f ) − Xt (f )|p ≤ c|h|1+ε , (B.5)
E

for p, ε, c positive constants. We have that it is possible to choose as the


support of the functional measure dν (∞) (·) the space of the real continuous
function on Λ the well-known famous Brownian path space
 
dµ(ω)G(ω) = dν (∞) (f (x))G̃(f (x)). (B.6)
Ω C([a,b],R)

For rigorous proofs and precise formulations of the above-sketched deep


mathematical results, the interested reader should consult the basic math-
ematical book by Schwartz.10
August 6, 2008 15:46 9in x 6in B-640 ch03

Hilbert Space Integration Method for the Wave Equation 83

Let us finally give a concrete example of such results considering a gen-


erating functional Z[J(x)] of the exponential quadratic functional form on
L2 (RN )
   
1
Z[J(x)] = exp − N
d x N
d yj(x)K(x, y)j(y) (B.7)
2 RN RN

associated with a Gaussian random field with two-point correlation function


given by the kernel of Eq. (B.7),

E{v(x)v(y)} = K(x, y). (B.8)

In the*case of the above-written kernel being a class-trace operator on


L2 (RN ) ( RN dN xK(x, x) < ∞), one can represent Eq. (B.7) by means of
a Gaussian measure with the support of the Hilbert space L2 (RN ). This
result was called to the reader’s attention for the bulk of this paper.
August 6, 2008 15:46 9in x 6in B-640 ch04

Chapter 4

Nonlinear Diffusion and Wave-Damped Propagation:


Weak Solutions and Statistical Turbulence Behavior∗

4.1. Introduction

One of the most important problems in the mathematical physics of the


nonlinear diffusion and wave-damped propagation is to establish the exis-
tence and uniqueness of weak solutions in some convenient Banach spaces
for the associated nonlinear evolution equation (see Refs. 1, 3–5). Another
important class of initial-value problems for nonlinear diffusion-damping
came from statistical turbulence as modeled by nonlinear diffusion or
damped hyperbolic partial differential equations with random initial con-
ditions associated to the Gaussian processes sampled in certain Hilbert
spaces2,5 and simulating the turbulence physical phenomena.6
The purpose of this chapter is to contribute to such mathematical
physics studies by using functional spaces compacity arguments in order to
produce proofs for the existence and uniqueness of weak solutions for a class
of special nonlinear diffusion equations on a “smooth” C ∞ domain with
compact closure Ω ⊂ R3 with Dirichlet boundary conditions and initial
conditions belonging to the space L2 (Ω). This study is presented in Sec. 4.2.
In Sec. 4.3, we present a similar analysis for the wave equation with a
nonlinear damping, analogous in its form to the nonlinear term studied in
Sec. 4.2 for the diffusion equation.
In Sec. 4.4, we present in some details the solution to the asso-
ciated problem for random initial conditions in terms of our proposed
cylindrical functional measures representations previously proposed in the
literature6 and defined by a cylindrical measure on the Banach space
L∞ ((0, T ), L2 (Ω)), a new result on the subject.
Finally, in Sec. 4.5, we present a complementary semigroup analysis on
the very important problem of anomalous diffusion for random conditions
in the path-integral framework.

∗ Author’s original results.

84
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 85

4.2. The Theorem for Parabolic Nonlinear Diffusion

Let us consider the following nonlinear diffusion equation in some strip


Ω × [0, T ] ⊂ R4 :

∂U (x, t)
= (−AU )(x, t) + ∆(F (U (x, t))) + f (x, t) (4.1)
∂t

with the initial and Dirichlet boundary conditions

U (x, 0) = g(x) ∈ L2 (Ω) ⊂ L1 (Ω), (4.2)



U (x, t)∂Ω ≡ 0, (4.3)

where A denotes a second-order self-adjoint uniform elliptic positive dif-


ferential operator, F (x) is a real function continuously differentiable on
the extended real line (−∞, +∞) with its derivative F  (x) strictly positive
on (−∞, +∞). The external source f (x, t) is supposed to be on the space
L∞ ([0, T ] × L2 (Ω)) = L∞ ([0, T ], L2 (Ω)).
We now state the existence and uniqueness theorem of ours.

Theorem 4.1. In the initial-value nonlinear diffusion equations (4.1)–


(4.3), for any g(x) ∈ L2 (Ω), and A = −∆ (minus Laplacean) there exists a
unique solution Ū(x, t) on L∞ ([0, T ] × L2 (Ω)) satisfying this problem in a
certain weak sense with a test functional space given by C0∞ ([0, T ], H 2 (Ω) ∩
H01 (Ω)).

Proof. The existence proof will be given for a general A as stated below
Eq. (4.3). Let {ϕi (x)} be the spectral eigenfunctions associated with the
operator A. Note that each ϕi (x) ∈ H 2 (Ω) ∩ H01 (Ω) and this set is complete
on L2 (Ω) as a result of the Gelfand generalized spectral theorem applied
for A1 since H 2 (Ω) ∩ H01 (Ω) is compactly immersed in L2 (Ω).
Consider the following system of nonlinear ordinary differential equa-
tions associated with Eqs. (4.1)–(4.3) — the well-known Galerkin system:1
 
∂U (n) (x, t)
, ϕj (x) + (AU (n) (x, t), ϕj (x))L2 (Ω)
∂t 2
L (Ω)

= (∇ · [(F  (U (n) (x, t)))∇U (n) (x, t)], ϕj (x))L2 (Ω)

+ (f (n) (x, t), ϕj (x))L2 (Ω) (4.4)


August 6, 2008 15:46 9in x 6in B-640 ch04

86 Lecture Notes in Applied Differential Equations of Mathematical Physics

subject to the initial condition


n
U (n) (x, 0) = (g, ϕi )L2 (Ω) · ϕi (x), (4.5)
i=1

where the finite-dimensional Galerkin approximants are given exactly in


terms of the spectral basis {ϕi (x)} as


n
(n)
U (n) (x, t) = Ui (t)ϕi (x), (4.6)
i=1


n
f (n) (x, t) = (f (x, t), ϕi (x))L2 (Ω) ϕi (x), (4.7)
i=1

and (, )L2 (Ω) is the usual inner product on L2 (Ω). Note that

 
n

U (n) (x, t)∂Ω = Ui (t)(ϕi (x))∂Ω = 0.
(n)
(4.8)
i=1

Let us introduce the short notations

U (n) (x, t) ≡ U (n) , (4.9)


(n)
(g, ϕi )L2 (Ω) ≡ gi , (4.10)
(n)
(f (x, t), ϕi (x))L2 (Ω) = fi . (4.11)

By multiplying the Galerkin system equation (4.4) by U (n) as usual3 we


get the a priori identity

1 d
U (n) 2L2 (Ω) + (AU (n) , U (n) )L2 (Ω)
2 dt

(n)
+ dxF  (U (n) )(∇U (n) ∇U ) = (f, U (n) )L2 (Ω) . (4.12)

By a direct application of the Gärding–Poincaré inequality with the


quadratic form associated with the operator A, one has that there is a
positive constant γ(Ω) such that3

(AU (n) , U (n) )L2 (Ω) ≥ γ(Ω)U (n) 2L2 (Ω) . (4.13)
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 87

This yields the following estimate for any integer positive p:

1 d
(U (n) 2L2 (Ω) ) + γ(Ω)U (n) 2L2 (Ω) + (F  (U (n) ))(1/2) ∇U (n) 2L2 (Ω)
2 dt
 
1 2 1 (n) 2
≤ pf (x, t)L2 (Ω) + U L2 (Ω) . (4.14)
2 p

At this point we observe that the square root of the function F  (x)
makes sense since it is always positive: F  (x) > 0.
1
By choosing Eq. (4.14) p big enough such that ap ≡ γ(Ω) − 2p > 0, we
have that
1 d 1
(U (n) 2L2 (Ω) ) + ap U (n) 2L2 (Ω) ≤ pf 2L2(Ω) (t), (4.15)
2 dt 2
or by the Gronwall lemma, that there is a uniform constant M [even for
T = +∞ if f ∈ L2 ([0, ∞], L2 (Ω))] such that

sup (U (n) (t)2L2 (Ω) )


0≤t≤T
  T
−ap T
≤ sup e dsf 2L2(Ω) (s)e+ap s + g2L2(Ω) = M.
0≤t≤T 0

(4.16)

Note that the Galerkin system of (nonlinear) ordinary differential equa-


(n)
tions for {Ui (t)} has a unique global solution on the interval [0, T ] as a
(n)
consequence of Eq. (4.16) — which means that i=1 (Ui (t))2 ≤ M —
n

and the Peano–Caratheodory theorem since f (t, ·) ∈ L∞ ([0, T ]).


As a consequence of the Banach–Alaoglu theorem applied to the
bounded set {U (n) } on L∞ ([0, T ], L2 (Ω)), there is a subsequence weak-star
convergent to an element (function) Ū (t, x) ∈ L∞ ([0, T ], L2 (Ω)), and this
subsequence will still be denoted by {U (n) (x, t)} in the analysis that follows.
An important remark is at this point. Since F (x) is a Lipschitzian
function on the closed interval where it is defined we have that the set of
functions {F (U (n) )} is a bounded set on L∞ ([0, T ], L2(Ω)) if the set {U (n) }
has this property.
As a consequence of this remark and of a priori inequalities given in
detail in Appendix A, we can apply the famous Aubin–Lion theorem1 to
insure the strong convergence of the sequence {U (n) (x, t)} to the function
Ū (x, t). As a straightforward consequence, our special nonlinearity on
Eq. (4.1), namely F (x) is a Lipschitzian function; we obtain, thus, the
August 6, 2008 15:46 9in x 6in B-640 ch04

88 Lecture Notes in Applied Differential Equations of Mathematical Physics

strong convergence of the sequence {F (U (n) (x, t)} to the L2 (Ω)-function


F (Ū (x, t)).
As a consequence of the above-made comments and by noting that
L2 (Ω) is continuously immersed in H −2 (Ω) and the weak-star convergence
definition, we have the weak equation below in the test space function of
v(x, t) ∈ C0∞ ([0, T ], H 2 (Ω) ∩ H01 (Ω)):
 T  
(n) dv
lim dt −U , + (AU (n) , v)L2 (Ω) − (F (U (n) ), ∆v)L2 (Ω)
n→∞ 0 dt L2 (Ω)
 T
= lim dt(f (n) , v)L2 (Ω) , (4.17)
n→∞ 0

or by passing the weak-star limit


 T  
dv(x, t)
dt −Ū (x, t), + (Ū (x, t), (Av)(x, t))L2 (Ω)
0 dt L2 (Ω)
  T
− (F (Ū (x, t)), ∆v(x, t))L2 (Ω) = dt(f (x, t), v(x, t))L2 (Ω) . (4.18)
0

This concludes the existence proof of Theorem 4.1, since v(0, x) =


v(T, x) = 0. 
Let us apply this to a concrete problem of random exponential nonlinear
diffusion on a cube [0, L]3 ⊂ R3 , mainly for the explanation of the above-
written abstract results:
  
∂U (x, t) k0 k0
=∇ + e−U(x,t) ∇U (x, t) + ∆U (x, t) (4.19)
∂t 2 2
with

U (x, 0) = g(x) ∈ L2 (Ω), (4.20)



U (x, t)x∈[0,L]3 ≡ 0, (4.21)

where g(x) are the samples of a stochastic Gaussian process on Ω with


correlation function defining an operator of Trace class on L2 (Ω), namely,

E{g(x)g(y)} = K(x, y) ∈ (L2 (Ω)) (4.22)
1

or in terms of the spectral set associated with the Laplacian, i.e.



n
(n)
g(x) = lim gi ϕi (x) (4.23)
n→∞
i=0
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 89

with

E{gi gk } = Kik = dxdyK(x, y)ϕi (x)ϕk (y). (4.24)
Ω×Ω

As a consequence of Theorem 4.1, an explicit solution of Eq. (4.19) must


be taken in L∞ ([0, ∞), L2 (Ω)) in the weak sense of Eq. (4.18) and given
thus by the Trigonometrical series sequence of functions:

U (x, t, [g])
 
n1
,n2 ,n3      
iπ jπ kπ
= n lim di,k,j (t, [g ]) sin x sin y sin z ,
1 →∞
n2 →∞
 L L L 
i,j,k=1
n3 →∞
(4.25)

where the set of absolutely continuous function {di,j,k (t, [g0 ])} on [0, T ] sat-
isfies the Galerkin set of ordinary differential equations with “random”
Gaussian initial conditions
      
iπ jπ kπ
di,j,k (0, [g ]) = g(x), sin x sin y sin z = g .
L L L L2 (Ω)
(4.26)
Let us now comment on the uniqueness of solution problem for the
nonlinear initial-value diffusion equations (4.1)–(4.3).
In the case under study, the uniqueness comes from the following tech-
nical lemma.

Lemma 4.1. If Ū(1) and Ū(2) in L∞ ([0, T ] × L2 (Ω)) are two functions
satisfying the weak relationship, for any v ∈ C0∞ ([0, T ], H 2 (Ω) ∩ H01 (Ω))
 T  
∂v
dt Ū(1) − Ū(2) , − + (Ū(1) − Ū(2) , Av)L2 (Ω)
0 ∂t L2 (Ω)

− (F (Ū(1) ) − F (Ū(2) ); ∆v)L2 (Ω) ≡ 0, (4.27a)

then Ū(1) = Ū(2) a.e. in L∞ ([0, T ] × L2 (Ω)).5

Let us notice that Eq. (4.27a) is a weak statement on the assumption


of vanishing initial values and it is a consequence of
 
lim supp ess d3 x|Ū(1) (x, t) − Ū(2) (x, t)|2 = 0. (4.27b)
t→0+ Ω
August 6, 2008 15:46 9in x 6in B-640 ch04

90 Lecture Notes in Applied Differential Equations of Mathematical Physics

The proof of Eq. (4.27) is a direct consequence of the fact that F (x) is
a Lipschitz function satisfying an inequality of the form
 
|F (Ū(1) ) − F (Ū(2) )|(x, t) ≤ sup F  (x) |Ū(1) − Ū(2) |(x, t) (4.27c)
−∞<x<+∞

(for a technical proof see Ref. 5). However in the case of A = −∆ (minus
Laplacian), as stated in our theorem, the identity Eq. (4.27a) means that for
the functions of the form v(x, t) = e−λt vλ (x) ∈ C0∞ ([0, T ], H 2 (Ω) ∩ H01 (Ω))
with ∆vλ (x) = −λvλ (x), the following identity must hold true:
 T
dt(F (Ū(1) ) − F (Ū(2) ), vλ )L2 (Ω) ≡ 0, (4.28)
0

which means that F (Ū(1) ) = F (Ū(2) ) a.e., and thus, Ū(1) = Ū(2) a.e., since
F (x) satisfies the lower-bound estimate
 
|F (x) − F (y)| ≥ inf |F  (x)| |x − y|.
−∞<x<∞

This result produces a rigorous uniqueness result for A = −∆.


Another important example of nonlinear diffusion equation somewhat
related to Eq. (4.1) is the density equation for a gas in a porous medium
with physical saturation. In this case, the law of conservation of mass
∂ρ
∇ · (ρV
)+ = 0, (4.29)
∂t

where V is the velocity of gas, ρ is the density of the gas, and P is the
pressure, together with Darcy’s law and the isothermic equation of state,
namely,

P = cργ (4.30)
 = −k∇P
V

leads to the following nonlinear diffusion equation:



 ∂ρ(x, t) = −γck∇(ργ ∇ρ)(x, t)
∂t (4.31)

ρ(x, 0) = f (x) ∈ L2 (Ω).

Since it is showed that for finite volume open, convex with C ∞ -boundary
domain Ω as in our study, the function ρ(x, t) which satisfies Eq. (4.31)
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 91

should be a bounded function for any t ≥ t0 > 0 (with t0 fixed),5 it is clear


that one can replace the above-written mathematical gas porous medium
equation, by a more physical equation of the form, taking into account the
physical phenomena of saturation,

 ∂ρ(x, t) = −γc∇(F (ρ)∇ρ)(x, t)
(ε)
∂t (4.32)
ρ(x, 0) = f (x) ∈ L2 (Ω)

with F(ε) (x) denoting a differentiable regularizing function of the function


xγ on the interval [0, M ], where M is the global upper-bound of the function
ρ(x, t).
One can thus apply our Theorem 4.1 to obtain an explicit set of functions
(n)
{ρ(ε) (x, t)} converging in the weak-star topology of L∞ ([t0 , ∞), L2 (Ω)) to
the solution ρ(ε) (x, t) of the problem equation (4.32). It is a reasonable
conjecture that in the nonsaturation case, one should take the ε → 0 limit
on Eq. (4.32), namely, ρ(x, t) = lim ρ(ε) (x, t), since F(ε) (x) converge to xγ
ε→0
in the C ∞ -topology of C([0, M ]). As a consequence, it is expected that a
solution for the physical equation (4.31) should be produced. A full technical
description of this limiting process will appear in an extended mathematics-
oriented paper.

4.3. The Hyperbolic Nonlinear Damping

We aim in this third section state a theorem analogous to Theorem 4.1


of Sec. 4.2 on diffusion, but now for the important case of existence of
nonlinear damping on the hyperbolic initial-value problem on Ω × [0, T ]
with imposed Dirichlet boundary conditions, including the new global case
of T = +∞ (see Sec. 4.2) and a damping positive constant ν,
  
∂ 2 U (x, t) ∂U (x, t) ∂U
+ (AU )(x, t) = −ν +∆ F (x, t) + f (x, t).
∂t2 ∂t ∂t
(4.33)
2
The L (Ω)-initial conditions are given by

U (x, 0) = g(x) ∈ L2 (Ω), (4.34a)

Ut (x, 0) = h(x) ∈ L2 (Ω), (4.34b)

U (x, t)|∂Ω = 0, (4.34c)


August 6, 2008 15:46 9in x 6in B-640 ch04

92 Lecture Notes in Applied Differential Equations of Mathematical Physics

and now the nonhomogenous term f (x, t) is considered to be a function


belonging to the functional space
L2 ([0, T ] × Ω) ∩ L∞ ([0, T ], L2 (Ω)). (4.35)
We thus have the following theorem of existence (without uniqueness).

Theorem 4.2. There exists a solution Ū (x, t) on L∞ ([0, T ] × L2 )Ω)) for


Eqs. (4.33)–(4.35) in the weak sense with a test functional space as given
by C0∞ ([0, T ], H 2 (Ω) ∩ H01 (Ω)).
In order to arrive at such a theorem, let us consider the analogous of
the estimated Eq. (4.12) for Eq. (4.33) with U̇ (n) ≡ ∂t

(U (n) (x, t)), namely,
1 d
U̇ (n) 2L2 (Ω) + νU̇ (n) 2L2 (Ω)
2 dt
+ (AU̇ (n) , U (n) )L2 (Ω) + (F  (U̇ (n) ))(1/2) ∇U̇ (n) 2L2 (Ω)
= (f, U̇ (n) )L2 (Ω) , (4.36)
or equivalently
1 d  (n) 2 
U̇ L2 (Ω) + (AU (n) , U (n) )L2 (Ω)
2 dt
 
+ ν U̇ (n) 2L2 (Ω) + (AU (n) , U (n) )L2 (Ω)

+ (F  (U̇ (n) ))(1/2) ∇U̇ (n) 2L2 (Ω) ≤ ν(AU (n) , U (n) )
 
1 1
+ pf 2L2(Ω) + U̇ (n) 2L2 (Ω) . (4.37)
2 p
1
If one chooses here the integer p such that 2p = ν, we have the simple
bound
1 d  (n) 2  1
U̇ L2 (Ω) + (AU (n) , U (n) )L2 (Ω) ≤ f 2L2(Ω) . (4.38)
2 dt 4ν
As a consequence of Eq. (4.38), there is a constant M such that the
uniform bounds hold true (even if for the case T = +∞ for the case of
f ∈ L2 ([0, ∞), L2 (Ω))):
sup ess0≤t≤T U̇ (n) 2L2 (Ω) ≤ M, (4.39)

sup ess0≤t≤T U (n) 2L2 (Ω) ≤ M. (4.40)


As a consequence of the bounds Eqs. (4.39) and (4.40), there are two
functions Ū(x, t) and P̄ (x, t) such that we have the weak-star convergence
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 93

on L∞ ([0, T ] × L2 (Ω)):
weak-star lim (U (n) (x, t)) = Ū (x, t), (4.41)
n→∞
 
∂ (n)
weak-star lim U (x, t) = P̄ (x, t). (4.42)
n→∞ ∂t
We thus have that the relationship below hold true for any test function
v(x, t) ∈ Cc∞ ([0, T ] × H 2 (Ω) ∩ H01 (Ω)) obviously satisfying the relations
v(x, 0) = v(x, T ) = ∆v(x, 0) = ∆v(x, T ) = vt (x, 0) = vt (x, T ) = vtt (x, 0) =
vtt (x, T ) ≡ 0 as a consequence of applying the Aubin–Lion theorem in a
similar way as it was used on Eq. (4.17):
 T  
d2 v
dt Ū , 2 + (Ū , Av)L2 (Ω)
0 d t L2 (Ω)
  
dv
+ ν Ū , − − (F (P̄ ), ∆v)L2 (Ω)
dt L2 (Ω)
 T
= dt(f, v)L2 (Ω) , (4.43)
0

with the initial conditions


Ū (x, 0) = g(x) ∈ L2 (Ω), (4.44)

P̄ (x, 0) = h(x) ∈ L2 (Ω). (4.45)


Let us now show that
 t
Ū (x, t) = dsP̄ (x, s). (4.46)
0

First, let us remark that integrating on the interval 0 ≤ t ≤ T , in the


relationship Eq. (4.36), one obtains the following estimate:
1 n 
U̇ (t)2L2 (Ω) − U̇ n (0)2L2 (Ω)
2
  t 
1
+ ν dsU̇n L2 (Ω) (s) + (AU (n) (t), U (n) (t))L2 (Ω)
2
0 2
 t
1
− (AU (n) (0), U (n) (0)) + ds(F  (U̇ (n) ))(1/2) ∇U̇ (n) 2L2 (Ω)
2 0
  t
p t 1
≤ dsf 2L2(Ω) +  U̇n (s)2 ds. (4.47)
2 0 2p 0
August 6, 2008 15:46 9in x 6in B-640 ch04

94 Lecture Notes in Applied Differential Equations of Mathematical Physics

Since the operator A satisfies the Gärding–Poincaré inequality on L2 (Ω),

(AU (n) , U (n) )L2 (Ω) (t) ≥ γ(Ω)U (n) 2L2 (Ω) (t), (4.48)

one can see straightforwardly from Eq. (4.47) by choosing 2p > ν1 and the
previous bounds Eq. (4.39) that there is a positive constant B such that
 T
dsU̇ (n) (s)2L2 (Ω) ≤ B = M T, (4.49)
0

which by its turn yields that dUndt(t,x) is weakly convergent on P̄ (x, t) in


L2 ([0, T ], L2(Ω)).
As a consequence of general theorems of function convergence on the
space of integrable functions (Aubin–Lion theorem1 again), one has that
P̄ (x, t) is the time-derivative of the function Ū (x, t) almost everywhere on Ω,
since it is expected that U (n) (x, t) should be a strongly convergent sequence
to Ū (x, t) on the separable and reflexive Banach space L∞ ([0, T ] × L2 (Ω)).1
(See Appendix B for mathematical details.)

4.4. A Path-Integral Solution for the Parabolic Nonlinear


Diffusion

Let us start the physicist-oriented section of our note by writing the abstract
scalar parabolic equations (4.1)–(4.3) in the integral (weak) form
 t
U (t) = e−At U (0) + ds e−(t−s)A ∆F (U (s)), (4.50)
0

where F (x) is a general nonlinear scalar functional such that F : L2 (Ω) →


H 2 (Ω).
Let us suppose that the initial conditions for Eq. (4.1) are samples of
Gaussian stochastic processes belonging to the L2 (Ω) space. For instance,
this mathematical fact is always true when the processes correlation
function defines an integral operator of the trace class on L2 (Ω),

E{U (0, x)U (0, y)} = K(x, y), (4.51)

where

dxK(x, x) < ∞. (4.52)

August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 95

At a formal path-integral method , one aims to compute the initial con-


dition average of arbitrary product of the function U (x, t) ≡ Ux (t) for arbi-
trary space–time points. This task is achieved by considering the associated
characteristic functional for Eq. (4.51), namely,
   
T
1
Z[j(t)] = E exp i dt j(t), U (t) L2 (Ω) . (4.53)
Z(0) 0

In order to write a path-integral representation for Eq. (4.53), we follow


our previous studies on the subject6 by realizing the random initial condi-
tions U (0) = U0 average as a Gaussian functional integral on L2 (Ω), namely,
  
   
1 
Z[f (t)] = dµ[U0 ] × dU (x, t)
Z(0)  L2 (Ω) L2 ([0,T ],L2 (Ω)) x∈Ω,t∈[0,T ]

  t 
× δ (F ) U (t) − e−At U0 + ds e−(t−s)A ∆F (U (s))
0

  ∞ 
× exp i dt j(t), U (t) L2 (Ω) , (4.54)
0 

where formally
    
 1 −1
dµ[U0 ] = (dU0 (x)) × exp − dxdyU0 (x)K (x, y)U0 (y) .
2 Ω
x∈Ω
(4.55)
By using the well-known Fourier functional integral representation for
the delta-functional inside the identity equation (4.54)6 (see this old idea
in A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics, Vol. 2 (MIT
Press, Cambridge, 1971),6 one can rewrite Eq. (4.54) in the following form:
 
1 
Z[j(t)] = (dU (x, t)) × dλ(t)
Z(0) L∞ ([0,T ],L2 (Ω)) L2 ([0,T ],L2 (Ω))
  !  t "
t
−(t−s)A
× exp i λ(t), U (t) + ds e ∆F (U )(s)
0 0 L2 (Ω)
  
1 t # $
× exp − ds λ(s)((e−As K −1 e+As )λ)(s) L2 (Ω) . (4.56)
2 0
August 6, 2008 15:46 9in x 6in B-640 ch04

96 Lecture Notes in Applied Differential Equations of Mathematical Physics

After evaluating the Gaussian cylindrical measure associated with the


Lagrange multiplier fields λ(x, t), one obtains our formal path-integral rep-
resentation, however, with a weight well-defined mathematically as showed
in Sec. 4.2, as the main conclusion of Secs. 4.2–4.4

1 
Z[j(t)] = dU (x, t)
Z(0) L∞ ((0,T ),L2 (Ω))

  T
1 T
× exp − ds ds
2 0 0
! 
∂U
× + AU − ∆F (U ) (s);
∂s
  "
∂U 
× + AU − ∆F (U ) (s )
∂s L2 (Ω)
  
T
× exp i ds j(s), U (s) L2 (Ω) . (4.57)
0

It is worth rewriting the result equation (4.57) in the usual physicist


notation of Feynman path-integrals6

1
Z[f (x, t)] = DF [U (x, s)]
Z(0) L∞ ((0,T ),L2 (Ω))
     
T T
1 ∂U
× exp − ds ds
+ AU − ∆F (U ) (x, s)
dx
02 0 Ω ∂t
  
∂U
× + AU − ∆F (U ) (x, s )
∂t
  
T
× exp i ds dx j(s, x)U (s, x) . (4.58)
0 Ω

4.5. Random Anomalous Diffusion, A Semigroup Approach

Recently, it has been an important issue on mathematical physics of dif-


fusion on random porous medium, the study of the anomalous diffusion
equation on the full space RD but under the presence of a positive random
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 97

potential as modeled by the parabolic quasilinear equation (0 < t < ∞) as


written below7 :
∂U (x, t)
= −D0 (−∆)α U (x, t) + V (x)U (x, t) + F (U (x, t)), (4.59a)
∂t
U (x, 0) = f (x), (4.59b)
where U (x, t) is the diffusion field, D0 is the medium diffusibility constant,
V (x) denotes the stochastic samples (positive functions) associated with
a general random field process with realizations on the space of square-
integrable functions (by the Minlo’s theorem — see Appendix D), F (x)
denotes the problem’s nonlinearity represented by a Lipschitzian function
as in the previous sections and, finally, (−∆)α represents the effect of the
anomalous diffusion of the field U (x, t) on the ambient RD where the dif-
fusion takes place and it is represented here by a fractional power of the
Laplacean operator. The constant α here is called the anomalous diffusion
exponent.
In order to give a precise mathematical meaning we will present a dif-
ferent mathematical scheme for the previous sections for Eq. (4.59). Let us,
thus, proceed for a moment by rewriting it formally in the weak-integral
form for those initial-values f (x) ∈ L2 (RD ) as in Eq. (4.54), namely,
U (t, [V ]) = exp[−t(D0 (−∆)α + V )]f
 t
+g ds{exp[−(t − s)(D0 (−∆)α + V )]}F (U (s)), (4.60)
0
where we have used a notation emphasizing the stochastic variable nature of
the diffusion field U (x, t, [V ]) as a functional of the samples V (x) ∈ L2 (RD )
associated with our random diffusion potential.
Physical quantities are functionals of the diffusion field U (t, [V ]), and
after its determination one should average over these random potential
samples. The whole averaging information is contained in the space–time
characteristic functional as pointed out in Sec. 4.4:
   ∞ 
Z[j(x, t)] = dµ[V ] × exp i dx dt j(x, t) · U (x, t, [V ]) .
L2 (RD ) RD 0
(4.61)

Here, dµ[V ] means the random potential cylindrical measure associated


with the V (x)-characteristic functional

Z[h(x)] = dµ[V ] exp i( h, v L2 (RD ) ). (4.62)
L2 (RD )
August 6, 2008 15:46 9in x 6in B-640 ch04

98 Lecture Notes in Applied Differential Equations of Mathematical Physics

Note that we do not suppose the Gaussianity of the field statistics in


Eq. (4.62).
In order to write a functional integral representation for the character-
istic functional equation (4.61) as much as similar representation obtained
in Sec. 4.4, let us rewrite Eq. (4.61) into the weak-integral form as done in
Eq. (4.56):
  
Z[j(x, t)] = dµ[V ] DF [λ] DF [U ]
L2 (RD ) L2 (RD ) L2 (RD )

× exp i{(λ, U + I(F (U )) + e−tA f )L2 (RD ) } (4.63)


where I(F (U )) and e−tA denote the objects on the right-hand side of
Eq. (4.60) and the (formal at this point of our study) contractive generator
semigroup below:
A = D0 (−∆)α + V. (4.64)
At this point one could proceed in a physicist way by rewriting Eq. (4.63)
as a path-integral associated with the dynamics of three fields (the well-
known Martin–Siggin–Rose component path-integral — Ref. 6):
Z[j(x, t)]
  
= dµ[V ] DF [λ] DF [U ]
L2 (RD ) L2 (RD ) L2 (RD )
    
1 ∂
  0 − D0 (−∆)α  
× exp i     2 ∂t  λ
[λ, U ], 1 ∂ U
− − D0 (−∆)α 0
2 ∂t L2 (RD )

× exp i g(λ, F (U ))L2 (RD )


  
0 V λ
× exp i [λ, U ], . (4.65)
V 0 U L2 (RD )

It is worth remarking that in the usual case of the cylindrical mea-


sures dµ[V ] is a purely Gaussian measure, formally written in the physicist
notation as
  
1
dµ[V ] = D [V ] exp −
F
dx dyV (x)K −1 (x, y)V (y)
2 L2 (RD ) L2 (RD )

(4.66)
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 99

with K(x, y) defining an integral operator of the trace class on L2 (RD ), one
can easily make a further simplification by integrating out exactly the V (x)-
Gaussian functional integral and producing the effective quartic interaction
term as written below:
   
0 V λ
dµ[V ] exp i [λ, U ],
L2 (RD ) V 0 U L2 (RD )
   ∞   ∞
1
= exp − dx dt dy ds
2 RD 0 RD 0

× (λ(x, t)U (x, t))K(x, y)(λ(y, s)U (y, s)) . (4.67)

Let us call the reader’s attention that at this point one can straightfor-
wardly implement the usual Feynman–Wild–Martin–Siggia–Rosen diagra-
matics with the free propagator given explicitly in the momentum space
by6
1
[∂t − D0 (−∆)α ]−1 ↔ . (4.68)
iw − D0 (|k|2α )

It is important to remark that all the above analyses are still somewhat
formal at this point of our study since it is based strongly on the hypothesis
that the operator A as given by Eq. (4.64) is a C0 -generator of a contractive
semigroup on L2 (RD ), which straightforwardly leads to the existence and
uniqueness of global solution for Eq. (4.59). Let us give a rigorous proof of
ours of such a self-adjointness result, which by its turn will provide a strong
connection between the parameter α of the Laplacean power, related to the
anomalous diffusion coefficient, and the underline dimension D of the space
where the anomalous diffusion is taking place.
Let us first recall the famous Kato–Rellich theorem on self-adjointness
perturbative of self-adjoint operators.

Theorem of Kato–Rellich. Suppose that A is a self-adjoint operator on


L2 (RD ), B is a symmetric operator with Dom(B) ⊃ Dom(A), such that
(0 ≤ a < 1),

BϕL2 (RD ) ≤ aAϕL2 (RD ) + bϕL2 (RD ) , (4.69)

then, A + B is self-adjoint on Dom(A) and essentially self-adjoint on any


core of A.
August 6, 2008 15:46 9in x 6in B-640 ch04

100 Lecture Notes in Applied Differential Equations of Mathematical Physics

In order to apply this powerful theorem to our general case under study,
let us recall that the fractional power of the Laplacean operator, (−∆)α ,
has as an operator domain the (Sobolev) space of all square-integrable
functions ϕ(x) such that its Fourier transform ϕ̃(k) satisfies the condition
|k|α ϕ̃(k) ∈ L2 (RD ) and as a core domain the Schwartz space S(RD ).8
As a consequence, we have the straightforward estimate for ϕ ∈ S(RD )
as
 
 1 

ϕ̃L2 (RD ) ≤   (k 2α + 1)ϕ̃L2 (RD )
1 + k 2α L2 (RD )

≤ C(α, D)(k 2α ϕ̂L2 (RD ) + ϕ̂L2 (RD ) ), (4.70)

where the finite constant



dk
C(α, D) = < ∞, (4.71)
RD (1 + k 2α )2

if α > D4 , a condition relating the “anomalous” diffusion coefficient α and


the intrinsic space–time dimensionality D as said before in the introduction.
Note that for the physical case of D = 3, we have that α = 34 +ε with ε > 0.
Let us rescale the function ϕ̃(k) (r > 0) (for β > 0 arbitrary)

ϕ̃r (k) = rβ ϕ̃(rk). (4.72)

We thus have

ϕ̃r L1 (RD ) = dkrβ ϕ̃(rk)
RD
= rβ−D ϕ̃L1 (RD ) (4.73)

and

ϕ̃r L2 (RD ) = r(2β−D)/2 ϕ̃L2 (RD ) (4.74)

together with

K 2α ϕ̃r L2 (RD ) = r(2β−D−4α)/2 K 2α ϕ̃L2 (RD ) . (4.75)

Let us substitute Eqs. (4.73)–(4.75) into Eq. (4.70). We arrive at the


estimate
-
ϕ̃r L1 (RD ) ≤ C(α, D) r(2β−D−4α)/2 K 2α ϕ̃L2 (RD )
.
+ r(2β−D)/2 ϕ̃L2 (RD ) (4.76)
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 101

or equivalently
/ D
ϕ̃L1 (RD ) ≤ C(α, D) r 2 −2α K 2α ϕ̃L2 (RD )
D
0
+ r 2 ϕ̃L2 (RD ) . (4.77)

Let us now estimate the function V ϕ on the space L2 (RD ):

V ϕL2 (RD ) ≤ V L2 (RD ) ϕL∞ (RD )


≤ V L2 (RD ) ϕ̃L1 (RD )
D
≤ (V L2 (RD ) C(α, D) · r 2 −2α )K 2α ϕ̃L2 (RD )
D
+ (V L2 (RD ) C(α, D) · r 2 )ϕ̃L2 (RD ) . (4.78)

Now one can just choose r such that

V L2 (RD ) × C(α, D) × r(D−4α)/2 < 1, (4.79)

in order to obtain the validity of the bound a < 1 on Eq. (4.69) and, thus,
concluding that (−∆)α +V is an essentially self-adjoint operator on S(RD ).
So, its closure on L2 (RD ) produces the operator used in the above-exposed
semigroup C0 -construction for V (x) being a positive function almost every-
where. This result of ours is a substantial generalization of Theorem X.15
presented in Ref. 8. Note that the V (x) sample positivity is always the
physical case of random porosity in a Porous medium. In such a case, the
random potential is of the exponential form V (x) = V0 (exp{ga(x)}), where
g is a positive small parameter and V0 is a background porosity term. The
effective cylindrical measure on the random porosity parameters is written
as (see Eq. (4.66))
   
∼ 1 ga(x) −1
dµ[V ] = D [a(x)] exp −
F
dx dye K (x, y)e ga(y)
.
g1 2 RD RD
(4.80)

The above-exposed C0 -semigroup study complements the study on


purely nonlinear diffusion made in the previous sections purely by com-
pacity arguments.
Finally, let us briefly sketch on the important case of wave propagation
U (x, t) in a random medium described by a small stochastic damping pos-
2
itive L2 (RD ) function ν(x), such that ν 4(x) is a Gaussian process sampled in
the L2 (RD ) space. The hyperbolic linear equation governing such physical
August 6, 2008 15:46 9in x 6in B-640 ch04

102 Lecture Notes in Applied Differential Equations of Mathematical Physics

wave propagation is given by (see Sec. 4.3)


∂ 2 (U (x, t)) ∂U (x, t)
+ ν(x) = −(−∆)α U (x, t), (4.81)
∂2t ∂t

U (x, 0) = f (x) ∈ L2 (RD ), (4.82)


2
Ut (x, 0) = g(x) ∈ L (R ), D
(4.83)
where
  
ν 2 (x) ν 2 (y)
E = K(x, y) ∈ (L2 (RD )), (4.84)
4 4 1
One can see that after the variable change
ν(x)
U (x, t, [ν]) = e− 2 t · Φ(x, t, [ν]), (4.85)
the damped-stochastic wave equation takes the suitable form of a wave in
the presence of a random potential
 2 
∂ 2 Φ(x, t) ν (x)
= −(−∆)α
Φ(x, t) + Φ(x, t),
∂2t 4
Φ(x, 0) = f (x), (4.86)
ν(x)
Φt (x, 0) = f (x) + g(x) = ḡ(x).
2
Equation (4.86) has an operatorial scalar weak-integral solution for t > 0
of the following form:

√ sin(t A)
Φ(t) = cos(t A)f + √ ḡ, (4.87)
A
where the nonpositive self-adjoint operator A is given explicitly by
ν 2 (x)
A = (−∆)α − . (4.88)
4
As a conclusion, one can see that anomalous diffusion can be handled
mathematically in the framework of our previous technique of path-integrals
with constraints.

References

1. J. Wloka, Partial Differential Equations (Cambridge University Press, 1987).


2. Y. Yamasaki, Measures on Infinite Dimensional Spaces, Series in Pure Math-
ematics, Vol. 5 (World Scientific, 1985); B. Simon, Functional Integration and
Quantum Physics (Academic Press, 1979); J. Glimn and A. Jaffe, Quantum
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 103

Physics — A Functional Integral Point of View (Springer Verlag, 1981); Xia


Dao Xing, Measure and Integration Theory on Infinite Dimensional Space
(Academic Press, 1972); L. Schwartz, Random Measures on Arbitrary Topo-
logical Space and Cylindrical Measures (Tata Institute, Oxford University
Press, 1973); Ya G. Sinai, The Theory of Phase Transitions: Rigorous Results
(London, Pergamon Press, 1981).
3. O. A. Ladyzenskaja, V. A. Solonninkov and N. N. Ural’ceva, Linear and
Quasilinear Equations of Parabolic Type (Amer. Math. Soc. Transl., A.M.S.,
Providence, RI, 1968).
4. S. G. Mikhlin, Mathematical Physics, An Advanced Course, Series in Applied
Mathematics and Mechanics (North-Holland, 1970).
5. L. C. L. Botelho, J. Phys. A: Math. Gen., 34, L131–L137 (2001); Mod. Phys.
Lett. 16B(21), 793–8-6, (2002); A. Bensoussan and R. Teman, J. Funct. Anal.
13, 195–222 (1973); A. Friedman, Variational Principles and Free-Boundary
Problems, Pure & Applied Mathematics (John-Wiley & Sons, NY, 1982).
6. L. C. L. Botelho, Il Nuovo Cimento, 117B(1), 15 (2002); J. Math. Phys.
42, 1682 (2001); A. S. Monin and A. M. Yaglon, Statistical Fluid Mechanics,
Vol. 2 (MIT Press, Cambridge, 1971); G. Rosen, J. Math. Phys. 12, 812
(1971); L. C. L. Botelho, Mod. Phys. Lett. 13B, 317 (1999); A. A. Migdal,
Int. J. Mod. Phys. A 9, 1197 (1994); V. Gurarie and A. Migdal, Phys. Rev.
E 54, 4908 (1996); U. Frisch, Turbulence (Cambridge Press, Cambridge,
1996); W. D. Mc-Comb, The Physics of Fluid Turbulence (Oxford University,
Oxford, 1990). L. C. L. Botelho, Mod. Phys. Lett. B 13, 363 (1999).
7. L. C. L. Botelho, Nuov. Cimento 118B, 383 (2004); S. I. Denisov and
W. Horsthemke, Phys. Lett. A 282(6), 367 (2001); L. C. L. Botelho, Int. J.
Mod. Phys. B 13, 1663 (1999); L. C. L. Botelho, Mod. Phys. Lett. B 12, 301
(1998); J. C. Cresson and M. L. Lyra, J. Phys. Condens. Matter 8(7), L83
(1996); L. C. L. Botelho, Mod. Phys. Lett. B 12, 569 (1998); L. C. L. Botelho,
Mod. Phys. Lett. B 12, 1191 (1998); L. C. L. Botelho, Mod. Phys. Lett. B 13,
203 (1999); L. C. L. Botelho, Int. J. Mod. Phys. B 12, 2857 (1998); L. C. L.
Botelho, Phys. Rev. 58E, 1141 (1998).
8. M. Reed and B. Simon, Methods of Modern Mathematical Physics, Vol 2
(Academic Press, New York, 1980).

Appendix A

In this mathematical appendix, we intend to give a rigorous mathematical


proof of the result used on Sec. 4.2 about parabolic nonlinear diffusion
in relation to the weak continuity of the nonlinearity ∆F (U ) used in the
passage of Eq. (4.17) for Eq. (4.18).
First, we consider the physical hypothesis on the nonlinearity of the
parabolic equation (4.1) that the Laplacean operator has a cutoff in its
August 6, 2008 15:46 9in x 6in B-640 ch04

104 Lecture Notes in Applied Differential Equations of Mathematical Physics

spectral range, namely,


∆F (U (x, t)) → ∆(Λ) F (U (x, t)), (A.1)
(Λ)
where the regularized Laplacean ∆ means the bounded operator of norm
Λ, i.e. in terms of the spectral theorem of the Laplacean
 +∞
∆= λ dEs (λ), (A.2)
−∞
it is given by

∆(Λ) = λdEs (λ). (A.3)
|λ|<Λ
Let us impose either the well-known path-integral Sturm–Liouville time-
boundary conditions imposed on the elements of the path-integral domain
(Eq. (4.58)), when one is defining the path-integral by means of fluctuations
around classical configurations
Ū(x, T ) = Ū (x, 0) = 0. (A.4)
It is straightforward to see the validity of the following chain of inequal-
ities with a > 1:
 T 1 1
1 dUn (t) 12
dt 1
1 dt 1 2
1 ≤ Real((AUn (t), Un (t)) − (AUn (0), Un (0)))
0 L (Ω)

 1 1
T 1 (Λ) dUn 1
+ 1
dt 1∆ F (Un (t)) · 1
0 dt 1 L2 (Ω)
 
T
2
≤0+0+a ∆F (Un (t))L2 (Ω) dt
0
 1 1 
1 T 1 dUn 12
+ dt 1 1
1 dt 1 2 . (A.5)
a 0 L (Ω)

In other words,
   T 1 1 
1 1 dUn 12
1− dt 1 1
1 dt 1 2
a 0 L (Ω)

 T
≤ caΛ2 · Un (t)2L2 (Ω)
0

2
≤ CaΛ T M, (A.6)
where c = supa≤x≤b |F  (x)|, and M is the bound given by Eq. (4.16).
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 105

Now, by a direct application of the well-known Aubin–Lion theorem,1


one has that the sequence {Un } is a compact set on L2 (Ω). So, it con-
verges strongly to Ū on L2 (Ω) as a consequence of the Lipschitzian property
of F (x).
We have either the strong convergence of F (Un ) on F (Ū ), since for each
t we have that the inequality below holds true:

dx|F (Un ) − F (Ū )|2L2 (Ω) (t)

 2  
 2
≤ sup (F (x)) dx|Un − Ū|L2 (Ω) (t) → 0. (A.7)
−∞<x<+∞ Ω

It is, thus, an immediate result of the validity of the weak convergence


on L2 (Ω) used on Eq. (4.18):
 T  T
lim dt(F (Un ), ∆v)L2 (Ω) = dt(F (Ū ), ∆v)L2 (Ω) . (A.8)
n→∞ 0 0

Appendix B

In this appendix, we show that the function U (x, t) ∈ L∞ ([0, T ], L2(Ω)),


the weak solution of Eq. (4.1) in the test space D((0, T ), L2 (Ω)) (the
Schwart space of L2 (Ω)-valued test functions), satisfies the initial condition
(Eq. (4.2)). In order to show such a result, let us consider a test  function

(ε)
possessing the following form for each ε > 0: vi (x, t) = ϕi (x) χ[0,ε] (t)/ε ,
where ϕi (x) is a member of the spectral set associated with the self-adjoint
operator A : H 2 (Ω) ∩ H01 (Ω) → L2 (Ω). The notation χ[a,b] (t) denotes the
characteristic function of the interval [a, b].
Since Um (x, t) converges to U (x, t) in the weak-star topology of
L∞ ((0, T ), L2 (Ω)) and D((0, T ), L2 (Ω)) ⊂ L1 ((0, T ), L2 (Ω)), we have the
(ε)
relation below for each vi (x, t) that holds true
 T  T
(ε) (ε)
dt Um (x, t), vi (x, t) L2 (Ω) → dt U (x, t), vi (x, t) L2 (Ω) , (B.1)
0 0

or equivalently
 ε  ε
1 1
dt Um (x, t), ϕi (x) L2 (Ω) → dt U (x, t), ϕi (x) L2 (Ω) . (B.2)
ε 0 ε 0
August 6, 2008 15:46 9in x 6in B-640 ch04

106 Lecture Notes in Applied Differential Equations of Mathematical Physics

By means of the mean-value theorem applied to both sides of Eq. (A.2),


and taking the limit ε → 0, we have for each i ∈ Z that
Um (x, 0), ϕi (x) L2 (Ω) → U (x, 0), ϕi (x) L2 (Ω) (B.3)
as a direct consequence of our choice of the Galerkin elements Eqs. (4.6)
whose functional form does not change on the function of the order m ∈ Z.
As a consequence of Eqs. (4.5) and (4.6), it yields the result
g(x), ϕi L2 (Ω) = U (x, 0), ϕi (x) L2 (Ω) , (B.4)
or by means of Parserval–Fourier theorem,
lim U (x, t) − g(x)L2 (Ω) = 0 (B.5)
t→0

result, which by its turn, means that the function U (x, t) obtained by means
of our compacity technique satisfies the initial-condition Eq. (4.2) in the
L2 (Ω)-mean sense.

Appendix C

Let us show that the functions Ū(x, t) ∈ L∞ ([0, T ], L2(Ω)) given by


Eq. (4.41) and P̄ (x, t) ∈ L∞ ([0, T ], L2 (Ω)) — Eq. (4.42) are coincident as
elements of the above-written functional space of L2 (Ω) — valued essential
bounded functions on (0, T ).
To verify such a result, let us call the reader’s attention that since
Um (x, t) is weakly convergent to U (x, t) in L∞ ([0, T ], L2 (Ω)), we have that
the set {Um (x, t)} is convergent to the function U (x, t) as a Schwartz dis-
tribution on L2 (Ω) since D([0, T ], L2 (Ω)) ⊂ L1 ([0, T ], L2 (Ω)).
This means that
Um (x, t) → U (x, t) in D ([0, T ], L2 (Ω)). (C.1)
Analogous result holds true for the time-derivative of the above-written
equation as a result of U (x, t) being a function
∂ ∂
Um (x, t) → U (x, t) in D ([0, T ], L2 (Ω)). (C.2)
∂t ∂t
 
On the other side, the set ∂Um∂t(x,t) converges weakly in L1 ([0, T ],
L2 (Ω)) to P̄ (x, t) ∈ L∞ ([0, T ], L2(Ω)), which, by its turn, means that
∂Um (x, t)
→ P̄ (x, t) in D ([0, T ], L2 (Ω)). (C.3)
∂t
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 107

By the uniqueness of the limit on the distributional space


D ([0, T ], L2 (Ω)), we have the coincidence of ∂U(x,t)

∂t and P̄ (x, t) as ele-
ments of D ([0, T ], L2 (Ω)). However, P̄ (x, t) is a function;
 so by general
theorems on Schwartz distribution theory ∂U(x,t) ∂t must be a function
either since L2 (Ω) is a separable Hilbert space. As a consequence we have
(x,t)
that ∂U∂t = P̄ (x, t) as elements of L∞ ([0, T ], L2(Ω)), which is the result
searched:

∂U (x, t)
= P̄ (x, t) a.e. in ([0, T ] × Ω. (C.4)
∂t

Appendix D. Probability Theory in Terms of Functional


Integrals and the Minlos Theorem — An Overview

In this complementary appendix, we discuss briefly for the mathematics


oriented reader the mathematical basis and concepts of functional integrals
formulation for stochastic process and the important Minlos theorem on
Hilbert space support of probabilistic measures.
The first basic notion of Kolmogorov’s probability theory framework is
to postulate a convenient topological space (Polish spaces) Ω formed by
the phenomena random events. The chosen topology of Ω should possess a
rich set of nontrivial compact subsets. After that, we consider the σ-algebra
generated by all open sets of this — the so-called topological sample space,
denoted by FΩ . Thirdly, one introduces a regular measure dµ on this set
algebra FΩ , assigning values in [0, 1] to each member of FΩ .
The (abstract) triplet (Ω, FΩ , dµ) is thus called a probability space in
the Kolmogorov–Schwartz scheme.10
Let {Xα }α∈Λ be a set of measurable real functions on Ω, which may be
taken as continuous injective functions of compact support on Ω (without
loss of generality).
It is a standard result that one can “immerse” (represent) the
abstract space Ω on the “concrete” infinite product compact space R∞ =
2
α∈Λ (Ṙα ), where Ṙ is a compactified copy of the real line. In order to
achieve such result we consider the following injection of Ω in R∞ defined
by the family {Xα }α∈Λ :

I : Ω → R∞
(D.1)
ω → {Xα (ω)}α∈Λ .
August 6, 2008 15:46 9in x 6in B-640 ch04

108 Lecture Notes in Applied Differential Equations of Mathematical Physics

A new σ-algebra of events on Ω can be induced on Ω and affiliated to


the family {Xα }α∈Λ .
It is the σ-algebra generated by all the “finite-dimensional cylinders”
subsets, which are explicitly given in its generic formulas by

CΛfin = ω ∈ Ω | Λfin is a finite subset of Λ
 
and [(Xα )α∈Λfin (Ω)] ⊂ [aα , bα ] .
α∈Λfin

The measure restriction of the initial measure µ on this cylinder sets of


Ω will be still denoted by µ on what follows. Now it is a basic theorem of
2
Probability theory that µ induces a measure ν (∞) on α∈Λ (Ṙ)α , which can
be identified with the space of all functions of the index set Λ to the real
compactified line Ṙ; F (Λ, R) with the topology of pontual convergence.
At this point, it is straightforward to see that the average of any Borelian
(mensurable) function G(ω) (a random variable) is given by the following
functional integral on the functional space F (Λ, R):
 
dµ(ω)G(ω) = dν (∞) (f ) · G(I −1 (f )). (D.2)
Ω F (Λ,R)

It is worth recalling that the real support of the measure dν (∞) (f ) in


most practical cases is not the whole space F (Λ, R) but only a functional
subspace of F (Λ, R). For instance, in the most practical cases Λ is the
2
index set of an algebraic vectorial base of a vector space E, and α∈Λ (R)α
(without the compactification) is the space of all sequences with only a finite
number of nonzero entries as it is necessary to consider in the case of the
famous extension theorem of Kolmogorov. It turns out that one can consider
the support of the probability measure as the set of all linear functionals
on E, the so-called algebraic dual of E. This result is a direct consequence
of considering the set of random variables given by the family {e∗α }α∈Λ . For
applications, one is naturally lead to consider the probabilistic object called
characteristic functional associated to the measure dν (∞) (f ) when F (Λ, R)
is identified with the vector space of all algebraic linear functionals E alg of
a given vector space E as pointed out above, namely,

Z[j] = dν (∞) (f ) exp(if (j)), (D.3)
E alg

where j are elements of E written as j = α∈Λfin xα eα with {eα }α∈Λ


denoting a given Hammel (vectorial) basis of E.
August 6, 2008 15:46 9in x 6in B-640 ch04

Nonlinear Diffusion and Wave-Damped Propagation 109

One can show that if there is a subspace H of E with a norm  H


coming from an inner product (, )H such that

dν (∞) (f ) · f 2H < ∞ (D.4)
E∩H

one can show that the support of the measure of exactly the Hilbert space
(H, (, )H ) is the famous Minlos theorem.
Another important theorem (the famous Wiener theorem) is when we
have the following condition for the family of random variables with the
index set Λ being a real interval of the form Λ = [a, b]

dν (∞) (f )|Xt+h (f ) − Xt (f )|p ≤ c|h|1+ε (D.5)
E

for p, ε, c positive constants. We have that it is possible to choose as the


support of the functional measure dν (∞) (·) the space of the real continuous
function on Λ the well-known famous Brownian path space
 
dµ(ω)G(ω) = dν (∞) (f (x))G̃(f (x)). (D.6)
Ω C([a,b],R)

For rigorous proofs and precise formulations of the above-sketched deep


mathematical results, the interested reader should consult the basic math-
ematical book on Random Measures on Arbitrary Topological Spaces and
Cylindrical Measures by L. Schwartz (Tata Institute, Oxford University
Press, 1973).10
Let us now finally give a concrete example of such results considering a
generating functional Z[J(x)] of the exponential quadratic functional form
on L2 (RN ):
   
1
Z[J(x)] = exp − dN x dN yj(x)K(x, y)j(y) (D.7)
2 RN RN

and associate to a Gaussian random field with two-point correlation


function given by the kernel of Eq. (D.7):

E{v(x)v(y)} = K(x, y). (D.8)

In the3case of the above-written kernel being a class-trace operator on


2
L (RN ) ( RN dN xK(x, x) < ∞), one can represent Eq. (B.7) by means of
a Gaussian measure with the support of the Hilbert space L2 (RN ). This
result was called on for the reader’s attention for the bulk of this paper.
August 6, 2008 15:46 9in x 6in B-640 ch05

Chapter 5

Domains of Bosonic Functional Integrals and Some


Applications to the Mathematical Physics of
Path-Integrals and String Theory∗

5.1. Introduction

Since the result of R. P. Feynman on representing the initial value


solution of Schrodinger equation by means of an analytically time con-
tinued integration on an infinite-dimensional space of functions, the subject
of Euclidean functional integrals representations for quantum systems has
become the mathematical-operational framework to analyze quantum phe-
nomena and stochastic systems as shown in the previous decades of research
on theoretical physics.1−3
One of the most important open problem in the mathematical theory
of Euclidean functional integrals is related to the implementation of sound
mathematical approximations to these infinite-dimensional integrals by
means of finite-dimensional approximations outside of the always used [com-
puter oriented] space-time lattice approximations (see Refs. 2 and 3 —
Chap. 9). As a first step to tackle the above-cited problem, there is a need to
mathematically characterize the functional domain where these functional
integrals are defined.
The purpose of this chapter is to present, in Sec. 5.2, the formulation of
Euclidean quantum field theories as functional Fourier transforms by means
of the Bochner–Martin–Kolmogorov theorem for topological vector spaces
(Refs. 4 and 5 — Theorem 4.35) and suitable to define and analyze rigor-
ously the functional integrals by means of the well-known Minlos theorem
(Ref. 5, Theorem 4.31 and Ref. 6, Part 2) and presented in full details in
Sec. 5.3.
In Sec. 5.4, we present new results on the difficult problem of defining
rigorously infinite-dimensional quantum field path-integrals in general space
times Ω ⊂ Rν (ν = 2, 4, . . . ) by means of the analytical regularization
scheme.
∗ Author’s original results.
110
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 111

5.2. The Euclidean Schwinger Generating Functional as


a Functional Fourier Transform

The basic object in a scalar Euclidean quantum field theory in RD is the


Schwinger generating functional (see Refs. 1 and 3).
     
 
Z[j(x)] = ΩV AC exp i dD xj(x, it)φ(m) (x, it)  ΩV AC , (5.1)

where φ(m) (x, it) is the supposed self-adjoint Minkowski quantum field ana-
lytically continued to imaginary time and j(x) = j(x, it) is a set of func-
tions belonging to a given topological vector space of functions denoted
by E which topology is not specified yet and will be called the Schwinger
classical field source space. It is important to remark that {φm (x, it)} is a
commuting algebra of self-adjoints operators as Symanzik has pointed out.7
In order to write Eq. (5.1) as an integral over the space E alg of all linear
functionals on the Schwinger source space E (the so-called algebraic dual
of E), we take the following procedure, different from the usual abstract
approach (as given — for instance — in the proof of IV-11-2), by making
the hypothesis that the restriction of the Schwinger generating functional
equation (5.1) to any finite-dimensional RN of E is the Fourier transform
of a positive continuous function, namely
  

N 
N
Z Cα j̄α (x) = exp i Cα Pα g̃(P1 , . . . , PN )dP1 , . . . , dPN .
α=1 RN α=1
(5.2)

Here, {jα (x)}α=1,...,N is a fixed vectorial basis of the given finite-


dimensional subspace (isomorphic to RN ) of E.
As a consequence of the above-made hypothesis (based physically
on the renormalizability and unitary of the associated quantum field
theory), one can apply the Bochner–Martin–Kolmogorov theorem (Ref. 5,
Theorem 4.35) to write Eq. (5.1) as a functional Fourier transform on the
space E alg (see Appendix A)

Z[j(x)] = exp(ih(j(x))dµ(h), (5.3)
E alg

where dµ(h) is the Kolmogorov cylindrical measure on E alg = Πλ∈A (Rλ )


with A denoting the index set of the fixed Hamel vectorial basis used in
August 6, 2008 15:46 9in x 6in B-640 ch05

112 Lecture Notes in Applied Differential Equations of Mathematical Physics

Eq. (5.2) and h(j(x)) is the action of the given linear (algebric) functional
(belonging to E alg ) on the element j(x) ∈ E.
At this point, we relate the mathematically non-rigorous physicist point
of view to the Kolmogorov measure dµ(h) (Eq. (5.3)) over the algebraic
linear functions on the Schwinger source space. It is formally given by the
famous Feynman formulae when one identifies the action of h on E by
means of an “integral” average


h(j) = dxD j(x)h(x). (5.4)
RD

Formally, we have the equation


dµ(h) = dh(x) exp{−S(h(x))}, (5.5)
x∈RD

where S is the classical action of the classical field theory under quanti-
zation, but with the necessary coupling constant renormalizations needed
to make the associated quantum field theory well-defined.
Let us outline these proposed steps on a λφ4 -field theory on R4 .
At first we will introduce the massive free field theory generating func-
tional directly into the infinite volume space R4 .


1
Z[j(x)] = exp − d4 xd4 x j(x)((−∆)α + m2 )−1 (x, x )j(x ) , (5.6)
2

where the free field propagator is given by

 
eik(x−x )
((−∆)α + m2 )−1 (x, x ) = d4 k , (5.7)
k 2α + m2

with α a regularizing parameter with α > 1.


As the source space, we will consider the vector space of all real
sequences on Πλ∈(−∞,∞) (R)λ , but with only a finite number of non-zero
components. Let us define the following family of finite-dimensional pos-
itive linear functionals {LΛf } on the functional space C( λ∈(−∞,∞) Rλ ; R)
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 113

  
LΛf g(Pλs1 , . . . , PλsN ) =  Q  g(Pλs1 , . . . , PλsN )

λ∈Λf
 
 1  
× exp − (λ2α + m2 )(Pλ )2
 2 
λ∈Λf
 

× d(Pλ π(λ2α + m2 )) . (5.8)
λ∈Λf

Here, Λf = {λs1 , . . . , λsN } is an ordered sequence of real number of the real


line which is the index set of the Hamel basis of the algebraic dual of the
proposed source space.
Note that we have the generalized eigenproblem expansion

((−∆)α + m2 )eiλx = (λ2α + m2 )eiλx . (5.9)

By the Stone–Weierstrass theorem or the Kolmogoroff theorem applied


to the family of finite-dimensional measure in Eq. (5.8), there is a unique
extension measure dµ({Pλ }λ∈(−∞,∞) ) to the space Πλ∈(−∞,∞) Rλ = E alg
and representing the infinite-volume generating functional on our chosen
source space (the usual Riesz–Markov theorem applied to the linear func-
tional L = lim{Λf } sup LΛf on C(Πλ∈(−∞,∞) Rλ , R) leads to this extension
measure).10

Z[j(x)] = Z[{jλ }λ∈Λf ]


 
 
= ! dµ(0)·(α) ({P } 
λ λ∈(−∞,∞) ) exp i jλ Pλ 
Q
Rλ λ∈(−∞,∞)
λ∈(−∞,∞)

 
 1  (jλ )2 
= exp − . (5.10)
 2 λ2α + m2 
λ∈Λf

At this point it is very important to remark that the generating func-


tional equation (5.10) has continuous natural extension to any test space
(S(RN ), D(RN ), etc.) which contains the continuous functions of compact
support as a dense subspace.
At this point we consider the following quantum field interaction func-
tional which is a measurable functional in relation to the above-constructed
Kolmogoroff measure dµ(0)·(α) ({Pλ }λ∈(−∞,∞) ) for α non-integer in the
August 6, 2008 15:46 9in x 6in B-640 ch05

114 Lecture Notes in Applied Differential Equations of Mathematical Physics

original field variable φ(x)

1 (α) 1 (α)
V (α) (φ) = λR φ4 + (Zφ (λR , M ) − 1)φ((−∆)α )φ − [(m2 Zφ (λR , m) − 1
2 2
(α)
− (δm2 )(α) (λR )]φ2 − [Zφ (λR , m)(δ (α) λ)(λR , m)]φ4 . (5.11)

Here, the renormalization constants are given in the usual analytical finite-
part regularization form for a λφ4 -field theory. It is still an open problem
in the mathematical-physics of quantum fields to prove the integrability
in some distributional space of the cut-off removing α → 1 limit of the
interaction lagrangean exp(−V (α) (φ)) (see Sec. 5.4 for the analysis of this
cut-off removing on space functions).

5.3. The Support of Functional Measures — The


Minlos Theorem

Let us now analyze the measure support of quantum field theories gener-
ating functional equation (5.3).
For higher dimensional space-time, the only available result in this
direction is the case that we have a Hilbert structure on E.4−6
At this point, we introduce some definitions. Let ϕ : Z+ → R be an
increasing fixed function (including the case ϕ(∞) = ∞). Let E be denoted
by H and H Z be the subspace of H alg = (Πλ∈A=[0,1] Rλ ) (with A being the
index set of a Hamel basis of H), formed by all sequences {xλ }λ∈A ∈ H alg
with coordinates different from zero at most a countable number

H Z = {(xλ )λ∈A|xλ = 0 for λ ∈ {λµ }µ∈Z }. (5.12)

Consider the following weighted subset of H alg :


Z
H(e) = {{xλ }λ∈A ∈ H Z }

and
 
1 
N
lim (xλσ(µ) )2 <∞
N →∞ ϕ(N ) n=1

for any σ : N → N , a permutation of the natural numbers.


We now state our generalization of the Minlos theorem.
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 115

Theorem 5.1. Let T be an operator with domain D(T ) ⊂ H, and


T ; D(T ) → H such that for any finite-dimensional space H N ⊂ H, the sum
is bounded by the function ϕ(N )
 
N
 T ei , T ej   ≤ ϕ(N ).
(0)
(5.13)
(i,j)=1

Here , (0) is the inner product of H and {ep }1≤p≤N is a vectorial basis of
the subspace H N with dimension N .
Suppose that Z[j(x)] is a continuous function on D(T ) = (D(T ), , (1) ),
where , (1) is a new inner product defined by the operator T (j, j̄(1) =
T j, T j̄(0) ). We have, thus, that the support of the cylindrical measure
equation (5.3) is the measurable set HeZ .

Proof. Following closely Refs. 1, Theorems 2.2 and 4, let us consider the
following representation for the characteristic function of the measurable
set HeZ ⊂ H alg


·XHeZ ({xλ }λ∈A )

  
 1α  2
N
1  2
N
= lim lim exp − xλl = 1 if lim xλl < ∞

 α→0 N →∞ 2ϕ(N ) N →∞ ϕ(N )

 =1 =1

0 otherwise
(5.14)

Now, its measure satisfies the following inequality:



dµ(h) = µ(H alg ) = 1 > µ(HeZ ). (5.15)
H alg

But
  
α  2
N
µ(HeZ ) = lim lim dµ(h) exp − xλ
α→0 N →∞ H alg 2ϕ(N )
=1
 
1
= lim lim 2πα N/2 dj1 , . . . , djN
α→0 N →∞ ( ϕ(N )) RN
  
N
1 ϕ(N )
× exp − j2 Z̃(j1 , . . . , JN ), (5.16)
2 α
=1
August 6, 2008 15:46 9in x 6in B-640 ch05

116 Lecture Notes in Applied Differential Equations of Mathematical Physics

where
 

N
Z̃(j1 , . . . , jN ) = dµ({xλ }) exp i xλ j
πRλ
λ∈A =1

 

N
= dµ(h) exp i xλ j . (5.17)
H alg =1

Now due to the continuity and positivity of Z[j] in D(T ); we have that
for any ε > 0 → ∃δ such that the inequality below is true since we have
that: Z(j1 , . . . , jN ) ≥ 1 − ε − δ22 (j, j)(1)

 
1 ϕ(N )  2
N
1
 N/2 dj1 · · · djN exp − j Z̃(j1 , . . . , jN )
2πα RN 2 α
=1
ϕ(N )


2  
N
1
≥1−ε− 2   dj1 · · · djN
δ  2πα N/2 RN
(m,n)=1 ϕ(N )

  N 
1 ϕ(N )  2 (1)
× exp − j jm jn em , en 
2 α
=1
 
    N 
2 α
=1−ε− 2 δmn T en , T em (0)
δ  ϕ(N ) 
(m,n)=1
 
2 α 2
≥1−ε− ϕ(N ) ≥ 1 − ε − α. (5.18)
δ2 ϕ(N ) δ2

By substituting Eq. (5.18) into Eq. (5.15), we get the result

2  
1 ≥ µ(HeZ ) ≥ 1 − ε − lim α = 1 − ε. (5.19)
δ 2 α→0

Since ε was arbitrary we have the validity of our theorem.


As a consequence of this theorem in the case of ϕ(N ) being bounded
(so T T ∗ is an operator of Trace Class), we have that HeZ = H which is the
usual topological dual of H.
At this point, a simple proof may be given to the usual Minlos theorem
on Schwartz spaces.5,6
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 117

Let us consider S(RD ) represented as the countable normed spaces of


sequences8

2
S(RD ) = m, (5.20)
m=0

where
  
N
2 
= (xn )n∈Z , xn ∈ R  (xn )2 nm < ∞ (5.21)
m

n=0

The topological dual is given by the nuclear structure sum8



! ∞
!
S  (RD ) = 2
−n = ( 2n )∗ (5.22)
n=0 n=0

We, thus, consider E = S(RD ) in Eq. (9.3) and Z[j(x)] = Z[{jn }n∈Z ] is
"∞ "∞
continuous on n n . Since Z[{jn }n∈Z ] ∈ C( n=0 2n , R) we have that for
any fixed integer p, Z[{jn }n∈Z ] is continuous on the Hilbert space 2p which,
by its turn, may be considered as the domain of the following operator:
2
Tp : pc 0 → 0

{jn } → {np/2 jn }. (5.23)

It is straightforward to have the estimate


 
 N 
  
 (0) 
Tp em , Tp en   ≤ N (Bp ) (5.24)

(m,n)=1 

for some positive integer B and {ei } being the canonical orthonormal basis
of l02 . By an application of our theorem for each fixed p; we get that the
support of measure is given by the union of weighted spaces

! ∞
!
supp dµ(h) = ( 2p )∗ = 2
−p = S  (RD ). (5.25)
p=0 p=0

At this point we can suggest, without a proof of straightforward (non-
topological) generalization of the Minlos theorem.
August 6, 2008 15:46 9in x 6in B-640 ch05

118 Lecture Notes in Applied Differential Equations of Mathematical Physics

Theorem 5.2. Let {Tβ }β∈C be a family of operators satisfying the hypothe-
#
sis of Theorem 5.3. Let us consider the Locally Convex space β∈C Dom(Tβ )
(supposed non-empty) with the family of norms ψ β = Tβ ψ, Tβ ψ1/2 .
If the Functional Fourier transform is continuous on this locally convex
space, the support of the Kolmogoroff measure equation (5.3) is given by the
#
following subset of [ β∈C Dom(Tβ )]alg , namely
!
supp dµ(h) = Hϕ2 β , (5.26)
β∈C

where ϕβ are the functions given by Theorem 5.1. This general theorem will
not be applied in what follows.

Let us now proceed to apply the above displayed results by consid-


ering the Schwinger generating functionals for two-dimensional Euclidean
quantum eletrodynamics in Bosonized parametrization9
    −1 
1 2 2 2 e2
Z[j(x)] = exp − d x d yj(x) (−∆) + (−∆) (x, y)j(y) ,
2 R2 R2 π
(5.27)

where in Eq. (5.27), the electromagnetic field has the decomposition in


Landau Gauge

Aµ (x) = (εµν ∂ν φ)(x) (5.28)

and j(x) is, thus, the Schwinger source for the φ(x) field taken as a basic
dynamical variable.9
Since Eq. (5.27) is continuous in L2 (R2 ) with the inner product defined
2
by the trace class operator ((−∆)2 + eπ (−∆))−1 , we conclude on the basis
of Theorem 5.1 that the associated Kolmogoroff measure in Eq. (5.3) has
its support in L2 (R2 ) with the usual inner product. As a consequence,
the quantum observable algebra will be given by the functional space
L1 (L2 (R2 ), dµ(h)) and usual orthonormal finite-dimensional approxima-
tions in Hilbert spaces may be used safely, i.e. if one considers the basis
$∞
expansion h(x) = n=1 hn en (x) with en (x) denoting the eigenfunctions of
the operator in Eq. (10.27) we get the result

!
L1 (RN , dµ(h1 , . . . , hN )) = L1 (L2 (R2 ), dµ(h)). (5.29)
n=1
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 119

It is worth mentioning that if one uses the Gauge vectorial field


parametrization for the (Q.E.D)2 — Schwinger functional

Z[j1 (x), j2 (x)]


    −1 
1 2 2 e2
= exp − d x d yji (x) −∆ + (x, y)δil jl (x) (5.30)
2 R2 R2 π

the associated measure support will now be the Schwartz space S  (R2 )
2
since the operator (−∆ + eπ )−1 is an application of S(R2 ) to S  (R2 ). As a
consequence, it will be very cumbersome to use Hilbert finite-dimensional
approximations8 as in Eq. (5.29).
An alternative to approximate tempered distributions is the use of its
Hermite expansion in S  (R) distributional space associated with the eigen-
#
functions of the Harmonic-oscillator V (x) ∈ L∞ (R) L2 (R) potential per-
tubation (see Ref. 3 for details with V (x) ≡ 0).
 
d2
− 2 + x2 + V (x) Hn (x) = λn Hn (x). (5.31)
dx

Another important class of Bosonic functionals integrals are those asso-


ciated with an Elliptic positive self-adjoint operator A−1 on L2 (Ω) with
suitable boundary conditions. Here Ω denotes a D-dimensional compact
manifold of RD with volume element dν(x).
 
1
Z[j(x)] = exp − dν(x) dν(y)j(x)A+1 (x, y)j(y) . (5.32)
2 Ω Ω

If A is an operator of trace class on (L2 (Ω), dν) we have, thus, the validity
of the usual eigenvalue functional representation

Z[{jn }µ∈Z ]
 ∞  ∞
 ∞
 1 
= d(c λ ) exp − λ c2 X2 ({cn }n∈Z ) exp i c j
2
=1 =1 =1

with the spectral set

A−1 σ = λ σ ,
(5.33)
j = j, σ 
August 6, 2008 15:46 9in x 6in B-640 ch05

120 Lecture Notes in Applied Differential Equations of Mathematical Physics

and the characteristic function set


 ∞
 
1 if c2n < ∞,
X2 ({cn }n∈Z ) = (5.34)


n=0
0 otherwise.

It is instructive to point out the usual Hermite functional basis (see 5.4,
Ref. 5) as a complete set in L2 (E alg , dµ(h)), only if the Gaussian Kol-
mogoroff measure dµ(h) is of the class above studied.
A criticism to the usual framework to construct Euclidean field theories
is that it is very cumbersome to analyze the infinite volume limit from the
Schwinger generating functional defined originally on compact space times.
In two dimensions the use of the result that the massive scalar field theory
generating functional
 
1
exp − d2 x d2 yj(x)(−∆ + m2 )−1 (x, y)j(y) (5.35)
2 R2 R2

with j(x) ∈ S(R2 ) is given by the limit of finite volume Dirichlet field
theories
    L
1 L 0 T 1
lim exp − dx dx dy 0
L→∞ 2 −L −T −L
T →∞
 T 
1 0 1 2 −1 1 1 0 0 0 1
× dy j(x , x )(−∆D + m ) (x , y , x , y )j(y , y )
−T
(5.36)
may be considered, in our opinion, as the similar claim made that is possible
from a mathematical point of view to deduce the Fourier transforms from
Fourier series, a very, difficult mathematical task (see Appendix B).
Let us comment on the functional integral associated with Feynman
propagation of fields configurations used in geometrodynamical theories in
the scalar case

G[β in (x); β out (x), T ](j)


    +∞     
1 T d2
= exp − dt d x φ − 2 + A φ (x, t)
ν
φ(x,0)=β in (x) 2 0 dt
out −∞
φ(x,T )=β (x)
  
T +∞
× exp i dt dν xj(x)t)φ(x, t) . (5.37)
0 −∞
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 121

If we define the formal functional integral by means of the eigenfunctions


of the self-adjoint Elliptic operator A, namely:

φ(x, t) = φk (t)ψk (x), (5.38)
{k}

where

Aψk (x) = (λk )2 ψk (x) (5.39)

it is straightforward to see that Eq. (5.36) is formally exactly evaluated in


terms of an infinite product of usual Feynman Wiener path measures

G[β in (x); β out (x), T ](j)



= DF [ck (t)]
ck (0)=φk (0)
{k} ck (T )=φk (T )
        
1 T
d2 2
T
× exp − ck − 2 + λk ck (t)dt exp i dtjk (t)ck (t)
2 0 dt 0
% &
λk λk
= + exp − (φ2k (T )
sin(λk T ) 2 sin(λk T )
{k}

+ φ2k (0)) cos(λk T ) − 2φk (0)φk (T )
 
2φk (T ) T 2φk (0) T
− dtjk (t) sin(λk t) − dtjk (t) sin(λk (T − t))
λk 0 λk 0
  T  t '
2
× − dt dsjk (t)jk (s) sin(λk (T − t)) sin(λk s) .
(λk )2 0 0
(5.40)

Unfortunately, our theorems do not apply in a straightforward way to


infinite (continuum) measure product of Wiener measures in Eq. (5.40) to
produce a sensible measure theory on the functional space of the infinite
product of Wiener trajectories {ck (t)} (note that for each fixed x, a sample
field configuration φ(t, 0) in Eq. (5.36) is a Hölder continuous function,
result opposite to the usual functional integral representation for the
Schwinger generating functional equations (5.1)–(5.5)), where it does not
make mathematical sense to consider a fixed point distribution φ(t, 0) —
see Sec. 5.4, Eq. (5.74).
August 6, 2008 15:46 9in x 6in B-640 ch05

122 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let us call to attention that there is still a formal definition of the above
Feynman path propagator for fields equation (5.37) which at large time
T → +∞ gives formally the quantum field functional integral equation
(5.5) associated with the Schwinger generating functional.
We thus consider the functional domain for Eq. (5.37) as composed of
field configurations which has a classical piece added with another fluctu-
ating component to be functionally integrated out, namely
σ(x, t) = σCL (x, t) + σq (x, t). (5.41)
Here, the classical field configuration problem (added with all zero modes
of the free theory) defined by the kinetic term £
 
d2
− 2 + £ σ CL (x, t) = j(x, t), (5.42)
dt
with
σ CL (x, −T ) = β1 (x); σ CL (x, T ) = β2 (x), (5.43)
namely,
 −1
d2
σCL (x, t) = − 2 + £ j(x, t) + (all projection on zero modes of £).
dt
(5.44)
As a consequence of the decomposition equation (5.43), the formal geomet-
rical propagator with an external source is given by

G[β1 (x), β2 (x), T, [j]] = D[σ(x, t)]
σ(x,−T )=β1 (x)
σ(x,+T )=β2 (x)
   
1 T
d2
× exp − dtd xσ(x, t) − 2 + £ σ(x, t)
ν
2 −T dt
  
T
× exp i dt dν xj(x, t)σ(x, t) (5.45)
−T

may be defined by the following mathematically well-defined Gaussian func-


tional measure
   
1 T
exp − dt dν xj(x, t)σ CL (x, t) σq (x,−T )=0
dσq (x, t)
2 −T
σq (x,+T )=0
     
1 T
d2
× exp − dt d xσq (x, t) − 2 + £ σq (x, t) .
ν
(5.46)
2 −T dt
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 123

The above claim is a consequence of the result below


      
1 T d2
σq (x,−T )=0
D[σq (x, t)] exp − dt d xσq (x, t) − 2 + £ σq (x, t)
D
2 −T dt
σq (x,T )=0
( d2 )
−1
= detDir2 − 2 + £ , (5.47)
dt
where the subscript Dirichlet on the functional determinant means that one
must impose formally the Dirichlet condition on the domain of the operator
d2 
(− dt2 +£) on D (R
D
×[−T, T ]) (or L2 (RD ×[−T, T ] if £−1 belongs to trace
class). Note that the operator £ in Eq. (5.46) does not have zero modes by
the construction of Eq. (5.41).
At this point, we remark that at the limit T → +∞, Eq. (5.45) is
exactly the quantum field functional equation (5.5) if one takes β1 (x) =
β2 (x) = 0 (note that the classical vacuum limit T → ∞ of Wiener measures
is mathematically ill-defined (see Theorem 5.1 of Ref. 1).
It is an important point to remark that σCL (x, t) is a regular

C ([−T, T ] × Ω) solution of the Elliptic problem equation (5.42) and
the fluctuating component σq (x, t) is a Schwartz distribution in view of
d2
the Minlos-Dao Xing Theorem 1, since the elliptic operator − dt 2 + £ in

Eq. (5.47) acts now on D ([−T, T ] × Ω) with the range D([−T, T ] × Ω),
which by its turn shows the difference between this framework and the pre-
vious one related to the infinite product of Wiener measures since these
objects are functional measures in different functional spaces.
Finally, we comment that functional Schrodinger equation, may be
mathematically defined for the above displayed field propagators equation
(5.37) only in the situation of Eq. (5.40). For instance, with £ = −∆ (the
Laplacean), we have the validity of the Euclidean field wave equation for
the geometrodynamical path-integral equation (5.37)


G[β1 (x), β2 (x), T, [j]]
∂T
 ( )
δ2
= dν x + 2 − |∇β2 (x)|2 + j(x, T ) G[β1 (x), β2 (x), T, [j]]
Ω δ β2 (x)
(5.48)

with the functional initial condition

lim G[β1 (x), β2 (x), T ] = δ (F ) (β1 (x) − β2 (x)). (5.49)


T →0+
August 6, 2008 15:46 9in x 6in B-640 ch05

124 Lecture Notes in Applied Differential Equations of Mathematical Physics

5.4. Some Rigorous Quantum Field Path-Integral


in the Analytical Regularization Scheme

In this core section, we address the important problem of producing


concrete non-trivial examples of mathematically well-defined (in the
ultra-violet region!) path-integrals in the context of the exposed theorems
in the previous sections of this paper, especially Sec. 5.2, Eq. (5.11).
Let us thus start our analysis by considering the Gaussian measure
associated with the (infrared regularized) α-power (α > 1) of the Laplacean
acting on L2 (R2 ) as an operational quadratic form (the Stone spectral
theorem)

α
(−∆)ε = (λ)α dE(λ), (5.50a)
εIR ≤λ

(0) 1
Zα,ε IR
[j] = exp − j, (−∆)−α ε jL2 (R2 )
2

= d(0)
α,ε µ[ϕ] exp(ij, ϕL2 (R2 ) ). (5.50b)

Here, εIR > 0 denotes the infrared cut-off.


It is worth to call the readers’ attention that due to the infrared regu-
larization introduced in Eq. (5.50a), the domain of the Gaussian measure
is given by the space of square integrable functions on R2 by the Minlos
theorem of Sec. 3, since for α > 1, the operator (−∆)−α εIR defines a class
trace operator on L2 (R2 ), namely

1
Tr 1 ((−∆)εIR ) = d2 k
H −α
< ∞. (5.50c)
(|K|2α + εIR )
This is the only point of our analysis where it is needed to consider the
infrared cut-off considered on the spectral resolution equation (5.50a). As
a consequence of the above remarks, one can analyze the ultraviolet renor-
malization program in the following interacting model proposed by us and
defined by an interaction gbare V (ϕ(x)) with V (x) denoting a function on
R such that it posseses an essentially bounded Fourier transform and gbare
denoting the positive bare coupling constant.
Let us show that by defining a renormalized coupling constant as (with
gren < 1)
gren
gbare = (5.51)
(1 − α)1/2
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 125

one can show that the interaction function



exp −gbare (α) d2 xV (ϕ(x)) (5.52)

(0)
is an integrable function on L1 (L2 (R2 ), dα,εIR µ[ϕ]) and leads to a well-
defined ultraviolet path-integral in the limit of α → 1.
The proof is based on the following estimates:
Since almost everywhere we have the pointwise limit

exp −gbare (α) d2 xV (ϕ(x))
 
N
(−1)n (gbare (α))n
× lim dk1 · · · dkn Ṽ (k1 ) · · · Ṽ (kn )
N →∞ n! R
n=0

× dx1 · · · dxn eik1 ϕ(x1 ) · · · eikn ϕ(xn ) (5.53)
R2

we have that the upper-bound estimate below holds true


 
 ∞
(−1)n (gbare (α))n

|ZεIR [gbare ]| ≤ 
α
dk1 · · · dkn Ṽ (k1 ) · · · Ṽ (kn )
 n! R2
n=0
  
P 
(0) i N ) 
× dx1 · · · dxn dα,εIR µ[ϕ](e =1 k ϕ(x 
).
R2 
(5.54a)

with
 
ZεαIR [gbare ] = d(0)
α,εIR µ[ϕ] exp −gbare (α) d2 xV (ϕ(x)) (5.54b)

(0)
we have, thus, the more suitable form after realizing d2 ki and dα,εIR µ[ϕ]
integrals, respectively

(gbare (α))n  n
∞
|ZεαIR =0 [gbare ]| ≤ ||Ṽ ||L∞ (R)
n=0
n!
 
 
×  dx1 · · · dxn det− 2 [Gα (xi , xj )] 1≤i≤N  .
1
(N )
(5.55)
1≤j≤N
August 6, 2008 15:46 9in x 6in B-640 ch05

126 Lecture Notes in Applied Differential Equations of Mathematical Physics

(N )
Here, [Gα (xi , xj )] 1≤i≤N denotes the N × N symmetric matrix with (i, j)
1≤j≤N
entry given by the Green-function of the α-Laplacean (without the infrared
cut-off here! and the needed normalization factors!).

Γ(1 − α)
Gα (xi , xj ) = |xi − xj |2(1−α) . (5.56)
Γ(α)

At this point, we call the readers’ attention that we have the formulae on
the asymptotic behavior for α → 1.
    1
 (N − 1)(−1)N  − 2
lim det − 12 (N )
[Gα (xi , xj )] ∼ (1 − α) N/2 
×   .
α→1 π N/2 
α>1
(5.57)

After substituting Eq. (5.57) into Eq. (5.55) and taking into account the
hypothesis of the compact support of the nonlinear V (x) (for instance:
supp V (x) ⊂ [0, 1]), one obtains the finite bound for any value grem > 0,
and producing a proof for the convergence of the perturbative expansion in
terms of the renormalized coupling constant.

 n  n
( Ṽ L∞ (R) ) gren (1n )
lim |ZεαIR =0 [gbare (α)]| ≤ 1 √ (1 − α)n/2
α→1
n=0
n! (1 − α) 2 n

≤ egren Ṽ L∞ (R) < ∞. (5.58)

Another important rigorously defined functional integral is to consider the


following α-power Klein Gordon operator on Euclidean space time

L = (−∆)α + m2 (5.59)

with m2 a positive “mass” parameters.


Let us note that L−1 is an operator of class trace on L2 (Rν ) if and only
if the result below holds true
 *π
1 ν νπ +
TrL2 (Rν ) (L−1 ) = dν k 2α = C̄(ν)m (α −2)
cosec < ∞,
k + m2 2α 2α
(5.60)
namely if
ν
α> . (5.61)
2
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 127

In this case, let us consider the double functional integral with functional
domain L2 (Rν )
 
(0) (0)
Z[j, k] = dG β[v(x)] d(−∆)α +v+m2 µ[ϕ]

× exp i dν x(j(x)ϕ(x) + k(x)v(x)) , (5.62)

where the Gaussian functional integral


, on the fields V (x) has a Gaussian
generating functional defined by a 1 -integral operator with a positive
defined kernel g(|x − y|), namely
 
(0)
Z (0) [k] = dG β[v(x)] exp i dν xk(x)v(x)
 
1
= exp − dν x dν y(k(x)g(|x − y|)k(x)) . (5.63)
2

By a simple direct application of the Fubbini–Tonelli theorem on the


exchange of the integration order in Eq. (5.62), lead us to the effective λϕ4 -
like well-defined functional integral representation

(0)
Zeff [j] = d((−∆)α +m2 ) µ[ϕ(x)]

1
× exp − dν xdν y|ϕ(x)|2 g(|x − y|)|ϕ(y)|2
2

× exp i dν xj(x)ϕ(x) . (5.64)

Note that if one introduces from the begining a bare mass parameters m2bare
depending on the parameters α, but such that it always satisfies Eq. (5.60)
one should obtain again Eq. (5.64) as a well-defined measure on L2 (Rν ).
Of course, that the usual pure Laplacean limit of α → 1 in Eq. (5.59), will
need a renormalization of this mass parameters (limα→1 m2bare (α) = +∞!)
as much as its done in the previous example.
Let us continue our examples by showing again the usefulness of
the precise determination of the functional-distributional structure of the
domain of the functional integrals in order to construct rigorously these
path-integrals without complicated limit procedures.
August 6, 2008 15:46 9in x 6in B-640 ch05

128 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let us consider a general Rν Gaussian measure defined by the generating


functional on S(Rν ) defined by the α-power of the Laplacean operator −∆
acting on S(Rν ) with a of small infrared regularization mass parameter µ2
1- .
Z(0) [j] = exp − j, ((−∆)α + µ20 )−1 j L2 (Rν )
2

= d(0)
α µ[ϕ] exp(iϕ(j)). (5.65)
E alg (S(Rν ))

An explicit expression in momentum space for the Green function of the


α-power of (−∆)α + µ20 is given by
  
+α 2 −1 dν k ik(x−y) 1
((−∆) + µ0 ) (x − y) = e . (5.66)
(2π)ν k 2α + µ20
Here, C̄(ν) is a ν-dependent (finite for ν-values!) normalization factor.
Let us suppose that there is a range of α-power values that can be
chosen in such a way that one satisfies the constraint below

d(0)
α µ[ϕ]( ϕ L2j (Rν ) )
2j
<∞ (5.67)
E alg (S(Rν ))

with j = 1, 2, . . . , N and for a given fixed integer N , the highest power


of our polynomial field interaction. Or equivalently, after realizing the ϕ-
Gaussian functional integration, with a space-time cut-off volume Ω on the
interaction to be analyzed in Eq. (5.70)
  j
ν α 2 −j dν k
d x[(−∆) + µ0 ] (x, x) = vol(Ω)
Ω k 2α + µ20
ν
π νπ 
= Cν (µ0 )( α −2) cosec < ∞. (5.68)
2α 2α
For α > ν−1
2 , one can see by the Minlos theorem that the measure support
of the Gaussian measure equation (5.65) will be given by the intersection
Banach space of measurable Lebesgue functions on Rν instead of the pre-
vious one E alg (S(Rν ))
N
L2N (Rν ) = (L2j (Rν )). (5.69)
j=1

In this case, one obtains that the finite-volume p(ϕ)2 interactions


 
  N  
exp − λ2j (ϕ2 (x))j dx ≤ 1 (5.70)
 Ω 
j=1
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 129

is mathematically well-defined as the usual pointwise product of measurable


functions and for positive coupling constant values λ2j ≥ 0. As a conse-
(0)
quence, we have a measurable functional on L1 (L2N (Rν ); dα µ[ϕ]) (since
it is bounded by the function 1). So, it would make sense to consider math-
ematically the well-defined path-integral on the full space Rν with those
values of the power α satisfying the contraint equation (5.67).
 
   N  
Z[j] = d(0) µ[ϕ] exp − λ2j ϕ2j
(x)dx
L2N (Rν )
α
 Ω 
j=1
  
× exp i j(x)ϕ(x) . (5.71)

Finally, let us consider an interacting field theory in a compact space-time


Ω ⊂ Rν defined by an integer even power 2n of the Laplacean operator with
Dirichlet boundary conditions as the free Gaussian kinetic action, namely
1
Z (0) [j] = exp − j, (−∆)−2n jL2 (Ω)
2

(0)
= d(2n) µ[ϕ] exp(ij, ϕL2 (Ω) ). (5.72)
W2n (Ω)

Here, ϕ ∈ W2n (Ω) is the Sobolev space of order n which is the functional
domain of the cylindrical Fourier transform measure of the generating func-
tional Z (0) [j], a continuous bilinear positive form on W2−n (Ω) (the topo-
logical dual of W2n (Ω)).
By a straightforward application of the well-known Sobolev immersion
theorem, we have that for the case of
ν
n−k > (5.73)
2
including k a real number of the functional Sobolev space W2n (Ω) is con-
tained in the continuously fractional differentiable space of functions C k (Ω).
As a consequence, the domain of the Bosonic functional integral can be
further reduced to C k (Ω) in the situation of Eq. (5.73)

(0)
Z (0) [j] = d(2n) µ[ϕ] exp(ij, ϕL2 (Ω) ). (5.74)
C k (Ω)

That is our new result generalizing the Wiener theorem on Brownian paths
in the case of n = 1 , k = 12 , and ν = 1.
August 6, 2008 15:46 9in x 6in B-640 ch05

130 Lecture Notes in Applied Differential Equations of Mathematical Physics

Since the Bosonic functional domain in Eq. (5.74) is formed by real


functions and not by distributions, we can see straightforwardly that any
interaction of the form

exp −g F (ϕ(x))dν x (5.75)

with the non-linearity F (x) denoting a lower bounded real function (γ > 0)

F (x) ≥ −γ (5.76)

is well-defined and is an integrable function on the functional space


(0)
(C k (Ω), d(2n) µ[ϕ]) by the direct application of the Lebesque theorem
  
 
 exp −g F (ϕ(x))dν x  ≤ exp{+gγ}. (5.77)
 

At this point we make a subtle mathematical remark that the infinite


volume limit of Eqs. (5.74) and (5.75) is very difficult, since one loses the
Garding–Poincaré inequality at this limit for those elliptic operators and,
thus, the very important Sobolev theorem. The probable correct procedure
to consider the thermodynamic limit in our Bosonic path-integrals is to
consider solely a volume cut-off on the interaction term Gaussian action as
in Eq. (5.71) and there search for vol(Ω) → ∞.
As a last remark related to Eq. (5.73) one can see that a kind of “fishnet”
exponential generating functional

1
Z (0) [j] = exp − j, exp{−α∆}jL2 (Ω) (5.78)
2

has a Fourier transformed functional integral representation defined on


the space of the infinitely differentiable functions C ∞ (Ω), which physically
means that all field configurations making the domain of such path-integral
has a strong behavior-like purely nice smooth classical field configurations.
As a general conclusion, we can see that the technical knowledge of
the support of measures on infinite-dimensional spaces — specially the
powerful Minlos theorem of Sec. 3 is very important for a deep mathe-
matical physical understanding into one of the most important problem
is quantum field theory and turbulence which is the problem related to
the appearance of ultraviolet (short-distance) divergences on perturbative
path-integral calculations.
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 131

5.5. Remarks on the Theory of Integration of Functionals


on Distributional Spaces and Hilbert–Banach spaces

Let us first consider a given vector space E with a Hilbertian structure , ,


namely H = (E,  ), where   means a inner product and H a complete
topological space with the metrical structure induced by the given , . We
have, thus, the famous Minlos theorem on the support of the cylindrical
measure associated with a given , quadratic form defined by a positive def-
inite class trace operator A ∈ 1 (H, H) (see Appendix for a discussion on
Fourier transforms in vector spaces of infinite-dimension)

1 1 1 1
exp − b, Ab = exp − |A| 2 b, |A| 2 b = dA µ(v) · exp(iv, b)
2 2 H
(5.79)

since any given class trace operator can always be considered as the
composition of two Hilbert–Schmidt, each one defined by a function on
L2 (M × M, dν ⊗ dν).
Here, the cylindrical measure dA µ(v), defined already on the vector
space of the linear forms of E, with the topology of pontual convergence —
the so-called algebraic dual of E — has its support concentrated on the
Hilbert spaces H, through the isomorphism of H and its dual H by means
of the Riesz theorem.
This result can be understood more easily, if one represents the given
Hilbert space H as a square-integrable space of measurable functions on a
complete measure space (M, dν)L2 (M, dν). In this case, the class trace pos-
itive definite operator is represented by an integral operator with a positive-
definite Kenel K(x, y). Note that it is worth to re-write Eq. (5.79) in the
Feynman path-integral notation as

1
exp − dν(x)dν(y)f (x)K(x, s)f (s)
2 M×M
 
1
= dϕ(x)
Z(0) L2 (M,dν)
x∈M
&  '
1
× exp − dν(x)dν(y)ϕ(x)K −1 (x, y)ϕ(y)
2 M×M

× exp dν(x)f (x)ϕ(x) . (5.80)
M
August 6, 2008 15:46 9in x 6in B-640 ch05

132 Lecture Notes in Applied Differential Equations of Mathematical Physics

Here, the “inverse Kenel” of the operator A is given by the relation



dν(y)K(x, y)K −1 (y, x ) = identity operator (5.81)
M
and the path-integral normalization factor is given by the functional deter-
1 1
minant Z(0) = det− 2 (K) = det 2 (K −1 ).
A more invariant and rigorous representation for the Gaussian
path-integral equations (5.79) and (5.80) can be exposed through an
eigenfunction–eigenvalue harmonic expansion associated with our given
class trace operator A, namely
Aβn = λn βµ , (5.82)

 ∞

v= vn βn ; ϕ= ϕn βn , (5.83)
n=0 n=0
 ∞

1
exp − λn |vn |2
2 n=0

= lim sup dϕ|ϕ1  . . . dϕ|ϕN 
N RN

  1  
1  |ϕ|ϕn |2
N N
1
2

N 
× exp − exp i ϕn v̄n .
2 n=0 λn 2πλn 
n=0 n=0

(5.84)
The above cited theorem for the support characterization of Gaussian
path-integrals can be generalized to the highly non-trivial case of a non-
linear functional Z(v) on E, satisfying the following conditions:
(a) Z(0) = 1,


N
(b) Z(vj − vk )zj z̄k ≥ 0, for any {zi }1≤i≤N : zi ∈ N
j,k

and {vi }1≤i≤N : vi ∈ E. (5.85)

(c) there is an H-subspace of E with an inner product , , such that Z(v)


is continuous in relation to a given inner product , A coming from
a quadratic form defined by a positive definite class trace operator A
on H. In others words, we have the sequential continuity criterium
(if H is separable):
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 133

lim Z(vn ) = 0, (5.86)


n→∞

if
 / 
lim vn , Avn  = 0 A ∈ (H) (5.87)
n→∞ 1

We have, thus, the following path-integral representation:



Z(v) = dµ(ϕ) exp(iv, ϕ), (5.88)
H

where

dµ(ϕ) = 1. (5.89)
H

Another less mathematically rigorous result is that the one related to an


inversible self-adjoint positive-definite operator A in a given Hilbert space
(H, , ) — not necessarily a bounded operator in the class trace operator as
considered previously. In order to write (more or less) formal path-integrals
representations for the Gaussian functional

1
Z(j) = exp − j, A−1 j (5.90)
2

with f ∈ Dom(A−1 ) ⊂ H, we start by considering the usual spectral


expression for the following quadratic form:

ϕ, Aϕ = λϕ, dE(λ)ϕ (5.91)
σ(A)

with σ(A) denoting the spectrum of A (a subset of R+ !) and dE(λ) are the
spectral projections associated with the spectral representation of A.
In this case one has the result for the path-integral weight as
  
1 1
exp − λϕ, dE(λ)ϕ = exp − ϕ, Aϕ
2 σ(A) 2
 
  
1
= lim sup exp − ϕ, λdE(λ)ϕ . (5.92)
 2 
λ∈σFin (A)

Here σFin (A) denotes all subsets with a finite number of elements of σ(A).
August 6, 2008 15:46 9in x 6in B-640 ch05

134 Lecture Notes in Applied Differential Equations of Mathematical Physics

As a consequence, one should define formally the generating


functional as
1
Z(j) = exp − j, A−1 j
2
 0 
  +∞
1 2 λ
= lim sup dxλ · e− 2 λ(xλ ) eijλ xλ (5.93)
 −∞ 2π 
λ∈σFin (A)

associated with the self-adjoint operator A acting on a Hilbert space H



Z(j) = dA µ(ϕ)eij,ϕ . (5.94)
H

Otherwise, one should introduce formal redefinitions of parameters entering


in the definition of our action operator A, in such a way to render finite the
functional determinant in Eq. (5.80). Let us exemplify such calculational
point with the operator (−∆ + m2 ), acting an L2 (RN ) (with domain being
given precisely by the Sobolev space H 2 (RN )). We note that

TrL2 (RN ) (exp(−t(−∆ + m2 )))


 2 & +∞ '
1 2 2
= √ dn ke−tk e−tm
2π −∞
 0

 (N − 2)!! π

 N −1 if N − 1 is even,
2(2t) 2 t
= C(N ) (5.95)

 ((N − 2/2))! 1

 N if N − 1 is odd.
2 (t) 2

with C(N ) denoting an N -dependent constant.


It is worth to note that one must introduce in the path-integral equation
(5.93), some formal definition for the functional determinant of the self-
adjoint operator A, which by its term leads to the formal process of the
“infinite renormalization” in quantum field path-integrals
 
   +∞
dxλ − 1 λ(xλ )2 
dA µ(ϕ) = lim sup √ e 2
H (R-valued net)  −∞ 2π 
λ∈σFin (A)
 
1  −1
= lim sup  √ = detF 2 [A]
λ
λ∈σFin (A)
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 135

 1 2
 
1 1/ε
dt −tλ 
= lim+ exp + e
ε→0  2 ε t 
λ∈σFin (A)
   
1/ε
1 dt
= lim exp + T r(e−tA ) . (5.96)
ε→0+ 2 ε t

In the case of the finitude of the right-hand side of Eq. (5.94) (which
means that e−tA is a class trace operator and the finitude up the proper-
time parameter t), one can proceed as in the appendix to define mathemat-
ically the Gaussian path-integral.
Let us now consider the proper-time (Cauchy principal value sense)
integration process as indicated by Eq. (5.95), for the case of N an even
space-time dimensionality
 1/ε 2  1/ε 2
dt e−tm e−tm
I(m2 , ε) = = dt N +2 (5.97)
ε t tN/2 ε t 2

The whole idea of the renormalization/regularization program means a


(non-unique) choice of the mass parameter as a function of the proper-time
cut-off ε in such a way that the infinite limit of ε → 0+ turns out to be
finite, namely

lim I(m2 (ε), ε) < ∞. (5.98)


ε→0+

A slight generalization of the above exposed Minlos–Bochner theorem in


Hilbert spaces is the following theorem:

Theorem 5.3. Let (H, , 1 ) be a separable Hilbert space with the inner
product , 1 . Let (H0 , , 0 ) be a subspace of H, so there is a trace class
operator T : H → H, such that the inner product , 0 is given explicitly
by g, h1 = g, T h0 = T 1/2 g, T 1/2 h0 . Let us, thus, consider a positive
1
definite functional Z(j) ∈ C((H, , 1 ), R) [if jn −→ j, then Z(jn ) → Z(j)
on R+ ]. We obtain that the Bochner path-integral representation of Z(j)
is given by a measure supported at these linear functionals, such that their
restrictions in the subspace H0 are continuous by the norm induced by the
“trace-class” inner product , 1 .

With this result in our hands, it became more or less straight-


forward to analyze the cylindrical measure supports in distributional spaces.
August 6, 2008 15:46 9in x 6in B-640 ch05

136 Lecture Notes in Applied Differential Equations of Mathematical Physics

For instance, the basic Euclidean quantum field distributional spaces of


tempered distributions in RN : S  (RN ), can always be seen as the strong
topological dual of the inductive limit of Hilbert spaces below considered



sp = (xn ) ∈ C  lim np xn
n→∞




2p
= 0, with the inner product (xn ), (yn )sp = n xn ȳn . (5.99)
n=1

We note now that



S(RN ) = sp (5.100)
p≥1

and

!
 N
S (R ) = s−p . (5.101)
p≥1

An important property is that sp ⊃ sp+1 and they satisfy the hypothesis


of Theorem 5.3, since

 ∞
 1 2
n2p |xn |2 = n(2p+2) |x | (5.102)
n=1 n=1
n2 n

and
∞
1 π2
= . (5.103)
n=1
n2 6

As a consequence
π2
||(xn )||sp ≤ ||(xn )||sp+1 (5.104)
6
If one has an arbitrary continuous positive-definite functional in S(RN ),
necessarily its measure support will always be on the topological dual of
s , for each p. As a consequence, its support will be on the union set
#p+1

p=1 s−p (since s−p ⊂ s−(p+1) ) and thus it will be the whole distribution
space S  (RN ).
Similar results hold true in other distributional spaces.
The application of the above-mentioned result in Gaussian path-
integrals is always made with the use of the famous result of the kernel
theorem of Schwartz–Gelfand.
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 137

Theorem 5.4 (Gelfand). Any continuous bilinear form B(j, j) defined in


the test space of the tempered distribution S  (RN ) has the following explicit
representation:

B(j, j) = dN xdN yj(x)(Dxm Dyn F )(x, y)j̄(y) (5.105)

with j ∈ S(RN ), F (x, y) is a continuous function of polynomial growth and


Dxm , Dyn are the distributional derivatives of order m and n, respectively.

In all cases of application of this result to our study presented in the


previous chapters were made in the context that (Dxm Dyn F )(x, y) is a fun-
damental solution of a given differential operator representing the kinetic
term of a given quantum field Lagrangean.
As a consequence, we have the basic result in the Gaussian path-integral
in Euclidean quantum field theory

− 12 B(j,j)
e = dµ(T ) exp{i(T (j))}, (5.106)
T ∈S  (RN )

where T (j) denotes the action of the distribution T on the test function
j ∈ S(RN ).
At this point of our exposition let us show how to produce a fundamental
solution for a given differential operator P (D) with constant coefficients,
namely

P (D) = ap D p . (5.107)
|ρ|≤m

A fundamental solution for Eq. (5.107) is given by a (numeric) distribution


E ∈ S  (RN ) such that for any ϕ ∈ S(RN ), we have:

(P (D)E)(ϕ) = δ(ϕ) = ϕ(0) (5.108)

or equivalently

E(t P (D)ϕ) = δ(ϕ), (5.109)

where t P (D) is the transpose operator through the duality of S(RN ) and
S  (RN ).
August 6, 2008 15:46 9in x 6in B-640 ch05

138 Lecture Notes in Applied Differential Equations of Mathematical Physics

By means of the use of a harmonic–Hermite expansion for the searched


fundamental solution

 ∞

S  (RN )
E = (E, Hp )Hp = Ep Hp (5.110)
p=1 p=1

with Hp denoting the appropriate Hermite polynomials in RN , together


with the test function harmonic expansion

 ∞

S(RN )
ϕ = (ϕ, Hp )Hp = ϕp Hp (5.111)
p=1 p=1

and the use of the relationship between Hermite polynomials



t
P (D)Hp = Mpq Hq (5.112)
|q|≤(p)

with (p) depending on the order of Hp and the order of the differential
operator t P (D). (For instance in S(R): (d/dx)Hn (x) = 2nHn−1 (x);

d2 d
2
Hn (x) = 2n Hn−1 (x) = 4n(n − 1)Hn−2 (x), etc.,
d x dx
one obtains the recurrence equations for the searched coefficients Ep in
Eq. (5.110)
 
  
ϕn  Mnq Eq  = ϕn Hn (0) (5.113)
n |q|≤(n) n

2
for any (ϕn ) ∈ or equivalently

Mnq Eq = Hn (0) (5.114)
|q|≤(n)

the solution of the above written infinite-dimensional system produces a set


of coefficients {Ep }p=1,...,∞ satisfying a condition that it belongs to some
space s−r , where r is the order of the fundamental solution being searched
(the rigorous proof of the above assertions is left as an exercise to our
mathematically oriented reader!).
Finally, let us sketch the connection between path-integrals and the
operator framework in Euclidean quantum field theory, both still mathe-
matically non-rigorous from a strict mathematical point of view. In other
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 139

former approach, one has a self-adjoint operator H̄(j) indexed by a set


of functions (the field classical sources of the quantum field theory under
analysis) belonging to the distributional space S(RN ). This self-adjoint
operator is formally given by the space-time integrated Lagrangean field
theory and the basic object is the generating functional as defined by the
vacuum–vacuum transition amplitude
Z(j) = eiH(j) ΩVAC , ΩVAC  (5.115)
with ΩVAC denoting the theory vacuum state. It is assumed that Z(j) is
a continuous positive definite functional on S(RN ). As a consequence of
the above exposed theorems of Minlos and Bochner, there is a cylindrical
measure dµ(T ) on S  (RN ) such that the generating functional Z(j) is repre-
sented by the quantum field path-integral defined by the above-mentioned
measure:

Z(j) = dµ(T ) exp{i(T (j)} (5.116)
S  (RN )

in which the Feynman symbolic notation express itself in the following


symbolic-operational Feynman notation:

1
Z[j(x, t)] = DF [T (x, t)]
Z(0) S  (RN )
R +∞ R +∞
1
dn−1 xdtL(T,∂t T,∂x T ) i dn−1 xdtj(x,t)T (x,t)
× e− 2 −∞ e −∞ , (5.117)
with L(T, ∂t T, ∂x T ) means generically the Lagrangean density of our field
theory under quantization.
Let us exemplify the Feynman symbolic Euclidean path-integral as given
by Eq. (5.117) in the Gaussian case (free Euclidean field massless theory in
RN , N ≥ 2)
  +∞  
1 +∞ N (−1)
exp − d x dN yj(x) j(y)
2 −∞ −∞ (n − 2)Γ(n)|x − y|n−2

= DF [T (x)]
S  (RN )
 +∞  +∞
1
× exp − N
d x dN yT (x)((−∆)x δ (N ) (x − y)T (y)
2 −∞ −∞
 +∞
+ 12 (N )
× det ((−∆)xδ (x − y)) exp i dN xj(x)T (x) .
−∞
(5.118)
August 6, 2008 15:46 9in x 6in B-640 ch05

140 Lecture Notes in Applied Differential Equations of Mathematical Physics

Or for the heat differential operator in S  (RN × R+ )


  ∞  +∞  ∞
1 +∞ N
exp − d x dt dN y dt
2 −∞ 0 −∞ 0
   
1 |x − y|2  
×  exp − θ(t − t )j(y, t )
( 2π(t − t )))N 4(t − t )

= DF [T (x, t)]
S  (RN ×R+ )
 +∞  ∞ &  '
1 ∂
× exp − dN x dtT (x, t) ∆x − T (x, t)
2 −∞ 0 ∂t
& '  +∞  ∞

× det1/2 ∆x − × exp i dN x dtJ(x, t)T (x, t) .
∂t −∞ 0

(5.119a)

It is worth to point out that the fourth-order path-integral in S  (R4 ):


  +∞  
1 +∞ 4 4 1 2
exp − d x d (x)j(x) |x − y| n(x − y) j(y)
2 −∞ −∞ 8π
 R +∞ 4
1 2
= DF [ϕ(x)]e− 2 −∞ d x(ϕ(x)(∆ )ϕ(x))
S  (R4 )
 R +∞
d4 xj(x)ϕ(x)
= d∆2 µ(ϕ)eiϕ(j) ei −∞ (5.119b)
R (R4 )

1
holds mathematically true since the locally integrable function 8π |x|2 n|x|
is a fundamental solution of the differential operator ∆ : S (R ) → S(R4 )
2  4

when acting on distributional spaces.


As a last important point of this section, we present an important result
on the geometrical characterization of massive free field on a Euclidean
space-time.
Firstly, we announce a slightly improved version of the usual Minlos
theorem.

Theorem 5.5. Let E be a nuclear space of test functions and dµ a given


σ-measure on its topologic dual with the strong topology. Let , 0 be an inner
product in E, inducing a Hilbertian structure on H0 = (E, , 0 ), after its
topological completation.
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 141

We assume the following:


(a) There is a continuous positive definite functional in H0 , Z(j), with an
associated cylindrical measure dµ.
(b) There is a Hilbert–Schmidt operator T : H0 → H0 ; invertible, such
that E ⊂ Range (T ), T −1 (E) is dense in H0 and T −1 : H0 → H0 is
continuous.
We have, thus, that the support of the measure satisfies the relationship

supp dµ ⊆ (T −1 )∗ (H0 ) ⊂ E ∗ . (5.120)

At this point we give a non-trivial application of Theorem 5.5.


Let us consider a differential inversible operator L : S  (RN ) → S(R),
together with a positive inversible self-adjoint elliptic operator P : D(P ) ⊂
L2 (RN ) → L2 (RN ). Let Hα be the following Hilbert space:
* +
Hα = S(RN ), P α ϕ, P α ϕL2 (RN ) = , α , for α a real number .
(5.121)

We can see that for α > 0, the operators below

P −α : L2 (RN ) → H+α
(5.122)
ϕ → (P −α ϕ)

P α : H+α → L2 (RN )
(5.123)
ϕ → (P α ϕ)

are isometries among the following subspaces

D(P −α ), , L2 ) and H+α

since

P −α ϕ, P −α ϕH+α = P α P −α ϕ, P α P −α ϕL2 (RN ) = ϕ, ϕL2 (RN )


(5.124)

and

P α f, P α f L2 (RN ) = f, f H+α . (5.125)


August 6, 2008 15:46 9in x 6in B-640 ch05

142 Lecture Notes in Applied Differential Equations of Mathematical Physics

If one considers T a given Hilbert–Schmidt operator on Hα , the com-


posite operator T0 = P α T P −α is an operator with domain being D(P −α )
and its image being the Range(P α ). T0 is clearly an invertible operator
and S(RN ) ⊂ Range(T ) means that the equation (T P −α )(ϕ) = f
has always a non-zero solution in D(P −α ) for any given f ∈ S(RN ).
Note that the condition T −1 (f ) be a dense subset on Range(P −α ) which
means that

T −1 f, P −α ϕL2 (RN ) = 0 (5.126)

has a unique solution of the trivial solution f ≡ 0.


Let us suppose that T −1 : S(RN ) → Hα be a continuous application
and the bilinear term (L−1 (j))(j) be a continuous application in the Hilbert
L2 L2
spaces H+α ⊃ S(RN ), namely: if jn −→ j, then L−1 : P −α jn −→
L−1 P −α j, for {jn }n∈Z and jn ∈ S(RN ).
By a direct application of Theorem 5.5, we have the result

1
Z(j) = exp − [L−1 (j)(j)] = dµ(T ) exp(iT (j)). (5.127)
2 (T −1 )∗ Hα

Here, the topological space support is given by

 
(T −1 )∗ Hα = [(P −α T0 P α )−1 ]∗ (P α (S(RN )))
= [(P α )∗ (T0−1 )∗ (P −α )∗ ]P α (S(RN ))
= P α T0−1 (L2 (RN )). (5.128)

2 
, +2 m N) : S (R ) → S(R
In the important case of L = (−∆ N N
)
∗ 2 −2β ∗
and T0 T0 = (−∆ + m ) ∈ 1 (L (R )) since Tr(T 0 T 0 ) =
2 N Γ( N )Γ(2β− N )
1
2(m2 )β
( m1 ) 2 2 Γ(β) 2 ∞ for β N4 with the choice P = (−∆ + m2 ), we
can see that the support of the measure in the path-integral representation
of the Euclidean measure field in RN may be taken as the measurable subset
below

supp {d (−∆ + m2 )u(ϕ)} = (−∆ + m2 )−α · (−∆ + m2 )+β (L2 (RN ))


(5.129)
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 143

since L−1 P −α = (−∆ + m2 )−1−α is always a bounded operator in L2 (RN )


for α > −1.
As a consequence each field configuration can be considered as a kind
of “fractional distributional” derivative of a square integrable function
written as
( N
)
ϕ(x) = (−∆ + m2 ) 4 +ε−1 f (x) (5.130)

with a function f (x) ∈ L2 (RN ) and any given ε > 0, even if originally
all fields configurations entering into the path-integral were elements of
the Schwartz tempered distribution spaces S  (RN ) certainly very “rough”
mathematical objects to characterize from a rigorous geometrical point of
view.
We have, thus, made a further reduction of the functional domain of the
free massive Euclidean scalar field of S  (RN ) to the measurable subset as
given in Eq. (5.130) denoted by W (RN )
1
exp − [(−∆ + m2 )−1 j](j)
2

= d(−∆+m2 ) µ(ϕ)eiϕ(j)
S  (RN )

2 N
) 4 +ε−1 f
= d(−∆+m2 ) µ̃(f )eif,(−∆+m L2 (RN ) . (5.131)
W (RN )⊂S  (RN )

References

1. B. Simon, Functional Integration and Quantum Physics (Academic Press,


1979).
2. B. Simon, The P (φ)2 Euclidean (quantum) field theory, in Princeton Series
in Physics (1974).
3. J. Glimm and A. Jaffe, Quantum Physics, — A Functional Integral Point of
View (Springer Verlag, 1988).
4. Y. Yamasaki, Measure on Infinite Dimensional Spaces, Vol. 5 (World Scien-
tific, 1985).
5. D.-X. Xia, Measure and Integration Theory on Infinite Dimensional Space
(Academic Press, 1972).
6. L. Schwartz, Random Measure on Arbitrary Topological Space and Cylindrical
Measures (Tata Institute, Oxford University Press, 1973).
7. K. Symanzik, J. Math. Phys. 7, 510 (1966).
8. B. Simon, J. Math. Phys. 12, 140 (1971).
9. L. C. L. Botelho, Phys. Rev. 33D, 1195 (1986).
August 6, 2008 15:46 9in x 6in B-640 ch05

144 Lecture Notes in Applied Differential Equations of Mathematical Physics

10. E. Nelson, Regular probability measures on function space, Ann. Math. 69(2),
630–643 (1959); V. Rivasseau, From Perturbative to Constructive Renormal-
ization (Princeton University, 1991).
11. L. C. L. Botelho, Il Nuovo Cimento 117B, 37 (2002); Il Nuovo Cimento
117B, 331 (2002); J. Phys. A: Math. Gen. 34, 131–137 (2001); Mod.
Phys. Lett. 173, Vol. 13/14, 733 (2003); Il Nuovo Cimento 117B, 15
(2002); Ya. G. Sinai, Topics in Ergodic Theory (Princeton University Press,
1994).
12. R. Teman, Infinite-Dimensional Systems in Mechanics and Physics, Vol. 68
(Springer-Verlag, 1988); J. M. Mourão, T. Thiemann and J. M. Velhinho, J.
Math. Phys. 40(5), 2337 (1995).
13. A. M. Polyakov, Phys. Lett. 103B, 207 (1981); L. C. L. Botelho, J. Math.
Phys. 30, 2160 (1989).
14. L. C. L. Botelho, Phys. Rev. 49D, 1975 (1994).
15. A. M. Polyakov, Phys. Lett. 103B, 211 (1981).
16. L. Brink and J. Schwarz, Nucl. Phys. B121, 285 (1977).
17. L. C. L Botelho, Phys. Rev. D33, 1195 (1986).
18. B. Durhuus, Quontum Theorem of String, Nordita Lectures (1982).
19. C. G. Callan Jr., S. Coleman, J. Wess and B. Zumine, Structure of Phe-
nomenological Lagrangians — II, Phys. Rev. 177, 2247 (1969).
20. L. C. L. Boteho, Path integral Bosonization for a non-renormalizable
axial four-dimensional. Fermion Model, Phys. Rev. D39(10), 3051–3054
(1989).

Appendix A

In this appendix, we have given new functional analytical proofs of the


Bochner–Martin–Kolmogorov theorem of Sec. 5.2.
Bochner–Martin–Kolmogorov Theorem (Version I). Let f : E → R
be a given real function with domain being a vector space E and satisfying
the following properties:
(1) f (0) = 1.
(2) The restriction of f to any finite-dimensional vector subspace of E is
the Fourier Transform of a real continuous function of compact support.
Then there is a measure dµ(h) on σ-algebra containing the Borelians if
the space of linear functionals of E with the topology of pontual convergence
denoted by E alg such that for any y ∈ E

f (g) = exp(ih(g))dµ(h). (A.1)
E alg
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 145

Proof. Let {êλ∈A } be a Hamel (vectorial) basis of E, and E (N ) a given


subspace of E of finite-dimensional. By the hypothesis of the theorem, we
have that the restriction of the functions to E (N ) (generated by the elements
of the Hamel basis {êλ1 , . . . , êλN } = {eλ }λ∈ΛF ) is given by the Fourier
transform
N 

f σλ êλ = Q (dPλ1 · · · dPλN )
=1 λ∈ΛF Rλ
1N 2

× exp aλ Pλ ĝ(Pλ1 , . . . , PλN ), (A.2)
=1

with ĝ(Pλ1 , . . . , PλN ) ∈ Cc ( λ∈ΛF Rλ ). 


As a consequence of the above result, we consider the following well-
defined family of linear positive functionals on the space of continuous
function on the product space of the Alexandrov compactifications of R
denoted by Rw :
1  2Dual
LλF ∈ C w λ
(R ) ; R , (A.3)
λ∈ΛF

with
 
LΛF [ĝ(Pλ1 , . . . , PλN )] = Q
ĝ(Pλ1 , . . . , PλN )(dPλ1 · · · dPλN ).
λ∈ΛF (Rw )λ

(A.4)

Here, ĝ(Pλ1 , . . . , PλN ) still denotes the unique extension of Eq. (A.2) to the
Alexandrov compactification Rw .
Now, we remark that the above family of linear continuous functionals
have the following properties:

(1) The norm of LλF is always the unity since



||Lλ1 || = Q ĝ(Pλ1 , . . . , PλN )dPλ1 · · · dPλN = 1. (A.5)
λ∈ΛF (Rw )λ

(2) If the index set ΛF , contains ΛF the restriction of the associated linear
functional ΛF , to the space C( λ∈Λf (Rw )λ , R) coincides with LΛF .
August 6, 2008 15:46 9in x 6in B-640 ch05

146 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now, a simple application of the Stone–Weierstrass theorem shows us


that the topological closure of the union of the subspace of functions of
finite variable is the space C( λ∈A (Rw )λ , R), namely
 
!
C (Rw )λ , R =C (Rw )λ , R , (A.6)
ΛF ⊂A λ∈ΛF λ∈A

where the union is taken over all family of subsets of finite elements of the
index set A.
As a consequence of the Remark 2 and Eq. (A.6) there is a unique
extension of the family of linear functionals {LΛF } to the whole space
C( λ∈A (Rw )λ , R) and denoted by L∞ . The Riesz Markov theorem gives
us a unique measure dµ̄(h) on λ∈A (Rw )λ representing the action of this
functional on C( λ∈A (Rw )λ , R).
We have, thus, the following functional integral representation for the
function f (g):

f (g) = Q exp(ih̄(g))dµ(h̄). (A.7)
( w λ
λ∈A (R ) )

$N
Or equivalently (since h̄(g) = i=1 pi ai for some {pi }i∈N < ∞), we have
the result

f (g) = Q (exp ih(g))dµ(h) (A.8)
( λ∈A Rλ )

which is the proposed theorem with h ∈ ( λ∈λF Rλ ) being the element


which has the image of h̄ on the Alexandrov compactification λ∈λF (Rw )λ .
The practical use of the Bochner–Martin–Kolmogorov theorem is dif-
ficult by the present day’ non-existence of an algoritm generating explicitly
a Hamel (vectorial) basis on function of spaces. However, if one is able to
apply the theorems of Sec. 5.3, one can construct explicitly the functional
measure by only considering the topological basis as in the Gaussian func-
tional integral equation (5.32).
Bochner–Martin–Kolmogorov Theorem (Version II). We now have
the same hypothesis and results of Theorem Version 1 but with the more
general condition.
(3) The restriction of f to any finite-dimensional vector subspace of E is the
Fourier Transform of a real continuous function vanishing at “infinite.”
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 147

For the proof of the theorem under this more general mathematical
condition, we will need two lemmas and some definitions.

Definition A.1. Let X be a normal space, locally compact and satisfying


the following σ-compacity condition:

!
X= Kk , (A.9)
n=0

with

Kn ⊂ int(Kn+1 ) ⊂ Kn+1 . (A.10)

We define the following space of continuous function “vanishing” at


infinite:
  

C̃0 (X, R) = f (x) ∈ C(X, R) lim sup |f (x)| = 0 . (A.11)
n→∞ c x∈(Kn )

We have, thus, the following lemma.

Lemma A.1. The topological closure of the functions of compact support


contains C̃0 (X, R) in the topology of uniform convergence.

Proof. Let f (x) ∈ C̃0 (X, R) and gµ ∈ C(X, R), the (Uryhson) functions
c
associated with the closed disjoints sets K̄n and (Kn+1 ). Now it is straight-
forward to see that (f · gn )(x) ∈ Ci (X, R) and converges uniformly to f (x)
by definition (A.11).
At this point, we consider a linear positive continuous functional
L on C̃0 (X, R). Since the restriction of L to each subspace C(Kn , R)
satisfy the conditions of the Riesz–Markov theorem, there is a unique
measure µ(n) on Kn containing the Borelians on Kn and representing
this linear functional restriction. We now use the hypothesis equation
(A.10) to have a well-defined measure on a σ-algebra containing the
Borelians of X
 
µ̄(A) = lim sup µ(n) A Kn . (A.12)

for A in this σ-algebra and representing the functional L on C̃0 (X, R)



L(f ) = f (x)dµ̄(x). (A.13)
X

Note that the normal of the topological space X is a fundamental


hypothesis used in this proof by means of the Uryhson lemma.
August 6, 2008 15:46 9in x 6in B-640 ch05

148 Lecture Notes in Applied Differential Equations of Mathematical Physics

λ
Unfortunately, the non-countable product space λ∈A R is not a
normal topological space (the famous Stone counter example) and we
cannot, thus, apply the above lemma to our vectorial case equation (A.8).
However, we can overcome the use of the Stone–Weierstrass theorem in the
proof of the Bochner–Martin–Kolmogorov theorem by considering directly
a certain functional space instead of that given by Eq. (A.6).
Thus, we define the following space of infinite-dimensional functions
vanishing at finite
 
def
!
C0 (R∞ , R) ≡ C0 Rλ , R = C̃0 Rλ , R , (A.14)
λ∈A ΛF ⊂A λ∈λF

where the closure is taken in the topology of uniform convergence.


If we consider a given continuous linear functional L on C0 ( λ∈A Rλ , R)
there is a unique measure µ∞ on the union of the Borelians λ∈ΛF Rλ
representing the action of L on C0 (R∞ , R).
Conversely, given a family of consistent measures {µΛF } on
the finite-dimensional spaces ( λ∈ΛF Rλ ) satisfying the property of
µΛF ( λ∈ΛF Rλ ) = 1, there is a unique measure on the cylinders λ∈A Rλ
associated with the functional L on C0 ( λ∈A Rλ , R).
By collecting the results of the above written lemmas we will get the
proof of Eq. (A.8) in this more general case. 

Appendix B. On the Support Evaluations of Gaussian Measures

Let us show explicitly by one example of ours of the quite complex behavior
of cylindrical measures on infinite-dimensional spaces R∞ .
Firstly, we consider the family of Gaussian measures on R∞ =
{(xn )1≤n≤∞ , xn ∈ R} with σn ∈ 2 .
N   
x2
(∞) 1 − 2σn2
d µ({xn }) = lim sup dxn √ e n . (B.1)
N
n=1
σn π

Let us introduce the measurable sets on R∞


 ∞
 ∞
 
E(αn ) = (xn ) ∈ R∞ ; x 2(xn) = α2n x2n < ∞ and α2n σn2 < ∞.
n=1 =1
(B.2)
2
Here, {αn } is a given sequence suppose to belonging to either.
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 149

Now, it is straightforward to evaluate the “mass” of the infinite-


dimensional set E(xn ) , namely
 & P
'
(∞) ∞ −ε( ∞ α2n x2n )
µ(E(αn ) ) = d µ({xn }) lim e n=1

R∞ ε→0+
 1 n
2
1
= lim lim sup (1 + 2ε α2n σn2 )− 2 (B.3)
ε→0+ 0≤≤n =1

Note that
 n
1 1
(1 + 2ε α2n σn2 )− 2 ≤ $n . (B.4)
1+ =1 α2n σn2
=1

As a consequence, one can exchange the order of the limits in Eq. (B.3)
and arriving at the result
 1 n 2
1
(∞)
µ(E(αn ) ) = lim sup lim (1 + 2ε α2n σn2 )− 2
0≤≤n ε→0+
=1

= lim sup {1} = 1. (B.5)


0≤≤n

So we can conclude, on the basis of Eq. (B.5) that the support of the
measure equation (B.1) is the set of E(αn ) for any possible sequence
{αn } ∈ 2 . Let us show that (E(αn ) )C ∩ E(βn ) = {φ}, so that these sets
are not coincident.
Let the sequences be

σn = n−σ ,
αn = nσ−1 , (B.6)
βn = nσ−λ ,

with γ > 1 and σ > 0.


We have that
  1 π2
α2n σn2 = = , (B.7)
n2 6

 
βn2 σn2 = n−2λ < ∞. (B.8)

So E{αn } and E{βn } are non-empty sets on R∞ .


August 6, 2008 15:46 9in x 6in B-640 ch05

150 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let us consider the point {x̄n } ∈ R∞ and defined by the relation

x̄2n = n−2(σ−1)−ε . (B.9)

We have that
  2 
(x̄n )2 α2n = n−2(σ−1)−ε · nn (σ−1)
= n−ε

and
  
(x̄n )2 βn2 = n−2(σ−1)−ε · n2(σ−λ) = n2−ε−2λ

If we choose ε = 1; γ > 1 (γ = 32 !), we obtain that the point {x̄n }


$ 2 2
belongs to the set E{βn } (since n = π6 ), however it does not belongs
$∞
to E{xn } (since n=0 n−1 = +∞), although the support of the measure
equation (B.1) is any set of the form E{γn } with {γn } ∈ 2 .

Appendix C. Some Calculations of the Q.C.D. Fermion


Functional Determinant in Two-Dimensions
and (Q.E.D.)2 Solubility

Let us first define the functional determinant of a self-adjoint, positive def-


inite operator A (without zero modes) by the proper-time method
 ∞
dt
log detF (A) = − lim Trf (e−tA ) , (C.1)
β→0+ ε t

where the subscript f reminds us the functional nature of the objects under
study and so its trace.
It is thus expected that the definition equation (C.1) has divergents
counter terms as ε → 0+ , since exp(−tA) is a class trace operator only
for t ≥ ε. Asymptotic expressions at the short-time limit t → 0+ are well-
known in mathematical literature. However, this information is not useful in
the first sight of Eq. (C.1) since one should know TrF (e−tA ) for all t-values
in [ε, ∞).
A useful remark on the exact evaluation is in the case where the operator
A is of the form A = B + m2 1 and one is mainly interested in the effective
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 151

asymptotic limit of large mass m2 → ∞. In this particular case, one can


use a saddle-point analysis of the expression in Eq. (C.1)
 ∞ & '
2 dt −m2 t −tB
lim [log detF (B + m 1)] = − lim e lim TrF (e ) .
m2 →∞ ε→0 ε t t∈0+

(C.2)

Another very important case is covered by the (formal) Schwarz–Romanov


theorem announced below.
Theorem C.1. Let A(σ) be a one-parameter family of positive-definite self-
adjoints operators and satisfying the parameter derivative condition (0 ≤
σ ≤ 1)
d
A(σ) = fA(σ) + A(σ)g, (C.3a)

where f and A are σ-independent objects (may be operators).
Then we have an explicit result
   1
detF (A(1)) (σ) 2
log = dσ lim TrF [f e−ε(A ) ]
detF (A(0)) 0 ε→0+

 1
(σ) 2
+ dσ lim TrF [ge−ε(A ) ] . (C.3b)
0 ε→0+

The proof of Eqs. (C.3a) and (C.3b) is based on the validity of the
differential equation in relation to the σ-parameter
d * 2 2
+
[log detF (A(σ))2 ] = lim+ 2 TrF (f e−ε(A(σ)) ) + Trε (ge−ε(A(σ)) )
dσ ε→0
(C.4)

which can be seen from the obvious calculations written down in the above
equation

[log detF (A(σ))2 ]


 ∞ &
= − lim+ dt TrF (f A(σ) + A(σ)g)A(σ) + A(σ)(f A(σ)
ε→0 ε
'
−2 d 2
+ A(σ)g)(−(A(σ)) ) (exp(−t(A(σ)) ))
dt
= Eq. (C.4). (C.5)
August 6, 2008 15:46 9in x 6in B-640 ch05

152 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let us apply the above formulae in order to evaluate the functional


determinant of the “Chrially transformed” self-adjoint Dirac operator in a
two-dimensional space-time

D(σ) = exp(σγ 5 ϕ2 (x)λa )|∂|(exp(σγ 5 ϕa (x)λa ). (C.6)

Here, the Chiral Phase W [ψ] in Eq. (C.6) takes value in SU (N ) for instance.
One can see that
7 1
( D(σ))2 = −(∂µ + iGµ (σ))2 1 − [γµ , γν ] Fµν (−iGµ (σ)) (C.7a)
4
with the Gauge field

γµ Gµ = γµ (W −1 ∂µ W ). (C.7b)

Hence, the asymptotics of the operator (C.7a) are easily evaluated (see
Chap. 1, Appendix E)

lim TrF (exp(−ε(D(σ))2 ))


ε→0+

1 g(iεµν γ5 )
= lim+ Tr 1+ Fµν (−iGµ (σ)) , (C.8a)
ε→0 4πε 2
where we have used the Seeley expansions below for the square of the Dirac
operator in the presence of a non-Abelian connection

A : C0∞ (R2 ) → C0∞ (R2 ), (C.8b)

∂ ∂
Aϕ = (−∆) − V1 (ξ1 , ξ2 ) − V2 (ξ1 , ξ2 ) − V0 ((ξ1 , ξ2 ), (C.8c)
∂ξ1 ∂ξ2

8 9
lim TrCc∞ (R2 ) (e−tA )
t→0+
  
1 1 ∂ ∂ 1 1
= + − V1 + V0 − (V 2 + V22 ) − V0
4πt 8π ∂ξ1 ∂ξ2 16π 1 4π
× (ξ1 , ξ2 ) + O(t), (C.8d)

7 7
( ∂ − ig Gµ )2 = (−∂ 2 )ξ + (2igGµ ∂m u)ξ
& '
igσ µν 2 2
+ ig(∂µ Gµ ) + Fµν (G) + g Gµ , (C.8e)
2 ξ
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 153

which leads to the following exact integral (non-local) representation for


the non-Abelian Dirac determinant

detF (D(1))
log
detF (∂)
 & 1 '
i 2 a µν
= d x TrSU(N ) ϕa (x)λ dσEµν F (−iGµ (σ)) .
2π 0
(C.9)

A more invariant expression for Eq. (C.9) can be seen by considering the
decomposition of the “SU (N ) gauge field” Gµ (σ) in terms of its vectorial
and axial components:

(W −1 (σ)∂µ W (σ)) = Vµ (σ) + γ 5 Aµ (σ) (C.10)

or equivalently (γµ γ5 = itEµν Yν , γs = iγ0 γ1 , [γµ , γν ] = −2iEµν γ5 ):

Gµ (σ) = Vµ (σ) + iEµν Aν (σ). (C.11)

At this point we point out the formula

Fµν (−iGµ (σ)) = {(iDαv (σ)Aα (σ))E µν − [Aµ (σ), Aν (σ)] + Fµν (Vµ (σ))}.
(C.12)

Here

DαV Aβ = ∂α Aβ + [Vα , Aβ ]. (C.13)

Note that Aµ (σ) and Vµ (σ) are not independent fields since the chiral
phase W (σ) satisfies the integrability condition

Fµν (W ∂µ W ) ≡ 0 (C.14)

or equivalently

Fµν (Vβ (σ)) = −[Aµ (σ), Aν (σ)], (C.15)

DµV Aν (σ) = DνV Aµ (σ). (C.16)


August 6, 2008 15:46 9in x 6in B-640 ch05

154 Lecture Notes in Applied Differential Equations of Mathematical Physics

After substituting Eqs. (C.12)–(C.16) in Eq. (C.9), one can obtain the
results
    1

 i
=
 d2
x Tr λa
φ (α) dσ(2iDµ (σ)Aµ (σ)


SU(N ) a
 2π
 0



 −[Aµ (σ),Aν (σ)]

 : ;< =

 E ) − [Aµ (σ), Aν (σ)] ,

 + µν ( F µν (Vα (σ)) (C.17)



 

 i

 d2 x TrSU(N )


=
detF (D(1))  2π
log   1 
detF (∂) 
 × λ φa (x)
a


dσ(2iDµ (σ)Aµ (σ))

 0

 

 i

 + d2 x TrSU(N )



 2π

   1 



 × a


λ φa (x) dσ(−2[A µ (σ), A ν (σ)])

 0



= I1 (φ) + I2 (φ). (C.18)

Let us now show that the term I1 (φ) is a mass term for the physical gauge
field Aµ (σ = 1) = Aµ .
First, we observe the result

Lµ (σ)
: ;< =
TrDirac ⊗ TrSU(N ) γ 5 Aµ (σ) dσ
d
(W ∂µ W )(σ)

= TrDirac ⊗ TrSU(N ) {γ 5 φa (x)λa (γ 5 (∂µ Aµ (σ)) − [γ 5 Aµ (σ), Lµ (σ)])(x)}


= 2 TrSU(N ) {λa φa (x)(∂µ Aµ (σ) + [Vµ (σ), Aµ (σ)](x))}
= 2 TrSU(N ) {λa φa (x)Dµ (σ)Aµ (σ)} = I1 (φ). (C.19)

By the other side

TrDirac ⊗ TrSU(N ) {γ 5 Aµ (σ)Lµ (σ)}


 
5 d d
= TrDirac ⊗ TrSU(N ) γ Aµ (σ) Vµ (σ) + γ5 Aµ (σ)
dσ dσ
 
d
= TrSU(N ) (Aµ (σ)Aµ (σ)) . (C.20)

August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 155

At this point it is worth to see the appearance of a dynamical Higgs mech-


anism for Q.C.D. in two-dimensions.
Let us now analyze the second term I2 (φ) in Eq. (C.18)
  1 
i
exp − d2 x TrSU(N ) dσEµν (2φa (x)λa [Aµ (σ), Aν (σ)](x))
2π 0
   1
1
i 2 :;<=
= exp − d x TrDirac⊗SU(N ) dσ(γ5 γ5 Eµν φa (x)λa
2π 0

−1 −1
× [γ5 W (σ)∂µ W (σ) − γ5 Vµ (σ), γ5 W (σ)∂ν W (σ) − γ5 Vν (σ)] .

(C.21)

Since we have the identity as a consequence of the fact that [σγ5 φa (x)λa ,
γ5 φb (x)λb ] = σ(φa (x)φb (x)[λa , λb ] ≡ 0;


W −1 (σ) W (σ) = γ5 φa (x)λa , (C.22)
∂σ
we can see the appearance of a term of the form of a Wess–Zumino–Novikov
topological functional for the Chiral group SU (N ), namely
 
i ∂
I2 (φ) = exp − d x TrColor⊗Dirac Eµν γ5 W −1 (σ) W (σ)
2
2π ∂σ

× [γ5 W −1 (σ)∂µ W (σ), γ5 W −1 (σ)∂ν W (σ)]

+ terms (φ, Vµ ), (C.23)

which after the one-point compactification of the space-time to S 3 and


considering only smooth phases (φ(x) ∈ C ∞ (S 3 )) one can see that the
Wess–Zumino–Novikov functional is a homotopical class invariant. For the
closed ball S 3 × [0, 1] = ({x̄ ≡ (x1 , x2 , σ)})

d3 x̄ TrSU(N )axial (γ5 (W −1 ∂α W )(x̄))(γ5 (W −1 ∂µ W )(x̄)
S 3 ×[0,1]

× (γ5 W −1 ∂ν W ))(x̄) = απn (C.24)

with α an over all factor and n ∈ Z+ .


August 6, 2008 15:46 9in x 6in B-640 ch05

156 Lecture Notes in Applied Differential Equations of Mathematical Physics

Finally, let us call our readers attention that the Dirac operator in the
presence of a non-Abelian SU (N ) gauge field Aµ (x) = Aaµ (x)λa , can always
be re-written in the “chiral phase” in the so-called Roskies gauge fixing

ea (x) ea (x) > [φ](iγ µ ∂µ )W


> [φ].
iγ µ (∂µ − gGµ ) = eiγs φ (iγ µ ∂µ )eiγs φ =W (C.25)

Here
* Rx +
> [φ] = eiγs φea (x) ≡ PDirac PSU(N ) eiγs −∞ dξ (εµν Gν)(x) .
µ
W (C.26)

Since

> )∂µ (W
(γµ Gµ )(x) = +γµ (W > )−1 (x). (C.27)

It is worth now to use the formalism of invariant functional inte-


gration — Appendix in Chap. 1 — to change the quantization variables of
> (φ). This task is easily
the gauge field Aµ (x) to the SU (N )-axial phases W
accomplished through the use of Riemannian functional metric on the man-
ifold of the gauge connections

dS 2 = d2 x TrSU(N ) (δGµ δGµ )(x)

1
= d2 x TrSU(N )⊗Dirac [(γµ δGµ )(γ µ δGµ )](x)
4

77
= detF [ D D∗ ]adg × > W −1 )(δ W
d2 x TrSU(N )Axial [(δ W > W −1 )] ,

(C.28)

since
* +
> )W
γµ (δGµ ) = γµ ∂µ (δ W > −1 + ∂µ W
> (−W > )W
> −1 (δ W > −1 )

>W
= γµ (∂µ − [Gµ , ])(δ W > −1 )

and

>W
[γµ , δ W > −1 ] = δ W
>{γµ , W
> −1 } + {γµ , δ W
> }W
> −1 ≡ 0 (C.29)
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 157

and leading to new parametrization for the gauge field measure


? 7 7 @1
> (x)].
DF [Gµ (x)] = detF,adj ( D D∗ ) 2 DHaar [W (C.30)

 1
Here, detF,adj φφ∗ 2 is the functional Dirac operator in the presence of the
gauge field and in the adjoint SU (N )-representation. Its explicit evaluations
are left to our readers.
The full gauge-invariant expression for the Fermion determinant is con-
jectured to be given on explicit integration of the gauge parameters con-
sidered now as dynamical variables in the gauge-fixed result. For instance,
in the Abelian case and in the gauge fixed Roskies gauge equation (C.6)
result, we have the Schwinger result

detF [iγ µ (∂µ − ieAµ )]


 
1 e2 1
= D [W (x)] exp −
F
d2 x (Aµ − ∂µ W (x))2
2 π 2
 
e2
= D [W (x)] exp −
F
d2 x[A2µ + (∂µ W )2 + 2Aµ ∂µ W ]

 &   '
e2 ∂µ ∂ν
= exp − d2 x Aµ δµν − Aν (x) . (C.31)
π (−∂ 2 )

Note that the Haar measure on the Abelian Group U (1) is



2
δSW = d2 x[δ(eiW ∂µ e−iW )δ(eiW ∂µ e−iW )](x)

= d2 x(δW (−∂ 2 )δW )(x). (C.32)

Let us solve exactly the two-dimensional quantum electrodynamics.


Firstly, Eq. (C.10) takes the simple form in term of the chiral phase in
the Roskies gauge

Gµ (x) = (εµν ∂ν φ)(x). (C.33)

Now, the somewhat cumbersome non-Abelian equations (C.3a) and (C.3b)


has a straightforward form in the Abelian case

DF [Gµ ] = detF (−∆)DF [φ] (C.34)


August 6, 2008 15:46 9in x 6in B-640 ch05

158 Lecture Notes in Applied Differential Equations of Mathematical Physics

and thus, we have the exactly soluble expression for the (Q.E.D.)2 -
generating functional (a non-gauge invariant object!)

1
Z[Jµ , η, η̄] = DF [φ]DF [χ]DF (χ̄]
2
 &   '
1 e2
× exp − d2 x φ − ∂ 4 + ∂ 2 φ + Eµν (∂ν Jµ )φ (x)
2 π
 & 7'  
1 0 i∂ χ
× exp − d2 x(χ, χ̄) 7
2 i∂ 0 χ̄
 & −igγ φ ' 
s
2 e 0 η
× exp − d x(χ, χ̄)
0 e−igγs φ η̄

For instance, correlations functions are exactly solved and possessing a


“coherent state” factor given by the φ-average below (after “normal orde-
nation” at the coincident points) and explicitly given a proof of the con-
finement of the fermionic fields since one cannot assign LSZ-scattering fields
configuration for them by the Coleman theorem since then grown as the
1
factor |x − y| 2g2 at large separation distance
 
e−i(γs )x φ(x) e−i(γs )y φ(y)
φ

=  cos(ϕ(x)) cos(ϕ(y))φ 1x ⊗ 1x + γs ⊗ 1 sin(ϕ(x)) cos(ϕ(y))φ

+ γs ⊗ 1 cos(ϕ(x)) sin(ϕ(y))φ − γs ⊗ γs  sin(ϕ(x)) sin(ϕ(s))φ


? g2
−1 @
2 2
−(−∂ 2 )−1 (x,y)
= e−g −∂ + π
  
π 1 g 1
= exp + 2 K0 √ |x − y| + g|x − y| (1x ⊗ 1y ). (C.35)
g 2π π 2π

The two-point function for the two-dimensional-electromgnetic field shows


clearly the presence of a massive excitation (Fotons have acquired a mass
term by dynamical means)

d2 k 1eik(x−s)
Gµ (x)Gν (g)φ = (Eµα Eνβ )(k α k β ) 2
(2π) k 2 (k 2 + eπ )
 δ µν
1 : ;< = δ αβ k 2 eik|x−y|
= d2 k [Eµα Eνβ ] 2 . (C.36)
2π k 2 (k 2 + eπ )
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 159

As an important point of this supplementary appendix, we wish to point


out that the chirally transformed Dirac operator equation (C.6) in four-
dimensions, still have formally an exactly integrability as expressed by the
integral representation equation (C.3) (see the asymptotic expansion equa-
tions (4.38) in Chap. 4). It reads as follows
1 7 2 
det( D(σ)2 ) i
log = − d4 x TrSU(N ) φa (x)λa
det(∂)2 2π 2
& 1 '
c
× c
dσFαβ (−iGµ (σ))Fµν (−iGµ (σ))εαβµν λc λc ,
0

(C.37)

where

−iγµ Gµ (σ) = exp (σγ 5 φa (x)λa )(i∂) exp (σγ 5 φa (x)λa ) (C.38)

and we have the formulae19

−iγµ Gµ (σ) = Vµ (σ) + γ 5 Aµ (σ), (C.39a)

Aµ (σ) = ∆−1
γs φa λa {sin h(∆γs φa λa ) ◦ ∂µ (γ5 φ λa )},
a
(C.39b)

Vµ (σ) = ∆−1
γs φa λa {(1 − cos h(∆γs φa λa )) ◦ ∂µ (γ5 φ λa )}
a
(C.39c)

with the matrix operation

∆X ◦ Y = [X, Y ] (C.39d)
(n)
and ∆X denoting its n-power.
For a complete quantum field theoretic analysis of the above formulae in
an Abelian (theoretical) axial model we refer our work to L. C. L. Botelho.20
Finally and just for completeness and pedagogical purposes, let us
deduce the formal short-time expansion associated with the second-order
positive differential elliptic operator in Rν used in the previous cited
reference

L = −(∂ 2 )x + aµ (x)(∂µ )x + V (x). (C.40)

Its evolution kernel k(x, y, t) = x| − exp(−t L) |y satisfies the heat-kernel
equation

K(x, y, t) = −Lx K(x, y, t), (C.41)
∂t
August 6, 2008 15:46 9in x 6in B-640 ch05

160 Lecture Notes in Applied Differential Equations of Mathematical Physics

K(x, y, 0) = δ (ν) (x − y). (C.42)


After substituting the asymptotic expansion below into Eq. (C.41) [with
K0 (x, y, t) denoting the free kernel (aµ ≡ 0 and V ≡ 0)]
1∞ 2
t→0+ 
K(x, y, t)  K0 (x, y, t) n
t Hn (x, y) (C.43)
n=0

and by taking into account the obvious relationship for t > 0


(∂µ K0 (x, y, t))|x=y = 0. (C.44)
We obtain the following recurrence relation for the coefficients Hn (x, x):
8 9
(n + 1)Hn+1 (x, x) = − (−∂x2 )Hn (x, x) + aµ (x) · ∂µ Hn (x, x) + V (x) .
(C.45)
For the axial Abelian case in R4 , we have the result:
  
1 s µ ν
(egγs φ (i∂)egγs φ )2 = (−∂ 2 )14×4 + gγ [γ , γ ] ∂µ φ(x) ∂ν
2
+ [−gγs ∂ 2 φ + (g)2 (∂µ φ)2 ]. (C.46)
Note that in R4
H0 (x, x) = 14×4 ,

H1 (x, x) = −V (x),
1
H2 (x, x) = [−∂ 2 V + aµ ∂µ V + V 2 ](x). (C.47)
2

Appendix D. Functional Determinants Evaluations


on the Seeley Approach

In this technical appendix, we intend to highlight the mathematical eval-


uation of the functional determinants involved in the theory of random
surface path-integral.18
Let us start by considering a differential elliptic self-adjoint operator
of second-order acting on the space of infinitely differentiable functions of
compact support in R2 ,Cc∞ (R2 , Cq ) with values in Cq

A= aα (x)Dxα (D.1)
|α|≤2
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 161

with α = (α1 , α2 ) multi-indexes and


 α1  α2
1 ∂ 1 ∂
Dxα = (D.2)
i ∂x1 i ∂x2

and Aα (x) ∈ Cc∞ (R2 , Cq ).


By introducing the usual square-integrable inner product in Cc∞ (R2 , Cq )
and making the hypothesis that A is a positive definite operator, one may
consider the (contractive) semi-group generated by A and defined by the
spectral calculus

1 e−tλ
e−tA = dλ (D.3)
2πi C (λ1q×q − A)

with C being an (arbitrary) path containing the positive semi-weis λ > 0


(the spectrum of the operator A) with a counter-clockwise orientation.
According to Seeley, one must consider the symbol associated with the
resolvent pseudo-differential operator (λ1q×q − A)−1 and defined by the
relation below
2

σ(A − λ1) = eixξ (A − λ1)eixξ = Aj (x, ξ, λ). (D.4)
|j|=0

with
 

Aj (x, ξ, λ) =  aj (x)(ξ1α1 )(ξ2α2 ) 0 ≤ j < 2, (D.5)
|α|=α1 +α2 =j

 

A2 (x, ξ, λ) = −λ1q×q +  aj (x)(ξ1α1 )(ξ2α2 ) j = 2. (D.6)
|α|=α1 +α2 =j

This is basic for symbols calculations, the important scaling properties as


written below

Aj (x, cξ, c2 λ) = (c)j Aj (x, ξ, λ). (D.7)

It too is a fundamental result of the Seeley’s theory of pseudo-differential


operators that the resolvent operator (A − λ1)−1 (the associated Green
August 6, 2008 15:46 9in x 6in B-640 ch05

162 Lecture Notes in Applied Differential Equations of Mathematical Physics

functions of the operator A) has an expansion of the form in a suitable


functional space


σ((A − λ1)−1 ) = C−2−j (x, ξ, λ) (D.8)
j=0

and satisfies the relations below

σ(A − λ1) · σ((A − λ1)−1 ) = 1, (D.9)

 
1   

[(Dξα (σ(A − λ1))(x, ξ)Dxα [C−2−j (x, ξ, λ)] = 1. (D.10)
a!  
|α|≤2
j=0

Recurrence relationships for the explicitly determination of the Seeley


coefficients of the resolvent operator equation (D.8) can be obtained
through the use of the scaling properties
1 1
ξ = pξ  ; λ 2 = p(λ ) 2 , (D.11)

1
C−2−j (x, pξ  , (p(λ ) 2 )2 ) = p−(2+j) C−2−j (x, ξ  , λ ). (D.12)

After using Eq. (D.11) in Eq. (D.10) and for each integer j, comparing the
resultant power series in the variable 1/p(1 = 1 + 0( p1 ) + · · · + 0( p1 )n + · · · ),
one can get the result as

C−2 (x, ξ) = (A2 (x, ξ)−1 ) (D.13)

8 9
>
> >
>
< X =
1
0 = a2 (x, ξ)C−2−j (x, ξ) + Dξα ak (x, ξ) ((iDxα C−2− (x, ξ, λ)))
>
> α! >
>
: <j ;
k−|α|−2−=−j

(D.14)

For the explicit operator in Cc∞ (R2 , Rq ) is given below


 
∂2 ø2
A = − g11 2 + g22 2 1q×q
∂x1 ∂x2
∂ ∂
− (A1 )q×q − (A2 )q×q − (A0 )q×q (D.15)
∂x1 ∂x2
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 163

with all the coefficients in Cc∞ (R2 , Rq ), one obtains the following results
after calculations:

A2 (x, ξ, λ) = (g11 (x)ξ12 + g22 ξ22 − λ)1q×q ,


A2 (x, ξ, λ) = −iA1 (x)ξ1 − iA2 (x)ξ2 , (D.16)
A0 (x, ξ, λ) = −A0 (x),

and

C−2 (x, ξ) = (g11 (x)ξ12 + g22 (x)ξ22 − λ)−1 ,


C−3 (x, ξ) = i(A1 (x)ξ1 + A2 (x)ξ2 )(C−2 (x, ξ))2
&    '
∂ ∂
−2ig11(x)ξ1 g11 (ξ1 )2 + g22 (ξ2 )2 (C−2 (x, ξ))3
∂x1 ∂x2
&    '
∂ ∂
−2ig22(x)ξ2 g11 (ξ1 )2 + g22 (ξ2 )2 (C−2 (x, ξ))3
∂x2 ∂x2
(D.17)

By keeping in view evaluations of the heat kernel of our given differential


operator, let us write its expansion in terms of the Seeley coefficients of
(D.8):
∞  2 
−tA 1
T r(e )= d2 xσ(e−tA )(x, ξ)
j=0
2π R2

∞  2  
1 1
=
j=0
2π 2πi
& −∞  '
× d(−is)eist d2 xd2 ξC−2−j (x, ξ, −is)
+∞ R2 ×R2

   & +∞ '
1 1 2 2 is −is
= d ξ d x e C−2−j (x, ξ, )ds
j=0
t (2π)3 R2 R2 −∞ t

  
1 1 2+j 1
= d2 ξ d2 xeis t( 2 )
C−2−j (x, t 2 ξ, −is)
j=0
t (2π)3 R2 R2


    +∞
(j−2)
−3 2 2
= t 2 (2π) d ξ d x C−2−j (x, ξ, −is) .
j=0 R2 R2 −∞

(D.18)
August 6, 2008 15:46 9in x 6in B-640 ch05

164 Lecture Notes in Applied Differential Equations of Mathematical Physics

By applying Eq. (D.17) to the differential operator as given in Eq. (D.15),


we obtain the short-time Seeley expansion as an asymptotic expansion in
the variable t
    
−tA q 2 √ q 2 √ 1
T r(e )∼ d x g11 g22 + d x g11 g22 − R
4πt 4π 6
  
(− 12 divcov A )
  :  ;< =
2  1 1  ∂ √ ∂ √ 
+ d x − √ Tr  ∂x1 ( g11 g22 A1 + ∂x2 ( g11 g22 A2 ) 
 2 g11 g22  

  & '
1 (A1 )2 (A2 )2
+ d2 x − Tr + + A0 + 0(t). (D.19)
4 g11 g22
By applying the above formula for the Polyakov’s covariant path-integrals,
let us first introduce the R2 complex structure

z = x1 + x2 , z̄ = x1 − ix2 , (D.20)

∂ ∂ ∂ ∂ ∂ ∂
= −i , = +i . (D.21)
∂z ∂x1 ∂x2 ∂ z̄ ∂x1 ∂x2
For each integer j, let us define two Hilbert spaces Hj and H̄j as follows:
(a) Hj is defined as the vector space of all complex functions f (z, z̄) =
f1 (x, y) + if2 (x, y) with the following tensorial behavior under the
action of a conformal tranformation

z = z(w), (D.22)

 −j
∂w
f (z, z̄) = f˜(w, w̄). (D.23)
‘∂z
Let us introduce into Hj the following inner product

(q, f )Hj = dz z̄(ρ(z, z̄)j+1 ḡ(z, z̄)f (z, z̄)), (D.24)
R2

with ρ(z̄) is a positive continuous real-valued function of compact


support in R2 and associated with a conformal metric.

ds2 = ρ(z, z̄) dz ∧ dz̄.


August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 165

(b) H̄j is defined same as above exposed with the following tensor
 −j
∂w
f (z, z̄) = f˜(w, w̄). (D.25)
∂z

At this point, we can verify that the above written inner products are
conformal invariant
      j+1
∂z ∂ z̄  ∂w 2
(g, f )Hj = dwdw̄   ρ̃(w, w̄)
∂w ∂w  ∂z 
R2
1 −j 2 1 −j 2
∂ ∂w
× w f˜(w, w̄) (g̃(w, w̄))
∂z ∂z

= dwdw̄(ρ̃(w, w̄))j+1 f˜(w, w̄)(g̃(w, w̄)). (D.26)
R2

Let us now introduce the following weighted Cauchy–Riemann operators


with a U (1)-real valued connection A = (Az , Az̄ ) in R2 ≡ C.

(a) Lj = Hj → H̄−(j+1)
f → (ρ(z, z̄))j (∂z̄ + Az̄ )f,
(D.27)
(b) L̄j = H̄j → H(j+1)
f → (ρ(z, z̄))j (∂z + Az )f

together with the adjoint operators (Lj ϕ, f )H̄−(j+1) = ϕ, L∗j f Hj , namely:

L∗j = −L̄−(j+1) : H̄−(j+1) → Hj (and L̄∗j = −L−(j+1) )


(D.28)
f → −(ρ(z, z̄))−(j+1) (∂z + Az ))f.

For simplicity, we consider the case of Az = Az ≡ 0.


The second-order positive definite operators are given by

Lj = L∗j Lj = −L̄−(j+1) Lj : Hj → Hj ,
(D.29)
L̄ = (L̄j )∗ L̄j : −L−(j+1) L̄j : H̄−(j+1) → H̄−(j+1) .

Possesses the explicitly expressions (for ρ(z, z̄) = eϕ(z,z) )

Lj = −e−(j+1)ϕ(z,z̄) ∂z ejϕ(z,z̄) ∂z̄ , (D.30)


−(j+1)ϕ(z,z̄)
L̄j = −e ∂z̄ e jϕ(z,z̄)
∂z .
August 6, 2008 15:46 9in x 6in B-640 ch05

166 Lecture Notes in Applied Differential Equations of Mathematical Physics

They have the following Seeley expansion:


  
−tLj ρ(z, z̄) (1 + 3j)
lim TrCc∞ (R2 ) (e ) = dz dz̄ − ∆ gρ(z, z̄) ,
t→0+ 2πt j2π
(D.31)
  
ρ(z, z̄) (2 + 3j)
lim TrCc∞ (R2 ) (e−tL̄j ) = dz dz̄ + ∆ gρ(z, z̄) .
t→0+ 2πt j2π
(D.32)
The above expressions are from the Seeley asymptotic expansion
lim TrCc∞ (R2 ) (e−tA )
t→0+
 √    
2 g 1 1 √
∼ d x Tr(1)2×2 − gR Tr(1)2×2
4π t 24π
1 √
gB0 + O(t),
+ (D.33)

where A is the elliptic second-order self-adjoint differential operator in the
presence of a Riemann metric ds2 = gµν dxµ dxν (in a tensorial notation in

the space L2 (R2 , gdx1 dx2 ):
 
1 √
A = − √ (∂µ 12×2 + Bµ ) gg µν (∂ν 12×2 + Bν ) − (B0 ) (D.34)
g
with Bµ (xν ) denoting Cc∞ (R2 , R2 ) functions.
After all these preliminaries discussions, we pass to the problem of eval-
uating functional determinant (without zero modes)
 ∞
dt
g det Lj = lim − TrCc∞ (R2 ) (e−tLj ) . (D.35)
ε→0 +
ε t
It is straightforward to verify that the following chain of equations related
to the functional variations of the conformal structure hold true [Herewith
Tr ≡ TrCc∞ (R2 ) ]
 δLj
 : ;< = 

dt Tr δϕ δ ϕe−tLj
δLj
δ g det Lj = lim+
ε→0 ε
 ∞
= lim dt Tr [(−(j + 1)δϕLj − j L̄(j+1) δϕLj )e−tLj ]
ε→0+ ε
 ∞
= lim dt Tr [−(j + 1)δϕLj e−tLj + jδϕL̄−(j+1) e−tL̄−(j+1) ] ,
ε→0+ ε

(D.36)
August 6, 2008 15:46 9in x 6in B-640 ch05

Domains of Bosonic Functional Integrals 167

where we have used the functional identity

Tr(−j L̄−(j+1) δϕLj e−tLj )


( )
= Tr −j L̄−(j+1) (L̄−(j+1) )−1 (δϕ)Lj L̄−(j+1) e−tL̄−(j+1)
( )
= Tr −j · 1δϕ(−L̄−(j+1) )e−tL̄−(j+1) (D.37)

since

e−tL̄−(j+1) = (L̄−(j+1) )−1 e−tLj L̄−(j+1) (D.38)

is a consequence of the operatorial relationship

L̄−(j+1) = (L̄−(j+1) )−1 Lj L̄−(j+1) . (D.39)

As a consequence, we obtain, the following results:

δ( g det Lj )

= −(j + 1) lim+ Tr(δϕe−εLj ) + j lim+ Tr(δϕe−εL̄−(j+1) ) (D.40)


ε→0 ε→0

& '
1 ϕ(x) (1 + 3j) 
= −(j + 1) 2
d xδϕ(x) e − ∆ϕ(x) 
2πε 12π 
R2 ε→0+

& '
1 ϕ(x) (2 + 3j) 
+j 2
d xδϕ(x) e + ∆ϕ(x)  . (D.41)
2πε 12π 
R2 ε→0+

Combining together, we obtain our final “basic-brick” formulae of the


quantum geometric path-integrals for random surfaces:
  & '
1 2
δ g det Lj = lim − d xδϕ(x)e ϕ(x)
ε→0+ 2πε R2
& '
(1 + 6j(j + 1)) 2
+ d xδϕ(x)∆ϕ(x) (D.42)
12π R2

which produces the “brick” result (Polyakov)


&   '
1 2
g det Lj = lim+ − d xe ϕ(x)
ε→0 2πε R2
&  '
1 + 6j(j + 1)) 1
× − d2 x (∂a ϕ)2 (x) (D.43)
12π R2 2
August 6, 2008 15:46 9in x 6in B-640 ch05

168 Lecture Notes in Applied Differential Equations of Mathematical Physics

The result in the presence of gauge fields can be obtained through


Bosonization techniques and only will lead to the additional term to be
added to Eq. (D.43) in its right-hand side
&  '
1 2
8 1 2
9
exp − d xe ϕ(x)
A1 (x)A (x) + A2 (x)A (x)
2π R2
&  '
1
= exp − (dz ∧ dz̄)(Az · Az̄ ) . (D.44)

August 6, 2008 15:46 9in x 6in B-640 ch06

Chapter 6

Basic Integral Representations in Mathematical Analysis


of Euclidean Functional Integrals∗

In this complementary chapter to Chap. 5, we expose additional rigorous


mathematical concepts and theorems behind Euclidean functional integrals
as proposed in Chap. 5 and used thoroughly in other chapters.
In Sec. 6.1, we present a pure topological proof of the basic measure
theory (Riesz–Markov theorem), mathematical concept basic to construct
rigorously functional integrals. In Sec. 6.2, we present analogous results on
the mathematical structure of the L. Schwartz distributions.
In Sec. 6.3, we present the important Kakutani theorem on the equiv-
alence of Gaussian measures in Hilbert spaces, the mathematical basis for
the rigorous framework for Jacobian transformations in Euclidean path-
integrals.

6.1. On the Riesz-Markov Theorem

The words set and function are not as simple as they may seem.
They are potent words. They are like seeds, which are prim-
itive in appearance but have the capacity for vast and intricate
developments — G.F. Simmons.

The Riesz representation theorem

Theorem 6.1. Let X be a topological compact Hausdorff space. Let L be a


positive linear functional on C(X). There exists a unique positive measure
dL µ and an associated σ-algebra on X which represents L in the sense that

L(f ) = f (x)dL u(x). (6.1)
X

Let us begin our proof by introducing the smallest ring containing the
compact subsets of X.
∗ Author’s original results.

169
August 6, 2008 15:46 9in x 6in B-640 ch06

170 Lecture Notes in Applied Differential Equations of Mathematical Physics

Another mathematical structure we need is the following Banach space.


Let p C(X) be the vector space formed by all linear combinations of the
elements of C0 (X) and the characteristic functions of the compact sets of X.
We introduce the sup norm on this vector space and take its completion
still denoted by p C(X) (which is a Banach space). It is a straightforward
consequence of the Hahn–Banach theorem that the given positive linear
functional L has a unique extension to p C(X) still denoted by L in what
follows.
We define now an equivalence relation on the algebra of sets introduced
above through the relationship

∀β, α ∈ A and α ∼ β ⇔ L(χα∆β ) = 0, (6.2)

here α∆β denotes the topological closure of the difference set α∆β = (α −
β) ∪ (β − α) = (α ∩ β  ) ∪ (β ∩ α ). On this coset algebra of sets A/∼, denoted
by EBaire (X), we introduce a metrical structure by means of the metric set
function

dL (A, B) = L((χA∆B )). (6.3)

By considering the topological completion of the metric space


(EBaire (X), dL ), we obtain our proposed σ-algebra on X and a measure
defined by the simple metrical relation

µL (α) = dL (α, φ). (6.4)

At this point, it is evident that the extension theorem of Caratheodory


is a simple rephrasing of Eq. (6.4), since for a given µL -measurable set
Ω ∈ (E Baire (X), µL ) and ε > 0, there exists a finite family of disjoint
compact sets on X : {Ke }e=1,...,N (c) such that
   
N (ε) N (ε)
 
d Ω, K  ≤ ε ⇔  µL (Ke ) − ε
=1 =1
 
N (ε)

≤ µL (Ω) ≤ ε +  µL (Ke ) (6.5)
=1

Let us introduce a large Banach space Cbounded (ĒBaire (X), R) with


the usual sup norm. It is straightforward to see that the given functional
L ∈ (p C(X))∗ has a unique extension L  to this new space of continuous
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 171

function on ĒBaire (X) (which is straightforwardly identified with the mea-


surable functions on (ĒBaire (X), uL )!). Since the characteristic functions of
compact sets are elements of Cbounded(ĒBaire (X), R) any f ∈ C(x) is the
limit on the topology of C(ĒBaire (X), R) of the simple functions (monotone
nondecreasing) sequence
 n 
n2
j−1 
f (X) = lim χ E (x) + nχ F (x) ≡ lim Sn (f ). (6.6)
n→∞   n→∞
n,j n
j=1
2n

Here, the (compact!) sets En,j and Fn in X are defined by


 
j−1 j
En,j = f −1 , , (6.7a)
2n 2n
 
Fn = f −1 [n, ||f ||C(x) ] . (6.7b)

Now the assertive expressed by Eq. (6.1) is a simple result of the defi-
nition of integration
 
 )=L
L(f ) = L(f  lim Sn (f ) = lim L(Sn (f ))
n→∞ n→∞
 n 

n2
j − 1
= lim  n
µL (En,j ) + nµL (Fn )
n→∞
j=1
2
  
def
= lim fn (x)dL µ(x) ≡ f (x)dL µ(x), (6.8)
n→∞ X X

which proves the Riesz–Markov theorem.


As a last point of this section let us give a criterion for the existence of
invariant sets in relation to a given (measurable) transformation

T : (ĒBaire (X), dL ) → (ĒBaire (X), dL ). (6.9)

If the measurable transformation is a contraction (or some of its power!)


between the above complete metric spaces, namely, if c < 1 and an integer
ρ ∈ Z+ such that

ρ ρ
dL (T A, T B) = dL µ(x)χ(T ρ A∆T ρ B) (x)
X
 
≤e dL µ(x)χ(A∆B) (x) , (6.10)
X
August 6, 2008 15:46 9in x 6in B-640 ch06

172 Lecture Notes in Applied Differential Equations of Mathematical Physics

then there exists a point-fixed set Ā, such that

T (Ā) = Ā, (6.11)

in the sense that



dL (T (Ā), Ā) = dL µ(x)X(T (Ā)∆Ā) (x) = 0. (6.12)
X

We now show a concrete version of the Riesz representation theorem.

Theorem 6.2. Let L be a continuous linear functional on C(Ω̄), the space


of the continuous function defined in a compact set Ω̄ ⊂ R4 and satisfying
the following property
 PN 
L ei( j=1 pj xj ) χΩ̄ (x) = fΩ (p1 , . . . , pn ) = fΩ (p) ∈ L1 (RN ) (6.13)

Then, there exists a (unique) function φL (x) ∈ C(Ω̄) representing the


action of the functional Eq. (6.13) on C(Ω̄) by the integral representation

L(g(x)) = dN xφL (x)g(x). (6.14)
Ω̄

Proof. Let us firstly consider the Fourier transform of the function fΩ (p).
Namely
 n 
1
φL (x) = √ dN p · eiP ·x fΩ (p). (6.15)
2π F N

Obviously φL (x) ∈ C0 (RN ).


Due to supposed continuity of the functional L, one can show
that fΩ (p) ∈ C ∞ (RN ), and we have the differentiability relation (M =
( 1 , . . . , N )) (exercise)
   
∂ |M|  N 
N    
(f Ω (p)) = L (ix1 )1
· · · (ixn ) exp i p x
j j ,
∂p11 · · · ∂pnN  
j=1

(6.16)

which means that the inversion Fourier transform theorem holds true
 N
1
L((x1 )1 · · · (xn )n ) = √ dN xe−ipx {(x1 )1 · · · (xn )n φL (x)}|p=0
2π RN
 N  
1
= √ dN x(x1 )1 · · · (xn ) φL (x) . (6.17)
2π RN
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 173

By the Weierstrass theorem, we have, finally, our envisaged result on


C(Ω̄)
 
N
L(g(x)) = d xg(x)φL (x)χΩ̄ (x) = dN xg(x)φL (x). (6.18)
RN Ω̄ 

6.2. The L. Schwartz Representation Theorem on C ∞ (Ω)


(Distribution Theory)

The Quantum and Random World is an application of Cantor


Set Theory in its developments — Luiz Botelho.

After having exposed the fundamental abstract result of Riesz–Markov on


the structure of the elements of the dual space of continuous linear func-
tionals in C(X), with X denoting a general compact topological space, we
pass on to the problem of describing continuous functionals on the vector
space C ∞ (Ω), with Ω denoting a open set of RN .
Let us, thus, start by considering a sequence of compact sets Kn , with
∞
the interior property (Kn+1 ) ⊃ Kn and such that Ω = n=1 Kn , together
with the complete metric space C ∞ (Kn ), defined by the vector space of
infinitely differentiable functions in Ω, with support in Kn with the Frechet
metric
∞
2−m ||f − g||m
d(f, g) = , where ||f ||m : sup sup |Dp f (x)|.
m=0
1 + ||f − g||m x∈Ω |p|≤m

The basic contribution of L. Schwartz is to consider the topology of


the inductive limit on C ∞ (Ω) as writing formally as topological spaces

∞
Cind (Ω) = n=0 C ∞ (Kn ) rigorously means that the topology in C ∞ (Ω) is
the weakest topology which makes all the canonical injections

DN : C ∞ (Kn ) → C ∞ (Ω), (6.19)

continuous applications.
A topological basis for the origin of C ∞ (Ω) is formed by all those convex
and barreleds sets U ⊂ C ∞ (Ω) such that U ∩ C ∞ (Kn ) is always a neigh-
borhood of the origin in C ∞ (Kn ). The main result and reason for intro-
ducing such inductive topology in C ∞ (Ω) is that it leads to the fundamental

result that Cind (Ω) is a sequentially complete topological vector space.
At this point, it is worth to call to the readers’ attention that the usual
∞
nondistributional topological definition of C ∞ (Ω) as m=0 C m (Ω) (always
August 6, 2008 15:46 9in x 6in B-640 ch06

174 Lecture Notes in Applied Differential Equations of Mathematical Physics

used in others approach of generalized functions) is stronger than the L.


Schwartz inductive topology introduced above.

We always rewrite Cind (Ω) in the well-known L. Schwartz notation as
D(Ω), the Schwartz test function space. The description of the notion of
convergence in D(Ω) is straightforward, since we have the sequential com-
pleteness topological property. Namely, a sequence ϕn (x) ∈ D(Ω) converges
in D(Ω) if there exists a set Kn such that ϕn → ϕ̄ in C ∞ (Kn ).
Another basic result as a consequence of the introduction of inductive
limit topology in C ∞ (Ω) is the straightforward description of the dual space
of D(Ω), denoted by D (Ω) and named as the L. Schwartz distribution space
in Ω
 ∗
 
D (Ω) = C ∞ (Kn ) = (C ∞ (Kn ))∗ . (6.20)
n n

Note that the structural description of (C ∞ (Kn ))∗ is expected to be


closely related to that of C(Kn )∗ (the Riesz–Markov theorems). In fact, we
have L. Schwartz generalization of the functional integral representation of
Riesz–Markov theorem.

Theorem 6.3 (Laurent Schwartz). Any given continuous linear func-


tional L ∈ D (Ω) may be represented by a sequence of complex Borel mea-
sures dn u(x) in Kn , a sequence of multi-indexes {Pn } = {p1n , . . . , PnN }
through the integral representation
∞  
L(ϕ) = dµn (x)(DPn ϕ)(x) , (6.21)
n=0 Kn

where
1 N
∂ (Pn +···+Pn )
(Dp ϕ)(x) = Pn1 PN
ϕ(x1 , . . . , xN ). (6.22)
∂x1 · · · ∂xNn
Proof. Let L ∈ D (Ω), but with compact support Ks ⊂ Ω. By the inductive
limit topology (exercise 1), there exists a constant cs > 0, and an integer
ms > 0, such that for any ϕ ∈ D(Ω), we have the estimate
 
|L(ϕ)| ≤ cs sup sup |Dp ϕ(x)| . (6.23)
x∈Ks |p|≤ms

Note that the triple (C, Ks , m) is not unique. We now consider the
∂ p1
following elliptic operator L = ( ∂x 1
) · · · ( ∂x∂ n )pn (p = p1 + · · · + pn ). Since
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 175

Lp is an injective application of C ∞ (Ω) into C ∞ (Ω), and this T restricts the


dense subspace Lp [D(Ω)] (range of Lp in D(Ω)) which satisfies the obvious
estimate for any

f ∈ C(Ks ) : L(L−1
p f ) ≤ cs · sup |f (x)| (6.24)
x∈Ks

we can apply the Riesz–Markov Theorem 6.1 to the composed functional


L ◦ L−1
p in C(Ks )

−1
L(Lp f ) = ds u(x)f (x), (6.25)
Ks

or equivalently (exercise) for any ϕ ∈ D(Ω), we have the functional integral


representation

L(ϕ) = ds µ(x)(Dps ϕ)(x). (6.26)
Ks

In general, we just consider a unity partition subordinate to a given

open cover of Ω (1 = n hn (x), Kn ⊂ supp hn ⊂ Kn+1 , Kn compact set of
Ω and U Ku = Ω and hn (x) = 1, for x ∈ Kn ):

 ∞

L= hn (x)L = Ln . (6.27)
n=1 n=1

At this point, we introduce the weak-* topology in D (Ω) through a


sequential criterion. A sequence of distributions Ln ∈ D (Ω) converges
weak-star if the sequence of measures in Eq. (6.25) converges in the weak-
star topology of C(Ks )∗ .
After the proof of L. Schwartz representation theorem, let us introduce
the operation of derivation in the distributional sense. First, let us recall
some definitions in the functional analysis of vector topological spaces. Let
E and F be two vector spaces with topologies compatible with its vectorial
structure and U : E → F a linear continuous application between them.
For any y  ∈ F ∗ (dual of F ), we can associate the element x of E  through
the definition (t U : F  → E  ):

(x ) = x (x) = y  (U (x)) = (t U (y)). (6.28)

It can be showed that if U is continuous, the t U remains continuous if E


and F are Frechet spaces like C ∞ (Ks ).
August 6, 2008 15:46 9in x 6in B-640 ch06

176 Lecture Notes in Applied Differential Equations of Mathematical Physics

As a consequence of the above remarks, the usual derivative operator is


a linear continuous application between D(Ω). Namely

D : D(Ω) → D(Ω).

By the duality equation (6.26) mentioned above, one has a natural deri-
vative application in D (Ω)
def def
(−DL) ≡ (t D)(L)(f ) ≡ L(Df ), (6.29)

D (Ω)
besides being always a continuous operation in D (Ω) if Ln −→ ⇔
D (Ω) t
t
DLn −→ DL.
At this point, we call our reader to show that the sequence of functions
fn (x) = n1 sin(nx) seen as kernels of distributions in D (R) obviously con-
verges to the zero distribution in D (R):
  
1
lim dx sin nx ϕ(x) = 0. (6.30)
n→∞ R n

As a consequence of the above remark, we have the validity of the result


called the Riemann–Lebesgue lemma

lim dx cos(nx)ϕ(x) = 0. (6.31)
n→∞ R

Namely
 D (R)
d sin nx) D (R)
= cos(nx) −→ 0. (6.32)
dx n

A further finer structural analysis can be implemented to Eq. (6.21)


by means of an application of the Radon–Nikodym theorem to the pair of
complex Borel measures (dus (x), dN x) on the Borelians of Ω.
 hs (x)
 
∞ 
  !
  du (x)  
  s  
L(ϕ) =  (Dps ϕ)(x)  n  dn x
s=0
 Ω  d x  

∞ 
 
+ (Dps ϕ)(x)dνrsing (x) , (6.33)
s=0 Ω
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 177

where hs (x) ∈ L1 (Ω, dN x), and dVnsing (x) is a singular measure (in relation
to the Lebesgue measure dN x in Ω) with support at points (Dirac delta
functions) and on sets of Lebesgue zero measure


dVssing (x) = a,s δ(x − x,s ) + dVs(continuous singular) (x). (6.34)
=0

Another important distributional space is the (topological) dual space


of test functions with polynomial decreasing S(RN ), a very basic object in
wave fields quantum path-integral (see Chap. 19)
 
S(RN ) = u ∈ C ∞ (RN ) | ||ϕ||n,r = sup |xn Dm ϕ(x)| < (∞) . (6.35)
x∈RN

We have the following structural theorem, analogous to Theorem 6.1 of


L. Schwartz.

Theorem 6.4. Given a functional L in (S(Rn )) , we can always rep-


resent L by a Borel complex measure du(x) in RN by means of (xp =
xp11 · · · xpNn , etc.)

L(ϕ) = dµ(x)(xp Dq ϕ)(x). (6.36)
RN

The proof of the above integral representation for distributions in


S q (RN ) is based on the fact that for a given L ∈ S q (RN ), these are multi-
indexes (p, q), such that is a positive constant c with

|L(ϕ)| ≤ c||ϕ||p,q . (6.37)

Note that the elliptic operator Lp = xp Dq = xp11 · · · xpnn ( ∂x ) ···


∂ q1
1
( ∂xN ) is a subjective application of S(R ) into S(R ), and Lp (S(RN ))
∂ qN N N

is dense in C0 (RN ) (continuous functions vanishing at ∞), the great use-


fulness of S 1 (RN ) in quantum field theoretic path-integrals is related by the
fact that the usual Fourier transform is a vectorial/topological isomorphism
in S(RN ). By duality, one straightforwardly define the Fourier transforms
in S 1 (RN ) which remains as a topological isomorphism in the distributional
space S 1 (RN )

F : S(RN ) → S(RN ), (6.38)


t
F : S q (RN ) → S q (RN ), (6.39)
t
F (L)(ϕ) = L(F (ϕ)). (6.40)
August 6, 2008 15:46 9in x 6in B-640 ch06

178 Lecture Notes in Applied Differential Equations of Mathematical Physics

The above equations are important results in the applications of distri-


bution theory of L. Schwartz given by the following result. Let A be a
continuous linear application between a locally convex topological vector
space E with value in the topological dual of another locally convex topo-
logical vector space F  . Then, the bilinear form in E × F defined by the
relation B(f, g) = (Af )(g) is continuous in E × F  , when one introduces
the weak topology on F  . As a consequence, every continuous bilinear form
on S(RN ) is of linear superposition of forms written below for a pair of
multi-indexes (m, n)

B(f, g) = F (x, y)(Dm f (x))(Dn g(x))dn xdn y. (6.41)
RN ×RN

Here, F (x, y) is a continuous function of polynomial grow in RN


 # 
#
+#
∃p ∈ Z # lim F (x, y)(|x|2 + |y|2 )−p = 0 .
|x|→∞
|y|→∞

6.3. Equivalence of Gaussian Measures in Hilbert Spaces


and Functional Jacobians

In this section, we present the mathematical analysis of the Jacobian change


of variable in Gaussian functional integrals in Hilbert spaces through the
formalism of the Kakutani theorem.
Let A−1 and B −1 be positive definite trace class operators in a given
Hilbert space (H, , ) and operators inverse of the operators A and B.
The spectral representations for theses operators
Aϕn = λn ϕn , (6.42a)
Bσn = αn σn , (6.42b)
define Gaussian measures dA−1 u(ϕ) and dB −1 u(ϕ) in the Borelian algebra
of the cylindrical sets in H and are defined by
$N   & '
% λn λn
2
dA−1 µ(ϕ) = lim sup dϕ, ϕn  exp − ϕ, ϕn  ,
N n=1
2 2π
(6.43a)
$ '
%
N ( α ) & α 
dϕ, σn  exp − ϕ, σn 2
n n
dB −1 ν(ϕ) = lim sup .
N n=1
2 2π
(6.43b)
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 179

We have, thus, the Kakutani theorem and the equivalence of the above
measures.
Theorem 6.5 (Kakutani). The two measures Eqs. (6.43a) and (6.43b)
are mutually equivalent or singular. In the first case, we have the criterion
that
  λ2 − α2 
n n
< ∞, (6.44)
n
λ n αn

and the Radon–Nykodin derivative of the above measures is given by


$N *   +'
dA−1 µ(ϕ) % λn 1 
2
= lim exp − (λn − αn ) (ϕ, σn ) .
dB −1 ν(ϕ) N →∞ αn 2 n
(6.45)

Note that in the case of the Radon–Nykodim derivative


 
dA−1 µ(ϕ) −1 1
= det(AB ) exp − ϕ, (A − B)ϕ . (6.46)
dB −1 ν(ϕ) 2

On the basis of Eqs. (6.45) and (6.46), one can show the Wienner result
about translation invariant of Gaussian measures. Let Th : H → H be the
translation operator in H. Let us consider the translated Gauss measure
∞ '
%
− 12 λn [ϕ,ϕn +h,ϕm ]2
dA−1 µ(Th ϕ) = dA−1 µ(ϕ + h) = dϕ|ϕn e .
n=1
(6.47)

By the Kakutani theorem, the translated measure dA−1 µ(Th ϕ) is equiv-


alent to the measure dA−1 µ(ϕ) if and only if
 |h, ϕn |2
1 = Ah, h < ∞, (6.48)
n λn

or equivalently, the translational-invariance of the measure is insured if h


belongs to the domain of the operator A.
At this point, we remark that Dom(A) is a set of zero measure for
(H, dA−1 µ(ϕ). (Finite action configurations smooth field configurations,
makes a set of zero functional measure.) Let us give a simple proof of such
important result in practical calculations with path-integrals.
August 6, 2008 15:46 9in x 6in B-640 ch06

180 Lecture Notes in Applied Differential Equations of Mathematical Physics

First, let us rewrite the finite action set of path integrated configurations
(= Dom(A)) in the following form (for ε > 0):

χDom(A) (ϕ) = lim+ lim exp{−αpN ϕ, APN ϕ}


α→0 N →∞
$
1, if ϕ, Aϕ < ∞,
= (6.49)
0, otherwise,

where the orthogonal projections Pn → 1 in the strong sense.


Let as evaluate formally in the “physical way” its functional measure
content
  
−αPN ϕ,APN ϕ
MA−1 (χDom(A) (ϕ)) = lim lim dA−1 µ(ϕ)e
α→0 N →∞ H
  
N
× lim lim exp − g(1 + α)
α→0+ N →∞ 2
N
= lim lim (e− 2 α ) = e−∞ = 0. (6.50)
α→0 N →∞

At this point, one can see that the usual Schwinger procedure to deduce
functional equations for the quantum field generating functional does not
make sense in the Euclidean framework of path-integrals since the measure
is not translational invariant. Namely, in the usual Feynmann notation

δ ( − 1 ϕ,Aϕ −V (ϕ) )
DF [ϕ] e 2 e = 0. (6.51)
H δϕ
As one can see from the above exposed result that the Minlos theorem is
a powerth “tool” in the functional integration theory in infinite dimension.
In this context, we have the converse of the Minlos theorem in Hilbert
spaces, the so called Sazonov theorem. Let us give a simple proof of this
result.

Theorem 6.6 (Sazonov). Given a Gaussian measure dµ(ϕ) in a sepa-


rable Hilbert space H such that

dµ(ϕ)||ϕ||2 < ∞. (6.52)
H

Then, the functional Fourier transforms



Z(f ) = dµ(ϕ) exp(if, ϕH ), (6.53)
H
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 181

is a continuous functional in relation to the Sazonov Hilbert–Schmidt inner


product

ϕ̄, ψ̄Sz = dµ(ϕ)ϕ, ϕ̄ϕ, ψ̄. (6.54)
H

Proof. First, let us remark that the above-defined Sazonov norm is of


Hilbert–Schmidt class by construction, since for any orthonormal basis (en )
in (H, , ), we have the relationship


N N 
 
(en , en )Sz = dµ(ϕ)ϕ, en ϕ, en 
n=1 n=1 H

≤ dµ(ϕ)||ϕ||2 < ∞. (6.55)
H

|| ||
It is obvious to see that if one has a sequence σn ∈ H and σn −→ 0,
then we have that

(σn , σn )Sz = dµ(ϕ)(ϕ, σn )2 → 0. (6.56)
H

Consequently by the Lebesgue theorem ϕ, σn  → 0 almost everywhere


in (H, dµ(ϕ)). Obviously (1 − exp i(ϕ, σn )) → 0 almost everywhere. As an
easy estimate, we have
# #
# #
#
|Z(σn ) − Z(0)| = # dµ(ϕ)(1 − e iϕ,σn  #
)# → 0. (6.57)
H

Result showing the (sequential) continuity of the functional Z(j) in


the new separable Hilbert space (H, , Sz ). Another important comment
related to the Radom–Nykodin derivative Eq. (6.45) is the case one has the
B-operator as a “formal” perturbation of the operator A. Namely

A = B + gC and [B, C] = 0. (6.58)

In this situation, we have the relationship


 
1
dµ(ϕ)
B+gC /dB ν(ϕ) = det[1 + gCB −1 ] exp − ϕ, Cϕ . (6.59)
2
August 6, 2008 15:46 9in x 6in B-640 ch06

182 Lecture Notes in Applied Differential Equations of Mathematical Physics

By using the proper time definition for evaluating the functional deter-
minant in Eq. (6.59)

det[1 + gCB −1 ] = exp Tr g[1 + gCB −1 ]


 g 
−1 −1 −1
= exp dλ Tr(CB (1 + λCB ) )
0
 g 
def
≡ exp Tr R(λ)dλ , (6.60)
0

where the resolvent operator


, R(λ) satisfies the operatorial equation, and
−1
we suppose that CB ∈ 1 (H) due to the obvious estimate
# #
#%∞ # P∞ P∞
# # −1
# (1 + λn )# = e[ n=1 n(1+λn )] ≤ e n=1 λn = eTr(CB )
# #
n=1

R(λ) = CB −1 − λ(CB −1 )R(λ). (6.61)

Sometimes perturbative evaluations can be implemented for Eq. (6.61).


Let us exemplify in this “toy model”

1 d2
x|A|x  = − δ(x − x2 ) + gV (x, x ) = B + gC. (6.62)
2 d2 x

We can see that

d2  d2
[B, C] = 0 ⇔ V (x, x ) = V (x, x ). (6.63)
d2 x d2 x

So the resolvent R(λ) ≡ R(x, x , λ) satisfies the integral equation below


 
1 d2
R(x, w, λ) − δ(w − x )
2 dw2
 +∞
= gV (x, x ) − λ V (x, y)R(y, x , λ)dy. (6.64)
−∞

As a geometrical comment, in this chapter, from a set of informal lecture


notes for graduate students’ seminars, we have a geometric stochastic cri-
terion for differentiability of the samples (class of functions) in the space
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 183

C 2 (L2 (Ω), dA−1 µ(ϕ)). Let us start by considering the estimate of the
Newton differential quotient
 # #
# ϕ(x + h) − ϕ(x) ##
dA−1 µ(ϕ) ## lim #
L2 (Ω) h→0 h
 1/2
[φ(x + h)φ(x + h) + φ(x)φ(x) − 2φ(x + h)φ(x)]
≤ dA−1 µ(ϕ)
L2 (Ω) h2
1 −1
≤ [A (x + h, x + h) + A−1 (x, x) − (A(x + h, x) + A(x, x + h))]1/2 ,
h
(6.65)

which insures the differentiability in the average of the path-integral


samples if we have the differentiability property for the correlation
function
#
∂ 2 A−1 (x, y) ##
# < ∞.
∂x∂y x=y


6.4. On the Weak Poisson Problem in Infinite Dimension

In this section, we intend to present a weak solution to the Poisson problem


in infinite dimensions (see Chap. 1) by means of the simple basic Hilbert
space techniques.
Let us start our analysis by considering a given separable Hilbert space
H and an associated Gaussian cylindrical measure dQ u(x) on the Borel
,cylinder of H, with Q denoting a class trace positive-definite operator (Q ∈
+
1
(H, H)). We intend now to show the existence of a real-valued weak
solution to the following (functional) Hilbert space-valued Poisson problem
in the Hilbert space L2 (H, dQ u)

− TrH [QD2 U (x)] + x, DU H = f (x)(χBR (0) (x)) (6.66)

Here, χBR (0) (x) is the characteristic function of the ball of radius R in
H (BR (0) = {x ∈ H | x, x ≤ R), u(x, t) ∈ C 2 (H, R) ∩ L2 (H, dQ u)
and- D2 the second-order
. (operator) Frechet derivative in H (for instance
2 iP,xH
Tr QD (e ) = −(p, QpH )eip,xH ).
August 6, 2008 15:46 9in x 6in B-640 ch06

184 Lecture Notes in Applied Differential Equations of Mathematical Physics

By using the following formula of integration by parts in infinite dimen-


sions (Qβk = λk βk )
 
dQ µ(x)Dϕ, DψH (x) = − (Tr[D2 ψ(x)], ϕ(x))dQ µ(x)
H H

+ x, (QDψ)(x)H ϕ(x)dQ µ(x), (6.67)
H

one can rewrite Eq. (6.66) into a functional integral form where it is possible
now to consider a rigorous mathematical meaning for weak solutions in
L2 (H, dQ µ)
 
dA µ(x)BDu, BDg = dA µ(x)(f (x)χBR (x))g(x), (6.68)
H H
,+
where B ∈ 2 (H), and BB ∗ = Q is the usual decomposition of a trace
class (positive-definite) operator in terms of Hilbert–Schmidt root square
operator B.
Since Q is positive-definite, there exists a positive constant c such that
for any x ∈ H, we have the lower bound

Qx, xH ≥ cx, x. (6.69)

Let us now introduce the following Hilbert–Sobolev space associated to


the cylindrical measure dQ µ(x)

H 1 (L2 (H, dQ µ)) = {f ∈ C 1 (H, R) ∩ L2 (H, dQ µ)


#  #
# #
#f, f H 1 ≡ d µ(x)BDf, BDf  < ∞ #. (6.70)
# Q H #
H

We now show that the real linear function associated to our class of cut-
off datum f (x)χBR (x) as expressed by the right-hand side of Eq. (6.132) is
continuous in the topology associated to Eq. (6.70). Since the above space is
separable, one can verify the continuity through a sequential argument. Let
us thus consider a sequence of elements {gn } in H 1 (L2 (H, dQ µ)) converging
to a given ḡ ∈ H 1 (L2 (H, DQ µ)).
We have the estimate below
# #
# #
# dQ µ(x)(f (x)χBR (x))(gn − g)(x)#
# #
H
 1/2
1/2
≤ ||f ||L2 (H,dQ µ) dQ µ(x)(gn − gn )2 (x)χBR (x) . (6.71)
H
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 185

At this point, we observe the validity of the estimate:


 
dQ µ(x)q 2 (x)χBR (x)
H
 - .   
4 · ( k λk )2
≤ dQ µ(x)Dg, Dg(x) . (6.72)
R2 − 2 TrH (Q) H
2
So, if one chooses R in a such way that TrH (Q) < R2 and observing
that (see Eq. (6.69))
  
1
dQ µ(x)Dg, Dg(x) ≤ dQ µ(x)BDg, BDg(x) , (6.73)
H c
one can see naturally that the functional linear below

Lf (g) = dQ µ(x)(f (x)χBR (x))g(x), (6.74)
H

belongs to dual space of our Sobolev space (Eq. (6.70)).


We now apply the Riesz theorem used in Sec. 6.4 to find an element in
H (L2 (H, dQ µ)) being the weak solution of the functional equation (6.68).
1

Namely

Lf (g) = dQ µ(x)BDg, BDu(x). (6.75)
H

One explicit construction of this weak-solution U (s) is given by the


series in H 1 (L2 (H, dA µ))
 ∞ ∞ 
 
U= Lf (en )en = dQ µ(y)f (y)en (y) · en (x), (6.76)
n=1 n=1 H

where en is any given orthonormal basis in H 1 (L2 (H, dQ µ)) (not in


L2 (H, dQ µ)!). Such basis are easily constructed by a simple application of
the Grham–Schmidt orthogonalization process to any set of linearly inde-
pendent elements {En (x)} in C 1 (H, R) ∩ L2 (dQ µ) like the Hermite func-
tionals. Namely
E1 (x)
e1 (x) =
||E1 ||H 1
..
.
n−1
En (x) − n=1 En , ek H 1 · en (x)
en (x) = n−1 . (6.77)
||En − n=1 En , ek H 1 · ek ||H 1
August 6, 2008 15:46 9in x 6in B-640 ch06

186 Lecture Notes in Applied Differential Equations of Mathematical Physics

The reader is invited to apply such procedure to the following Poisson


problem in loop space theory for quantum fields as follows:

H = L2 ([0, 2π])
 −1
 d2 2
Q(σ, σ ) = − 2 + m (σ, σ  ), (6.78a)
dx
Functional Laplacean equation
 2π  2π    
δ δ
dσ dσ  Q(σ, σ 
) ψ[f (σ)]
0 0 δf (σ  ) δf (σ)
= FR [f (σ)]χ||f ||H 1 ≤R (f (σ)). (6.78b)
0

The functional data FR (f (σ)] can be expressed as a functional (Minlos)


Fourier transform of any analytical function H(p) in H

FR [f ] = dQ µ(p) · H(p) exp(ip, f h ). (6.78c)
H

6.5. The Path-Integral Triviality Argument

We start our analysis by considering the chiral non-abelian SU (Nc ) thirring


model Lagrangean on the Euclidean space-time of finite volume Ω ⊂ R4 .
1 a − −−→ ←−− −
L(ψ, ψ̄) = [ψ̄ (iγµ ∂µ ψ a ) + (ψ̄ a iγµ ∂µ )ψ a ]
2
 2 
g µ 5 A 2
+ (ψ̄b γ γ (λ )bc ψc ) . (6.79)
2
Here, (ψ a , ψ̄ a ) are the euclidean four-dimensional chiral fermion fields
belonging to a fermionic fundamental representation of the SU (Nc )
nonabelian group with Dirichlet boundary condition imposed at the finite-
volume region Ω. In the framework of path-integrals, the generating func-
tional of the Green’s functions of the quantum field theory associated with
the Lagrangean equation (6.79) is given by ( ∂ = iγµ ∂µ )
 2
N% −N
1
Z[ηa , η̄a ] = D[ψa ]D[ψ̄a ]
Z(0, 0) a=1
$  * +  '
1 4 0 ∂ ψa
× exp − d x(ψa , ψ̄a ) ← −∗
2 Ω  ∂ 0 ψ̄a
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 187

 2 
g
× exp − d4 x(ψ̄b γ 5 γ µ (λA )bc ψc )2 (x)
2 Ω
  
4
× exp −i d x(ψ̄a ηa + η̄a ψa )(x) . (6.80)

In order to proceed with a bosonization analysis of the fermion field


theory described by the above path-integral, it appears to be convenient
to write the interaction Lagrangian in a form closely parallel to the usual
fermion–vector coupling in gauge theories by making use of an auxiliary
nonabelian vector field Aaµ (x), but with a purely imaginary coupling with
the axial vectorial fermion current (at the Euclidean world).

Z[ηa , η̄a ]
 N%
2
−N  N% 2
−N % 3
1
= D[ψa (x)]D[ψ̄a (x)] D[Aaµ (x)]
Z(0, 0) a=1 a=1 µ=0
     
1 4 0 
 ∂ + igγ5 A ψa
× exp − d x(ψa , ψ̄a ) (x)
2 Ω  )∗
( ∂ + igγ5 A 0 ψ̄a
     
1 4 4
× exp − d x(Aµ Aµ )(x) exp −i
a a
d x(ψ̄a ηa + η̄a ψa )(x) .
2 Ω Ω
(6.81)

At this point, we present our idea to bosonize (solve) exactly the above
fermion path-integral. The main point is to use the long-time ago suggestion
that at the strong coupling gbare → 0+ and at a large number of colors
(the t’Hooft limit2 ), one should expect a great reduction of the (continuum)
vector dynamical degrees of freedom to a manifold of constant gauge fields
living on the infinite-dimensional Lie algebra of SU (∞). In this approximate
leading t’ Hooft limit of large number of colors, we can evaluate exactly the
fermion path-integral by noting the Dirac kinetic operator in the presence
of constant SU (N ) gauge fields written in the following suitable form
     
1 4 0 V (ϕ)  ∂V (ϕ) ψa
exp − d x(ψa ψ̄a ) (x) ,
2 Ω V (ϕ)∗  ∂ ∗ V ∗ (ϕ) 0 ψ̄a
(6.82)

where the chiral phase-factor given by

V (ϕ) = exp[−gγ5 (Aaµ · xµ )λa ], (6.83)


August 6, 2008 15:46 9in x 6in B-640 ch06

188 Lecture Notes in Applied Differential Equations of Mathematical Physics

with the chiral SU (N ) valued phase defined by the constant gauge field
configuration

ϕ(xµ ) = ϕa λa = (Aaµ · xµ )λa . (6.84)

Note that due to the attractive coupling on the axial current-axial


current interaction of our thirring model (Eq. (6.80)), the axial vector cou-
pling is made through an imaginary-complex coupling constant ig.
Now we can follow exactly as in the well-known chiral path-integral
bosonization scheme3 to solve exactly the quark field path-integral
(Eq. (6.82)) by means of the chiral change of variables

ψ(x) = exp{−gγ5ϕ(x)}χ(x), (6.85)


ψ̄(x) = χ̄(x) exp{−gγ5 ϕ(x)}. (6.86)

After implementing the variable change equations (6.7) and (6.8), the
fermion sector of the generating functional takes to a form where the
independent Euclidean fermion fields are decoupled from the interacting–
intermediating non-Abelian constant vector field, Aaµ , namely
 2
N% −N
1
Z[ηa η̄a ] = D[χa (x)]D[χ̄a (x)]
Z(0, 0) a=1

 2
+∞ N% −N  
1
× d[Aaµ ] × exp − (vol(Ω))(Aaµ )2
−∞ a=1
2

× det+1
[( ∂ + igγ5 A
F  )( ∂ + igγ5 A )∗ ]
     
1 0  ∂ χa
× exp − d4 x(χa , χ̄a ) ∗ (x)
2 Ω ∂ 0 χ̄a
  
4 −gγ5 ϕ(x) −gγ5 ϕ(x)
× exp −i d x(χ̄a e ηa + η̄a e χa )(x) . (6.87)

Let us now evaluate exactly the fermionic functional determinant on


Eq. (6.9) which is exactly the Jacobian functional associated to the chiral
fermion field reparameterizations (Eqs. (6.7) and (6.8)).
In order to compute this fermionic determinant, n det+1 F [( ∂ + ig A)
( ∂ + ig A)∗ ], we use the well-known theorem of Schwarz–Romanov4 by
introducing a σ-parameter (0 ≤ σ ≤ 1) dependent family of interpolating
Dirac operators

 (σ) = ( ∂ + ig A
D  (σ) ) = exp{−gσγ5 ϕ(x)}( ∂) exp{−gσγ5 ϕ(x)}. (6.88)
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 189

Since we have the straightforward relationship


d (σ)

D  (σ) + D
= (−gγ5 ϕ) D  (σ) (−gγ5 ϕ), (6.89)

and the usual proper time definition for the involved functional determi-
nants on our study, namely
 ∞
ds (σ) (σ)∗
+1 (σ) (σ)∗
log detF (D D  ) = lim TrF (e−s(D D ). (6.90)
ε→0+ ε s
As a consequence, we get the somewhat expected result that the Fermion
functional determinant on the presence of constant gauge external fields
coincides with the free one, namely

detF [( ∂ + ig A  )∗ ]
 )( ∂ + ig A
= 1, (6.91)
detF [( ∂)( ∂)∗ ]

since W [Aaµ ] ≡ 0 on basis of Eq. (6.16).


Let us return to our “Bosonized” generating functional (Eq. (6.9)) (after
substituting the above results on it)
 2
N% −N
1
Z[ηa , η̄a ] = D[χa (x)]D[χ̄a (x)]
Z(0, 0) a=1

 2
+∞ N% −N  
1 2
× d[Aaµ ] exp
+ vol(Ω) TrSU(N ) (Aµ )
−∞ a=1
2
     
1 4 0 ∂ χa
× exp − d x(χa , χ̄a ) (x)
2 Ω  ∂∗ 0 χ̄a
  
4 −gγ5 (Aa λa )xµ −gγ5 (Aa λa )xµ
× exp −i d x(χ̄a e µ ηa + η̄a e µ χa )(x) .

(6.92)

Let us argue in favor of the theory’s triviality by analyzing the long-


distance behavior associated to the SU (N ) gauge-invariant fermionic com-
posite operator B(x) = ψa (x)ψ̄a (x). It is straightforward to obtain its exact
expression from the bosonized path-integral (Eq. (6.18))

B(x)B(y) = (χa (x)χ̄a (x))(χa (y)χ̄a (y))(0) G((x − y)), (6.93)

here, the reduced model’s Gluonic factor is given exactly in its


structural-analytical form by the path-integral (without bothering us with
August 6, 2008 15:46 9in x 6in B-640 ch06

190 Lecture Notes in Applied Differential Equations of Mathematical Physics

the γ5 -Dirac indexes)


 2
+∞ N% −N  
1 1
G((x − y)) ∼ d[Aaµ ] exp + vol (Ω) TrSU(N ) (Aµ )2
G(0) −∞ a=1
2
$ / '
× TrSU(Nc ) P exp −g Aα dxα , (6.94)
Cxy

with Cxy a planar closed contour containing the points x and y and pos-
sessing an area S given roughly by the factor S = (x − y)2 .
The notation  (0) means that the Fermionic average is defined solely
by the fermion-free action as given in the decoupled form (Eq. (6.18)).
Let us pass to the important step of evaluating the Wilson phase factor
average (Eq. (6.20)) on the coupling regime of gbare → 0+ and formally at
the limit of t’Hooft of large number of colors N → ∞. As the first step to
implement such evaluation, let us consider our loop Cxy as a closed contour
lying on the plane µ = 0, ν = 1 bounding the planar region S.
We now observe that the ordered phase-factor for constant gauge fields
can be exactly evaluated by means of a triangularization of the planar
region S, i.e.

M
S= ∆(i)
µν . (6.95)
l=1
(i)
Here, each counter-clock oriented triangle µν is adjacent to next one
(i) (i+1)
µν ∩ µν common side with the opposite orientations.
At this point, we note that
( −g R (i) Aα ·dxα )
P e ∆µν ∼ (1) (2) (3)
= e−gAα ·α · e−gAα ·α · e−gAα ·α , (6.96)
(i)
where { α }1=1,2,3 are the triangle sides satisfying the (vector) identity
(1) (2) (3)
α + α + α ≡ 0.
Since we have that
( H ) %n ( −g R (i) Aα dxα )
P e−g (x) Aα dxα = lim P e ∆µν
, (6.97)
n→∞
i=1

and by using the Campbel Hausdorff formulae to sum up the product limit
(Eq. (6.23)) at g 2 → 0 limit with X and Y denoting general elements of
the SU (N ) — Lie algebra:
1
ex · ey = eX+Y + 2 [X,Y ] + 0(g 2 ), (6.98)
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 191

one arrives at the non-Abelian stokes theorem for constant gauge fields
( R ) ( RR 01(s)
)
−g C Aα dxα
P e xy = P e−g S F01 dσ
( 2
)
= P e+(g) [A0 ,A1 ]·S . (6.99)

As a consequence, we have the effective approximate result (formally


exact at N → ∞) to be used in our analysis that follows
( R )
−g C Aα dxα
TrSU(N ) P e xy

   
(g 2 s)2 1
∼ exp + (TrSU(N ) [A0 , A1 ])2 + O . (6.100)
2 N

Let us now substitute Eq. (6.26) into Eq. (6.20) and taking into account
the natural two-dimensional degrees of freedom reduction on the average
Eq. (6.20)
 2
+∞ N% −N
1
G((x − y)) = d[Aa1 ]d[Aa0 ]
G̃(0) −∞ a=1

1
× exp + vol (Ω)[TrSU(N ) (A20 + A21 )]
2
 
(g 2 s)2
× exp + (TrSU(N ) [A0 , A1 ])2 , (6.101)
2

where G̃(0) is the normalization factor given explicitly by


 2
+∞ N% −N  
1 a 2 a 2
G̃(0) = − vol(Ω)[(A0 ) + (A1 ) ] .
d[Aa1 ]d[Aa0 ] exp
−∞ a=1
2
(6.102)
By looking closely at Eqs. (6.27) and (6.28), one can see that the
behavior at large N of the Wilson phase factor average is asymptotic to the
value of the usual integral (TrSU(N ) (λa λb = −δab )).
 +∞    
1 2 (g 2 s)2 4
G((x − y))N 1 ∼ da exp − vol(Ω)a exp − a
−∞ 2 2
 +∞  −1 'N 2 −N
1 2
× da exp − vol(Ω)a . (6.103)
−∞ 2
August 6, 2008 15:46 9in x 6in B-640 ch06

192 Lecture Notes in Applied Differential Equations of Mathematical Physics

By using the well-known result (see Ref. 5 — pp. 307, 323, Eq. (3))
 ∞   4  4 
3 γ γ γ
exp(−β 2 x4 − 2γ 2 x2 )dx = 2− 2 e 2β2 K 14 2
, (6.104)
0 β 2β

we obtain the closed result at finite volume


 
 0vol(Ω)N
G((x − y))N 1 ∼   2  
 2 · g √SN
2

 −1 N 2 −N
+ (vol(Ω))2
„ «   √ 

32 g2√SN 2 (vol(Ω))2 N 2
 π 
×e N 2 2 K 14   1   .
16g 4 N 2 S 2 vol(Ω) 2 
2· 2
(6.105)

Let us now give a (formal) argument for the theory’s triviality at infinite
volume vol(Ω) → ∞. Let us, first define the infinite-volume theory’s limit
by means of the following limit

vol(Ω) = S 2 , (6.106)

and consider the asymptotic limit of the correlation function at |x−y| → ∞


(S → ∞).
By using the standard asymptotic limit of the Bessel function
&
−z π
lim K 41 (z) ∼ e , (6.107)
z→∞ 2z

one obtains the result (limN →∞ g 2 N = g∞


2
< ∞) in four dimensions
$ & 'N 2 −N
N N 2S4 16π − S2 N 2
G((x − y)) N 1 ∼ lim · e 16S2 e 16
|x−y|→∞ S→∞ S N 2S22
1
∼ . (6.108)
|x − y|4(N 2 −N )

So, we can see that for big N , there exists a fast decay of Eq. (6.19)
without any bound on the power decay. However, in the usual L.S.Z.
framework for quantum fields, it would be expected a nondecay of such
factor as in the two-dimensional case (see Eq. (6.33) for vol(Ω) = S),
meaning physically that one can observe fermionic scattering free states at
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 193

large separation. Naively at N → ∞ where we expect the full validity of our


analysis, one could expect on the basis of the behavior of Eq. (6.33), the
vanishing of the above analyzed fermionic correlation function Eq. (6.19)
as faster as any power of |x − y| for large |x − y|. This result would imply
that g 2bare ∼ 0+ may be zero from the very beginning and making a sound
support to the fact that the chiral thirring model — at g 2 ∼ 0+ and for large
number of colors — may still remain a trivial theory, a result not expected
at all in view of previous claims on the subject that resummation may turn
nonrenormalizable field theories in nontrivial renormalizable useful ones.1
Finally as a last remark on our formulae (Eqs. (6.31)–(6.33)), let us
point out that a mathematical rigorous sense to consider them is by taking
the Eguchi–Kwaia continuum space-time Ω, a set formed of n hyper-four-
dimensional cubes of side a — the expected size of the nonperturbative
vacuum domain of our theory — and the surface S being formed, for
instance, by n squares on the Ω plane section contained on the plane µ = 0,
ν = 1. As a consequence of the construction exposed above, we can see that
the large behavior is given exactly by (N = Nc )

 (N 2 n2 a8 )
N 2 −N

 2 na2 2
!  

n→∞ N 32.
ḡ∞
√ N 2 (n2 a8 )
G(na)N 1 ∼ e 2
K 41
 ḡ
 ∞
2 · na2 2 )2 n2 a4 
16(g∞ 

 N 2 −N
1
∼ . (6.109)
na4

6.6. The Loop Space Argument for the Thirring


Model Triviality

In order to argue the triviality phenomenon of the SU (N ) non-Abelian


thirring model of Sec. 6.1 for finite N , let us consider the generating func-
tional equation (6.3) for vanished fermionic sources ηa = η̄a = 0, the so-
called vacuum energy theory’s content

 2
N% −N 3
% R
1
d4 x(Aa a
Z(0, 0) = D[Aaµ (x)]e− 2 Ω µ Aµ )(x)

a=1 µ=0

× detF [( ∂ + igγ5 A  )∗ ].


 )( ∂ + igγ5 A (6.110)
August 6, 2008 15:46 9in x 6in B-640 ch06

194 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point of our analysis, let us write the functional determinant on


Eq. (6.35) as a functional on the space of closed bosonic paths {Xµ (σ), 0 ≤
σ ≤ T , Xµ (0) = Xµ (T ) = xµ }, namely6

 )∗ ]
 )( ∂ + igγ5 A
g detF [( ∂ + igγ5 A
  /
= PSU(N ) · P Dirac exp − g Aµ (Xβ (σ)dXµ (σ)
Cxy Cxy
/ 
i α β
+ [γ , γ ] Fαβ (Xβ (σ))ds . (6.111)
2 Cxy

The sum over the closed loops Cxy with fixed end-point xµ is given by
the proper-time bosonic path-integral
   $  '
 ∞
dT 1 T
= d4 xµ DF [X(σ)] exp − Ẋ 2 (σ)dσ) .
0 T χµ (0)=xµ =χµ (T ) 2 0
Cxy
(6.112)

Note the symbols of the path ordination P of both, Dirac and


color indexes on the loop space phase factors on the expression
equation (6.36).
By using the Mandelstam area derivative operator δ/δσγρ (X(σ)),6 one
can rewrite Eq. (6.36) into the suitable form as an operation in the loop
space with Dirac matries bordering the loop Cxx , i.e.

g detF [( ∂ + igγ5 A  )∗ ]


 )( ∂ + igγ5 A
  /
i δ
× PDirac dσ [γ α , γ β ](σ)
Cxx 2 δσαβ (X(σ))
Cxy
  / 
× PSU(N ) exp −g Aµ (Xβ (σ)dXµ (σ)) . (6.113)
Cxx

In order to show the triviality of functional fermionic determinant


when averaging over the (white-noise!) auxiliary non-Abelian fields as in
Eq. (6.35), we can use a cummulant expansion, which in a generic form
reads off as
 
1
ef Aµ = exp f Aµ + (f 2 Aµ − f 2Aµ ) + · · · . (6.114)
2
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 195

So let us evaluate explicitly the first-order cummulant


 /
i δ
PDirac ds [γ α (σ), γ β (σ)]
Cxy 2 δσαβ (X(σ))
Cxy
2   / 3 
× PSU(N ) exp − g Aµ (Xβ (σ))dXµ (σ) ,
Cxx Aµ
(6.115)
with the average  Aµ defined by the path-integral equation (6.35).
By using the Grassmanian zero-dimensional representation to write
explicitly the SU (N ) path-order as a Grassmanian path-integral6
  / 
PSU(N ) exp − g Aµ (Xβ (σ))dXµ (σ)
Cxx
 
 −N
N% 2 2
N −N
= DF [θa (σ)]DF [θa∗ (σ)]  θa (0)θa∗ (T )
a=1 a=1
  
 2
N −N
i T
d ∗ d
× exp  dσ θa (σ) θa (σ) + θa∗ (σ) θa (σ) 
2 0 a=1
dσ dσ
  
T
× exp g dσ(Aaµ (X β (σ))(θb (λa )bc θc∗ )(σ)dX µ (σ)) , (6.116)
0

one can easily see that the average over the Aµ (x) fields is straightforward
and produced as a result of the following self-avoiding loop action
2   / 3
PSU(N ) exp −g Aµ (Xβ (σ))dXµ (σ)
Cxx Aµ
 
 −N
N% 2 2
−N
N
= DF [θa (σ)]DF [θa∗ (σ)]  θa (0)θa∗ (T )
a=1 a=1
  
 2
N −N
i T
d ∗ 
d
× exp  dσ θa (σ) θa (σ) + θa∗ (σ) θa (σ) 
2 0 a=1
dσ dσ
  
g2 T T
× exp dσ dσ  [(θb (λa )bc θc∗ )(σ)(θb (λa )bc θc∗ )(σ  )]
2 0 0

× δ (D) (Xµ (σ) − Xµ (σ  ))dXµ (σ)dXµ (σ  ) . (6.117)
August 6, 2008 15:46 9in x 6in B-640 ch06

196 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point one can use the famous probabilistic-topological Parisi


argument7 well-used to understand the λϕ4 triviality at the four-
dimensional space-time to see that due to the fact that Hausdorff dimension
of our Brownian loops {Xµ (σ)} is two, and the topological rule for con-
tinuous manifold holds true in the present situation, one obtains that for
ambient space greater than (or equal to) four, the Hausdorff dimension of
the closed-path intersection set of the argument of the delta function on
Eq. (6.42) is empty. So, we have as a consequence
2   / 3
PSU(N ) exp −g Aµ (Xβ (σ))dXµ (σ) = 1, (6.118)
Cxx Aµ

which in turn leads to the thirring model’s triviality for space-time RD with
D ≥ 4.

References

1. J. Zinn-Justin, Quantum Field Theory and Critical Phenomena (Clarendon,


Oxford, 2001); G. Broda, Phys. Rev. Lett. 63, 2709 (1989); L. C. L. Botelho,
Int. J. Mod. Phys. A 15(5), 755 (2000).
2. A. Di Giacomo, H. G. Dosch, V. S. Shevchenko and Yu A. Simonov, Phys.
Rep. 372, 319 (2002).
3. L. C. L. Botelho, Phys. Rev. 32D(6), 1503 (1985); Europhys. Lett. 11(4), 313
(1990).
4. L. C. L. Botelho, Phys. Rev. 30D(10), 2242 (1984); V. N. Romanov and
A. S. Schwarz, Theoreticheskaya i Matematichesnaya, Fizika 41(2), 190
(1979).
5. I. Gradshteym and I. M. Ryzhik, Tables of Integrals (Academic Press, New
York, 1980).
6. L. C. L. Botelho, J. Math. Phys. 30, 2160 (1989).
7. B. Simon, Functional Integration and Quantum Physics (Academic Press,
1979).
8. K. Fujikawa, Phys. Rev. D 21, 2848 (1980).
9. R. Roskies and F. A. Schaposnik, Phys. Rev. D 23, 558 (1981).
10. K. Furuya, R. E. Gamboa Saravi and F. A. Schaposnik, Nucl. Phys. B208,
159 (1982).
11. A. V. Kulikov, Theor. Math. Phys. (U.S.S.R.) 54, 205 (1983).
12. K. D. Rothe and I. O. Stamatescu, Ann. Phys. (N.Y.) 95, 202 (1975).
13. L. C. L. Botelho and M. A. Rego Monteiro, Phys. Rev. D 30, 2242
(1984).
14. J. D. Bjorken and S. D. Drell, Relativistic Quantum Fields (McGraw-Hill,
New York, 1965).
15. J. Goldstone and F. Wilczek, Phys. Rev. Lett. 47, 988 (1981).
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 197

16. K. Fujikawa, Phys. Rev. D 21, 2848 (1980); R. Roskies and F. Schaposnik,
ibid. 23, 518 (1981).
17. L. Bothelho, Phys. Rev. D 31, 1503 (1985); A. Das and C. R. Hagen, ibid.
32, 2024 (1985).
18. L. Botelho, Phys. Rev. D 33, 1195 (2986).
19. O. Alvarez, Nucl. Phys. B238, 61 (1984).
20. V. Romanov and A. Schwartz, Teor. Mat. Fiz. 41, 190 (1979).
21. K. Osterwalder and R. Schrader, Helv. Phys. Acata 46, 277
(1973).
22. P. B. Gilkey, in Proceeding of Symposium on Pure Mathematics, Stanford,
California, 1973, eds. S. S. Chern and R. Osserman (American Mathematical
Society, Poidence, RI, 1975), Vol. 127, Pt. II, p. 265.
23. B. Schroer, J. Math. Phys. 5, 1361 (1964); High Energy Physics and Ele-
mentary Particles (IAEC, Vienna, 1965).
24. K. Bardakci and B. Schroer, J. Math. Phys. 7, 16 (1966).
25. A. Jaffe, Phys. Rev. 158, 1454 (1966).

Appendix A. Path-Integral Solution for a Two-Dimensional


Model with Axial-Vector-Current-
Pseudoscalar Derivative Interaction

Analysis of quantum field models in two space-time dimensions has


proved to be a useful theoretical laboratory to understand phenomena like
dynamical mass generation, topological excitations, and confinement; all
features expected to be present in a realistic four-dimensional theory of
strong interactions. Recently, a powerful nonperturbative technique has
been used to analyze several two-dimensional fermionic models in the
(Euclidean) path-integral approach. This technique is based on a chiral
change of variables in the functional fermionic measure.8−11 It is the
purpose of this appendix to solve exactly another fermionic model by means
of this nonperturbative technique. The model to be studied describes the
interaction of a massive pseudoscalar field with massless fermion fields in
terms of a derivative coupling and was analyzed previously in Ref. 12 by
using the operator approach.
We start our study by considering the Euclidean Lagrangian of the
model:

 1 (ψ, ψ̄, φ)(x)


L
 
1 1
= −iψ̄γµ ∂µ ψ + (∂µ φ)(∂µ φ) + m2 φ2 + g ψ̄γ 5 γµ ψ∂µ φ (x), (A.1)
2 2
August 6, 2008 15:46 9in x 6in B-640 ch06

198 Lecture Notes in Applied Differential Equations of Mathematical Physics

where ψ = (ψ1 , ψ2 ) denotes a massless fermion, φ a massive pseudoscalar


field, and g is the coupling constant. The Lagrangian (A.1) is invariant
under the global Abelian and Abelian-chiral groups

ψ → eiα ψ, ψ → eiγ5 β ψ, (α, β) ∈ R,

with the formally Noether conserved currents

∂µ (ψ̄γ 5 γµ ψ) = 0, ∂µ (ψ̄γ µ ψ) = 0.

The Hermitian γ matrices we are using satisfy the (Euclidean) relations

{γµ , γν } = 2δµν , γµ γ5 = iεµν γν , γ5 = iγ0 γ1 ,


ε01 = −ε10 = 1 (µ, ν = 0, 2).

In the framework of path-integrals, the generating functional of the Green’s


functions of the quantum field theory associated with the Lagrangian (A.1)
is given by

Z[J, n, n̄] = D[φ]D[ψ]D[ψ̄]
  
2
× exp − d x[L1 (ψ, ψ̄, φ)(x) + φJ + n̄ψ + ψ̄n](x) . (A.2)

In order to perform a chiral change of variable in (A.2) it appears to be


convenient to write the interaction Lagrangian in a form closely parallel to
the usual fermion–vector coupling in gauge theories by making use of the
identity
  
exp − d2 xg(ψγ 5 γµ ψ∂µ φ)(x)
  
2
= D[Aµ ]δ(Aµ + iεµα ∂α φ) exp d xg(ψ̄γµ Aµ ψ)(x) , (A.3)

where Aµ (x) is an auxiliary Abelian vector field. Substituting (B.3) into


(B.2) we obtain a more suitable form for Z[J, n, n̄],

Z[J, n, n̄] = D[φ]D[ψ]D[ψ̄]D[Aµ ]δ(Aµ + iεµα ∂α φ)
  
× exp − d2 x[L1 (ψ, ψ̄, φ, Aµ ) + Jφ + n̄ψ + ψ̄n] , (A.4)
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 199

with
1 1
 2 (ψ, ψ̄, φ, Aµ ) = −iψ̄γµ (∂µ − igAµ )ψ + (∂µ φ)(∂µ φ) + m2 φ2 .
L (A.5)
2 2
Now we can proceed as in the case of gauge theories9,10 in order to
decouple the fermion fields from the auxiliary vector field Aµ (x) by making
the chiral change of variables

ψ(x) = eiγ5 β(x) X(x),

ψ̄(x) = χ̄(x)eiγ5 β(x) , (A.6)


i
Aµ (x) = − (εµν ∂ν β)(x).
g
We note the relation β = gΦ between the chiral phase β and the pseu-
doscalar field Φ due to the constraint δ(Aµ + iεµα ∂α φ) in (A.4). After the
change (A.6), the Lagrangian (A.5) takes a form where now the fermion
fields χ(x), χ̄(x) are free and decoupled from the auxiliary vector field
Aµ (x),
1 1
 3 (χ, χ̄, φ) = −iχ̄γµ ∂µ χ + (∂µ φ)(∂µ φ) + m2 φ2 .
L (A.7)
2 2
On the other hand, the fermionic measure DψDψ̄, defined in terms
of the normalized eigenvectors of the Hermitian operator [−iγµ (∂µ −
igAµ )][−iγµ (∂µ −igAµ )]∗ , is not invariant under the chiral change and given
the Jacobian
 2  
g
exp − d2 x(Aµ )(x)

= {Det[−iγµ (∂µ − igAµ )](−iγµ (∂µ − igAµ )}. (A.8)

In terms of the new fermionic field χ(x), χ̄(x) and taking into account the
Jacobian (A.8), the generating functional reads

Z[J, n, n̄] = D[χ]D[χ̄]D[φ]
 
1
× exp − d2 x[−iχ̄γµ ∂µ χ + (1 − g 2 /π)(∂µ φ)(∂µ φ)
2

1
+ m2 φ2 + Jφ + n̄eigγ5 φ χ + χ̄eigγ5 φ η](x) . (A.9)
2
August 6, 2008 15:46 9in x 6in B-640 ch06

200 Lecture Notes in Applied Differential Equations of Mathematical Physics

We shall now use the effective generating function (A.9) to compute


exactly the (bare) Green’s functions of the model. The two-point Green’s
function of the pseudoscalar field φ(x) is straightforwardly obtained,
 
1 m
(φ(x)φ(y)) = K0 |x − y| , (A.10)
2π (1 − g 2 /π)1/2

where K1 (w) denotes the Hankel function of an imaginary argument of


order zero. The two-point fermion Green’s function is easily computed by
noting that the fermion fields χ(x), χ̄(x) are free and decoupled in (A.9).
The result reads
1 (xµ − yµ )
(ψα (x)ψ̄β (y)) = (γµ )αβ
2π |x − y|2
  
g2 m
× exp K0 |x − y| ,
2π(1 − g 2 /π) 2π(1 − g 2 /π)1/2
(A.11)

where we have used the dimensional regularization scheme to assign the


value 1 to the “tadpole” contribution that appears in (A.11), i.e.
 
g2
exp K 0 (0) = 1.
1 − g 2 /π 2

It is also interesting to compute correlation functions involving fermions


ψ(x), ψ̄(x) and the pseudoscalar field φ(x). For instance, we get the fol-
lowing expression for the vertex Γαβ (x, y, z) = (ψα (x)ψ̄β (y)φ(z)):

Γαβ (x, y, z)
1 (xµ − yµ ) ig
= (γµ γ5 )αβ
2π |x − y|2 2π(1 − g 2 /π)1/2
    
m m
× K0 |x − z| − K 0 |y − z|
(1 − g 2 /π)1/2 (1 − g 2 /π)1/2
  
g2 m
× exp K 0 |x − y| . (A.12)
2π(1 − g 2 /π) (1 − g 2 /π)1/2

It is instructive to point out that in a perturbative analysis of the model,


the previous correlation functions correspond to the full sum of nonrenor-
malized Feynman diagrams involving all possible radiative corrections. The
perturbative renormalization analysis can be implemented by applying the
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 201

asymptotic Lehmann–Symanzik–Zimmermann conditions (or equivalently,


the Dyson prescription) in the (bare) propagators and the vertex function
of model.12–14 This analysis results in that the pseudoscalar field φ gets
(finite) wave-function and mass renormalizations, φ(R) = (Zφ )−1 φ and
m(R) = m2 (Zφ )2 ), respectively, with

1
Zφ = ,
(1 − g 2 /π)1/2

and the coupling constant gets a multiplicative renormalization g(R) =


gZφ . For instance, the renormalized two-point fermion Green’s function is
given by

(ψα (x)ψ̄β (y))(R)


 2

1 (xµ − yµ ) g(R)
= (γµ )αβ exp K0 (m((R)|x − y|) , (A.13)
2π |x − y|2 2π

which has the following short- and long-distance behavior:

(ψα (x)ψ̄β (y))|x−y|→0 ∼


(R) 2
= (γµ )αβ |x − y|−(1+g(R) /2w) , (A.14)

1 (xµ − yµ )
(ψα (x)ψ̄β (y))|x−y|→0 ∼
(R)
= (γµ )αβ · (A.15)
2π |x − y|2

The ultraviolet behavior (A.14) implies that the fermion field carries an
2
anomalous dimension γψ = g(R) /4π. We remark that the model displays
the appearance of the axial anomaly as a consequence of the presence of
the massive field π.8,12 For instance, in terms of the bare field φ, we have

g g m2
∂µ (ψ̄γµ γ5 ψ) = φ = φ. (A.16)
π π 1 − g 2 /π

Next, we shall consider the case of massive fermions in the model


 
1 1 2 2 5
 (ψ, ψ̄, φ) = −iψ̄γµ ∂µ ψ + µψ̄ψ + (∂µ φ)(∂µ ψ) + m φ + g ψ̄γ γµ ψ∂µ ψ .
L
2 2
(A.17)

In order to get an effective action as in (A.9), we make again the chiral


change (A.6) where it becomes important to remark that the fermion
measure DψDψ̄ is to be defined by the eigenvectors of the massless Her-
mitian Dirac operator ( D(A  D(A)∗ ) and thus yields the same Jacobian.15
August 6, 2008 15:46 9in x 6in B-640 ch06

202 Lecture Notes in Applied Differential Equations of Mathematical Physics

This fact is related to the fermion mass independence of the anomalous part
of the divergence of chiral current ∂µ (ψ̄γ µ γ 5 ψ) in fermion models inter-
acting with gauge fields. Proceeding as above we obtain the new effective
(bare) Lagrangian
L
 (χ, χ̄, φ)(x)
 
2igγ5 ψ 1 2 1 2 2
= −iχ̄γµ ∂µ χ + µχ̄e χ + (1 − g /π)(∂µ φ)(∂µ φ) + m φ (x),
2 2
(A.18)
where now the fermion fields χ(x), χ̄(x) possess an interacting term
µχ̄[cos(2gφ) + iγ5 sin(2gφ)]χ which contributes to the phenomenon of for-
mation of fractionally fermionic solutions.15
As we have shown, chiral changes in path-integrals provide a quick,
mathematically and conceptually simple way to solve and analyze the model
studied in this appendix.
Note added. We could like to make some clarifying remarks on the analysis
implemented above. First, in order to implement the regularization rule
in Eqs. (A.3)–(A.11), we consider the associated perturbative power series
in the coupling constant g dimensionally regularized which automatically
takes into account the Wick normal-order operation in the correlation
functions (Eqs. (A.9)–(A.11)). Second, we observe that the model should
be redefined in the region g 2 > π since its associated Gell–Mann–Low
function has a nontrivial zero for g 2 = π (the model contains a tachyonic
excitation).12

Appendix B. Path-Integral Bosonization for an Abelian


Nonrenormalizable Axial Four-Dimensional
Fermion Model

B.1. Introduction
The study of two-dimensional fermion models in the framework of chiral
anomalous path-integral has been shown to be a powerful nonperturbative
technique to analyze the two-dimensional bosonization phenomenon.
It is the purpose of this appendix to implement this nonperturbative
technique to solve exactly a nontrivial and nonrenormalizable
four-dimensional axial fermion model which generalizes for four dimensions
the two-dimensional model studied in the previous appendix.
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 203

B.2. The Model


Let us start our analysis by considering the (Euclidean) Lagrangian of the
proposed Abelian axial model in R4
1
 1 (ψ, ψ̄, φ) = ψ̄γµ (i∂µ − gγ5 ∂µ φ)ψ + g 2 (∂µ φ)2 + V (φ),
L (B.1)
2
where ψ(x) denotes a massless four-dimensional fermion field, φ(x) a pseu-
doscalar field interacting with the fermion field through a pseudoscalar
derivative interaction, and V (φ) is a φ self-interaction potential given by

−g 4 g2
V (φ) = 4
φ(∂µ φ)2 (−∂ 1 φ) + 2 (−∂ 2 φ)(−∂ 2 φ). (B.2)
12π 4π
The presence of the above φ potential is necessary to afford the exact
solubility of the model as we will show later (Eq. (B.18)).
The Hermitian γ matrices we are using satisfy the (Euclidean) relations

[γµ , γν ] = 2δµν , γ5 = γ0 γ1 γ2 γ3 . (B.3)

The Lagrangian L 1 (ψ, ψ̄, φ) is invariant under the global Abelian and
chiral Abelian groups

ψ → eiα ψ, ψ → eiγ5 β ψ, (α, β) ∈ R, (B.4)

with the Noether conserved currents at the classical level:

∂µ (ψ̄γ 4 γ µ ψ) = 0, ∂µ (ψ̄γµ ψ) = 0. (B.5)

In the framework of path-integrals, the generating functional of the


correlation functions of the mathematical model associated with the
Lagrangian L  1 (ψ, ψ̄, φ) is given by

1
Z[J, η, η̄] = D[φ]D[ψ]D[ψ̄]
Z[0, 0, 0]
  
4
× exp − d x[L1 (ψ, ψ̄, φ) + Jφ + η̄ψ + ψ̄η](x) . (B.6)

In order to generalize for four-dimensions the chiral anomalous path-


integral bosonization technique, we first rewrite the full Dirac operator in
the following suitable form:

 D[φ] = iγµ (∂µ + igγ5 ∂µ φ) = exp(igγ5 φ)(iγµ ∂µ ) exp(igγ5 φ). (B.7)


August 6, 2008 15:46 9in x 6in B-640 ch06

204 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now, we proceed as in the two-dimensional case by decoupling the


 1 (ψ, ψ̄, φ)
fermion field from the pseudoscalar field φ(x) in the Lagrangian L
by making the chiral change of variables:

ψ(x) = exp[igγ5 ψ(x)]χ(x),


(B.8)
ψ̄(x) = χ̄(x) exp[igγ5 φ(x)].

On the other hand, the fermion measure D[ψ]D[ψ̄], defined by the eigen-
vectors of the Dirac operator  D[φ]  D[φ]∗ , is not invariant under the chiral
change and yields a nontrivial Jacobian, as we can see from the relationship
   
4
D[ψ]D[ψ̄] exp − d x(ψ̄  D[φ]ψ)(x)

= [Det( D[ψ]  D[ψ]∗ )]


   
4
= J[φ] D[χ]D[χ̄] exp − d x(χ̄iγµ ∂µ (χ)(x) . (B.9)
4
Here, J[φ] = Det( D[φ]  D[φ]∗ ) Det( D[φ = 0]  D[φ = 0]∗ ) is the explicit
expression for this Jacobian.
It is instructive to point out that the model displays the appearance
of the axial anomaly as a consequence of the nontriviality of J[φ], i.e.
∂µ (ψ̄γµ γ5 ψ)(x) = {(δ/δφ)J[ψ]}(χ).
So, to arrive at a complete bosonization of the model (Eq. (B.6)) we
face the problem of evaluating J[φ].
Let us, thus, compute the four-dimensional fermion determinant
det( D[φ]) exactly. In order to evaluate it, we introduce a one-parameter
family of Dirac operators interpolating the free Dirac operator and the
interacting one  D(ζ) [φ], namely,

 D(ζ) [φ] = exp(igγ5 ζφ)(iγµ ∂µ ) exp(igγ5 ζφ), (B.10)

with ζ ∈ [0, 1].


At this point, we introduce the Hermitian continuation of the operators
 D(ζ) [φ] by making the analytic extension in the coupling constant ḡ = ig.
This procedure has to be done in order to define the functional determinant
by the proper-time method since only in this way ( Dζ [φ])2 can be con-
sidered as a (positive) Hamiltonian ( Dζ [φ])2 ≡ ( Dζ [φ])( Dζ [φ])∗ .
The justification of this analytic extension in the model coupling con-
stant is due to the fact that typical interaction energy densities such as
ψ̄γ 4 ψ, ψ̄γµ ∆µ ψ, which are real in Minkowski space-time, become complex
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 205

after continuation in Euclidean space-time.21 As a consequence, the above


analytic coupling extension must be done in the proper-time regularization
for the Dirac functional determinant.
By using the property
d
 Dζ [φ] = ḡγ5 ψ  Dζ [φ] +  Dζ [φ]γ5 ḡφ, (B.11)

we can write the following differential equation for the functional
determinant19 :
d
n Det  Dζ [φ]


= 2 lim d3 x Trx|(ḡγ5 ψ) exp{−ε( Dζ [φ]2 }|x, (B.12)
ε→0+

where Tr denotes the trace over Dirac indices.


The diagonal part of exp{−ε( Dζ [φ]2 } has the asymptotic expansion20,22
 
1 ε2
x| exp{−ε( Dζ [φ])2 }|x ∼ 1 + εH 1 (φ) + H2 (φ) , (B.13)
ε→0+ 16π 2 ε2 2!
with the Seeley coefficients given by (see the complements)

H1 (φ) = ζ ḡγ5 ∂ 2 φ − ḡ 2 ζ 2 (∂µ φ)2 , (B.14)

and

H2 (φ) = 2∂ 2 H1 (φ) + 2ζ ḡγ4 (∂µ φ)∂µ H1 (φ) + 2[H1 (φ)]2 . (B.15)

By substituting Eqs. (B.14) and (B.15) into Eq. (B.12), we obtain finally
the result for the above-mentioned Jacobian:

J[φ] = J0 [φ, ε]J1 [φ]. (B.16)

Here, H0 [φ, ε] is the ultraviolet cut-off dependent Jacobian term


 2  
g 4 2
J0 [φ, ε] = exp d x[φ(−∂ )φ](x) (B.17)
4π 2 ε
and J1 (φ) is the associated Jacobian finite part
  
g2
J1 [φ] = exp − 2 d4 x(−∂ 2 φ)(−∂ 2 φ)(x)

 2  
g 4 2 2
× exp d x[φ(∂µ φ) (−ø φ)](x) . (B.18)
12π 2
August 6, 2008 15:46 9in x 6in B-640 ch06

206 Lecture Notes in Applied Differential Equations of Mathematical Physics

From Eqs. (B.17) and (B.1), we can see that the (bare) coupling constant
2
g gets an additive (ultraviolet) renormalization. Besides the Jacobian term
cancels with the chosen potential V (φ) in Eq. (B.2).
As a consequence of all these results, we have the following expression
for Z[J, η, η̄] with the fermions decoupled:

1
Z[J, η, η̄] = D(φ]D(χ]D(χ̄]J[φ](x)
Z[0, 0, 0]
  
1 1 1 2
× exp − d4 x χ̄(γµ ∂µ )χ + x̄(iγµ ∂µ )χ + gR (∂µ φ)2
2 2 2

+ η̄ exp(igR γ5 φ)χ + χ̄ exp(igR γ5 φ)η . (B.19)

This expression is the main result of this appendix and should be com-
pared with the two-dimensional analogous generating functional analyzed in
Appendix A, Eq. (A.9). Now, we can see that the quantum model given by
Eq. (6.6), although being nonrenormalizable by usual power counting and
Feynman-diagrammatic analysis, it still has nontrivial and exactly soluble
Green’s functions. For instance, the two-point fermion correlation function
is easily evaluated and produces the result

ψα (x1 )ψ̄β (x2 ) = Sαβ


F
(x1 −x2 ; m = 0) exp[−∆F (x1 , x2 .m = 0)]. (B.20)

Here, ∆(x1 − x2 ; m = 0) is the Euclidean Green’s function of the massless


free scalar propagator. We notice that correlation functions involving
fermions ψ(x), ψ̄(x) and the pseudoscalar field φ(x) are easily computed
too (see Ref. 17).
As a conclusion of our appendix let us comment to what extent our pro-
posed mathematical Euclidean axial model describes an operator quantum
field theory in Minkowski space-time.
In the operator framework the fields ψ(x) and φ(x) satisfy the following
wave equations:

iγµ ∂µ ψ = igγµ γ5 (∂µ φ)ψ,


δ∆
 φ = ∂µ (ψ̄γµ γ5 ψ) + · (B.21)
δφ

It is very difficult to solve exactly Eq. (B.12) in a pure operator


framework because the model is axial anomalous [∂µ (ψ̄γµ γ5 ψ) = 0].
August 6, 2008 15:46 9in x 6in B-640 ch06

Basic Integral Representations in Mathematical Analysis 207

However the path-integral study (Eq. (B.19)) shows that the operator
solution of Eq. (B.21) (in terms of free (normal-ordered) fields) is given by

ψ(x) = exp[iγ5 φ(x)]χ(x), (B.22)

since it is possible to evaluate exactly the anomalous divergence of the


above-mentioned axial-vector current and, thus, choose a suitable model
potential V (φ) which leads to the above simple solution.
It is instructive to point out that the operator solution (Eq. (B.22))
coincides with the operator solution of the U (1) vectorial model analyzed by
Schroer in Ref. 1 (the only difference between the model’s solutions being
the γ5 factor in the phase of Eq. (B.22)).
Consequently, we can follow Schroer’s analysis to conclude that the
Euclidean correlation function (Eq. (B.2)) defines Wightman functions
which are distributions over certain class of analytic test functions.23−25 But
the model suffers the problem of the nonexistence of time-ordered Green’s
functions which means that the proposed axial model in Minkowski space-
time does not satisfy the Einstein causality principle.23
Finally, we remark that the proposed model is to a certain extent less
trivial than the vectorial model since for nondynamical φ(x) field the asso-
ciated S matrix is nontrivial and is given in a regularized form by the result
   +∞ 
S = T exp i (ψ̄γµ γ5 ∂µ φψ)(x)d4 x
−∞
  +∞   
δ
= exp −i φ(x) Jε [φ] (x)d4 x , (B.23)
−∞ δφ(x)

with Jε [φ] expressed by Eq. (B.16).

B.3. Complement
We now briefly calculate the asymptotic term of the second-order positive
differential elliptic operator:

( Dζ [φ])2 = (−∂ 2 ) − (ḡγ5 ζ∂µ φ)∂µ − ḡζγ5 ∂ 2 φ + (∂µ φ)2 . (B.24)

For this study, let us consider the more general second-order elliptic four-
dimensional differential operator (non-necessarily) Hermitian in relation to
the usual normal in L2 (RD ) (Ref. 22):

 x = −(∂ 2 )x + aµ (x)(∂µ )x + V (x).


L (B.25)
August 6, 2008 15:46 9in x 6in B-640 ch06

208 Lecture Notes in Applied Differential Equations of Mathematical Physics


K(x, y.ζ) = −Lx K(x, y; ζ),
∂ζ
lim K(x, y; ζ) = δ (D) (x − y). (B.26)
η→0+

The Greeen’s function K(x, y, ζ) has the asymptotic expansion


* ∞ +

lim K(x, y; ζ) ∼ K0 (x, y; ζ) m
ζ Hm (x, y) , (B.27)
ζ→0+
m=0

where K0 (x, y, ζ) is the evolution kernel for the n-dimensional Laplacian


(−∂ 2 ). By substituting Eq. (B.27) into Eq. (B.26) and taking the coinciding
limit x → y, we obtain the following recurrence relation for the coefficients
Hm (x, x):

*∞ ∞
 1 n  ζn  ζn x
ζ Hn+1 (x, x) = − (−∂x2 )Hn (x, x) + an (x) ∂ Hn (x, x)
n=0
n! n=0
n! n=0
n! µ

+
 ζn
+ V (x) Hn (x, x) . (B.28)
n=0
n!

For D = 4, Eq. (A.5) yields the Seeley coefficients

H0 (x, x) = 14×4 ,
H2 (x) = −V (x),
H2 (x, x) = 2[−∂x2 V (x) + aµ (x)∂µ V (x) + V 2 (x)]. (B.29)

Now substituting the value aµ (x) = [−ζ ḡ∂µ φ]14×4 and V (x) =
[(−ḡζγ5 ∂x2 φ + ∂µ φ]14×4 into Eq. (A.6), we obtain the result (Eqs. (B.14)
and (B.15)) quoted in the main text of this appendix.
August 6, 2008 15:46 9in x 6in B-640 ch07

Chapter 7

Nonlinear Diffusion in RD and Hilbert Spaces:


A Path-Integral Study∗

7.1. Introduction

The deterministic nonlinear diffusion equation is one of the most important


topics in mathematical physics of the nonlinear evolution equation
theory.1–3 An important class of initial-value problem in turbulence has
been modeled by nonlinear diffusion stirred by random sources.4
The purpose of this complementary chapter for Chap. 4, in mathe-
matical methods for physics, is to provide a model of nonlinear diffusion
were one can use and understand the compacity functional analytic argu-
ments to produce the theorem of existence and uniqueness on weak solu-
tions for deterministic stirring in L∞ ([0, T ] × L2 (Ω)). We use these results
to give a first step “proof” for the famous Rosen path-integral represen-
tation for the Hopf characteristic functional associated to the white-noise
stirred nonlinear diffusion model. These studies are presented in Sec. 7.2.
In Sec. 7.3, we present a study of a linear diffusion equation in a Hilbert
space, which is the basis of the famous loop wave equations in string and
polymer surface theory.3,4

7.2. The Nonlinear Diffusion

Let us start our chapter by considering the following nonlinear diffusion


equation in some strip Ω × [0, T ] with Ω̄ denoting a C ∞ -compact domain
of RD .

∂U (x, t)
= (+∆U )(x, t) + ∆(∧) F (U (x, t)) + f (x, t), (7.1)
∂t

∗ Author’s original results.

209
August 6, 2008 15:46 9in x 6in B-640 ch07

210 Lecture Notes in Applied Differential Equations of Mathematical Physics

with initial and Dirichlet boundary conditions as

U (x, 0) = g(x) ∈ L2 (Ω), (7.2)


U (x, t) |∂Ω ≡ 0 (for t > 0). (7.3)

We note that the nonlinearity of the diffusion-spatial term of the


parabolic problem (Eq. (7.1)) takes into account the physical properties
of nonlinear porous medium’s diffusion saturation physical situation where
this model is supposed to be applied1 — by means of the hypothesis that the
regularized Laplacean operator ∆(∧) in the nonlinear term of the governing
diffusion equation (7.1) has a cut-off in its spectral range. Additionally, we
make the hypothesis that the nonlinear function F (x) is a bounded real
continuously differentiable function on the extended interval (−∞, ∞) with
its derivative F  (x) strictly positive there. The external source f (x, t) is
supposed to belong to the space L∞ ([0, T ] × L2 (Ω)) or to be a white-noise
external stirring of the form (Ref. 2, p. 61) when in the random case
 
d  d
f (·, t) = βn (t)ϕn (·) = w(t). (7.4)
dt dt
n∈Z

Here, {ϕn } denotes a complete orthonormal set on L2 (Ω), and βn (t), n ∈ Z


are independent Wiener processes.
Let us show the existence and uniqueness of weak solutions for the
above diffusion problem stated by means of Galerking method for the case
of deterministic f (x, t) ∈ L∞ ([0, T ] × L2 (Ω)).
Let {ϕn (x)} be spectral eigen-functions associated to the Laplacean ∆.
Note that each ϕn (x) ∈ H 2 (Ω) ∩ H01 (Ω).3 We introduce now the (finite-
dimensional) Galerkin approximants

n
(n)
U (n) (x, t) = Ui (t)ϕi (x),
i=1


n
f (n) (x, t) = (f (x, t), ϕi (x))L2 (Ω) ϕi (x), (7.5)
i=1

subject to the initial-conditions



n
U (n) (x, 0) = (g(x), ϕi )L2 (Ω) ϕi (x), (7.6)
i=1

here (, )L2 (Ω) denotes the usual inner product on L2 (Ω).


August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 211

After substituting Eqs. (7.5) and (7.6) in Eq. (7.1), one gets the weak
form of the nonlinear diffusion equation in the finite-dimension approxi-
mation as a mathematical well-defined systems of ordinary nonlinear differ-
ential equations, as a result of an application of the Peano existence-solution
theorem.
 
∂U (n) (x, t)
, ϕj (x) + (−∆U (n) (x, t), ϕj (x))L2 (Ω)
∂t 2
L (Ω)

= (∇(∧) · [(F  (U (n) (x, t))∇(∧) U (n) (x, t)], ϕj (x))L2 (Ω)

+ (f (n) (x, t), ϕj (x))L2 (Ω) . (7.7)

By multiplying the associated system (Eq.(7.7)) by U (n) we get the diffusion


equation in the finite-dimensional Galerking sub-space in the integral form:

1 d
U (n) 2L2 (Ω) + (−∆U (n) , U (n) )L2 (Ω)
2 dt

+ d3 x(F  (U (n) )(∇(∧) U (n) · ∇(∧) U (n) )(x, t) = (f, U (n) )L2 (Ω) . (7.8)

This result in turn yields a prior estimate for any positive integer p:

1 d 1
(U (n) 2L2 (Ω) ) + γ(Ω)U (n) 2L2 (Ω) + (F  (U (n) )) 2 (∇U (n) )2L2 (Ω)
2 dt
 
1 1 (n) 2
≤ pf (x, t)L2 (Ω) + U L2 (Ω) .
2
(7.9)
2 p

Here, γ(Ω) is the Garding–Poincaré constant on the inequality of the


quadratic form associated to the Laplacean operator defined on the domain
H 2 (Ω) ∩ H01 (Ω).

U (n) 2H 1 (Ω) = (−∆U (n) , U (n) )L2 (Ω) ≥ γ(Ω)U (n) ||2L2 (Ω) . (7.10)

By choosing the integer p big enough and applying the Gronwall


lemma, we obtain that the set of function {U (n) (x, t)} forms a bounded
set in L∞ ([0, T ], L2 (Ω)) ∩ L∞ ([0, T ], H01 (Ω)) and in L2 ([0, T ], L2 (Ω)). As
a consequence of this boundedness property of the function set {U (n) },
there is a sub-sequence weak-star convergent to a function Ū (t, x) ∈
L∞ ([0, T ], L2(Ω)), which is the candidate for our “weak” solution of
Eq. (7.1).
August 6, 2008 15:46 9in x 6in B-640 ch07

212 Lecture Notes in Applied Differential Equations of Mathematical Physics

Another important estimate is to consider again Eq. (7.9), but now


considering the Sobolev space H01 (Ω) on this estimate (Eq. (7.9)), namely,
 T
1
(U (T )L2 (Ω) − U (0) ) + C̄0
(n) 2 (n) 2
dt||U (n) ||2H 1 (Ω)
2 0
0

 T  T
1 1
≤ p ||f ||2L2 (Ω) dt + U (n) 2L2 (Ω) dt < M̄ < ∞,
2 0 2p 0
(7.11)

since we have the coerciveness condition for the Laplacean operator

(−∆U (n) , U (n) )L2 (Ω) ≥ C̄0 (U (n) , U (n) )H01 (Ω) . (7.12)

Note that U (n) (0)2 ≤ 2g(x)2L2 (Ω) (see Eq. (7.8)), and {U (n) (T )2L2 (Ω) }
is a bounded set of real positive numbers.
As a consequence of a prior estimate of Eq. (7.11), one obtains that the
previous sequence of functions {U (n) } ∈ L∞ ([0, T ], H01(Ω) ∩ H 2 (Ω)) forms
a bounded set on the vector-valued Hilbert space L2 ([0, T ], H01 (Ω)) either.
Finally, one still has another a prior estimate after multiplying the
Galerkin system (Eq. (7.7)) by the time-derivatives U̇ (n) , namely
 T 2
dUn (t)
dt ≤ Real(∆Un (T ), Un (T ))
0 dt L2 (Ω)
 T  
dUn
− (Un (0), Un (0)) + (∧)
dt ∆ F (Un (t)),
0 dt L2 (Ω)
 T 2
 T 2
1 1 dUn
≤ p ∆ (∧)
F (Un (t)) dt + dt .
2 0 L2 (Ω) 2p 0 dt L2 (Ω)

(7.13)

By noting that
 2
T sup {F (x)}
∆(∧) F (Un (t))2L2 (Ω) dt ≤ ∆(∧) 2op
0 x ∈ [−∞, ∞]
 T
× dtUn (t)2L2 (Ω) < ∞, (7.14)
0

dt } is
one obtains as a further result that the set of the derivatives { dU n

2 2 2 −1
bounded in L ([0, T ], L (Ω)) (so in L ([0, T ], H (Ω)).
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 213

At this point, we apply the famous Aubin–Lion theorem3 to obtain


the strong convergence on L2 (Ω) of the set of the Galerkin approximants
{Un (x, t)} to our candidate Ū (x, t), since this set is a compact set in
L2 ([0, T ], L2(Ω)) (see Appendix A).
By collecting all the above results we are lead to the strong conver-
gence of the L2 (Ω)-sequence of functions F (Un (x, t)) to the L2 (Ω) function
F (Ū (x, t)).
We now assemble the above-obtained rigorous mathematical results to
obtain Ū (x, t) as a weak solution of Eq. (7.1) for any test function v(x, t) ∈
C0∞ ([0, T ]), H 2 (Ω) ∩ H01 (Ω))
 T  
dv
lim dt U , −(n)
+ (−∆U (n) , v)L2 (Ω) (F (U (n) ), −∆(∧) v)L2 (Ω)
n→∞ 0 dt L2 (Ω)
 T
= lim dt(f (n) , v), (7.15)
n→∞ 0

or in the weak-generalized sense mentioned above


 T  
dv(x, t)
dt Ū (x, t), − + (Ū (x, t), (−∆v)(x, t))L2 (Ω)
0 dt L2 (Ω)

+ (F (Ū (x, t), −(∆(∧) v(x, t))L2 (Ω)


 T
= dt(f (x, t), v(x, t))L2 (Ω) , (7.16)
0

since v(0, x) = v(T, x) ≡ 0 by our proposed space of time-dependent test


functions as C0∞ ([0, t], H 2 (Ω) ∩ H01 (Ω)), suitable to be used on the Rosens
path-integrals representations for stochastic systems (see Eqs. (7.22a) and
(7.22b) in what follows).
The uniqueness of our solution Ū (x, t), comes from the following
lemma.4

Lemma 7.1. If Ū(1) and Ū(2) in L∞ ([0, T ] × L2 (Ω)) are two functions
satisfying the weak relationship below

 T  
∂v
dt Ū(1) − Ū(2) , − + (Ū(1) − Ū(2) , +∆v)L2 (Ω)
0 ∂t L2 (Ω)

× (F (Ū(1) − F (Ū(2) ); + ∆v)L2 (Ω) ≡ 0, (7.17)
August 6, 2008 15:46 9in x 6in B-640 ch07

214 Lecture Notes in Applied Differential Equations of Mathematical Physics

then Ū(1) = Ū(2) a.e in L∞ ([0, T ] × L2(Ω)). The proof of Eq. (7.17) is easily
obtained by considering the family of test functions on Eq. (7.16) of the
following form vn (x, t) = g(ε) (t)e+αn t ϕn (x) with −∆ϕn (x) = αn ϕn (x) and
g(t) = 1 for (ε, T − ε) with ε > 0 arbitrary. We can see that it reduces to
the obvious identity (αn > 0).
 T −ε
dt exp(αn t)(F (Ū(1) ) − F (Ū(2) ), ϕn )L2 (Ω) ≡ 0, (7.18)
ε

which means that F (Ū(1) ) = F (Ū(2) ) a.e on (0, T )×Ω since ε is an arbitrary
number. We have thus Ū(1) = Ū(2) a.e, as F (x) satisfies the lower bound
estimate by our hypothesis on the kind of nonlinearity considered in our
nonlinear diffusion equation (7.1).

inf(F  (x))
|F (x) − F (y)| ≥ |x − y|. (7.19)
−∞ < x < +∞

Let us now consider a path-integral solution of Eq. (7.1) (with g(x) = 0)


for f (x, t) denoting the white-noise stirring4

E(f (x, t)f (x , t )) = λδ (D) (x − x )δ(t − t ), (7.20)

where λ is the noise strength.


The first step is to write the generating process stochastic functional
through the Rosen–Feynman path-integral identities4
    
T
Z[J(x, t)] = Ef exp i dt dD xU (x, t, [f ])J(x, t) , (7.21a)
0 Ω

Z[J(x, t)] = Ef DF [U ]δ (F ) (∂t U − ∆U − ∆(∧) (F (U )) − f )
   
T
× exp i dt D
d xU (x, t)J(x, t) , (7.21b)
0 Ω
   T 
Z[J(x, t)] = Ef DF [U ]DF [λ] exp i dt dD xλ(x, t)
0 Ω

× (∂t U − ∆U − ∆ (F (U ) − f ))
(∧)

   
T
× exp i dt D
d xU (x, t)J(x, t) , (7.21c)
0 Ω
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 215

   T 
1
Z[J(x, t)] = DF [U ] exp − dt dD x
2λ 0 Ω

× (∂t U − ∆U − ∆(∧) (F (U ))2 (x, t))
   
T
× exp i dt dD xU (x, t)J(x, t) . (7.21d)
0 Ω

The important step made rigorous mathematically possible on the above


(still formal) Rosen’s path-integral representation by our previous rig-
orous mathematical analysis is the use of the delta functional identity on
Eq. (7.21b) which is true only in the case of the existence and uniqueness
of the solution of the diffusion equation in the weak sense at least for mul-
tiplier Lagrange fields λ(x, t) ∈ C0∞ ([0, T ], H 2 (Ω) ∩ H01 (Ω)).
As an important mathematical result to be pointed out is that in general
case of a nonporous medium4 in R3 , where one should model the diffusion
nonlinearity by a complete Laplacean ∆F (U (x, t)), one should observe that
the set of (cut-off) solutions {Ū (∧) (x, t)} of Eq. (7.1) still remains a bounded
set on L∞ ([0, T ], L2 (Ω)). Since we have the a priori estimate uniform bound
for the U (n) -derivatives below in D = 3 (with G (x) = F (x)). Namely
    
 T dU (n)
2   T 
   dU (n) (t) 
 dt ≤ dt d x(∆F (U (t))) ·
3 (n)

 0 dt L2 (Ω)   0 Ω dt 
 
  T   
 d(U (n)
(x, t))   d 
+  d xf (x, t)
3  ≤  dt Real
  d x∆G(U (t) 
3 (n)
Ω dt 0 dt Ω 
 
1 1 · (n)
+ sup pf (x, t)2L2 (Ω) + U 2L2 (Ω)
2 0≤t≤T p
  
 

≤ Real d x(∆G(U (T, x)) − ∆G(U (0, x)) 
3 (n) (n)

 
1 1 · (n)
+ sup pf (x, t)2L2 (Ω) + U L2 (Ω)
2 0≤t≤T p
1 1 · (n)
≤ pf 2L∞ ((0,t),L2 (Ω)) + U (t)L∞ ((0,t),L2 (Ω)) < ∞, (7.22)
2 2p
 
 
where U (n) (T, x) = U (n) (0, x) = 0 (see Eq. (7.3)). The uniform
∂Ω ∂Ω
1
bound for the derivatives is achieved by choosing 2p < 1.
August 6, 2008 15:46 9in x 6in B-640 ch07

216 Lecture Notes in Applied Differential Equations of Mathematical Physics

As another point worth to call to attention is that the above considered


function space is the dual of the Banach space L1 ([0, T ], L2 (Ω)). So, one
can extract from the above set of cut-off solutions a candidate Ū (∞) (x, t),
in the weak-star topology of L∞ ([0, T ], L2 (Ω)) for the above cited case of
cut-off removing ∧ = +∞.6 However, we will not proceed thoroughly in
this straightforward technical question of cut-off removing our model of
nonlinear diffusion in this chapter for general spaces RD .
Finally, we remark that in the one-dimensional case Ω ∈ R1 , one
can further show by using the same compacity methods the existence
and uniqueness of the diffusion equation added with the hydrodynamic

advective term 12 dx (U (x, t))2 , which turns the diffusion equation (7.1) as a
kind of nonlinear Burger equation on a porous medium.
It appears very important to remark that Galerking methods applied
directly to the finite-dimensional stochastic equation (7.7) (see Eq. (7.4))
may be saving-time computer simulation candidates for the “turbulent”
path-integral (Eqs. (6.22a)–(6.22d)) evaluations by approximate numerical
methods (Ref. 2).

7.3. The Linear Diffusion in the Space L2 (Ω)

Let us now present some mathematical results for the diffusion problem
in Hilbert spaces formed by square-integrable functions L2 (Ω),5 with the
domain Ω denoting a compact set of RD .
The diffusion equation in the infinite-dimensional space L2 (Ω) is given
by the following functional differential equation (see Ref. 5 for mathematical
notation).

∂ψ[f (x); t] 1
= TrL2 (Ω) ([QDf2 ψ[f (x, t)]) (7.23a)
∂t 2

ψ[f (x), t → 0+ ] = Ω[f (x)]. (7.23b)

Here, ψ[f (x), ·] is a time-dependent functional to be determined through


the governing equation (7.23) and belonging to the space L2 (L2 (Ω), dQ µ(f ))
with dQ µ(f ) denoting the Gaussian measure on L2 (Ω) associated to Q —
a fixed positive self-adjoint trace class operator 1 (L2 (Ω)) — and Df2 is the
second — Frechet derivative of the functional ψ[f (x), t] which is given by
a f (x)-dependent linear operator on L2 (Ω) with associated quadratic form
(Df2 ψ[f (x), t] · g(x), h(x))L2 (Ω) .
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 217

By considering explicitly the spectral base of the operator Q on L2 (Ω)

Qϕn = λn ϕn , (7.24)
The L2 (Ω)-infinite-dimensional diffusion equation takes the usual form:
 

Ψ fn ϕn , t = ψ (∞) [(fn ), t], (7.25a)
n
 

Ω fn ϕn = Ω(∞) [(fn )], (7.25b)
n

∂ψ (∞) [(fn ), t] 
= [(λn ∆fn )ψ (∞) [(fn ), t]], (7.25c)
∂t n

ψ (∞) [(fn ), 0] = Ω(∞) [(fn )], (7.25d)

or in the Physicist’ functional derivative form (see Ref. 5).


 
∂ 1 δ2
ψ[f (x), t] = D
d x dD x Q(x, x ) ψ[f (x), t], (7.26a)
∂t 2 Ω Ω δf (x )δf (x)
ψ[f (x), 0] = Ω[f (x)]. (7.26b)

Here, the integral operator Kernel of the trace class operator is explicitly
given by

Q(x, x ) = (λn ϕn (x)ϕn (x )). (7.26c)
n

A solution of Eq. (7.26a) is easily written in terms of Gaussian path-


integrals5 which reads in the physicist’s notations

1 1 −1
ψ[f (x), t] = DF [g(x)]Ω[f (x) + g(x)] × det 2 Q
2
L (Ω) t
   
1 D  −1  
× exp − D
d x d x g(x) · Q (x, x )g(x ) . (7.27)
2t Ω Ω

Rigorously, the correct functional measure on Eq. (7.27) is the nor-


malized Gaussian measure with the following generating functional
   
Z[j(x)] = dtQ µ[g(x)] exp i j(x)g(x)dD x
L2 (Ω) Ω
   
t
= exp − dD x dD x j(x)Q+1 (x, x )j(x ) . (7.28)
2 Ω Ω
August 6, 2008 15:46 9in x 6in B-640 ch07

218 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point, it becomes important to remark that when writing the


solution as a Gaussian path-integral average as done in Eq. (7.27), all the
L2 (Ω) functions in the functional domain of our diffusion functional field
ψ[f (x), t] belongs to the functional domain of the quadratic form asso-
ciated to the class trace operator Q the so-called reproducing kernel of the
operator Q which is not the whole Hilbert space L2 (Ω) as naively indicated
on Eq. (7.27), but the following subset of it:
1 ⊂
Dom(ψ[·, t]) = {f (x) ∈ L2 (Ω)|Q− 2 f ∈ L2 (Ω)} = L2 (Ω). (7.29)
The above result gives a new generalization of the famous Cameron–
Martin theorem that the usual Wienner measure (defined by the one-
dimensional Laplacean with Dirichlet conditions on the interval
 end-points)

Wien
is translation invariant, i.e. dWien µ[f + g] = dWien µ[f ] × d dWienµ[f +g]
µ[f ] , if
and only if the shift function g(x) is absolutely continuous with derivative
 
d2
on L ([a, b]). In other words, g ∈ H0 ([a, b]) = Dom
2 1
− d2 x .
Another important point to call to the readers’ attention is that one
can write Eq. (7.27) in the usual form of diffusion in finite-dimensional case
(see Appendix B):
  
√ 1 1
ψ[f (x), t] = D [g(x)]Ω[f (x) + tg(x)] × det
F 2 Q
L2 (Ω) t
   
1 D  −1 
× exp − D
d x d x g(x)Q (x, x )g(x) , (7.30)
2 Ω Ω

At this point it is worth to call to the readers’ attention that dtQ µ and
dQ µ Gaussian measures are singular to each other by a direct application of
Kakutani theorem for Gaussian infinite-dimensional measures for any time
t > 0:
dtQ µ[g(x)]/dQ µ[g(x)] = +∞. (7.31)
Let us apply the above results for the physical diffusion of polymer
rings (closed strings) described by periodic loops X(σ) ∈ RD , 0 ≤ σ ≤

 A = X(σ) with a nonlocal diffusion coefficient Q(σ, σ ) (such that
A,AX(σ+A)
0
dσ 0 dσ Q(σ, σ ) = Tr[Q] < ∞). The functional governing equation in
loop space (formed by polymer rings) is given by
 A  A
∂ψ (ε) [X(σ); A] δ2
dσ  Qij (σ, σ  )
(ε)
= dσ ψ (ε) [X(σ), A],
∂A 0 0 δ Xi (σ)δ Xj (σ)
(7.32a)
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 219

   
A A
λ   
ψ (ε)
[X(σ); 0] = exp − dσ dσ Xi (σ)Mij (σ, σ )Xj (σ ) .
2 0 0

(7.32b)

Here, the ring polymer surface probability distribution ψ (ε) [X(σ), A]


depends on the area parameter A, the area of the cylindrical polymer surface
of our surface-polymer chain. Note that the presence of a parameter ε on
the above objects takes into account the local (the integral operator kernel)
case Q(σ, σ  ) = δ(σ − σ  ) as a limiting case of the rigorously mathematical
well-defined (class trace) situation on the end of the observable evaluations
 
(ε)  1 (σ − σ  )2
Qij (σ, σ ) = exp − . (7.33)
πε ε2
The solution of Eq. (7.32a) is straightforwardly written in the case of a
self-adjoint kernel M on L2 ([0, A]):
   A 
λ A 1
exp − dσ dσ  Xi (σ)Mij (σ, σ  )Xj (σ  ) det− 2 [1 + AλM (Q(ε) )]
2 0 0
   A  −1 
λ2 A  (ε) −1 1 
× exp + dσ dσ (M X)i (σ) (λM + (Q ) · (M X)j (σ ) .
2 0 0 A

(7.34)

The functional determinant can be reduced to the evaluation of an


integral equation
1
det− 2 [1 + AλM (Q(ε) )]
 
1
= exp − TrL (Ω) lg(1 + λAM (Q )
2
(ε)
2
  λ 
1   −1
= exp − TrL2 (Ω) (ε) (ε)
dλ [(Q )M )(1 + λ A(Q )M ) ]
2 0
  λ 
1
= exp − TrL2 (Ω) dλ R(λ ) . (7.35)
2 0

Here, the kernel operator R(λ ) satisfies the integral equation (accessible
for numerical analysis)

R(λ )(1 + λ A(Q(ε) )M = (Q(ε) )M. (7.36)


August 6, 2008 15:46 9in x 6in B-640 ch07

220 Lecture Notes in Applied Differential Equations of Mathematical Physics

Which in the local case of ε → 0+ , when considered in the final result


(Eqs. (7.34) and (7.35)), produces the explicitly candidate solutions for our
polymer-surface probability distribution with M a class trace operator on
the loop space, L2 ([0, A]):
  λ 
A  (ε)−1  −1
ψ[X(σ), A] = exp − TrL2 (Ω) dλ [M (Q + λ AM ) ]
2 0
   A 
λ A
× exp − dσ dσ  Xi (σ) · Mij (σ, σ  )Xj (σ  )
2 0 0
   A
λ2 A
× exp + dσ dσ  (M X)i (σ)
2 0 0
 −1 
(ε) −1 1  
× λM + (Q ) · (σ, σ )(M X)j (σ ) .
A

(7.37)

It is worth to call to the readers’ attention that if A ∈ 1 and B is
a bounded operator — so A · B is a class trace operator. The functional
determinant det[1 + AB] is a well-defined object as a direct result of the
obvious estimate, the result which was used to arrive at Eq. (7.37).


N 
N
lim (1 + λn ) ≤ exp λn = exp (Tr AB).
N →∞
n=0 n=0

As a last comment on the linear infinite-dimensional diffusion problem


(Eq. (7.23)), let us sketch a (rigorous) proof that Eq. (7.27) is the unique
solution of Eq.(7.23). First, let us consider the initial
.. 2 condition on Eq. (7.23)
as belonging to the space of all mappings G.L (Ω) → R that are twice
Fréchet differentiable on L2 (Ω) with uniformly continuous and bounded
second derivative Df2 G (a bounded operator of L(L2 (Ω)) with norm C̄).
This set of mappings will be denoted by uC 2 [L2 (Ω), R]. It is, thus, straight-
forward to see through an application of the mean-value theorem that the
following estimate holds true

sup |ψ[f (x), t] − G[f (x)]|


1
f (x)∈Q 2 L2 (Ω)

≤ |G(f (x) + g(x)) − G(g(x))|dtQ µ[g(x)]
L2 (Ω)
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 221

  
 1
≤ DG(f (x), g(x))L2 (Ω) + dσ(1 − σ)(D2 G[f (x)

L2 (Ω) 0


+ σg(x)]g(x), g(x)L2 (Ω)  dtQ µ[g(x)]
 1 
≤ 0 + C̄ dσ(1 − σ) g(x)2L2 (Ω) dtQ µ[g(x)]
0 L2 (Ω)

≤ C̄ g(x)2L2 (Ω) dtQ µ[g(x)]
L2 (Ω)

≤ C̄ Tr(tQ) = (C̄ Tr(Q)t → 0 as t → 0+ . (7.38)

We have thus defined a strongly continuous semi-group on the Banach


space U C 2 [L2 (Ω), R] with infinitesimal generator given by the infinite-
1
dimensional Laplacean Tr[QD2 ] acting on the space L2 (Q 2 (L2 (Ω)), R).
By the general theory of semi-groups on Banach spaces, we obtain that
Eq. (7.27) satisfies the infinite-dimensional diffusion initial-value problem
Eq. (7.23), at least for initial conditions on the space uC 2 [L2 (Ω), R]. Since
purely Gaussian functionals belong to uC 2 [L2 (Ω), R] and they form a dense
set on the space L2 (L2 (Ω), dQ µ), we get the proof of our result for general
initial condition on L2 (L2 (Ω), dQ µ).
Finally, we point out that the general solution of the diffusion problem
on Hilbert space with sources and sinks, namely

∂ 1
ψ[f (x), t] = TrL2 (Ω) [QDf2 ψ[f (x), t]] − V [f (x)]ψ[f (x), t] (7.39)
dt 2

with

ψ[f (x), t → 0+ ] = Ω[f (x)], (7.40)

possesses a generalized Feynman–Wiener–Kac Hilbert L2 (Ω) space-valued


path-integral representation, which in the Feynman physicist formal
notation reads as

ψ[h(x), T ] = DF [X(σ)]
C([0,T ],L2 (Ω))
    
T
1 dX −1 dX
× exp − dσ ,Q (σ)
2 0 dσ dσ L2 (Ω)
August 6, 2008 15:46 9in x 6in B-640 ch07

222 Lecture Notes in Applied Differential Equations of Mathematical Physics

  
T
×Ω X(σ)dσ + X(0)
0
    
T T
 
× exp − dσV X(σ )dσ + X(0) , (7.41)
0 0

where the paths satisfy the end-point constraint X(T ) = h(x) ∈


L2 (Ω); X(0) = f (x) ∈ L2 (Ω).

References

1. I. Sneddon, Fourier Transforms (McGraw-Hill Book Company, INC, USA,


1951); P. A. Domenico and F. W. Schwartz, Physical and Chemical Hydroge-
ology, 2nd edn. (Wiley, 1998); P. L. Schdev, Nonlinear Diffusive Waves (Cam-
bridge University Press, 1987); W. E. Kohler; B. S. White (ed.), Mathematics
of Random Media, Lectures in Applied Mathematics, Vol. 27 (American Math.
Soc., Providence, Rhode Island, 1980); R. Z. Sagdeev (ed.), Non-Linear and
Turbulent Processes in Physics, Vols. 1–3 (Harwood Academic Publishers,
1987); Geoffrey Grimmett, Percolation (Springer-Verlag, 1989).
2. G. Da Prato and J. Zabczyk, Ergodicity for Infinite Dimensional System,
London Math. Soc. Lect. Notes, Vol. 229 (Cambridge University Press, 1996);
R. Teman, C. Foias, O. Manley and Y. Treve, Physica 6D, 157 (1983).
3. R. Dortray and J.-L. Lion, Analyse Mathématique et Calcul Numérique,
Masson, Paris, Vols. 1–9, A. Friedman, Variational Principles and Free-
Boundary Problems — Pure and Applied, Math. (Wiley, 1988).
4. L. C. L. Botelho, J. Phys. A: Math. Gen. 34, L31–L37 (2001); L. C. L. Botelho,
II Nuovo Cimento 117 (B1), 15 (2002); J. Math. Phys. 42, 1682 (2001); L. C.
L. Botelho, Int. J. Mod. Phys. B 13 (13), 1663 (1999); Int. J. Mod. Phys. 12;
Mod. Phys. B 14(3), 331 (2002); A. S. Monin and A. M. Yaglon, Statistical
Fluid Mechanics, Vol. 2 (Mit Press, Cambridge, 1971); G. Rosen, J. Math.
Phys. 12, 812 (1971); L. C. L. Botelho, Mod. Phys. Lett 13B, 317 (1999); V.
Gurarie and A. Migdal, Phys. Rev. E 54, 4908 (1996); U. Frisch, Turbulence
(Cambridge Press, Cambridge, 1996); W. D. Mc-Comb, The Physics of Fluid
Turbulence (Oxford University, Oxford, 1990); L. C. L. Botelho, Mod. Phys.
Lett. B 13, 363 (1999); S. I. Denisov and W. Horsthemke, Phys. Lett. A 282
(6), 367 (2001); L. C. L. Botelho, Int. J. Mod. Phys. B 13, 1663 (1999); L. C.
L. Botelho, Mod. Phys. Lett. B 12, 301 (1998); L. C. L. Botelho, Mod. Phys.
Lett. B 12, 569 (1998); L. C. L. Botelho, Mod. Phys. Lett. B 12, 1191 (1998);
L. C. L. Botelho, II Nuovo Cimento 117 B(1), 15 (2002); J. Math. Phys. 42,
1682 (2001).
5. G. da Prato and J. Zabczyk, Second Order Partial Differential Equations in
Hilbert Spaces, Vol. 293 (London Math. Soc. Lect. Notes, Cambridge, UK,
2002).
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 223

6. J. Wloka, Partial Differential Equations (Cambridge University Press, 1987).


7. Soo Bong Chae, Holomorphy and Calculus in Normed Spaces, Pure and
Applied Math. Series (Marcel Denner Inc, NY, USA, 1985).

Appendix A. The Aubin–Lion Theorem

Just for completeness in this mathematical appendix for our mathematical


oriented readers, we intend to give a detailed proof of the basic result on
compacity of sets in functional spaces of the form L2 (Ω) and throughout
used in Sec. 2. We have, thus, the Aubin–Lion theorem3 in the Gelfand
triplet H01 (Ω) → L2 (Ω) → H −1 (Ω) = (H01 (Ω))∗
Aubin–Lion Theorem. If {Un (x, t)} is a sequence of time-differentiable
functions in a bounded set of L2 ([0, T ], H01 (Ω)) such that its time derivatives
forms a bounded set of L2 ([0, T ], H0−1(Ω)), we have that {Un (x, t)} is a
compact set on L2 ([0, T ], L2 (Ω)).
Proof. The basic fact we are going to use to give a mathematical proof
of this theorem is the following identity (Ehrling’s lemma). For any given
ε > 0, there is a constant C(ε) such that

Un L2 (Ω) ≤ εUn H01 (Ω) + C(ε)Un 2H −1 (Ω) . (A.1)

As a consequence, we have the following estimate


 T
Un − Um 2L2 ([0,T ]),L2 (Ω)
0
 T
≤ dt(εUn − Um H01 (Ω) + C(ε)Un − Um H −1 (Ω) )2
0
 T  T
≤ ε2 dtUn − Um 2H 1 (Ω) + (C(ε)2 dtUn − Um 2H −1 (Ω)
0
0 0

 T
+ 2εC(ε) dt(Un − Um H01 (Ω) × Un − Um H −1 (Ω) )
0
0

 T  T
≤ ε2 dtUn − Um 2H 1 (Ω) + C(ε)2 dtUn − Um 2H −1 (Ω)
0
0 0

 T
1
2  T
1
2

+ 2εC(ε) dtUn − Um 2H 1 (Ω) + dtUn − Um 2H −1 (Ω)


0
0 0
August 6, 2008 15:46 9in x 6in B-640 ch07

224 Lecture Notes in Applied Differential Equations of Mathematical Physics

 T
1
2
1
≤ 2ε2 M + 2εC(ε)M 2 dtUn − Um 2H −1 (Ω)
0

 T
+ (C(ε)2 ) dtUn − Um 2H −1 (Ω) . (A.2)
0

At this point, we use the Arzela–Ascoli theorem to see that {Un (x, t)} is
a compact set on the space C([0, T ], H −1 (Ω)) since we have the set equi-
continuity:
 t
Un (t) − Um (s)H −1 (Ω) ≤ Un (τ )H −1 (Ω) dτ
0
s
 T
1
2
(1− 12 )
≤ |t − s| Un (τ )2H −1 (Ω) dτ
0
1
≤ M̄ |t − s| 2 . (A.3)

It is a crucial step now by remarking that H01 (Ω) is compactly immersed


in L2 (Ω) (Rellich theorem). Let us note that for each t (almost every-
where in [0, T ]), Un (x, t) is a bounded set on H01 (Ω) since Un (x, t) belongs
to a bounded set L2 ([0, T ], H01 (Ω)) by hypothesis. As a consequence,
{Unk , (x, t)} is a compact set on L2 (Ω) (Rellich theorem) and so in H −1 (Ω)
almost everywhere in [0, T ] since L2 (Ω) → H −1 (Ω). By an application of
the Arzela–Ascoli theorem, there is a sub-sequence {Unk (x, t)} of {Un (x, t)}
(and still denoted by {Un (x, t)} such that it converges uniformly to a given
function Ū (x, t) ∈ C([0, T ], H −1(Ω)). As a direct result of this fact we have
that (for T < ∞!) for (n, m) → ∞.

 T
1
2

Un − Um 2H −1 (Ω)


0

 T
1
2

≤ (sup |Un − Um |C([0,T ],H −1 (Ω) ) 1 · dt → 0. (A.4)


0

Returning to our estimate (Eq. (A.2)), we see that this sub-sequence is


a Cauchy sequence in L2 ([0, T ], L2(Ω)). As a consequence, for each fixed
t ∈ [0, T ] (almost everywhere), Un (x, t) converges to Ū (x, t) in L2 (Ω). 
August 6, 2008 15:46 9in x 6in B-640 ch07

Nonlinear Diffusion in RD and Hilbert Spaces 225

Appendix B. The Linear Diffusion Equation

Let us show mathematically the basic functional integral representation


(Eq. (7.30)) for the L2 (Ω)-space diffusion equation (7.23).
As a first step for such proof, let us call the readers’ attention that
one should consider the second-order (Laplacean) D2 U (x, t) as a bounded
operator in L2 (Ω) in order to the operatorial composition with the positive
definite class trace operator Q still be a class trace operator as it is explicitly
supposed in the right-hand side of Eq. (7.23).
We thus impose as the sub-space of initial condition the diffusion
equation (7.23) for the (dense) vector sub-space of C(L2 (Ω), R) composed
of all functionals of the form

f (x) = dQ µ(p)F (p) exp(i p, xL2 (Ω) ), (B.1)
L2 (Ω)

with F (p) ∈ L2 (L2 (Ω), dQ µ).


By substituting the initial condition (Eq. (B.1)) into the integral rep-
resentation equation (7.30) and by using the Fubbini–Toneli theorem to
exchange the needed integrations order in the estimate below, we get:


U (x, t) = f (x + tξ)dQ µ(ξ)
L2 (Ω)
  √

= dQ µ(ξ) dQ µ(p)F (p)ei p,x+ tξ L2

L2 L2

1
= dQ µ(p)F (p) · ei p,x L2 e− 2 t p,Qp L2 . (B.2)
L2

Note that we have already proved that U (x, t) is a bounded functional of


C(L2 (Ω) × [0, ∞]; R) on the basis of our hypothesis on the initial functional
date (Eq. (B.1)).
At this point, we observe that the second-order Frechet derivatives of
the functional, exp i p, xL2 , are easily (explicitly) evaluated as7

  ∞
 ∂2  P∞ 
QD2 ei p,x L2 = λ ei( n=1 pn xn ) = − ( p, QpL2 ) ei p,x L2 .
∂ 2 x
=1
(B.3)
August 6, 2008 15:46 9in x 6in B-640 ch07

226 Lecture Notes in Applied Differential Equations of Mathematical Physics

We have thus a straightforward proof of our claim cited above on the


basis again of the chosen initial date sub-space

Tr[QD2 U (x, t)]



≤ dQ µ(p)|F (p)| p, QpL2
L2 (Ω)

 1
2  1
2

≤ dQ µ(p)|F (p)| 2
dQ µ(p)| p, QpL2 (Ω) | 2
L2 (Ω) L2 (Ω)

≤ (Tr Q)2 F 2L2 (L2 (Ω),dQ µ) < ∞. (B.4)

Now, it is a simple application to verify that Eq. (B.2) satisfies the dif-
fusion equation in L2 (Ω) (or in any other separable Hilbert space). Namely
  
∂U (x, t) 1
= dQ µ(p)F (p)ei p,x L2 − p, QpL2 (Ω)
∂t L2 (Ω) 2
t
× e− 2 , p,Qp L2 (Ω) , (B.5)


t
2
TrL2 (Ω) [QD U (x, t)] = dQ µ(p)F (p){D2 ei p,x
}e− 2 p,Qp L2 (Ω)

L2 (Ω)

t
= dQ µ(p)F (p){− p, QpL2 }ei p,x
e− 2 p,Qp L2 (Ω) ,
L2 (Ω)
(B.6)

with
  
i p,x − 2t p,Qp L2 (Ω)
U (x, 0) = dQ µ(p)F (p)e lim e = f (x). (B.7)
L2 (Ω) t→0+
August 6, 2008 15:46 9in x 6in B-640 ch08

Chapter 8

On the Ergodic Theorem∗

8.1. Introduction

One of the most important phenomenon in numerical studies of the non-


linear wave motion, especially in the two-dimensional case, is the existence
of overwhelming majority of wave motions that wandering over all possible
system’s phase space and, given enough time, coming as close as desired
(but not entirely coinciding) to any given initial condition.1
This phenomenon is signaling certainly that the famous recurrence
theorem of Poincaré is true for infinite-dimensional continuum mechanical
systems like that one represented by a bounded domain when subject to
nonlinear vibrations.
A fundamental question appears in the context of this recurrence phe-
nomenon and concerned to the existence of the time average for the asso-
ciated nonlinear wave motion and naturally leading to the concept of an
infinite-dimensional invariant measure for the nonlinear wave equation,1 the
mathematical phenomenon subjacente to the Poincaré recurrence theorem.
In this chapter, we intend to give applied mathematical arguments for
the validity of the famous Ergodic theorem in a class of polynomial non-
linear and Lipschitz wave motions. Our approach is fully based on Hilbert
space methods previously used to study dynamical systems in classical
mechanics which in turn simplifies enormously the task of constructing
explicitly the associated invariant measure on certain Sobolev spaces — the
infinite-dimensional space of wave motion initial conditions. This study is
presented in the main section of this chapter, namely Sec. 8.4.
In Sec. 8.2, we give a very detailed proof of the RAGE’s theorem, a
basic functional analysis rigorous method used in our proposed Hilbert

∗ Author’s original results.

227
August 6, 2008 15:46 9in x 6in B-640 ch08

228 Lecture Notes in Applied Differential Equations of Mathematical Physics

space generalization of the usual finite-dimensional Ergodic theorem and


fully presented in Sec. 8.3.
In Sec. 8.5, we complement our studies by analyzing the important case
of wave-diffusion under random stirring.
In Appendices A and B, we additionally present important technical
details related to Sec. 8.4, and they contain material as important as those
presented in the previous sections of our work.

8.2. On the Detailed Mathematical Proof of the RAGE


Theorem

In this purely mathematical section, we intend to present a detailed math-


ematical proof of the RAGE theorem2 used on the analytical proof of the
Ergodic theorem on Sec. 8.3 for Hamiltonian systems of N -particles.
Let us, thus, start our analysis by considering a self-adjoint operator L
on a Hilbert space (H, (, )), where Hc (L) denotes the associated continuity
sub-space obtained from the spectral theorem applied to L. We have the
following result.

Proposition (RAGE theorem). Let ψ ∈ Hc (L) and ψ̃ ∈ H. We have


the Ergodic limit
 T
1
lim |(ψ̃, exp(itL)ψ)|2 dt = 0. (8.1)
T →∞ T 0

Proof. In order to show the validity of the above ergodic limit, let us
rewrite Eq. (8.1) in terms of the spectral resolution of L, namely
 T
1
I= |(ψ̃, exp(itL)ψ)|2 dt
T 0
 T  +∞   +∞ 
1 +itλ −itµ
= e dλ (ψ̃, Ej (λ)ψ) e dµ (ψ̃, Ej (µ)ψ) .
T 0 −∞ −∞
(8.2)

Here, we have used the usual spectral representation of L


 +∞
(g, Lh) = λdλ (g, Ej (λ)h), (8.3)
−∞

with g ∈ H and h ∈ Dom (L).


August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 229

Let us remark that the function exp(i(λ − µ)t) is majorized by the


function 1 which in turn is an integrable function on the domain [0, T ] ×
R × R with the product measure
dt
⊗ dλ (ψ̃, Ej (λ)ψ) ⊗ dµ (ψ̃, Ej (µ)ψ), (8.4)
T
since
 T
dt
⊗ dλ (ψ̃, Ej (λ)ψ) ⊗ dµ (ψ̃, Ej (µ)ψ) = (ψ̃, ψ)2 < ∞. (8.5)
0 T
At this point, we can safely apply the Fubbini theorem to interchange
the order of integration in relation to the t-variable which leads to the
partial result below
 +∞ +∞  i(λ−µ)T 
e −1
I= dλ ψ̃, Ej (λ)ψ ⊗ dµ ψ̃, Ej (µ)ψ. (8.6)
−∞ −∞ i(λ − µ)T

Let us consider two cases. First, we take ψ̃ = ψ. In this case, we have


the bound
  
 i(λ−µ)T   2 sin (λ−µ) T 
e 
− 1   
 2
 ≤ 1.
 i(λ − µ)T  ≤  (λ − µ)T  (8.7)
 

If we restrict ourselves to the case of λ = µ, a direct application of the


Lebesgue convergence theorem on the limit T → ∞ yields that I = 0.
On the other hand, in the case of λ = µ, we intend now to show that
the set µ = λ is a set of zero measure with R2 in respect to the measure
dλ ψ, Ej (λ)ψ ⊗ dµ ψ, Ej (µ)ψ. This can be seen by considering the real
function on R

f (λ) = (ψ, Ej (λ)ψ). (8.8)

We observe that this function is continuous and nondecreasing since


ψ ∈ Hc (L) and range Ej (λ1 )C range Ej (λ2 ) for λ1 ≤ λ2 . Now, f (λ) is
¯
a uniform continuous function on the whole real line R, since for a given
ε > 0, there is a constant λ̃ such that
ε
(ψ, Ej (−∞, −λ̃)ψ) < , (8.9)
2
ε
ψ 2 − (ψ, Ej (−∞, λ̃)ψ) < . (8.10)
2
August 6, 2008 15:46 9in x 6in B-640 ch08

230 Lecture Notes in Applied Differential Equations of Mathematical Physics

Additionally, f (λ) is uniform continuous on the closed interval [−λ̃, λ̃], and
f (λ) cannot make variations greater than 2ε on (−∞, −λ̃] and [+λ̃, ∞),
besides being a monotonic function on R. These arguments show the
uniform continuity of f (λ) on whole line R. Hence, we have that for a given
ε > 0, there exists a δ > 0 such that
ε
|(ψ, Ej (λ )ψ) − (ψ, Ej (λ )ψ)| ≤ , (8.11)
ψ 2
for

|λ − λ | ≤ δ. (8.12)

In particular, for λ = λ + δ e λ = λ − δ


ε
(ψ, Ej (λ + δ)ψ) − (ψ, Ej (λ − δ)ψ) ≤ . (8.13)
ψ 2
As a consequence, we have the estimate
 +∞  λ+δ
dλ (Ej (λ)ψ, ψ) dµ (ψ, Ej (µ)ψ)
−∞ λ−δ

≤ ψ ((ψ, Ej (λ + δ)ψ) − (ψEj (λ − δ)ψ))


2

ε
≤ ψ 2
2
≤ ε. (8.14)
ψ

Let us note that for each ε > 0, there exists a set Dδ = {(λ, µ) ∈ R2 ;
|λ − µ| < δ} which contains the line λ = µ and has measure less than
ε in relation to the measure dλ (Ej (λ)ψ, ψ) ⊗ dµ (Ej (µ)ψ, ψ) as a result of
Eq. (8.14). This shows our claim that I = 0 in our special case.
In the general case of ψ̃ = ψ, we remark that solely the orthogonal
component on the continuity sub-space Hc (L) has a nonvanishing inner
product with exp(itL)ψ. By using now the polarization formulae, we reduce
this case to the first analyzed result of I = 0.
At this point, we arrive at the complete RAGE theorem.2
RAGE Theorem. Let K be a compact operator on (H, (, )). We have thus
the validity of the Ergodic limit

1 T
lim K exp(itL)ψ 2H dt = 0, (8.15)
T →∞ T 0

with ψ ∈ Hc (L).
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 231

We leave the details of the proof of this result for the reader, since any
compact operator is the norm operator limit of finite-dimension operators
and one only needs to show that
 N
1 T
lim cn (eitL ψ, en )gn dt = 0, (8.16)
T →∞ T 0
n=0 H

for cn constants and {en }, {gn } a finite set of vector of (H, (, )) .

8.3. On the Boltzmann Ergodic Theorem in Classical


Mechanics as a Result of the RAGE Theorem

One of the most important statement in physics is the famous zeroth law
of thermodynamics: “any system approaches an equilibrium state.” In the
classical mechanics frameworks, one begins with the formal elements of the
theory. Namely, the phase-space R6N associated to a system of N -classical
particles and the set of Hamilton equations
∂H ∂H
ṗi = − ; q̇i = , (8.17)
∂qi ∂pi
where H(q, p) is the energy function.
The above cited thermodynamical equilibrium principle becomes the
mathematical statement that for each compact support continuous func-
tions Cc (R6N ), the famous ergodic limit should hold true.3

1 T
lim f (q(t); p(t))dt = η(f ), (8.18)
T →∞ T 0

where η(f ) is a linear functional on Cc (R6N ) given exactly by the


Boltzmann statistical weight, and {q(t), p(t)} denotes the (global) solutions
of the Hamilton equation (8.12).
In this section, we point out a simple new mathematical argument of
the fundamental equation (8.18) by means of Hilbert space techniques and
the RAGEs theorem. Let us begin by introducing for each initial con-
dition (q(0), p(0)), a function ωq0 ,p0 (t) ≡ (q(t), p(t)), here q(t), p(t) is the
assumed global unique solution of Eq. (8.17) with prescribed initial condi-
tions. Let Ut : L2 (R6N ) → L2 (R6N ) be the unitary operator defined by
(Ut f )(q, p) = f (ωq0 p0 (t)). (8.19)
We have the following theorem (the Liouville’s theorem).5
August 6, 2008 15:46 9in x 6in B-640 ch08

232 Lecture Notes in Applied Differential Equations of Mathematical Physics

Theorem 8.1. Ut is a unitary one-parameter group whose infinitesimal


generator is −iL̄, where −iL is the essential self-adjoint operator acting on
C0∞ (R6N ) defined by the Poisson bracket.

3N  
∂f ∂H ∂f ∂H
(Lf )(p, q) = {f, H}(q, p) = − (q, p). (8.20)
i=1
∂qi ∂pi ∂pi ∂qi

The basic result we are using to show the validity of the Ergodic limit
equation (8.18) is the famous RAGEs theorem exposed in Sec. 8.1.

Theorem 8.2. Let φ ∈ Hc (−iL̄)(L2 (R6N ), here Hc (−iL̄) is the continuity


sub-space associated to self-adjoint operator −iL̄. For every vector β ∈
L2 (R6N ), we have the result
 T
1
lim |β, Ut ψ|2 dt = 0, (8.21)
T →∞ T 0

or equivalently for every ψ ∈ L2 (R6N )


 T
1
lim β, Ut ψdt = β, Pker(−iL̄) ψ, (8.22)
T →∞ T 0

where Pker(−iL̄) is the projection operator on the (closed) sub-space


ker(−iL̄).

That Eq. (8.22) is equivalent to Eq. (8.21), is a simple consequence of


the Schwartz inequality written below
    12    12
 T  T T
 (β, Ut ψ)dt ≤ (β, Ut ψ)2 dt 1dt , (8.23)

0 0 0

or
  T    T  12
1  1
 
β, Ut ψdt ≤ 2
(β, Ut ψ) dt (8.24)
T T 0
0

As a consequence of Eq. (8.23), we can see that the linear functional


η(f ) of the Ergodic theorem is exactly given by (just consider β(p, q) ≡ 1
on supp of Pker(−iL̄) (ψ))

η(f ) = Pker(−iL̄) (f )(q, p). (8.25)


August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 233

By the Riesz’s theorem applied to η(f ), we can rewrite (represent)


Eq. (8.25) by means of a (kernel)-function hη(H) (q, p), namely


η(f ) = d3N qd3N p(f (p, q))hη(H) (p, q), (8.26)
R6N

where the function hη(H) (q, p) satisfies the relationship

3N  
∂hη ∂H ∂H ∂hη
{hη(H) , L} = − = 0, (8.27)
i=1
∂qi ∂pi ∂qi ∂pi

or equivalently hη(H) (q, p) is a “smooth” function of the Hamiltonian


function H(q, p), by imposing the additive Boltzmann behavior for
hη(H) (q, p) namely, hη(H1 +H2 ) = hη(H1 ) · hη(H2 ) , one obtains the famous
Boltzmann weight as the (unique) mathematical output associated to the
Ergodic theorem on classical statistical mechanics in the presence of a
thermal reservoir.4

exp{−βH(q, p)}
hη(H) (q, p) =  , (8.28)
d3N qd3N p exp{−βH(q, p)}

with β a (positive) constant which is identified with the inverse macroscopic


temperature of the combined system after evaluating the system internal
energy in the equilibrium state. Note that hη(H) L2 = 1 since η(f ) = 1.
A last remark should be made related to Eq. (8.28). In order to obtain
this result one should consider the nonzero value in ergodic limit

 T
1
lim dt(Ut ψ)(q, p)hη(H) (q, p) = hη(H) , Pker(−iL̄) (ψ), (8.29)
T →∞ T 0

or by a point-wise argument (for every t)

(U−t hη(H) ) ∈ Pker(−iL̄) , (8.30)

that is

hη(H) ∈ Pker(−iL̄) ⇔ Lh = 0. (8.31)


August 6, 2008 15:46 9in x 6in B-640 ch08

234 Lecture Notes in Applied Differential Equations of Mathematical Physics

8.4. On the Invariant Ergodic Functional Measure for Some


Nonlinear Wave Equations

Let us start this section by considering the discreticized (N -particle) wave


motion Hamiltonian associated to a nonlinear polynomial wave equation
related to the vibration of a one-dimensional string of length 2a
N   
2a p2i 1
H(pi , qi ) = + (qi+1 − qi ) + g(qi )
2 2k
, (8.32)
N 2 2
l=1

with the initial one boundary Dirichlet conditions

qi (0) = qi ,

pi (0) = pi ,

qi (−a) = qi (a) = 0. (8.33)

Here k denotes a positive integer associated to the nonlinearity power and


g > 0 the nonlinearity positive coupling constant.
Since one can argue that the above Hamiltonian dynamical system
possess unique global solutions on the time interval t ∈ [0, ∞) one can, thus,
straightforwardly write the associated invariant (Ergodic) measure on the
basis of Ergodic theorem exposed in Sec. 8.2 restricted to the configuration
space after integrating out the term involving the canonical momenta
 T 
1
lim dtF (q1 (t), . . . , qN (t)) = F (q1 , . . . , qN )d(inv) µ(q1 , . . . , qN )
T →∞ T 0 RN
(8.34)
here the explicit expression for the Hamiltonian system invariant measure
in configuration space is given by [with β = 1]
(inv)
di µ(q1 , . . . , qN )

N     
1 1 2a (qi+1 − q1 )2 g
= exp − + (qi )2k
Zn
(0) 2 i=1
N a2 2k

× (dq1 · · · dqN ), (8.35a)


Zn(0) = d(inv) µ(q1 , . . . , qN ). (8.35b)
RN
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 235

Now it is a nontrivial result on the theory of integration on infinite-


dimensional spaces that the cylindrical measures (normalized to unity !)
converges in a weak (star) topology sense to the well-defined nonlinear
N
polynomial Wiener measure at the continuum limit N → ∞, or l=1 () →
L
−L dx() and qi (t) → U (x, t) (see Appendix B for the relevant mathematical
arguments supporting the mathematical existence of such limit).
As a consequence of Eqs. (8.35) at the continuum limit, we have the
infinite-dimensional analogous of Ergodic result for the nonlinear poly-
nomial wave equation (with the Boltzmann constant equal to unity — see
Eq. (8.28) — Sec. 8.3 of this study)
 T 
1
lim F (U (x, t))dt = dWiener µ[U (σ)] × F (U (σ))
T →∞ T 0 C α (−a,a)
  a 
g
× exp − 2k
dx(U (σ)) , (8.36)
2k −a

where the wave field U (x, t) ∈ L∞ ([0, ∞), L2k−1 ((−a, a))) satisfies in a gen-
eralized weak sense the nonlinear wave equation below (see Appendix A):

∂ 2 U (x, t) ∂ 2 U (x, t)
2
= + g(U (x, t))2k−1 , (8.37)
∂ t ∂2x
U (x, 0) = U (0) (x) ∈ H 1 ((−a, a)), (8.38)

Ut (x, 0) = 0, (8.39)

U (−a, t) = U (a, t) = 0, (8.40)

dWiener µ[U (0) (x)] denotes the Wiener measure over the unidimensional
Brownian field end-points trajectories {U (0) (x)| with U (0) (−a) = U (0) (a) =
0}, and F (U (0) (x)) denotes any Wiener-integrable functional.
Note that C α (−a, a) denotes the Banach space of the α-Holder con-
tinuous function (for α ≤ 12 ) which is obviously the support of the Wiener
measure dWienerµ[U (0) (x)].
It is worth to remark the normalization factor Z of the nonlinear Wiener
measure implicitly used on Eq. (8.36).
   a 
1 g
dWiener µ[U (σ)] exp − dx(U (σ))2k = 1. (8.41)
Z C α (−a,a) 2k −a

Let us pass to the important case of existence of a dissipation term on


the wave equation problem Eqs. (8.37)–(8.40) with ν the positive viscosity
August 6, 2008 15:46 9in x 6in B-640 ch08

236 Lecture Notes in Applied Differential Equations of Mathematical Physics

parameter and leading straightforwardly to solutions on the whole time


propagation interval [0, ∞):

∂ 2 U (x, t) ∂ 2 U (x, t) ∂U (x, t)


2
= −ν + g(U (x, t))2k−1 , (8.42)
∂ t ∂2x ∂t
U (x, 0) = U (0) (x) ∈ H 1 ((−a, a)), (8.43)

Ut (x, 0) = U (1) (x) ∈ L2 ((−a, a)). (8.44)

It is a simple observation that there is a bijective correspondence


between the solution of Eq. (8.42) with the Klein–Gordon like wave equation
ν
for the rescaled wave field β(x, t) = e 2 t U (x, t), namely
 2
∂ 2 β(x, t) ∂ 2 β(x, t) ν (2k−1)νt
− = β(x, t) + ge− 2 (β(x, t))2k−1 , (8.45)
∂ 2t ∂2x 4

β(x, 0) = U (0) (x) = H 1 ((−a, a)), (8.46)


ν
βt (x, 0) = U (1) (x) − U (0) (x) = 0, (8.47)
2
β(−a, t) = β(a, t) = 0. (8.48)

As a result we have the analogous of the Ergodic theorem applied to


Eq. (8.45) and a sort of Caldirola–Kanai action is obtained,5 a new result of
ours in this nonlinear dissipative case (with the path-integral identification
x → σ; U (0) (x) = X(σ))
 T 
1 + ν2 t
lim dtF (e U (x, t)) = 1
dWienerµ[X(σ)]F (X(σ))
T →∞ T 0 C 2 (−a,a)
  a 
g − (2k−1)νσ
× exp − dσe 2 × (X(σ))2k
2k −a
 2   a 
ν
× exp dσ(X(σ)) . 2
(8.49)
4 −a

Concerning the higher-dimensional case, let us consider A a strongly


positive elliptic operator of order 2m associated to the free vibration of a
domain Ω in a general space Rν , and satisfying the Garding coerciviness
condition

A= (−1)|α| Dxα (aαβ (x))Dxβ , (8.50)


|α|≤m
|β|≤m
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 237

with the Sobolev spaces operator’s domain

D(A) = H 2m (Ω) ∩ H0m (Ω). (8.51)

Note there exists a constant C0 (Ω) > 0(C0 (Ω) → 0 if vol (Ω) → ∞) such
that the Garding coerciviness condition holds true on D(A)

Real(AU, U )L2 (Ω) ≥ C0 (Ω) f H 2m (Ω) . (8.52)

We can thus associate a Lipschitz nonlinear external vibration on the


domain Ω as governed by the wave equation below

∂ 2 U (x, t)
= (AU )(x, t) + G (U (x, t)), (8.53)
∂2t
U (x, 0) = U (0) (x) ∈ H 2m (Ω) ∩ H0m (Ω), (8.54)

Ut (x, 0) = g(x) ∈ L2 (Ω). (8.55)

Here G(z) denotes a differentiable Lipschitizian function, like G(z) =


A sin (Bz), e−γz , etc. and thus, leading to the uniqueness and existence of
global weak solutions on C([0, ∞), H 2m (Ω) ∩ H0m (Ω)) by the fixed point
theorem.6
Analogously to the discretization/N -particle technique of the 1 + 1 case,
one obtains as a mathematical candidate for the Ergodic invariant measure,
the following rigorously measure on the Hilbert space of functions H 2m (Ω)
(with β = 1):
R
1
dν x G(φ(x))
d(inv) µ(φ(x)) = dA µ(φ(x)) × e− 2 Ω . (8.56)

Here dA µ(φ(x)) denotes the Gaussian measure generated by the


quadratic form associated to the operator A on its domain D(A) for real
initial conditions functions φ(x):

dA µ(φ(x))(φ(x1 )φ(x2 )) = (A−1 )(x1 , x2 ). (8.57)
L2 (Ω)

Note that the Green function of the operator A belongs to the trace class
operators on H 2m (Ω), which results on Eq. (8.57) through the application
of the Minlo’s theorem.7
August 6, 2008 15:46 9in x 6in B-640 ch08

238 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point, we remark that dinv µ(φ(x)) defines a mathematical rig-


orous Radon measure in H 2m (Ω) since the function (functional):
  
1
F (φ(x)) = exp − dν xG(φ(x)) , (8.58)
2 Ω

belongs to L1 (L2 (Ω), dA µ(φ(x))) due to the Lipschitz property of the non-
linearity G(z), i.e. there exist constants c+ and c− , such that

c− φ(x) ≤ G(φ(x)) ≤ c+ φ(x). (8.59)

As a consequence, we have the obvious estimate below


+
R R R
dn uxφ(x) 1
dν xG(φ(x)) −
dν xφ(x)
e−c Ω ≤ e− 2 Ω ≤ e−c Ω , (8.60)

so by the Lebesgue comparison theorem, we get the functional integrability


of the nonlinearity term of the Ergodic invariant measure Eq. (8.56):
 
 R 
 − 12 Ω dν xG(φ(x))  + 1 (c− +c+ )2 TrH 2m (Ω) (A−1 )
 dA µ(φ(x))e ≤e 2 < ∞,
 L2 (Ω) 
(8.61a)
and leading thus to the mathematical validity of Eq. (8.56).
At this point, it is worth to remark that one can easily generalize the
type of our allowed Lipschitz nonlinearity for those of the following form
     
γ(x) 2
dν xG(φ(x)) ≥ + dν x (φ) (x) + dν xC(x)φ(x) ,
Ω Ω 2 Ω
(8.61b)

with γ(x) and C(x) real positive L∞ (Ω) functions.


This claim is a result of the straightforward Gaussian-infinite-
dimensional integration
    
 ν γ(x)

 d µ(φ) exp − d x (φφ̄)(x) + dν
C(x) φ̄ 
 2m A
2
(x) 
H (Ω) Ω Ω
   
 −1 γ(y)

≤ det 2 −1
1 + d yA (x − y) ·
ν
2
   
1 
× exp + dν x dν yC(x) · A−1 (x, y)C(y)  < ∞. (8.61c)
2 Ω Ω
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 239

In the usual (Euclidean) Feynman path-integral formal notation, the


Ergodic theorem takes the physicist’s form after reintroducing the reservoir
temperature parameter β = 1/kT
  R
1 T 1 1 ν
lim F (U (x, t)) = DF [ϕ(x)]e− 2 β Ω d x(ϕ(x)·(Aϕ(x)))
T →∞ T 0 Z L2 (Ω)
R
dν xG(ϕ(x))
× F (ϕ(x))e−β Ω , (8.62)

with the normalization factor


 R R
1 ν ν
Z= DF [ϕ(x)]e 2 β Ω (ϕ(x)(Aϕ)(x))d x e−β Ω d xG(ϕ(x)). (8.63)
H 2m (Ω)

The functionals F (ϕ(x)) are objects belonging to the space


L (L2 (Ω), dA (ϕ(x))) ⊂ C(L2 (Ω), R).
1

At this point, let us remark that by a direct application of the mean


value theorem there is a sequence of growing times {tn }n∈Z (with tn → ∞)
such that
 R
1 1 ν
lim F (φ(x, tn )) = DF [ϕ(x)]e− 2 β Ω d x(ϕ·Aϕ)(x)
n→∞ 2 L2 (Ω)
R
dν xG(ϕ(x))
× F (ϕ(x))e−β Ω . (8.64)

It is worth to recall that the (positive) parameter β is not deter-


mined directly from the parameters of the underlying mechanical system
and can be thought of as a remnant of the initial conditions “averaged
out” by the Ergodic integral (see Sec. 8.3) and sometimes related to the
presence of an intrinsic dissipation mechanism involved on the Ergodic-
time average process responsible for the extensivity of the system’s ther-
modynamics (see Eq. (8.28) — Sec. 8.3). For instance, at the limit of
vanishing temperature (absolute Newtonian–Boltzmann–Maxwell zero tem-
perature β → ∞ and somewhat related to the so called Ergodic mixing
microconomical ensemble), one can see the appearance of an “atractor” on
the space H 2m (Ω) ∩ H0m (Ω) and exactly given by the stationary strong
unique solution of the wave equation (8.53) since, we always have a soluble
elliptic problem on H 2m (Ω)∩H0m (Ω) by means of the Lax–Milgram–Gelfand
theorem.6

(Aϕ∗ )(x) = G (ϕ∗ (x)) (8.65)

ϕ∗ (x)|∂Ω = 0. (8.66)
August 6, 2008 15:46 9in x 6in B-640 ch08

240 Lecture Notes in Applied Differential Equations of Mathematical Physics

Note that the stationary strong solution ϕ∗ (x) is not a real atractor in the
sense of the theory of dynamical systems, since we can only say that there
is a specific sequence of growing {tn }n∈Z , with tn → ∞, such that at the
leading asymptotic limit β → ∞ we have the time-infinite limit for every
solution of Eqs. (8.65) and (8.66):

lim F (ϕ(x, tn )) = F (ϕ∗ (x)). (8.67)


tn →∞

A (somewhat) formal path-integral calculation of the next term on the


result (Eq. (8.41)) for F (z) being smooth real-valued functions can be
implemented by means of the usual path-integrals Saddle-point techniques
applied to the (normalized) functional integral equation (8.62).8

lim F (ϕ(x, tn ))
tn →∞
 2
∗ 1δ F ∗
= F (ϕ (x)) + (ϕ (x))
β δ2ϕ
   2   
d δ G ∗ 
× lg det 1 + A−1 (ϕ ) + α  + O(β −2 ) (8.68)
dα δ2 ϕ 
α=0
 2
1 δ F ∗
= F (ϕ∗ (x)) + (ϕ (x))
β δ2 ϕ
  2 −1 
−1 δ G ∗
× TrH 2m (Ω) 1+A (ϕ ) + O(β −2 )
δ2 ϕ

∗ 1 δ2 F ∗
= F (ϕ (x)) + (ϕ (x))
β δ2 ϕ
∞    2 n
δ G
× (−1)n TrH 2m (Ω) A−1 2 (ϕ∗ ) + O(β −2 ) (8.69)
n=0
δ ϕ

Another important remark to be made is related to the existence of time-


independent simple constraints on the discreticized Hamiltonian equation
(8.32), namely

H (a) (q1 , . . . , qN ) = 0; a = 1, . . . , n. (8.70)

In this case, one should consider that the continuum version of Eq. (8.70)
leads to (strongly) continuos functionals on H 2m (Ω) ∩ H0m (Ω):

H (a) (φ(x, t)) = 0; a = 1, . . . , n. (8.71)


August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 241

They have a direct consequence of restricting the wave motion to a


closed nonlinear manifold of the original vector Hilbert space H 2m (Ω) ∩
H0m (Ω).
Now it is straightforward to apply the Ergodic theorem on the effective
mechanical configuration space and obtain, thus, as the Ergodic invariant
measure the constraint path-integral.
 T
1
lim dt F (φ(x, t))
T →∞ T 0
 R R
1 1
dν x(ϕ(x)·Aϕ(x)) −β dν xG(ϕ(x))
= DF [ϕ(x)]e− 2 β Ω e Ω
Z H 2m (Ω)
 

n
× F (ϕ(x)) δ (F ) (H (a) (ϕ(x))) , (8.72)
a=1

where δ (F ) (·) denotes the usual formal delta-functionals.

8.5. An Ergodic Theorem in Banach Spaces


and Applications to Stochastic–Langevin
Dynamical Systems

In this complementary Sec. 8.5, we intend to present our approach to study


long-time/Ergodic behavior of infinite-dimensional dynamical systems by
analyzing the somewhat formal diffusion equation with polynomial terms
and driven by a white noise stirring.4
In order to implement such studies, let us present the author’s general-
ization of the RAGE theorem for a contraction self-adjoint semi-group T (t)
on a Banach space X. We have, thus, the following theorem.

Theorem 8.3. Let f ∈ X. We have the Ergodic generalized theorem


 T
1
lim dt(e−tA f ) = Pker(A) (f ), (8.73)
T →∞ T 0

where A is the infinitesimal generator of T (t), supposed to be self-adjoint.

Let us sketch the proof of the above claimed theorem of ours.


As a first step, one should consider Eq. (8.73) rewritten in terms
of the “resolvent operator of A” by means of a Laplace transform (the
August 6, 2008 15:46 9in x 6in B-640 ch08

242 Lecture Notes in Applied Differential Equations of Mathematical Physics

Hile–Yosida–Dunford spectral calculus)


   i∞ 
1 T 1 zt −1
lim dt dze ((z + A) f ) . (8.74)
T →∞ T 0 2πi −i∞
Now it is straightforward to apply the Fubbini theorem to exchange the
order of integrations (dt, dz) in Eq. (8.74) and get, thus, the result

1 T
lim dt(e−tA f ) = lim− +((z + A)−1 f ). (8.75)
T →∞ T 0 z→0

The z → 0− limit of the integral of Eq. (8.75) (since Real (z) ⊂ ρ(A) ⊂
(−∞, 0)) can be evaluated by means of saddle point techniques applied to
Laplace’s transforms. We have the following result
 ∞
lim ((z + A)−1 f ) = lim e−zt (e−tA f ) dt
z→0− z→0− 0
−tA
= lim (e f ) = Pker(A) (f ). (8.76)
t→∞

Let us apply the above theorem to the Lengevin equation. Let us con-
sider the Fokker–Planck equation associated to the following Langevin
equation:
dq i ∂V j
(t) = − (q (t)) + η i (t), (8.77)
dt ∂qi
where {η i (t)} denotes a white-noise stochastic time process representing the
thermal coupling of our mechanical system with a thermal reservoir at tem-
perature T . Its two-point function is given by the “fluctuation–dissipation”
theorem
η i (t)η j (t ) = kT δ(t − t ). (8.78)
The Fokker–Planck equation associated to Eq. (8.74) has the following
explicit form
∂P i i
(q , q̄ ; t) = +(kT )∆qi P (q i , q̄ i ; t) + ∇qi (∇qi V · P (q i , q̄ i ; t)) (8.79)
∂t
lim P (q i , q̄ i ; t) = δ (N ) (q i − q̄ i ). (8.80)
t→0+

By noting that we can associate a contractive semi-group to the initial


value problem Eq. (8.80) in the Banach space L1 (R3N ), namely:

P (q i , q̄ i , t)f (q̄ i )dq̄ i = (e−tAf ). (8.81)
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 243

Here, the closed positive accretive operator A is given explicitly by

−A(·) = kT ∆qi (·) + ∇qi [(∇qi V )(·)], (8.82)

and acts first on X = C∞ (R3N ). It is instructive to point out that the


perturbation accretive operator B = ∇qi · [(∇qi V ) · (·)] on C∞ (R3N ) with
V (q i ) ∈ Cc∞ (R3N ) is such that it satisfies the estimate on S(R3N ) :
Bf S(R3N ) ≤ a ∆f S(R3N ) + b f S(R3N ) with a > 1 and for some b. As a
consequence A(·) = kT ∆qi (·) + (B(·)) generates a contractive semi-group
on C∞ (R3N ) or by an extension argument on the whole L1 (R3N ) since the
L1 -closure of C∞ (R3N ) is the Banach space L1 (R3N ).
At this point, we may apply our Theorem 8.3 to obtain the Langevin–
Brownian Ergodic theorem applied to our Fokker–Planck equation
 T  
1 −tA
lim dt dq(e f )(q) = Pker(A) (f )(q)
T →∞ T 0 R3N

= dN q̄ i f (q̄ i )Peq (q̄ i ), (8.83)
RN

where the equilibrium probability distribution is given explicitly by the


unique normalizable element of the closed sub-space ker(A).

O = +(kT )∆qi Peq (q̄ i ) + ∇q̄i [(∇q̄i V )P(q̄ i )] (8.84)

or exactly, we have the Boltzmann’s weight for our equilibrium Langevin–


Brownian probability distribution
 
1
P (q̄ ) = exp −
eq i i
(V (q̄ )) . (8.85)
kT

For the general Langevin equation in the complete phase space {qi , pi }
as in the bulk of this note, one should reobtain the complete Boltzmann
statistical weight as the equilibrium ergodic probability distribution
 N

1 1 1 i 2
P (q̄ , p̄ ) = exp −
eq i i
(p̄ ) + V (q̄) , (8.86)
Z kT 2
l=1

with the normalization factor


 
Z= dp̄i dq̄ i Peq (q̄ i , p̄i ). (8.87)
RN RN
August 6, 2008 15:46 9in x 6in B-640 ch08

244 Lecture Notes in Applied Differential Equations of Mathematical Physics

Let us apply the above exposed result for the following nonlinear (poly-
nomial) stochastic diffusion equation on a one-dimensional domain Ω =
(−a, a) with all positive coefficients {λj } on the nonlinear polynomial term
 
1 d2 U (x, t)  
2k−1
∂U (x, t) 
=+ −  λj U j 
 (x, t) + η(x, t), (8.88)
∂t 2 d2 x j=0
j=odd

U (x, t)|∂Ω = 0, (8.89)

U (x, 0) = f (x) ∈ L2 (Ω). (8.90)

Here, {η(x, t)} are samplings of a white-noise stirring.


In this case, one can show by standard techniques that the global
solution U (x, t) ∈ C([0, ∞), L2 (−a, a)) and by using the discreticized
approach of Sec. 8.4, the invariant Ergodic measure associated to the non-
linear diffusion equation is given by the mathematically functional equi-
librium Langevin–Brownian probability distribution
  

  a 

  2k
λj j  
P (U (x)) = exp −
eq 
dx  
U (x) × d− 1 d2 µ(U (x)).

 −a j 
 2 dx2
 j=0 
jeven

(8.91)

Here, d− 1 d2 µ(U (x)) denotes the Gaussian functional measure (nor-


2 dx2
malized to unity) on the Wiener space C α (−a, a). In the Feynman path-
integral notation (with β = 1)
  R a dU 2
1 T
DF [U (x)]e− 2 −a | dx | (x)
1
lim dt F (U (x, t)) =
T →∞ T 0 C α ((−a,a))
#  $
Ra P2k λj
− −a
dx j=0 j U j (x)
×e j=even × F [U (x)].
(8.92)

Note that Eq. (8.91) defines a rigorous mathematical object since it


does not need to be renormalized from a calculational point of view on the
Wiener space C α (−a, a)10 as its nonlinear exponential term is bounded by
the unity (Lebesgue theorem). It is worth to point out that for nonlinearities
of Lipschitz type on general domains ΩCRν , one can see that the support
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 245

of the Gaussian measure is readily L2 (Ω) and the equilibrium Langevin–


Brownian probability distribution
  
P (U (x)) = exp − d xG(U (x)) × d(− 12 ∆) µ, (U (x)),
eq D
(8.93)

is a perfect well-defined random measure on C α ((−a, a)) by the same


Lebesgue convergence theorem.

8.6. The Existence and Uniqueness Results for Some


Nonlinear Wave Motions in 2D

In this technical section, we give an argument for the global existence and
uniqueness solution of the Hamiltonian motion equations associated first to
Eq. (8.32) — Sec. 8.3 and by secondly to Eq. (8.53) — Sec. 8.3. Related to
the two-dimensional case, let us equivalently show the weak existence and
uniqueness of the associated continuum nonlinear polynomial wave equation
in the domain (−a, a) × R+ .

∂ 2 U (x, t) ∂ 2 U (x, t)
− + g(U (x, t))2k−1 = 0, (8.94)
∂2t ∂ 2x
U (−a, t) = U (a, t) = 0,
(8.95)
U (x, 0) = U0 (x) ∈ H 1 ([−a, a]),

Ut (x, 0) = U1 (x) ∈ L2 ([−a, a]). (8.96)

Let us consider the Galerkin approximants functions to Eqs. (8.95) and


(8.96) as
n  

Ūn (t) ≡ U (t) sin x . (8.97)
a
=1

Since there exists a γ0 (a) positive such that


 
d2
− 2 Ūn , Ūn ≥ γ0 (a)(Ūn , Ūn )H 1 ([−a,a]) , (8.98)
d x L2 ([−a,a])

we have the a priori estimate for any t

0 ≤ ϕ(t) ≤ ϕ(0), (8.99)


August 6, 2008 15:46 9in x 6in B-640 ch08

246 Lecture Notes in Applied Differential Equations of Mathematical Physics

with
1 ˙ 2 2 1 2k
ϕ(t) = Ūn (t) L2 + γ0 (a) Ūn H1 + Ūn L2k . (8.100)
2 2k

As a consequence of the bound equation (8.100), we get the bounds for


any given T (with Ai constants)

sup ess Ūn 2


H 1 (−a,a) ≤ A1 , (8.101)
0≤t≤T

sup ess Ū˙ n 2


L2 (−a,a) ≤ A2 , (8.102)
0≤t≤T

sup ess Ūn 2k


L2k (−a,a) ≤ A3 . (8.103)
0≤t≤T

By usual functional-analytical theorems on weak-compactness on


Banach–Hilbert spaces, one obtains that there exists a sub-sequence Ūn (t)
such that for any finite T

−−−−−−−−−→
Ūn (t)(weak − star) Ū (t) in L∞ ([0, T ], H 1 (−a, a)), (8.104)

−−−−−−−−−→
Ū˙ n (t)(weak − star) v̄(t) in L∞ ([0, T ], L2(−a, a)), (8.105)

−−−−−−−−−→
Ūn (t)(weak − star) p̄(t) in L∞ ([0, T ], L2k (−a, a)). (8.106)

At this point, we observe that for any p > 1 (with Ãi constants) and
T < ∞, we have the relationship
 T  T
p
p p
Ūn H 1 (−a,a) dt ≤ T (A1 ) 2 ⇔ Ūn L2 (−a,a) ≤ Ã1 , (8.107)
0 0

 T  T
p
Ū˙ n p
L2 (−a,a) dt ≤ T (A2 ) 2 ⇔ Ū˙ p
L2k (−a,a)
dt ≤ Ã2 , (8.108)
0 0

since we have the continuous injection below

H 1 (−a, a) → L2 (−a, a), (8.109)

L2k (−a, a) → L2 (−a, a). (8.110)


August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 247

As a consequence of the Aubin–Lion theorem,6 one obtains straightfor-


wardly the strong convergence on LP ((0, T ), L2 (−a, a)) together with the
almost everywhere point-wise equality among the solutions candidate
 t
Ūn → Ū (t) = p̄(t) = v̄(s)ds. (8.111)
0

By the Holder inequality applied to the pair (q, k)

Un − Ū Lq (−a,a) ≤ Un − Ū 1−θ
L2 (−a,a) Un − Ū θ
L2k (−a,a) , (8.112)

with 0 ≤ θ ≤ 1
1 1−θ θ
= + , (8.113)
q 2 2k

in particular with q = 2k − 1, one obtains the strong convergence of Un (t)


in the general Banach space L∞ ((0, T ), L2k−1 (−a, a)), with θ̄ = 1−k
k
( 3−2k
2k−1 ).
As a consequence of the above results, one can safely pass the weak
limit on

d2
C ∞ ((0, T ), L2 (−a, a)) lim (Ūn , v)L2 (−a,a)
n→∞ d2 t
 
d2
+ − 2 Ūn , v + g(Ūn2k−1 , v)L2 (−a,a)
d x L2 (−a,a)
 
d2 d2
= 2 (Ū , v)L (−a,a) + − 2 Ū , v
2 + g(Ū 2k−1 , v)L2 (−a,a) = 0,
d t d x L2 (−a,a)
(8.114)

for any v ∈ C ∞ ((0, T ), L2 (−a, a)).


At this point, we shetch a rigorous argument to prove the problem’s
uniqueness.
Let us consider the hypothesis that the finite function

((Ū )2k+1 (x, t) − (v̄)2k+1 (x, t))


a(x, t) = , (8.115)
(Ū (x, t) − v̄(x, t))

is essentially bounded on the domain [0, ∞) × (−a, a) where Ū (x, t) and


v̄(x, t) denote two hypothesized different solutions for the 2D-polynomial
wave equations (8.94)–(8.96). It is straightforward to see that its difference
August 6, 2008 15:46 9in x 6in B-640 ch08

248 Lecture Notes in Applied Differential Equations of Mathematical Physics

W (x, t) = (Ū − v̄)(x, t) satisfies the “Linear” wave equation problem


∂2W ∂2W
− + (aW )(x, t) = 0, (8.116)
∂2t ∂x

W (0) = 0, (8.117)

Wt (0) = 0. (8.118)

At this point, we observe the estimate (where H 1 → L2 !)


 
1 d d 2 2
W L2 (−a,a) + W H 1 (−a,a)
2 dt dt
 
dW
≤ a L∞ ((0,T )(−a,a)) W L2(−a,a) ×
dt L2 (−a,a)
 
dW 2
≤M + W 2L2 (−a,a)
dt L2 (−a,a)
 
dW 2
≤M + W 2
H (−a,a) ,
1 (8.119)
dt L2 (−a,a)
which after the application of the Gronwall’s inequality give us that
 
d 2
W 2 + W 2H 1 (−a,a) (t)
dt L (−a,a)
 
dW 2
≤ (0) + W 2H 1 (−a,a) (0) = 0, (8.120)
dt L2 (−a,a)
which proves the problem’s uniqueness under the not proved yet hypothesis
that in the two-dimensional case (at least for compact support infinite dif-
ferentiable initial conditions)

sup a(x, t) ≤ M. (8.121)


x∈(−a,a)
t∈[0,∞)

As a last comment on the 1 + 1-polynomial nonlinear wave motion, we


call the readers’ attention that one can easily extend the technical results
analyzed here to the complete polynomial nonlinearity
j<2k
∂2U ∂2U
2
− 2 = cj U 2k−j = C1 U 2k−1 + C3 U 2k−3 + · · · , (8.122)
∂ t ∂ x j=1
j,odd
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 249

with the set of couplings {Cj } belonging to a positive set of real number.
In this case, we have that the general solution belongs a priori to the space
%
of functions j=1 L∞ ((0, T ), L2k−j (−a, a)).
j<2k

References

1. W. Humzther, Comm. Math. Phys. 8, 282–299 (1968).


2. V. Enss, Comm. Math. Phys. 61, 285–291 (1978).
3. Ya. G. Sinai, Topics in Ergodic Theory (Princeton University Press, 1994).
4. A. B. Soussan and R. Teman, J. Functional Anal. 13, 195 (1973); L. C. L.
Botelho, J. Math. Phys. 42, 1612 (2001); Il Nuovo Cimento, Vol. 117B(1),
15 (2002).
5. L. C. L. Botelho, Phys. Rev. E 58, 1141 (1998); R. Teman, Infinite-
Dimensional Systems in Mechanics and Physics, Vol. 68 (Springer Verlag,
1988).
6. O. A. Ladyzhenskaya, The Boundary Value Problems of Mathematical
Physics, Applied Math Sciences, Vol. 49 (Springer Verlag, 1985).
7. Y. Yamasaki, Measures on Infinite-Dimensional Spaces, Series in Pure Math-
ematics, Vol. 5, (World Scientific, 1985).
8. B. Simon, Functional Integration and Quantum Physics (Academic Press,
1979); J. Glimm and A. Jaffe, Quantum Physics — A Functional Integral
Point of View (Springer Verlag, 1981).
9. V. Rivasseau, From Perturbative to Constructive Renormalization (Princeton
University, 1991).
10. C. L. Siegel and J. K. Moser, Lectures on Celestial Mechanics (Springer-
Verlag, Berlin, Heidelberg, New York, 1971).
11. S. Sternberg, Lectures on Differential Geometry (Chelsea Publishing
Company, New York, 1983).
12. V. Guillemin and A. Pollack, Differential Topology (Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1974).

Appendix A. On Sequences of Random Measures


on Functional Spaces

Let us consider a completely regular topological space X and the Banach


space of continuous function space
&
C(X) = f : X → R, f continuous on the X-topology with f ∞
'
= sup |f (x)| < ∞ . (A.1)
x∈X
August 6, 2008 15:46 9in x 6in B-640 ch08

250 Lecture Notes in Applied Differential Equations of Mathematical Physics

Associated to the Banach space C(X), there exists its dual M(X) which
is the set of random measures on X with norm given by its total variation
µ = |µ|(X), where µ ∈ M(X).
It is easy to see that if β(X) is the Stone-chech compactification of X,
we have the isometric immersion M(X) → M(β(X)).
We have, thus, the important theorem of Prokhorov on the accumulation
points of sets of random measures.7
Theorem (Prokhorov). If a family of measures {µα } on M(X) is such
that it satisfies the following conditions
(a) Bounded uniformly

µα  = |µα |(X) ≤ M. (A.2)

(b) Uniformly regular — For any given ε > 0, there exists a compact set
K(ε) ⊂ X such that
 
 
 f (x)dµα (x) ≤ ε f ∞ ,
 
X

for any f (x) ∈ C(X) satisfying the condition

(supp f (x)) ∩ K(ε) = {φ}. (A.3)

We have that the set of random measures is a relatively compact set on


the weak-star topology on M(X). Namely, for any f (x)C(X), we have that
there exists a measure ν ∈ M(X) such that for a given sequence {αk }
 
lim sup f (x)dµαK (x) = f (x)dν(x). (A.4)
{αK } X X

In order to apply such theorem to the set of random measures as given


by Eq. (8.35), we introduce the completely regular space formed by the
one-point compactification of the real line
 
 
X= (Ṙ)i , or X = β(R) , (A.5)
i∈Z i∈Z

and note the strict positivity of the integrand (the path-integral weight)
on Eq. (8.35) and its boundedness by the Gaussian free term since
g( 2a
N )
N (N
exp{− 21 (2k) l=1 (qi ) } < 1 for {qi } ∈
2k
l=1 (Ṙ)i .
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 251

We note either the chain property of the invariant measure set ensuring
us the measure uniqueness limit on Eq. (B.4) when applied to Eq. (8.35) —
Sec. 8.4, namely:

|f (q1 , . . . , qN −1 )|d(inv) µ(q1 , . . . , qN )
RN

≥ |f (q1 , . . . , qN −1 )|d(inv) µ(q1 , . . . , qN −1 ). (A.6)
RN −1

Finally, we remark that the well-defined nonlinear Wiener functional


measure coincides with those finite-dimensional as given by Eqs. (8.35a)
and (8.35b) when it restricts to the set of continuous polygonal paths and,
thus, ensuring the convergence of the measure set Eqs. (8.35a) and (8.35b)
to the full Wiener path measure Eq. (8.36), since its functional support is
contained on the set of continuous functions on the interval [−a, a].

Appendix B. On the Existence of Periodic Orbits in a Class


of Mechanical Hamiltonian Systems —
An Elementary Mathematical Analysis

B.1. Elementary May Be Deep — T. Kato

One of the most fascinating study in classical astronomical mechanics is that


one related to the inquiry on how energetic-positions configurations in a
given mechanical dynamical system can lead to periodic systems trajectories
for any given period T , a question of basic importance on the problem of
the existence of planetary systems around stars in dynamical astronomy.10
In this short note, we intend to give a mathematical characterization
of a class of N -particle Hamiltonian systems possessing intrinsically the
above-mentioned property of periodic trajectories for any given period T ,
a result which is obtained by an elementary and illustrative application of
the famous Brower fixed point theorem.
As a consequence of this mathematical result, we conjecture that the
solution of the existence of periodic orbits in classical mechanics of N -body
systems moving through gravitation should be searched with the present
status of our mathematical knowlegdment.
Let us, thus, start our note by considering a Hamiltonian function
H(xi , pi ) ≡ H(z i ) ∈ C ∞ (R6N ) of a system of N -particles set on motion in
R3 . Let us suppose that the Hessian matrix determinant associated to the
August 6, 2008 15:46 9in x 6in B-640 ch08

252 Lecture Notes in Applied Differential Equations of Mathematical Physics

given Hamiltonian is a positive-definite matrix in the (6N − 1)-dimensional


energy-constant phase space z i = (xi , pi ) of our given mechanical system
∂2H
> 0. (B.1a)
∂z k ∂z s H(z s )=E

In the simple case of N -particles with unity mass and a positive two-
body interaction with a potential V (|xi − xj |2 ), such that the set in Rn ,
{xi | V (xi ) < E} is a bounded set for every E > 0. Our condition Eq. (B.1a)
takes a form involving only the configurations variables ({xp }p=1,...,N ; xp ∈
R3 ). Namely,

||Mks ({xp })|| > 0, (B.1b)

with
N
Mks ({xp }) = [V  (|xi − xj |2 )4(xi − xj )2 (δis − δjs )(δik − δjk )
i<j

+ 2V  (|xi − xj |2 )(δik − δjk )(δis − δjs ))]. (B.1c)

The class of mechanical systems given by Eqs. (B.1b) and (B.1c) is


such that the energy-constant hypersurface H(z s ) = E has always a pos-
itive Gaussian curvature and — as a straightforward consequence of the
Hadamard theorem — such energy-constant hypersurface is a convex set of
R6N , besides obviously being a compact set. As a result of the above-made
remarks, we can see that our mechanical phase space is a convex-compact
set of R6N .
The equations of motion of the particles of our given Hamiltonian system
is given by
dxi (t) ∂H i
= (x (t), pi (t)),
dt ∂pi
(B.2)
dpi (t) ∂H
= − i (xi (t), pi (t)),
dt ∂x
with initial conditions belonging to our phase space of constant energy
HE = {(xi , pi ); H(xi , pi ) = E}, namely:

(xi (t0 ), pi (t0 )) ∈ HE . (B.3)

The existence of periodic solutions to the system of ordinary differential


equations Eqs. (B.2) and (B.3); for any given period T ∈ R+ is a con-
sequence of the fact that the Poincaré recurrent application associated to
August 6, 2008 15:46 9in x 6in B-640 ch08

On the Ergodic Theorem 253

the (global) solutions Eq. (B.2) of the given mechanical system. Note that
H(xi (t), pi (t)) = E

P : HE → HE ,
(B.4)
(xi (t0 ), pi (t0 )) → (xi (T ), pi (T )),

has always a fixed point as a consequence of the application of the Brower


fixed point theorem.12 This means that there is always initials conditions in
the phase space Eq. (B.3) that produces a periodic trajectory of any period
T in our considered class of mechanical N -particle systems.
As a physical consequence, we can see that the existence of planetary
systems around stars may not be a rare physical astronomical event, but
just a consequence of the mathematical description of the nature laws and
the fundamental theorems of differential topology and geometry in RN .
August 6, 2008 15:46 9in x 6in B-640 ch09

Chapter 9

Some Comments on Sampling of Ergodic Process:


An Ergodic Theorem and Turbulent
Pressure Fluctuations∗

9.1. Introduction

A basic problem in the applications of stochastic processes is the estimation


of a signal x(t) in the presence of an additive interference f (t) (noise). The
available information (data) is the sum S(t) = x(t) + f (t), and the problem
is to establish the presence of x(t) or to estimate its form. The solution of
this problem depends on the state of our prior knowledge concerning the
noise statistics. One of the main results on the subject is the idea of max-
imize the output signal-to-noise ratio,1 the matched filter system. One of the
most important results on the subject is that if one knows a priori the signal
in the frequency domain X(w) and, most importantly, the frequency domain
expression for the noise correlation statistics function Sf f (w), one has, at
least in the theoretical grounds, the exact expression for the optimum filter
transference function Hopt (w) = k(X ∗ (w))(Sf f (w))−1 ·e−iwt̄ , here t̄ denotes
certain time on the observation interval process [−A, A], supposed to be
finite here.
As a consequence, it is important to have estimators and analytical
expressions for the correlation function for the noise, specially in the
physical situation of finite-time duration noise observation. Another very
important point to be remarked is that in most of the cases of observed
noise (as in the turbulence research2); one should consider the noise (at
least in the context of a first approximation) as an Ergodic random
process.
In Sec. 9.2 we aim in this note to present (of a more mathematical ori-
ented nature) a rigorous functional analytic proof of an Ergodic theorem
stating the equality of time-averages and Ensemble-averages for wide-
sense mean continuous stationary random processes. In Sec. 9.3 somewhat
∗ Author’s original results.

254
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 255

of electrical engineering oriented nature, we present a new approach for


sampling analysis of noise, which leads to canonical and invariant ana-
lytical expressions for the Ergodic noise correlation function Sf f (w) already
taking into account the finite time-observation parameter which explicitly
appears in the structure formulae, a new result on the subject, since it
does not require the existence of Nyquist critical frequency on the sampling
rate, besides removing, in principle, the aliasing problem in the computer-
numerical sampling evaluations. In Sec. 9.4, we present a mathematical-
theoretical application of the above theoretical results for a model of
turbulent pressure fluctuations.

9.2. A Rigorous Mathematical Proof of the Ergodic


Theorem for Wide-Sense Stationary Stochastic Process

Let us start our section by considering a wide-sense mean continuous sta-


tionary real-valued process {X(t), −∞ < t < ∞} in a probability space
{Ω, dµ(λ), λ ∈ Ω}. Here, Ω is the event space and dµ(λ) is the underlying
probability measure.
It is well-known1 that one can always represent the above-mentioned
wide-sense stationary process by means of a unitary group on the Hilbert
space {L2 (Ω), dµ(λ)}. Namely, in the quadratic-mean sense in engineering
jargon

 +∞
X(t) = U (t)X(0) = eiwt d(E(w)X(0)) = eiHt (X(0)), (9.1)
−∞

here we have used the famous spectral Stone-theorem to rewrite the


associated time-translation unitary group in terms of the spectral
process dE(w)X(0), where H denotes the infinitesimal unitary group
operator U (t). We also supposed that the σ-algebra generated by the
X(t)-process is the whole measure space Ω, and X(t) is a separable
process.
Let us, thus, consider the following linear continuous functional on the
Hilbert (complete) space {L2 (Ω), dµ(λ)} — the space of the square inte-
grable random variables on Ω

 T
1
L(Y (λ)) = lim dtE{Y (λ)X(t, λ)}. (9.2)
T →∞ 2T −T
August 6, 2008 15:46 9in x 6in B-640 ch09

256 Lecture Notes in Applied Differential Equations of Mathematical Physics

By a straightforward application of the RAGE theorem,3 namely:


   T 
1 −iwt
L(Y (λ)) = dµ(λ)Y (λ) lim dte dE(w)X0 (λ)
Ω T →∞ 2T −T

 T
1
= lim E{Y e−iHt X̄}dt
T →∞ 2T −T

= dµ(λ)Y (λ)dE(0)X(0, λ)

= E{Y (λ)PKer(H) (X(0, λ))}. (9.3)

where PKer(H) is the (orthogonal projection) on the kernel of the unitary-


group infinitesimal generator H (see Eq. (9.1)).
By a straightforward application of the Riesz-representation theorem for
linear functionals on Hilbert spaces, one can see that PKer(H) (X(0))dµ(λ)
is the searched time-independent Ergodic-invariant measure associated
to the Ergodic theorem statement, i.e. for any square integrable time-
independent random variable Y (λ) ∈ L2 (Ω, dµ(λ)), we have the Ergodic
result (X(0, λ) = X(0)
 T
1
lim dtE{X(t)Ȳ } = E{PKer(H) (X(0))Ȳ }. (9.4)
T →∞ 2T −T

In general grounds, for any real bounded borelian function it is expected


that the result (not proved here)
 T
1
lim dtE{f (X(t))Ȳ } = E{PKer(H) f (X(0)) · Ȳ }. (9.5)
T →∞ 2T −T

For the auto-correlation process function, we still have the result for the
translated time ζ fixed (the lag time) as a direct consequence of Eq. (9.1)
or the process’ stationarity property
 T
1
lim dtE{X(t)X(t + ζ)} = E{X(0)X(ζ)}. (9.6)
T →∞ 2T −T

It is important to remark that we still have the probability average


inside the Ergodic time-averages (Eqs. (9.4)–(9.6)). Let us call the readers’
attention that in order to have the usual Ergodic like theorem result without
the probability average E on the left-hand side of the formulae, we proceed
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 257

(as it is usually done in probability text-books1 ) by analyzing the prob-


ability convergence of the single sample stochastic-variables below (for
instance)
 T
1
ηT = f (X(t))dt, (9.7)
2T −T
 T
1
RT (ζ) = dtX(t)X(t + ζ). (9.8)
2T −T
It is straightforward to show that if E{f (X(t))f (X(t+ζ))} is a bounded
function of the time-lag, or, if the variance written below goes to zero at
T →∞
 T  T
1
σT2 = lim dt 1 dt2 [E{X(t1 )X(t1 + ζ)X(t2 )X(t2 + ζ))
T →∞ 4T 2 −T −T

− E{X(t1 )X(t1 + ζ)}E{X(t1 )X(t1 + ζ)))] = 0, (9.9)

one has that the random variables as given by Eqs. (9.7) and (9.8) converge
at T → ∞ to the left-hand side of Eqs. (9.4)–(9.6) and producing thus
an Ergodic theorem on the equality of ensemble-probability average of the
wide sense stationary process {X(t), −∞ < t < ∞} and any of its single-
sample {X̄(t), −∞ < t < +∞} time average
 T
1
lim dtf (X̄(t)) = E{PKer(H) (f (X0 ))} = E{f (X(t))},
T →∞ 2T −T

(9.10)
 T
1
lim dtX̄(t)X̄(t + ζ) = E{X(0)X(ζ)}
T →∞ 2T −T

= E{X(t)X(t + ζ)}
= RXX (ζ). (9.11)

The above formulae will be analyzed in the next electrical engeneering


oriented section.

9.3. A Sampling Theorem for Ergodic Process

Let us start this section by considering F (w) as an entire complex-variable


function such that

|F (w)| ≤ ceA|w|, for w = x + iy ∈ C = R + iR, (9.12)


August 6, 2008 15:46 9in x 6in B-640 ch09

258 Lecture Notes in Applied Differential Equations of Mathematical Physics

and its restriction to real domain w = x + i0 satisfies the square integrable


condition
 ∞
|F (x)|2 dx < ∞. (9.13)
−∞

By the famous Wienner theorem (Ref. 4) there exists a function f (t) ∈


L2 (−A, A), vanishing outside the interval [−A, A], such that
 A
1
F (w) = √ dtf (t)eitw . (9.14)
2π −A
The time-limited function f (t) can be considered as a physically finite-
time observed (time-series) of an Ergodic process on the sample space Ω =
L2 (R) as considered in Sec. 9.1 (since RXX (0) < +∞). In a number of cases,
one should use interpolation formulae in order to analyze several statistics
aspects of such observed random process. One of the most useful result
in this direction, however, based on the very deep theorem of Carleson on
the pontual convergence of Fourier series of a square integrable function1,5
is the well-known Shanon (in the frequency-domain) sampling theorem for
time-limited signals, namely1,6
+∞
  πn  sin T w − πn
F (w) = F  T
. (9.15)
n=−∞
T T w − πn
T

Note in Eq. (9.15), however, the somewhat restrictive sampling con-


dition hypothesis of the Nyquist interval condition T ≥ A. Although quite
useful, there are situations where its direct use may be cumbersome due to
the Nyquist condition on the signal sampling besides the explicitly necessity
of infinite time observations { πn
T }n∈Z .
At this point, let us propose the following more invariant sampling–
interpolation result of ours. Let f (t) be an observed time-finite random
Ergodic signal (sample) as mathematically considered in Eqs. (9.12)–(9.14).
It is well-known that we have the famous uniform convergent Fourier
expansion for t = A sin θ, with − π2 ≤ θ ≤ + π2 for the Fourier kernel in
terms of Bessel functions
+∞

e+iwt = e+iw(A sin θ) = Jn (wA)einθ . (9.16)
n=−∞

As much as in the usual proof of the Shanon results Eq. (9.15), we introduce
Eq. (9.16) into Eq. (9.14) and by using the Lebesgue convergence theorem,
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 259

since f (t) ∈ L1 (−A, A) either, we get the somewhat (canonical) interpo-


lating formula without any Nyquist-like restrictive sampling frequency con-
dition on the periodogran F (w):
+∞
 
 π/2
inθ
F (w) = AJn (wA) dθe f (A sin θ) cos θ . (9.17)
n=−∞ −π/2

It is worth to call the readers’ attention that the sampling coefficients


as given by Eq. (9.17) are exactly the values of the Fourier transform of the
signal g(θ) = f (A sin θ) cos θ for − π2 < θ < θ2 , i.e.
 π/2
1
gn = dθ einθ f (A sin θ) cos θ
2π −π/2
 A
1 t
= dt einarc sin( A ) f (t)
2π −A


N
1 ti
 exp inarc sin · f (ti ) , (9.18)
i=1
2π A

which may be easily and straightforwardly evaluated by FFT algorithms,


from arbitrary — finite on their number — chosen sampling values
{f (A sin θi )}i=1,...,N with θi = arsin( tAi ) (here ti are the time obser-
vation process signal). Monte–Carlo integration techniques7 can be used on
approximated evaluations of Eq. (9.18) too. Note that f (w) is completely
determined by its samples Eq. (9.18) taken at arbitrary times ti ∈ [−A, A].
In the general case of a quadratic mean continuous wide-sense stationary
process {Xt , 0 < t < ∞},1 we still have the quadratic mean result analogous
to the above exposed result for those process with time-finite sampling1
 +∞
Xt = eiwt dXw , (9.19)
−∞

with the spectral process possessing the canonical form (in the quadratic
mean sense)
+∞
 
 π/2
X(w) = Jn (wA) einθ d(Xt=A sin θ ) . (9.20)
n=−∞ −π/2

Let us expose the usefulness of the sampling result Eqs. (9.17) and (9.18)
for the very important practical engineering problem of estimate the self-
correlation function of a time-finite observed sampling of an Ergodic process
August 6, 2008 15:46 9in x 6in B-640 ch09

260 Lecture Notes in Applied Differential Equations of Mathematical Physics

already taking into consideration the time-limitation of the sampling obser-


vation in the estimate formulae. We have, thus, to evaluate the (formal)
time-average with time-lag ζ > 0
 T
1
Rf f (ζ) = lim dtf (t)f (t + ζ). (9.21)
T →∞ 2T −T

Since we have observed the sampling-continuous function {f (t)} only for


the time-limited observation interval ([−A, A]), it appears quite convenient
at this point to rewrite Eq. (9.21) in the frequency domain (as a somewhat
generalized process) (see Ref. 1) where T → ∞ limit is already evaluated
 +∞   T 
1 iw1 ζ 1 it(w1 +w2 )
Rf f (ζ) = dw1 dw2 e lim dte F (w1 )F (w2 )
2π −∞ T →∞ 2T −T

 +∞
1
=√ dw|F (w)|2 eiwζ . (9.22)
2π −∞

It is worth to point out that Eq. (9.22) has already “built” T in


its structure, the infinite time-Ergodic evaluation {limT →∞ T1 −T (...)},
regardless of the time-limited finite duration nature of the observed sample
f (t)[θ(t + A) − θ(t − A)].
Now an expression for Eq. (9.22), after substituting Eq. (9.17) on
Eq. (9.22) (without the use of somewhat artificial aliased A-periodic
extensions of the observed sampling f (t), as it is commonly used in the
literature),6 can straightforwardly be suggested possessing the important
property of already taking into account (in an explicit way) the presence of
the observation time A in the result for the self-correlation function in the
frequency domain
+∞
 +∞

1
Sf f (w) = √ {[A2 (gn ḡm )] × Jn (wA)Jm (wA)}, (9.23)
2π n=−∞ m=−∞

with
 +∞
1
Rf f (ζ) = √ dw eiwζ Sf f (w). (9.24)
2π −∞

The above equations are the main results of this section.


It is worth to call the readers’ attention that after passing the random
signal f (t) by a causal linear system with transference function Hf f (w),6
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 261

one obtains the output self-correlation function statistics in the standard


formulae

Syy (w) = |Hyf (w)|2 × Sf f . (9.25)

Finally, let us comment on the evaluation of Eq. (9.23) on the time-


domain. This step can be implemented throughout the use of the well-
known formula for Bessel functions (see Ref. 8, Eq. (6.626)), analytically
continued in the relevant parameters formulae
 ∞
dxe−ax Jµ (bx)Jν (cx)dx
0
b µ cν
= Iµ,ν (a) = 2−µ−ν (a)−1−µ−ν
Γ(ν + 1)
 

 Γ(1 + µ + ν + 2)  c b2

× F −, −µ − , ν + 1, 2 − 2 .
!Γ(µ +  + 1) b 4a
=0

(9.26)

We have, thus, the result:


 +∞  
1 t t
√ dwe Jn (Aw) · Jm (wA) = (−1)n+m In,m −
iwt
+ In,m ,
2π −∞ A A
(9.27)

where
1
In,m (t) = 2−m−n (−it)−1−m−n
Γ(m + 1)
∞ 
 Γ(1 + n + m + ) 1

× F (−, −n − , m + 1, 1) − 2 .
!Γ(n +  + 1) 4t
=0

(9.28)

Another point to be called to attention and related to the integral eval-


uations of Fourier transform of Bessel functions are the recurrence set of
Fourier integrals of Bessel functions
 +∞  ∞ 
dt eiwt Jn (w) = (1 + (−1)n ) dt cos(wt)Jn (w)
−∞ 0
 ∞ 
+ i(1 + (−1)n+1 ) dt sin(wt)Jn (w) , (9.29)
0
August 6, 2008 15:46 9in x 6in B-640 ch09

262 Lecture Notes in Applied Differential Equations of Mathematical Physics


with, here, the explicitly expressions Tn (t) = 2−n [(t +
√ t2 − 1)n + (t −
t2 − 1)n ].
 ∞
cos(wt)Jn (w)dw
0
 1

(−1)k √ Tn (t), for 0 < t < 1 and n = 2k,

 1 − t2




0, for 0 < 1 < t,
=  ∞ 4k (9.30)

 cos(wt) Jn−1 (w)dw,

 w

 0 ∞


− cos(wt)Jn−2 (w)dw, for n = 2k + 1,
0

 ∞
sin(wt)Jn (w)dw
0
 1

 (−1)k √ Tn (t), for 0 < t < 1 and n = 2k + 1,

 − t2

 1


0, for 0 < 1 < t,
=  ∞
 sin(wt)

 2(2k + 1) Jn−1 (w)

  ∞ 0 w



− sin(wt)Jn−2 (w)dw, for n = 2k + 2.
0
(9.31)

9.4. A Model for the Turbulent Pressure Fluctuations


(Random Vibrations Transmission)

One of the most important studies of pressure turbulent fluctuations


(random vibrations) is to estimate the turbulent pressure component trans-
mission inside fluids.9,10 In this section, we intend to propose a simple
analysis of a linear model of such random pressure vibrations.
Let us consider an infinite beam backed on the lower side by a space of
depth d which is filled with a fluid of density ρ2 and sound speed v2 . On the
upper side of the beam there exists a supersonic boundary-layer turbulent
pressure P (x, t). The fluid on the upper side of the beam which is on the
turbulence steadily regime is supposed to have a free stream velocity U∞ ,
density ρ1 , and sound speed v1 .
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 263

The effective equation governing the “outside” pressure in our model is


thus given by (d ≤ z < ∞):
2
∂ ∂ ∂2 ∂2
+ U∞ p1 (x, z, t) − v12 + p1 (x, z, t) = 0 (9.32)
dt ∂x ∂ 2 x ∂ 2z

with the boundary condition (Newton’s second law)


2 
∂ ∂ ∂p1 (x, z, t) 
−ρ1 + U∞ W (x, t) =  . (9.33)
∂t ∂x ∂z z=d

Here, the beam’s deflection W (x, t) is assumed to be given by the beam


small linear deflection equation

∂ 4 W (x, t) ∂ 2 W (x, t)
B + m = P (x, t) + (p1 − p2 )(x, d, t), (9.34)
∂4x ∂2t
with B denoting the bending rigidity, m the mass per unit length of the
beam, and p(x, z, t) the (supersonic) boundary-layer pressure fluctuation of
the exterior medium.
The searched induced pressure p2 (x, z, t) in the fluid interior medium
(0 ≤ z ≤ d) is governed by

∂ 2 p2 ∂2 ∂2
(x, z, t) − (v2 )2 + p2 (x, z, t) = 0, (9.35)
∂2t ∂2x ∂2z
∂p2 (x, d, t) ∂2
= −ρ2 2 W (x, t). (9.36)
∂z ∂ t
The solution of Eqs. (9.32) and (9.33) is straightforward obtained in the
Fourier domain
 
v1 (w + U∞ k)2
p̃1 (k, z, w) = −ρ1  2 w̃(k, w)
v1 k 2 − (w + U∞ k)2

1
× exp − × v12 k 2 − (w + U∞ k)2 (z − d), (9.37)
v1

where the deflection beam W̃ (k, w) is explicitly given by

W̃ (k, w) = (P̃ (k, w) − p̃2 (k, w, d)


 −1
4 2 ρ1 (w + U∞ k)2 v1
× Bk − mw +  2 . (9.38)
v1 k 2 − (w + U∞ · k)2
August 6, 2008 15:46 9in x 6in B-640 ch09

264 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point, we solve our problem of determining the pressure


p̃2 (k, w, z) in the interior domain Eqs. (9.35) and (9.36) if one knows the
pressure p̃2 (k, w, d) on the boundary z = d. Let us, thus, consider the
Taylor’s series in the z-variable (0 ≤ z ≤ d) around z = d, namely

∂ p̃2 (k, w, z) 
p̃2 (k, w, z) = p̃2 (k, w, d) +
∂z  (z − d)
z=d

1 ∂ k p̃2 (k, w, z)
+ ···+ (z − d)k + · · · . (9.39)
k! ∂kz

From the boundary condition Eq. (9.36), we have the explicitly


expression for the second-derivative on the depth z

∂ p̃2 (k, w, z) 
 = +ρ2 w2 W̃ (k, w) = +ρ2 w2 [P̃ (k, w) − p̃2 (k, w, d)]
∂z z=d
 −1
4 2 ρ1 (w + U∞ k)2 v1
× Bk − mw +  . (9.40)
v1 k 2 − (w + U∞ k)2

The second z-derivative of the interior pressure (and the higher ones!)
is easily obtained recursively from the wave equation (9.35) (k ≥ 0, k ∈ Z+ )
and Eq. (9.40)

∂ 2+k 
2+k
p̃2 (k, z, w)
∂ z z=d

1 ∂k 
= × (−w2 + (v2 )2 k 2 ) p̃ (k, z, w)  . (9.41)
(v2 )2 k
∂ z
2 
z=d

Let us finally make the connection of this random transmission vibration


model with Sec. 9.3 by calling the readers’ attention that the general
turbulent pressure is always assumed to be expressible in an integral
form
 +∞  +∞
P (x, t) = dw dk eikx ei(w−ku)t G(k) · F (w), (9.42)
−∞ −∞

where F (w) is the Fourier transform of a time-limited finite duration sample


function of an Ergodic process1 simulating the stochastic-turbulent nature
of the pressure field acting on the fluid with the exactly interpolating for-
mulae Eqs. (9.17) and (9.18).
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 265

References

1. E. Wong and B. Hajek, Stochastic Processes in Engineering System (Springer-


Verlag, 1985); E. Wong, Introduction to Random Processes (Springer-Verlag,
1983); H. Cramér and M. R. Leadbetter, Stationary and Related Stochastic
Processes (Wiley, USA).
2. G. L. Eyink, Turbulence Noise J. Stat. Phys. 83, 955 (1996); V. Yevjevich,
Stochastic Processes in Hydrology (Water Resources Publications, Fort
Collins, Colorado, USA, 1972); H. Tennekes and J. L. Lunley, A First Course
in Turbulence (The MIT Press, 1985); L. C. L. Botelho, Int. J. Mod. Phys.
B 13(11), 363–370 (1999); S. Fauve et al., Pressure fluctuations in swirling
turbulent flows, J. Phys. II-3, 271–278 (1993).
3. L. C. L. Botelho, Mod. Phys. Lett. B 17(13–14), 733–741 (2003).
4. W. Ruldin, Real and Complex Analysis (Tata McGraw-Hill Publishing,
New Delhi, 1974).
5. Y. Katznelson, An Introduction to Harmonic Analysis (Dover Publishing,
1976).
6. A. Papoulis, A note on the predictability of band limited processes, Pro-
ceedings of the IEEE, Vol. 73 (1985); E. Marry, Poisson sampling and spectral
estimation of continuous-time processes, IEEE (Trans. Information Theory),
Vol. II-24 (1978); J. Morlet, Sampling theory and wave propagation, Proc.
51st. Annu. Meet. Soc. Explan. Geophys. Los Angeles (1981).
7. M. Kalos and P. Whitlock, Monte Carlo Methods, Vol. 1 (Wiley, 1986);
W. Press, S. Teukolsky, W. Vetterling and B. Flannery, Numerical Recipes
in C, 2nd ed. (Cambridge Univ. Press., 1996).
8. I. S. Gadshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products
(Academic Press, Inc., 1980).
9. D. J. Mead, Free wave propagation in periodically supported infinite beams
J. Sound Vib, 11, 181 (1970); R. Vaicaitis and Y. K. Lin, Response of finite
periodic beam to turbulence boundary-layer pressure excitation, AIAAJ 10,
1020 (1972).
10. L. Maestrello, J. H. Monteith, J. C. Manning and D. L. Smith, Measured
response of a complex structure to supersonic turbulent boundary layers,
AIAA 14th Aerospace Science Meeting, Washington, DC (1976).
11. W. Rudin, Real and Complex Analysis (Tata Mc Graw-Hill, 1974).
12. W. Rudin, Functional Analysis (Tata Mc Graw-Hill, 1974).
13. L. Gillman and M. Lewison, Rings of Continuous Functions (D. Van Nostrand
Company, Inc., 1960).
14. S. M. Flatte, Sound Transmission Through a Fluctuation Ocean (Cambridge
University Press, Cambridge, 1979); N. S. Van Kampen, Stochastic Process
in Physics and Chemistry (North-Holland, Amsterdam, 1981); R. Dashen,
J. Math. Phys. 20, 892 (1979).
15. P. Sheng (ed), Scattering and Localization of Classical Waves in Random
Media, World Scientific Series on Directions in Condensed Matter Physics,
Vol. 8, (World Scientific, Singapore, 1990); L. C. L. Botelho et al., Mod. Phys.
August 6, 2008 15:46 9in x 6in B-640 ch09

266 Lecture Notes in Applied Differential Equations of Mathematical Physics

Lett. B 12, 301 (1998); L. C. L. Botelho et al., Mod. Phys. Lett. B 13, 317
(1999), J. M. F. Bassalo, Nuovo Cimento B 110 (1994).
16. L. C. L. Botelho, Phys. Rev. E 49R, 1003 (1994).
17. L. C. L. Botelho, Mod. Phys. Lett. B 13, 363 (1999).
18. B. Simon, Functional Integration and Quantum Physics (Academic,
New York, 1979).
19. B. Simon and M. Reed, Methods of Modern Mathematical Physics, Vol. 2
(Academic, New York, 1979, ).
20. B. Simon and M. Reed, Methods of Modern Mathematical Physics, Vol. 3
(Academic, New York, 1979).
21. T. Kato, Analysis Workshop (Univ. of California at Berkeley, USA, 1976).
22. Any standard classical book on the subject as for instance
E. C. Titchmarsh, The Theory of Functions, 2nd ed. (Oxford University
Press, 1968); H. M. Edwards, Riemann’s Zeta Function, Pure and Applied
Mathematics (Academic Press, New York, 1974); A. Ivic, The Riemann Zeta-
Function (Wiley, New York, 1986); A. A. Karatsuba and S. M. Doramin, The
Riemann Zeta-Function (Gruyter Expositions in Math, Berlin, 1992).

Appendix A. Chapters 1 and 9 — On the Uniform


Convergence of Orthogonal Series —
Some Comments

A.1. Fourier Series

It is well-known that point-wise convergence of the partial sums of the


Fourier series of a function in a given class (continuous, square integrable,
etc.) constitutes a cumbersome and difficult mathematical problem.
A useful elementary result in this direction is the well-known Tauberian
theorem of G. Hardy.

Hardy Theorem. Let f ∈ L1 ([0, 2π]) and its complex Fourier coefficients
satisfying an estimate as

1
|bn | ≤ O . (A.1)
n

Then, the usual partial sums Sn (f )(x) and σn (f )(x) (Fejer partial sum
of order n) converge to the sum limit.

We have now the following sorts of variants of the above enunciated


Hardy theorem.
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 267

Theorem A.1. Let f (x) ∈ L1 ([0, 2π]) and thus usual Fourier coefficients
obeying the estimates
ãn
|an | ≤ , (A.2)
n
b̃n
|bn | ≤ , (A.3)
n
where

n
(ãk )2 ≤ Ma (n)2−ε , (A.4)
k=0


n
(b̃k )2 ≤ Mb (n)2−ε , (A.5)
k=0

with (Ma , Mb ) denoting (positive) real n-independent constants and ε ∈


(0, 2].
Then, there is a subsequence Snk (f )(x) which converges almost every-
where to the given function f (x).
Proof. It is straightforward that the following estimate holds true
1 n
k2 2
||Sn (f )(x) − σn (f )(x)||2L2 = 2
(an + b2n ) ≤ (MA + MB )n−ε . (A.6)
π n=1
n

We have as well as


||Sn (f )(x) − f (x)||L1 ≤ (MA + MB )n−ε . (A.7)
π
L1
As a consequence Sn (f ) −→ f (x), so it exists a basic result that there
exists a subsequence Snh (f )(x) converges almost everywhere to f (x) in
[0, 2π]. 

Theorem A.2. Let f (x) ∈ L2 ([0, 2π]) and suppose that the estimate below
holds true
 


||f − Sn (f )(x)||2L2 =  |b |2  ≤ M · nε−2 , (A.8)
≥n

for M and ε denoting positive real numbers (0 ≤ ε < 2). Then, there
exists a subsequence Snk (f )(x) which converges uniformly to f (x) almost
everywhere.
August 6, 2008 15:46 9in x 6in B-640 ch09

268 Lecture Notes in Applied Differential Equations of Mathematical Physics

Proof. Let
  12



an = ||f − Sn (f )||L1 ≤ 2π  |b |2  (A.9)
≥n

and

gn (x) = Sn (f )(x) + an .

Let us apply the Markov–Tchebbicheff inequalite with m being the


Lebesgue measure

m{x ∈ [0, 2π] : |(Sn (f ) + an ) − f |(x) ≥ n−ε }


 π 1
2√
1 2
× ε dx|Sn (f ) + an − f | (x) 2π
n −π
√   π 1
2π √ 2 2
2
≤ ε 2 dx|Sn (f ) − f | + 2π|an |
n −π
 
C̄  


≤ ε |b |2 ≤ 2 , (A.10)
n   n
≥n

where C̄ is an over-all positive constant.


We have thus that Eq. (A.10) means that the Markov–Luzin condition
is satisfied, since
∞
1 C̄ π2
lim = 0 and = C̄ < ∞. (A.11)
n→∞ nε
n=1
n2 6

 L1
By the Luzin theorem and the obvious result that an → 0 Sn (f ) −→
f , we have that there is a subsequence Snk (f )(x) almost everywhere con-
verging to f (x) uniformly. 

Finally, we have the basic theorem of uniform convergence of the Fourier


series, very useful on linear partial differential equations.

Theorem A.3. Let f (x) be an absolutely continuous function such that


f  (x) ∈ L2 ([0, 2π]) (f (x) ∈ H1 ([0, 2π]), with exception of first-order dis-
continuities. Then, the Fourier series converges uniformly to f (x) in any
interval not containing the discontinuity points.
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 269

Proof. Without loss of generality, let us consider a unique interior point


x0 as a point of discontinuity of our given function f (x). Namely

lim (f (x0 + h) − f (x0 − h)) = [f (x0 )]− < ∞. (A.12)


h→0
h>0

We now have the estimates:


   
1  π 
 1  sin nx0  |b |
−
|an | =  f (x) cos(nx)dx ≤  [f (x0 )]  + n ,
π −π π n n
 π    (A.13)
1  1  cos nx0  |a |
|bn | =  f (x) sin(nx) ≤  [f (x0 )]−  + n .
π −π π n n
Here an and bn are the associated Fourier coefficients of the derivative of
the function f (x).
Now it a straightforward estimate to see that the Weierstrass uniform
convergence criterion holds true in our case, since
∞  
 cos(nx0 )  1 
= −n 2 sin x0  , (A.14)
n=1
n 2
∞
sin(nx0 ) x0
= − π2· (A.15)
n=1
n 2

As a consequence, we have the estimate


 
n 
 
 ak cos(kx) + bk sin(kx)
 
h=1

n 
n n
|bk |  |ak |
n
= |Sn (f )(x)| ≤ |ak | + |bk | ≤ +
k k
h=1 h=1 h=1 k=1
 
 f (x )]− n
sin kx0 + cos kx0 
 0
+ , (A.16)
 π k 
h=1
!∞
where n=1 Mn < ∞ (see Eqs. (A.12) and (A.13)).
On basis of the above result one can see that there exists a unique
solution of the “brick” linear string wave equation in C 2 ([0, [×(0, ∞)) for
initial conditions satisfying the above constraints:
(A) U (x, 0) = f (x) with f (3) (x) ∈ L2 ([0, ]) and f (x) and f (1) (x) abso-
lutely continuous functions and f (3) possessing first-order discontinu-
ities (piecewise continuous) (f () = f (−) = f (1) () = f (1) (−) =
f (2) () = f (2) (−) = f (3) () = f (3) (−) = 0).
August 6, 2008 15:46 9in x 6in B-640 ch09

270 Lecture Notes in Applied Differential Equations of Mathematical Physics

(B) Ut (x, 0) = g(x) with g (2) (x) ∈ L2 ([0, ]) and g(x) being an absolutely
continuous with g (2) (x) piecewise continuous in [0, ] (g() = g(−) =
g (1) () = g (1) (−) = g (2) (−) = g (2) () = 0).


Finally, let us point out the following L. Schwartz distributional Fourier


series theorem.

Theorem A.4. Let f (x) ∈ L1loc (R) and f (x) periodic of period T . Let its
formal Fourier complex coefficients cn satisfy the polynomial bound
  
 1 T /2 
 2πin 
|cn | =  f (x)e− T x dx ≤ M (n)L ,
 T −T /2 

for a fixed positive integer L and M ∈ R+ . Then, the associated partial



sums Sn (f )(x) converges in the distributional sense in Dperiodic (R).

On the light of this distributional sense, one can see that it is always
possible
  to write a solution of the string linear wave equation in the space
C Dperiodic (R), (0, ∞) . (Exercise for our diligent readers.) For any given

datum f (x) and g(x) in Dperiodic (R).

A.2. Regular Sturm–Liouville Problem

In the case of an orthonormal set associated to a regular self-adjoint Sturm–


Liouville problem in the open interval (a, b)

d dyn
− p(x) + g(x)yn (x) = λn yn (x),
dx dx
(A.17)
a11 yn (a) + a12 yn (a) = 0,
a21 yn (b) + a22 yn (b) = 0,

we have the following result on the uniform convergence of the eigen-


functions yn (x).

Theorem A.5. Let f (x) ∈ Dom(L), where L  is the second-order dif-


ferential self-adjoint operator associated to the Sturm–Liouville problem
(Eq. (A.17)). Then, its series orthogonal expansion in terms of the eigen-
functions yn (x) is a uniformly convergent series.
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 271

Proof. Since f ∈ Dom(f ), we have that


 b ∞

f, L
 f L2 = dxf (x)(Lf )(x) = λn |cn |2 < ∞. (A.18)
a n=0

Here
 b
cn = dxf (x)yn (x).
a

Now straightforward to see that Cauchy partial sum


" N #2
 
N  yn (x)
cn yn (x) = ( λn cn ) √
n=M n=M
λn
" N #" N #
  yn (x)ȳn (x)
2
≤ λn |cn |
λn
n=M n=M
" # " N #
 1
2
≤ sup |yn (x)| f, L  f L2 . (A.19)
x∈(a,b) λ
n=m n

By the usual Sturm comparison theorems, we have the well-known


asymptotic limits for large M , N
 N 
 1  
N
(b − a)2
 
 ≤ , (A.20)
 λn  n2
n=M n=M

which leads to our result on the uniform convergence as a consequence of


the Weierstrass M -text. In the case of an infinite interval [a, b], we need
that the Green function L−1 should be a compact-trace class operator
with all eigenfunctions satisfying a n-uniform supremum bound as in
Eq. (A.17). 
Our reader can apply our result to the following Sturm–Liouville
problem in (−1, 1)
 
d dyn m
(1 − x2 ) − yn (x) = −λn yn (x). (A.21)
dx dx 1 − x2
In this case, we have that the problem Green function has a kernel
bounded
∞ 
 y (x)y (x̄) 
 n n  1
 ≤ · (A.22)
 λn  2m
n=0
August 6, 2008 15:46 9in x 6in B-640 ch09

272 Lecture Notes in Applied Differential Equations of Mathematical Physics

As a consequence of Theorem A.5, we have the uniform convergence of


the orthogonal series expansion in terms of the eigenfunctions yn (x) if
 1  2
 df  m2 f 2
dx   + < ∞.
−1 dx 1 − x2

Appendix B. On the Müntz–Szasz Theorem


on Commutative Banach Algebras

B.1. Introduction

The classical theory of Müntz–Szasz states that the set of functions


! 1
{1, tα1 , tλ2 , . . . , tλn , . . .} span C([0, 1], R) if λn = +∞. In this short note,
we intend to point out that the usual elementary proof of such a theorem
works out with minor changes to produce an abstract result in certain com-
mutative self-adjoint Banach algebras.
Let us start our note by considering A be a semi-simple (self-adjoint)
commutative Banach algebra possessing a simple generator f such that its
Gelfand transform f ∈ C(∆, R) is a (continuous) function with the property
that the polynomial algebra generated by f satisfies those conditions of the
Stone–Weirstrass theorem on C(∆, R).
We now state our generalization of the Müntz–Szasz theorem for the
abstract case.

Theorem B.1. Let {λn } be a sequence of positive real numbers such that
! 1
λn = +∞. Then, the span of the set {1, f , f , . . . , f
λ1 λ2 λn
, . . .} is the
whole C(∆, R).

Proof. Through the Hahn–Banach theorem and the Riesz representation


theorem applied to the topological dual of C(∆), our result will be a
straightforward consequence of the following proposition (Ref. 11, Theorem
15.26):
! 1
If λn = +∞, and u is a complex Borel measure on the compact
maximal ideal space ∆ such that

(f )λn , dµ = 0, (n = 1, 2, 3, . . .), (B.1)

then also

(f )n dµ = 0, (n = 1, 2, 3, . . .), (B.2)

August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 273

a contradiction with our hypothesis about the denseness of polynomial


algebra generated by f on C(∆, R).
In order to show this result, let us consider without loss of generality
our function f such that ||f ||C(∆) ≤ 1. Now it is straightforward to see that
the function

h(z) = f z dµ, (B.3)

is a bounded holomorphic function in the right half-plane H + = {z = x +


iy, x > 0} by an application of Morera theorem. Note that g(z) = f ( 1+z
1−z ) ∈
H α (U ). Here U = {z| | |z| < 1} is the unit disc and g(αn ) = 0, with
αn = λλnn −1
+1 · Note that either g(z) ≡ 0 if Σ(1 − |αn |) = ∞ or equivalently:
! 1
λn = +∞. This means that h(z) ≡ 0. 
As an interesting consequence of Theorem B.1, we have the following
generalization of the famous Wiener theorem on inversion of absolute
summable trigono/metric polynomials on the torus T = {(eiθ ), 0 ≤ θ <
2π}, useful in linear system theory.

Theorem B.2. Suppose that


+∞
 +∞

(eiθ ) = cn einθ ; |cn | < ∞ (B.4)
n=−∞ −∞

and (eiθ ) = 0 for eiθ ∈ T . Then, for any Müntz–Szasz sequence |λn |, we
have the result
+∞ +∞

1

= γn eiλn θ with |γn | < ∞. (B.5)
(e ) −∞ n=−∞

Proof. The element f (θ) = eiθ satisfies the conditions of Theorem B.1.
In the case of the Müntz–Szasz sequence {λn } be the prime numbers
sequence or that one made up by their logarithms, the result may have
interesting applications in the theory of Zeta functions.13
In the case of the existence of a polynomial subalgebra generated by
a finite number of generators {f1 , . . . , fp } we can use the Hahn–Banach
theorem for complex homomorphism of our Banach algebra through the
Gleason, Kahane, Zelazuo theorem (Ref. 12, Theorem 10.9).
Finally in the Banach algebra of the bounded continuous functions
vanishing at the infinity in a locally compact space X, it is straight-
forward to see that if f (x) is an arbitrary continuous function satisfying
August 6, 2008 15:46 9in x 6in B-640 ch09

274 Lecture Notes in Applied Differential Equations of Mathematical Physics

the condition f (x) ≥ 1 and such that the polynomial algebra generated
by the bounded function h = f exp(−f ) is dense on C0 (X), then the set
{1, hλ1 , hλ2 , . . . , hλn , . . . } remains dense in C0 (X) the only difference with
the standard proof presented in Theorem B.1 is related to the estimate on
the boundedness of the function h(z) on H + , which now takes the following
form:
 
sup |h(z)| ≤ sup |f e−f |z (m)dµ(m)
z∈H + z∈H +
 
≤ µ(X) × sup |f |x · e−x|f |(m)
0<x<∞
f (m)∈[1,∞)
 
x(gf −f )
≤ µ(X) × sup |e |(m)
0<x<∞
f (m)∈[1,∞)
 
≤ µ(X) × sup |e−x | = µ(X) < ∞.
0<x<∞ 

Appendix C. Feynman Path-Integral Representations


for the Classical Harmonic Oscillator
with Stochastic Frequency

C.1. Introduction

The problem of the (random) motion of a harmonic oscillator in the


presence of a stochastic time-dependent perturbation on its frequency is
of great theoretical and practical importance.14,15 In this appendix, we
propose a formal path-integral solution for the above-mentioned problem by
closely following our previous studies.16,17 In Sec. C.2, we write a Feynman
path-integral representation for the external forcing problem. In Sec. C.3,
we consider a similar problem for the initial-condition case.

C.2. The Green Function for External Forcing

Let us start our analysis by considering the classical motion equation of a


harmonic oscillator subject to an external forcing
 2 
d 2
+ w 0 (1 + g(t)) x(t) = F (t). (C.1)
dt2
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 275

Here, w02 (1 + g(t)) is the time-dependent frequency with stochastic part


given by the random function g(t) obeying the Gaussian statistics
 t
x(t, [g]) = G(t, t , [g])F (t )dt , (C.3)
0

where G(t, t , [g]) denotes the Green problem functionally depending on the
random frequency g(t) and the notation emphasizes that the objects under
study are functionals of the random part g(t) of the harmonic oscillator
frequency.
In order to write a path-integral representation for the Green function
equation (C.3), we follow our previous study16 by using a “proper-time”
technique by introducing related Schrödinger wave equations with an initial
point source and −∞ ≤ t ≤ +∞:
 2 
∂ Ḡ(x; (t, t )) d 2
i =− + w0 (1 + g(t)) Ḡ(s; (t, t )), (C.4)
∂s dt2
lim Ḡ(s; (t, t )) = δ(t − t ), (C.5)
s→0

lim Ḡ(s; (t, t )) = 0. (C.6)


s→x

At this point, we note the following identity between the Schrödinger


equations (C.4)–(C.6) and the searched harmonic oscillator Green function:
 ∞
G[(t, t , [g])] = −i dsḠ(s; (t, t )). (C.7)
0

Let us, thus, write a path-integral for the associated Schrödinger equa-
tions (C.4)–(C.7) by considering Ḡ(s; (t, t )) in the operator form (the
Feynman–Dirac propagator):

Ḡ(s; (t, t )) = t| exp(isH)|t , (C.8)

where H is the differential operator


 2 
d 2
H=− + w0 (1 + g(t)) . (C.9)
dt2
As in quantum mechanics, we write Eq. (C.8) as an infinite product of
short-time s propagations (see Chap. 3)
$ N  +∞
&
% s 

t| exp(isH)|t = lim dti ti | exp i H ti−1 . (C.10)
N →x N
i=1 −∞
August 6, 2008 15:46 9in x 6in B-640 ch09

276 Lecture Notes in Applied Differential Equations of Mathematical Physics

The standard short-time expansion in the s-parameter for Eq. (C.10) is


given by

lim ti |eisH |ti−1 d = lim dwi exp{is[(wi )2 + (w0 )2 (1 + g 2 (ti−1 ))]}
s→0+ s→0+

× exp[iwi (ti − ti−1 )]. (C.11)

If we substitute Eq. (C.11) into Eq. (C.10) and take the Feynman limit
of N → ∞, we will obtain the following path-integral representation after
evaluating the wi -Gaussian integrals of the representation equation (C.8):
 
    ) *
%  i 2
2
   dt(σ)
Ḡ(s; t, t )) =  dt(σ) exp dσ
 0<σs
 2 0 dσ
t(0)=t :t(s)=t
  s 
2
× exp i [w0 (1 + g(t(σ)))]dσ . (C.12)
0

The averaged out Eq. (C.7) is thus given straightforwardly by the fol-
lowing Feynman polaron-like path-integral:
 ∞ 
2
G(t, t , [g]) g = −i ds(ei(w0 ) s ) DF [t(σ)]
0 t(0)=t :t(s)=t
  ) *
2
i s dt
× exp dσ
2 0 dσ
  s  s 
4  
× exp −w0 dσ dσ K(t(σ); t(σ )) . (C.13)
0 0

The two-point correlation function is still given by a two-full similar


path-integral, namely

G(t1 , t1 , [g])G(t2 , t2 , [g]) g


 ∞
2
= ds1 ds2 ei(w0 ) (s1 +s2 )
0

× DF [t1 (σ), t2 (σ)]
t1 (0)=t1 :t2 (s1 )=t1 :t2 (0)=t2 :t2 (s2 )=t2
  s1  s2 
i
× exp dσ(t1 (σ))2 + dσ(i2 (σ))2
2 0 0
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 277

  s1  s1
× exp −w04 dσ dσ  K(t1 (σ) · t1 (σ  ))
0 0
 s1  s2
+ dσ dσ  (K(t1 (σ) · t2 (σ  )) + K(t2 (σ) · t1 (σ  )))
0 0
 s1  s2 
+ dσ dσ  K(t2 (σ) · t2 (σ  )) . (C.14)
0 0
Similar N -iterated path-integral expressions hold true for the N -point
correlation function x(t1 , [g]) · · · x(tN , [g]) g . Explicit and approximate
evaluations of the path-integral equations (C.13) and (C.14) follow pro-
cedures similar to those used in the usual contexts of physics statistics,
quantum mechanics, and random wave propagation (last reference of
Ref. 14).
Let us show such exact integral representation for Eq. (C.13) in the case
of the practical case of a slowly varying (even function) kernel of the form:
1/2
0 2 K(0)
K(t) ∼ K(0) − |t| |t| + = L,
2 0 (C.15)
K(t) ∼ 0 |t|  L.
In this case, we have the following exact result for the path-integral in
Eq. (C.13):
G(t, t , [g]) g
 ∞  ) *
−s(−iw02 ) −(u40 K(0))s2 1/2 w02 (− 2 )1/2 S 3/2
= dse e (2πis)
0 sin[s3/2 w02 (− 20 )1/2 ]
") " #* #
1/2 1/2
iw02 1/2 0 0
× exp s − cot w02 − s3/2 (t − t )2 .
2 2 2
(C.16)
Another useful formula is that related to the “mean-field” averaged
path-integral when the kernel K(t, t ) has a Fourier transform of the general
form:
 +∞
1 
K(t, t ) = √ dp · eip(t−t ) K̃(p). (C.17)
2π −∞
The envisaged integral representation for Eq. (C.13) is, thus, given by
 ∞   +∞ 
 isu20 w04 
G(t, t , [g]) g = −i dse exp − √ K̃(p) × M (p, s, t, t ) ,
0 2π −∞
(C.18)
August 6, 2008 15:46 9in x 6in B-640 ch09

278 Lecture Notes in Applied Differential Equations of Mathematical Physics

where
   
 s  s  2 
  i s
F dt
M (p, s, t, k ) = dσ dσ D [t(σ)] exp
0 0  t(0)=t 2 0 dσ
t(s)=t


× exp[ip(t(σ) − t(σ  ))] . (C.19)

C.3. The Homogeneous Problem

Let us start this section by considering the problem of determining two


linearly independent solutions of the homogeneous harmonic oscillator
problem
 2 
d 2
+ w0 (1 + g(t)) x(t) = 0, (C.20)
dt2

with the initial conditions

x(0) = x0 x (0) = v0 . (C.21)

It is straightforward to see that two linearly independent solutions are


given by the following expressions:
 t 
2
x1 (t, [g]) = exp y (σ, [g])dσ , (C.22)
0
 t
x2 (t, [g]) = x(t, [g]) (x1 (σ, [g]))−2 dσ, (C.23)
0

where y(t, [g]) satisfies the first-order nonlinear ordinary differential


equation
dy
(t) + (y(t)2 = −w02 (1 + g(t)). (C.24)
dt
In order to obtain a path-integral representation for Eq. (C.24), we
remark that the whole averaging (stochastic) information is contained in
the characteristic functional
+   ∞ ,
Z[j(t)] = exp i dt y(t, [g])j(t) . (C.25)
0 g
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 279

In order to write a path-integral representation for the characteristic


functional equation (C.25) we rewrite (C.25) as a Gaussian functional
integral in g(t):
  ∞
1
Z[j(t)] = DF [g(t)] exp − dt dt g(t)K −1 (t, t )g(t )
2 0
  ∞ 
× exp i dty(t, [g]) . (C.26)
0

At this point, we observe the validity of the following functional integral


representation for the characteristic functional equation (C.26) after con-
sidering the functional change g(t) → y(t) defined by Eq. (C.24), namely

Z[j(t)] = DF [y(t)]
 ∞  
1 dy
× exp − dt dt
+ y 2 (t) + w02 K(t, t )
2(w0 )4 0 dt
    ∞ 
dy
× + y2  2
(t ) + w0 exp i dt j(t)y(t) , (C.27)
dt 0

where we have used the fact that the Jacobian associated with the functional
change g(t) → y(t) is unity:
 
d δg(t)
detF + 2y = = 1. (C.28)
dt δy(t)

At this point, it is instructive to remark that in the important case of


a white-noise frequency process with strength γ

K(t, t ) = γδ(t − t), (C.29)

the path-integral representation for the characteristic functional equation


(C.27) takes the more amenable form

    2 

γ dy
Z[j(t)] = D [y(t)] exp −
F
dt + y 2 (t) + wo2
2(w0 )4 0 dt
  ∞ 
× exp i dty(t)j(t) , (C.30)
0
August 6, 2008 15:46 9in x 6in B-640 ch09

280 Lecture Notes in Applied Differential Equations of Mathematical Physics

we obtain, thus, the standard λϕ4 zero-dimensional path-integral as a


functional integral representation for the characteristic functional equation
(C.25) in the white-noise case
   ∞   
γ dȳ 2 2 4
Z[j(t)] = DF [ȳ(t)] exp − dt + 2w 0 ȳ + ȳ (t)
2(w0 )4 0 dt
  ∞ 
× exp i dt ȳ(t)j(t) . (C.31)
0

In this paragraph, we discuss some mathematical points related to the


final condition on the quantum propagator equations (C.4)–(C.6) (its van-
ishing in the limit s → ∞). Let us first remark that by imposing
∞ the trace
class condition on the correlation equation (C.2) k(t, t )( 0 k(t, t)dt < ∞)
one has the result that all realizations (sampling) g(t) of the associated
stochastic process are square integrable functions by a direct application of
the Minlos theorem on the domain of functional integrals.18 At this point,
d2
we note that if one restricts g(t) further to be a dt 2 -perturbation, namely
2
with g(t) ∈ L (0, ∞):
- 2 -
-d -
||w0 (1 + g(t))h||L2 ≤ a -
2 -
- dt2 h- 2 + b||h||L2 (C.32)
L

and 0 < a < 1 and b arbitrary, one can apply the Kato–Rellich theorem19
d2 2 2
to be sure that the domain of the differential operator − dt 2 + w0 + w0 g(t) is
2
(at least) contained on the domain of − dt d
2 which in turn has a purely con-
2
tinuous spectrum on L (R). As a consequence of the above-exposed remarks
one does not have bound states on the Schrödinger operator spectrum of
Eq. (C.4) for each realization of g(t) on the above-cited functional class. As
a consequence, we have that the evolution operator
 
d2
exp is − 2 + w02 (1 + g(t)) , (C.33)
dt
2
d 
is a unitary operator on Dom(− dt 2 ) and Ḡ(s, t, t ) in turn is expected to
d2
have the same behavior of exp(is[− dt2 ]) (asymptotic completeness20 ) at
s → ∞, which in turn vanishes at s → ∞, making our condition Eq. (C.6)
highly reasonable from a mathematical physicist’s point of view.
At this point, we remark that the different problem of determining the
quantum propagator of a particle under the influence of a harmonic potential
V (x) = 12 w02 x2 and a stochastic potential g(x) may be possible to handle
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 281

2 2
in our framework. Note that the “únperturbed” Hamiltonian − 2mdx h d
2 +
1 2 2
2 mw 0 x has a pure point spectrum (bound states) and its perturbation by
2
a g(x) ∈ L (R) potential does not alter the spectrum behavior. However,
one can proceed in a mathematical physicist’s (formal) way as exposed in
our letter and see that Eq. (C.3) adapted to this case is still (formally)
correct. Namely
 
   t ) *
 %  i dx(σ)
2

G(t, (x, x )) g =   
dx(σ) exp m dσ
0<σ<t
2 0 dσ
x(0)=x :x(t)=x
  t 
i 2 2
× exp mw0 dσ[x(σ)]
2 0
  t  t 
4 
× exp −w0 dσ dσ K(x(σ); x(σ)) . (C.34)
0 0

Finally, we want to point out that in the general case where k(t, t ) is not
defined on an operator of the trace class, the realizations g(t) will be dis-
tributional objects. Unfortunately in this case there is no rigorous spectral
perturbation mathematics for differential operators acting on nuclear spaces
(s (R), D (R)), etc, which is the natural mathematical setting to under-
stand Eq. (C.1) of this letter.
In this complementary paragraph, we complete our study by considering
the problem of a harmonic oscillator in the presence of a damping term
(0 < t < ∞)
 2 
d d 2
+ v + w0 (1 + g(t)) x(t) = F (t). (C.35)
dt2 dt
In order to map the above differential equation in the analysis pre-
sented in Sec. C.2, we implement in Eq. (C.35) the following time-variable
change:16
m v
ζ = (1 − e−t m u ),
v (C.36)
y(ζ) = x(t).
We obtain, thus, the following pure harmonic oscillator differential
equation without the damping term in place of the original equation (C.35),
namely
 2 
d 2
+ w0 [1 + g̃(ζ)] y(ζ) = F̃ (ζ). (C.37)
dζ 2
August 6, 2008 15:46 9in x 6in B-640 ch09

282 Lecture Notes in Applied Differential Equations of Mathematical Physics

Here
 m   v  
g̃(ζ) = g − lg 1 − ζ , (C.38)
v m
 m  v
− v lg 1 − m (ζ)
F̃ (ζ) = F  v 2 , (C.39)
1− m ζ
and the correlator stochastic frequency
"  ) *#
  m  (1 − ( m
v
)ζ) 
K(ζ, ζ ) = K((ζ, ζ )) = K lg  v  . (C.40)
v  1− m ζ 
From here on the analysis goes as in the bulk of this appendix under
the condition that the range associated with this new time variable is the
finite interval [0, m
v ].

Appendix D. An Elementary Comment on the Zeros of the


Zeta Function (on the Riemann’s Conjecture)

D.1. Introduction — “Elementary May Be Deep”


!∞ −x
The Riemann’s series ζ(x) = n=1 n converges uniformly for all real
numbers x greater than or equal to a given (fixed) abscissa x̄: x > x̄ > 1.
It is well-known that the complex-valued (meromorphic) continuation to
complex values (z = x+iy) throughout the complex plane z ∈ C is obtained
from standard analytic (finite-part) complex variables methods applied to
the integral representation.21

iΓ(1 − z) (−w)z−1
ζ(z) = dw
2π C ew − 1
 ∞  
1 1 1
= wz−1 × w − dw
Γ(z) 0 2 −1 w
 
iΓ(1 − z) 1
= (−w)z−2 − (−w)z−1
2π C 2
 (−1)n+z−2 Bn /wz+2n−2 

+ , (D.1)
n=1
(2n)!

here Bn are the Bernoulli’s numbers and C is any contour in the complex
plane, coming from positive infinity and encircling the origin once in the
positive direction.
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 283

An important relationship resulting from Eq. (D.1) is the so-called


functional equation satisfied by the zeta function, holds true for any
z∈C

ζ(z)  πz 
= 2z · π z−1 · sin · Γ(1 − z). (D.2)
ζ(1 − z) 2

In applications to number theory, where this special function plays a


special role, it is a famous conjecture proposed by B. Riemann (1856) that
the only nontrivial zeros of the zeta function lie in the so-called critical line
Real (z) = x = 12 .
In the next section, we intend to propose an equivalent conjecture,
hoped to be more suitable for handling the Riemann’s problem by the
standard methods of classical complex analysis,22 besides of proving
a historical clue for the reason that led B. Riemann to propose his
conjecture.22

D.2. On the Equivalent Conjecture D.1

Let us state our conjecture

Conjecture D.1. In each horizontal line of the complex plane of the form
Imaginary (z) = y = b = constant, the zeta function ζ(z) possess at most a
unique zero.

We show now that the above conjecture leads straightforwardly to a


proof of the Riemann’s conjecture.

Theorem D.1 (The Riemann’s Conjecture). All the nontrivial zeros


of the Riemann zeta function lie on the critical line Real (z) = x = 12 .

Proof. Let us consider a given nontrivial zero z̄ = x̄ + iȳ on the open strip
0 < x < 1, −∞ < y < +∞. It is a direct consequence of the Schwartz’s
reflection principle since ζ(x) is a real function in 0 < x < 1 that (z̄)∗ = x̄−
iȳ is another nontrivial zero of the Riemann function on the above pointed
out open strip. The basic point of our proof is to show that 1 − (z̄)∗ =
(1 − x̄) + iȳ is another zero of ζ(z) in the same horizontal line Im(z) = ȳ.
This result turns out that (1 − (z̄)∗ ) = z̄ on the basis of the validity of our
conjecture. As a consequence, we obtain straightforwardly that 1 − x̄ = x̄.
In others words, Real (z̄) = 12 .
August 6, 2008 15:46 9in x 6in B-640 ch09

284 Lecture Notes in Applied Differential Equations of Mathematical Physics

At this point, we call the readers’ attention, on the result that if z̄ is a


zero of the Riemann zeta function, then 1 − z̄ must be another nontrivial
zero is a direct consequence of the functional Eq. (D.2), since sin( πz2 ) and
Γ(1 − z) never vanish both on the open strip 0 < Real(z) < 1.
At this point of our note, we want to state clearly the significance of
replacing the Riemann’s original conjecture by our complex oriented con-
jecture rests on the possibility of progress in producing sound results for its
proof, which is not claimed in our elementary note. However, we intend to
point out directions (arguments) in its favor.
First, it is worth to call the readers’ attention on the validity of the
elementary expansion below on the open interval 0 < x < 1

 
1 d
ζ(x + ib) = exp ib (ζ(x)Γ(x) . (D.3)
Γ(x) dx

As a consequence, one can easily see that if ζ  (x + ib) (the zeta function
derivative on the horizontal line Im(z) = b = constant) does not possesses
zeros between two consecutive zeros of the zeta function ζ(x + ib) in 0 <
x < 1, then the Rieman’s conjecture is proved. 

At this point, we wish to remark the following weaker Conjecture D.2


is equivalent to our strong Conjecture D.1.

Conjecture D.2. If there is a zero of the Riemann’s zeta function t ∈ (0, 1)


n
such that the set of real number { ddn x ζ(t)}n=0,1,2,... is contained entirely on
the positive or negative real exist (these numbers all have the same signal),
then the Conjecture D.1 holds true.

On basis of the above-stated Conjecture D.2, we can easily show an


argument of our conjecture based on the power series expansion around the
fixed zero t (x̄ > t)

∞
ζ (n) (t)
ζ(x̄) = 0 = (x̄ − t)n > 0, (D.4)
n=1
n!

here x̄ is the supposed different zero of ζ(x) on the open interval (0, 1).
Finally, let us point out the following formula of ours, related to the
zeros of the Riemann zeta function on the complex plane C.
August 6, 2008 15:46 9in x 6in B-640 ch09

Some Comments on Sampling of Ergodic Process 285

Lemma D.1. Let {zn }n=0,1,2,... denote the zeros of the Riemann’s zeta
function ζ(z). We have thus the following result, for any integer p > 0.
 

 

∞
 

 1


n=1 (zn − zk )
p+1 


 

n=k
  2π 
1 1 ζ  (z)
= lim dθ + (zk + eiθ ) · (zk − z)p+1 e−ipθ .
z→zk 2π 0 z−1 ζ(z)
(D.5)

Proof. This result can be seen from the fact that h(z) = (z − 1)ζ(z) is
an integral function and from any integral function of order k, with zeros
z̄1 , z̄2 , . . . , (h(0) = 0)
   ∞

h (z)
p
d 1
lim (z − z̄k ) p+1
= −p! . (D.6)
z→zk dz h(z) (z̄n − z̄k )p+1
n=1
n=k

By rewriting the derivative term in the left-hand side of Eq. (D.6) by


means of the Cauchy theorem
  (p)     
h (z) p! h (w) −(p+1)
= dw [(w − z) ] , (D.7)
h(z) 2ni ∂D(z,1) h(w)

where D(z, 1) = {w | |w − z| ≤ 1}, and arranging terms of the form


(z − z̄k )p+1 /(w − z)(p+1) inside Eqs. (D.6) and (D.7), we obtain Eq. (D.5).


Lemma D.2. Let h(z) = ζ(z)(z − 1) be an (analytic) function on the


strip 0 < Real(z) < 1. Let us consider its expression on the unit disk con-
formally equivalent to the above considered strip, including the boundaries
correspondence.

1 i(i − w)
h lg = h(g(w)) = h̄(w). (D.8)
πi i+w

We have that, the generalized Jensen’s formula


 1  2π
η(x) 1  g
dx = dθ lg|h̄(eiθ )| (eiθ )dθ, (D.9)
0 x 2π 0 g
August 6, 2008 15:46 9in x 6in B-640 ch09

286 Lecture Notes in Applied Differential Equations of Mathematical Physics

where η(x) is the number of zeros of the zeta function on the region
g −1 (|w| < x), contained on the strip 0 < Real(z) < 1.
Proof. Since h(z) is an analytic function in the strip 0 < Real(z) < 1, we
are going to prove the general result for a region Ω conformally equivalent
to a disc of radius 1, including its boundary correspondence, namely

g : Ω → D(0, 1),
(D.10)
g(∂Ω) = ∂D(0, 1),

with w = g(z) a conformal mapping applying ∂Ω in ∂D(0, 1) diffeomorphi-


cally.
Let η(x) denotes the number of zeros of f (g −1 (w)) = h(w) on D(0, x).
By the usual Jensen’s formula for r < 1, with the hypothesis of h(w) has
no zero in ∂D(0, r).
 r 
η(x) 1 h (w)
dx = dw. (D.11)
0 x 2πi ∂D(0,r) h(w)

Since η(x) coincides with the number of zeros of the function f (z) in the
region g −1 (D(0, 1)), we obtain the Jensen’s formula in Ω in the general case
 r   
η(x) 1 g  (z)
dx = lg |l(z)| · dz − lg|f (0)|. (D.12)
0 x 2πi ∂Ω g(z)

Equation (D.8) comes from the fact that the strip 0 < Real(z) < 1 is
conformally mapped — with the boundary correspondence — on the unit
1
circle {w|j | |w| ≤ 1} by the function z = πi lg( i(i−w)
i+h ). 
August 6, 2008 15:46 9in x 6in B-640 ch10

Chapter 10

Some Studies on Functional Integrals Representations


for Fluid Motion with Random Conditions∗

10.1. Introduction

The main task of the statistical approach to random fluid dynamics1 is to


solve the set of infinite hierarchy equations for the random fluid velocity
correlation functions. One important scheme to solve these equations con-
sists in considering directly for the appropriate flux equation the random
conditions generating the flux stochasticity in the hope that the fluid tur-
bulence is appropriately described in this statistical approach at least as an
effective analytical theory.2
In this chapter, our aim is to present (formal) functional integral
representations for the Navier–Stokes equation in the following random
conditions:
(1) A pure white-noise initial fluid velocity condition in Sec. 10.2.
(2) A soluble Beltrami flux with appropriate Gaussian random stirrings in
Sec. 10.3.
(3) The Burger–Beltrami one-dimensional equation with a general
Gaussian random stirring in Sec. 10.4.
Finally, in Appendix A we show via path-integral techniques the
appearance of vortex phase factors as an important object in the advection
physics of scalars on fluid fluxes relevant to the studies presented in
Sec. 10.3.

10.2. The Functional Integral for Initial Fluid Velocity


Random Conditions

Let us start this section by writing the Navier–Stokes equation for


the velocity field of an incompressible fluid in the presence of a non-
random external force Fi (x, τ ) with a Gaussian (ultra-local) random initial
∗ Author’s original results.

287
August 6, 2008 15:46 9in x 6in B-640 ch10

288 Lecture Notes in Applied Differential Equations of Mathematical Physics

condition
 Tr
∂ ∂ i
vi − ν∆x vi + vk v = Fi , (10.1a)
∂τ ∂xk
vi (x, 0) = ϕi (x), (10.1b)
ϕi1 (x1 )ϕi2 (x2 ) = λδ (3) (x1 − x2 )δi1 i2 . (10.1c)

Let us remark that we have eliminated the pressure term − ρ1 ∆  · p by using


the incompressibility condition (∆v ) ≡ 0 which, in turn, leads us to consider
only the transverse part of the force and non-inertial field terms in Navier–
Stokes equation. The transverse part of a generical vector field W  (x, τ ) is
defined by the relations
 
1  −1  yW
(∆  )(y, ξ)
 Tr 
W (x, ξ) = W (x, ξ) + ∆ dy ,
4π x R3 |x − y|
 (x, ξ) = {W
W  i (x, ξ); i = 1, 2, 3}. (10.2)

Our task, now, is to compute the ϕ-averge of the N -point fluid velocity
field equation (10.1a) for an arbitrary space-time points, by means of a
functional integral representation for the characteristic functional of the
random fluid velocity fields Z[Ji (x, ξ)]; namely3

Vi1 (x1 , ξ1 , [ϕ]) · · · ViN (xN , ξN , [ϕ])ϕ


δ (N ) 
= (−1)N Z|Ji (x, ξ)|Ji (x,ξ)=0 , (10.3a)
δJi1 (x1 , ξ1 ) · · · δJiN (xN , ξN )
where
    ∞ 
Z[Ji (x, ξ)] = dµ[Vi ]exp − dx dξ(Vi · Ji )(x, ξ) . (10.3b)
M R3 0

The functional measure dµ[Vi ] in Eq. (10.3b) is defined over the func-
tional space M of all possible realizations of the random fluid motion defined
by Eq. (10.1). An explicit (formal) expression for the above functional
measure should be given by the product of the usual Feynman measure
weighted by a certain functional S[Vi ] to be determined,

dµ[Vi ] = DF [V i ]exp(S[V i ]), (10.4a)



DF [V i ] = (dVi (x, ξ)). (10.4b)
3
x∈R
0<ξ<∞
i=1,2,3
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 289

In order to determine the Weight Functional S[Vi ], we first rewrite the


Navier–Stokes equation as a pure integral equation which has an explicit
term taking into account the initial condition4

Ai [ṽ] = Bi [ϕ], (10.5)

with
 ∞ 
Ai [v] = Vi (x, ξ) − ds dyOijk (x − y, ξ − s)(Vj Vk )(y, s)
0 R3
 ∞ 
− ds dyH(1) (x − y), (ξ − s)Fi (y, s) (10.6a)
0 R3

Bi [ϕ] = dyH(0) (x − y, ξ)ϕi (y). (10.6b)
R3

Here, the Kernels Oijk , H(1) , H(0) are given, respectively, by


 
1 ∂ ∂ Ōi
Oik (z, ξ) = − Ōik + (z, ξ), (10.7a)
2 ∂z ∂zk
  
∂2 2νξ |z|  
Ōpq (z, ξ) = δpq θ(ξ)H(0) (|z|, ξ) + H0 (|z |, ξ)dz ,
∂zp ∂zq |z| 0
(10.7b)
 
1 |z|2
H(0) (|z|, ξ) = exp − , (10.7c)
(4πνξ)3/2 4πνξ
H(1) (|z|, ξ) = θ(ξ)H(0) (|z|, ξ). (10.7d)

Let us now introduce the following functional representation for the gener-
ating functional Z[Ji (x, ξ)]4,5

Z[Ji (x, ξ)] = DF [V i ]δ (F ) (V i − Ṽ i [ϕ])ϕ
   ∞ 
× exp − dx dξ(V i · J i )(x, ξ) , (10.8)
R3 0

where δ (F ) (·) denotes the delta-functional integral representation defined


by the rule

DF [Vi ]δ (F ) (Vi − Ai )Σ(Vi ) = Σ(Ai ), (10.9)
M

with Σ(Vi ) being an arbitrary functional defined on M .


By writing the ϕ-average in Eq. (10.9) by means of a Gaussian functional
integral in ϕ(x), we obtain the following functional integral representation
August 6, 2008 15:46 9in x 6in B-640 ch10

290 Lecture Notes in Applied Differential Equations of Mathematical Physics

for the weight S[V i ]:


  
1
exp(−S[V i ]) = DF [ϕi ] exp − dx(ϕi · ϕi )(x)
2λ R3
   ∞
× DF [K i ] exp i dx dξKi · (Ai [v] − B i |ϕ|) ,
R3 0
(10.10)
where we have used the Fourier functional integral representation for the
delta-functional in Eq. (10.8)
δ (F ) (V i − V i [ϕ]) = δ (F ) (Ai [v] − Bi [ϕ])
   
δ F i
= DETF Ak [v] D [K ] exp i dx
δVi R3
 ∞ 
× dξKi (A [v] − B [ϕ])(x, ξ) .
i i
(10.11)
0
It is worth to remark that the functional determinant in Eq. (10.11) is
unity as a straightforward consequence of the fact that the Green function
of the operator ∂/∂ξ is the step function (see Appendix B).
We, then, face the problem of evaluating ϕ and K functional integrals
in Eq. (10.11).
The ϕ-functional integral is of Gaussian type and easily evaluated
   
1
DF [ϕi ] exp − dx(ϕi · ϕi )(x)
2λ R3
    ∞ 
× exp − i dx dε(Ki · B [ϕ])(x, ε)
i
R3 0
   ∞  ∞
λ
= exp − dx1 dx2 dξ1 dξ2 K i (x1 , ξ1 )
2 R3 R3 0 0

× δ (3) (x1 − x2 )C(x1 , ξ1 ; x2 , ξ2 )K i (x2 , ξ2 ) , (10.12)

where the kernel C(x1 , ξ1 ; x2 , ξ2 ) is given by



C(x1 , ξ1 ; x2 , ξ2 ) = dzH(0) (x1 − z, ξ1 )H(0) (z − x2 , −ξ2 ) (10.13)
R3
and is the (formal) Green function of the self-adjoint extension of the square
B i∗ Bi diffusion operator
  
∂ ∂
(B i )∗ Bi = − − ν∆x1 − ν∆x1 (10.14)
∂ξ1 ∂ξ1
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 291

with the (well-posed) initial and boundary conditions

lim C(x1 , ξ1 ; x2 , ξ2 ) = δ (3) (x1 − x2 ). (10.15)


ξ1 →0+

Its explicit expression in K-momentum space is given by (see Ref. 6)


1 −νk2 |ξ1 −ξ2 | 2
C̃(k; ξ1 , ξ2 ) = − [e − e−νk (ξ1 +ξ2 ) ]. (10.16)
νk 2
As a consequence of Eq. (10.12), we have represented the weight S[v i ]
by a Gaussian functional integral in the Ki (x, ξ) field
    ∞
λ
exp(−S[v i ]) = DF [K i ] exp − dx1 dx2 dξ1 dξ2
2 R3 0

× K i (x1 , ξ1 )C(x1 , ξ1 ; x2 , ξ2 )δ (3) (x1 − x2 )Ki (x2 , ξ2 )


    ∞ 
× exp i dx dx i
dξ(K Ai [v])(x, ξ) (10.17)
R3 R3 0

By evaluating Eq. (10.17) we, thus, obtain the result


   ∞
1
exp(−S[v ]) = exp −
i
dx1 dx2 dξ1 dξ2 Ai [v](x1 , ξ1 )C −1
2λ R3 0

(3)
× (x1 , ξ1 ; x2 , ξ2 )δ (x1 − x2 )Ai [v](x2 , ξ2 ) (10.18)

By noting that (see Ref. 4)


 
−1 ∂
Bi [A[v]] = − ν∆k Ai [v]
∂ξ
   Tr
∂ ∂ i
= − ν∆x v + vk
i
v − FiTr . (10.19)
∂ξ ∂xk

We finally obtain the expression for the weight S[v i ]


  ∞   ∗
1 ∂
S[v i ] = dx dξdξ  − ν∆x Ai [v] (x, ξ)
2λ R3 0 ∂ξ
 

× − ν∆x Ai [v] (x, ξ  )
∂ξ 
  ∞  ∞
1
= dx dξ dξ 
2λ R3 0 0
August 6, 2008 15:46 9in x 6in B-640 ch10

292 Lecture Notes in Applied Differential Equations of Mathematical Physics

   Tr 
∂ ∂ i Tr
× − ν∆x v + vk
i
v + Fi (x, ξ)
∂ξ ∂xk
   Tr 
∂ ∂ i
× − ν∆x v + vk
i
v + Fi (x, ξ  ).
Tr
(10.20)
∂ξ  ∂xk
We obtain, thus, our proposed functional integral representation for
Eqs. (10.1a)–(10.1c)
    ∞ 
Z[Ji (x, ξ)] = D [v ] exp(−S[v ]) exp −
F i i
dx dξ(Ji · Vi )(x, ξ) .
R3 0
(10.21)
The above written functional integral is the main result of this section.
A perturbative analysis for Eq. (10.21) may be implemented by using
the free propagator Eq. (10.16) in the context of a background field decom-
position Vi = V̄i + βViq where V̄i satisfies the non-random Navier–Stokes
equation
 Tr
∂ ∂
V̄i = ν∆x V̄ − V̄k
i
V̄ i
+ F Tr , (10.22a)
∂ξ ∂xk
with
V̄ i (x, 0) ≡ 0, (10.22b)
with β being a coupling constant (β  1). It is worth remarking that the
cross term
  ∞
dx dξ(∂i v i ∆x vi )(x, ξ) (10.23a)
R3 0
i
vanishes in S[v ] as a result of the boundary condition
vi (x, 0) = vi (x, ∞) = 0 (10.23b)
for the pure random diffusion free propagator (Eq. (10.16)).
Finally, we point out that our proposed functional integral
equation (10.21) differs from that proposed in Ref. 7.

10.3. An Exactly Soluble Path-Integral Model for Stochastic


Beltrami Fluxes and its String Properties

In this section, our aim is to present, in our framework, an exactly soluble


path-integral model for stochastic hydrodynamic motions defined here to
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 293

be random regime of the physical Navier–Stokes equation (10.1a) in the


incompressible case dominated by generalized Beltrami fluxes defined by
the condition rot v = λv with λ a positive parameter.
Let us, thus, start with the usual Navier–Stokes equation (10.1a)
 
∂v 1 gradP
+ 2
grad(v ) − (v × rot v) = − + ν∆v + F ext , (10.24)
∂t 2 ρ
where the random stirring force is such that its satisfies the following spa-
tially non-local Gaussian statistics in our reduced model for turbulence, i.e.

(F ext )i (r, t)(F ext )j (r  , t ) = λ2 δij ((∆−1  


r )δ(r − r )δ(t − t ) (10.25)

where ∆−1r denotes the Laplacean Green function.


At this point we take curl of Eq. (10.24) and consider the already men-
tioned Beltrami flux condition and its direct consequence, namely:

λ2 v = rot(rot v) = grad(div v) − ∆v = −∆v, (10.26a)


v × rot v = v × (λv) = 0, (10.26b)

in order to replace the Navier–Stokes equation, Eq. (10.24) by the exact


soluble Langevin equation for the fluid flux stirred by the external force
Ωext = rot(F ext ) in our proposed model of Navier–Stokes turbulence dom-
inated by the generalized Beltrami fluxes
∂v(r, t) 1
= −νλ2 v(r, t) + Ωext (r, t) . (10.27)
∂t λ
The new external stirring Ωext = rot(F ext )(r, t) satisfies a Gaussian
process with the following two-point correlation function:
(r)    (r  )
Ωext ext  
 (r, t)Ω (r , t ) = (ε
jk
∂j )(ε j k
∂j  Fkext (r, t)Fkext  
 (r , t ))

    (r)
= λ2 (δ  δ jj − δ j δ  j )∂j ∂j  (∆−1
r δ
(3)
(r − r ) × δ(t − t ))
 (r) (r  )
= λ2 δ  δ (3) (r − r )δ(t − t ) − λ2 ∂ ∂ (∆−1  
r δ(r − r ))δ(t − t ).
(10.28)

It is obvious that Eq. (10.28) satisfies the incompressibility condition


necessary for the incompressibility consistency of our Brownian–Langevin
fluid equation (10.27) and its stochastic version below.
It is important to remark that the formal wave vectors of the Beltrami
hydrodynamical motions have eddies of a fixed scale |k| = λ in our reduced
August 6, 2008 15:46 9in x 6in B-640 ch10

294 Lecture Notes in Applied Differential Equations of Mathematical Physics

model. As a consequence of this fact, we assume implicitly the same wave


vector constraint in our random strings Eqs. (10.25) and (10.28).
Proceeding as in Sec. 10.2, it leads to the exact generating path
integral for our Brownian reduced model, where we have used the incom-
(r)
pressibility constraint ∂i v i (r, t) = 0 to see that the spatially non-local
piece of Eq. (10.28) does not contribute to the final path-integral weight
equation (10.29)
   +∞  ∞ 
3
Z[j(r, t)] = D[v(r, t)] exp i d r dt(j · v)(r, t)
−∞ 0
  
∂ 1 +∞ 3 3  +∞
× det − νλ2 δ (F ) (div v) exp − d rd r dtdt
∂t 2 −∞ 0
 
∂vi 
× + νλ vi (r − r )δ ii δ (3) (r − r  )δ(t − t )
2
∂t
 
(r) (r) −1   ∂vi
− ∂i ∂i (∆r δ(r − r ))δ(t − t )] + νλ vi (r − t )
2
∂t
   +∞  ∞ 
3
= D[(v(r, t)] exp i d r dt(j · v)(r, t)
−∞ 0
    2 
+∞ +∞
1 3 ∂v
× exp − d r dt + νλ2 v (r, t) . (10.29)
2 −∞ 0 ∂t

At this point it is worth to compare the exact soluble path-integral


written above (note the fixed wave vector |k| = γ imposed implicitly in
Eq. (10.29)) with that the one associated with the complete Navier–Stokes
equation for ultra-local random external stirring with strength D, namely:
Fi (r, t)Fj (r  , t ) = Dδ (3) (r−r )δ(t−t )δij and full range scale 0 ≤ |k| < ∞
(see Sec. 10.2)
  
∂ √ δ
Z[j(r, t)] = D[v(r, t)]det − ν∆ δk + D ((v · ∆)v)k
∂t δv
  +∞  √
1 +∞ 3 ∂
× − d r dt − ν∆v + D(v · ∇)v
2 −∞ 0 ∂t
2 
gradP √  +∞ 3  +∞
+ (r, t) exp i D d r (j · v)(v, t) .
ρ −∞ 0

(10.30a)
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 295

Let us remark that it is possible to eliminate the pressure term


−(1/ρ)grad P in this path-integral framework by using the incompress-
ibility condition div(v) = 0, which, by its turn leads one to consider only
the transverse part of the external force and of the nonlinear term in the
effective action in Eq. (10.30a) (see Sec. 10.2)
 
1 +∞ 3
Z[j(r, t)] = DF [v(r, t)] exp − d r
2 −∞
 +∞  √


× dt v − ν∆v + D((v · ∇v)tr )2 . (10.30b)
0 ∂t
Here the transverse part of a generic vector field is defined by the expression
(see Eq. (10.2))
1
(W (r, t))tr = W (r, t)− gradr (∆−1 (div W )). (10.31)

Note that now one should postulate the local two-point correlation
function in order to get Eq. (10.30b) Fitr (r, t)Fjtr ((r  , t )) = Dδ (3) (r −
r  )δ(t − t )δij .
It is worth to remark that Eqs. (10.3a) and (10.3b) applied to the Burger
equation lead to a different path-integral than that proposed in Ref. 8, since
in the path-integral framework
√ the viscosity is not a perturbative parameter
which, in our case, is D. Besides, the propagator in the free case for the
time parameter in the range 0 ≤ t ≤ ∞ is given by (see Eq. (10.16))
 −1  −1 
∂ ∂
− ν∆ · − − ν∆ (k, t, t )
∂t ∂t
1  2  2 

= − 2 e−νk |t−t | − e−νk |t+t | (10.32)
νk
and differing from the Dominicis–Martin propagator suitable for the range
−∞ ≤ t ≤ ∞ (Ref. 8)
 −1  −1 
∂ ∂
− ν∆ · − − ν∆ (k, t, t )
∂t ∂t
 +∞
 1
= dw(eiw(t−t ) ) 2 . (10.33)
−∞ w + ν 2 |k|4
Let us now evaluate the vortex phase factor defined by a fixed-time
spatial loop  = {(σ), a ≤ σ ≤ b} in our exactly soluble model
August 6, 2008 15:46 9in x 6in B-640 ch10

296 Lecture Notes in Applied Differential Equations of Mathematical Physics

equation (10.27) in order to see the connection with strings (random sur-
faces) (see Appendix A for the relevance of theses non-local objects for
advection phenomema)
   
exp i v((σ), t)d(σ)
v
     2 
1 +∞ 3 +∞
∂v
≡ DF [v(r, t)] exp − d r dt + νλ2 v (r, t)
2 −∞ 0 ∂t
  
× exp i v((σ), t)d(σ) . (10.34)

Since the flux is of a Beltrami type in our soluble model equation (10.29),
we propose to rewrite the circulation phase factor as a sum over all surfaces
bounding the fixed loop  by making use of Stokes theorem and by taking
into account again the Beltrami condition, i.e.
     +∞   
H
v·d 1 +∞ 3 ∂
ei c = DF [v(r, t)] exp − d r dt v − 2 + νλ2 v
2 −∞ 0 ∂t
     

× exp iλ v(x, t) · dA(x) . (10.35)
S S

By observing now that the two-point correlation of our Brownian–Beltrami


turbulent flux is exactly given by
 2 
e−νλ |t−t |
vi (r, t)vj (r , t )v = e−ik·(r−r) θ(t − t )δij , (10.36)
|k|=λ νλ2

with t, t ∈ [0, ∞] and θ(0) = 1/2 in this initial value problem, we can easily
evaluate the average equation (10.35) and producing a strongly coupled area
dependent functional for the spatial vortex phase factor in our proposed
turbulent flux regime
 H 
W [, v] ≡ ei  v(,t)·d
  
λ sin(λ|x − y|)
= exp − dA(y) dA(y) . (10.37)
ν S |x − y|
{S}

Just for completeness of our study and in order to generalize the Beltrami
flux turbulence analysis represented in the main text, for the physical case
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 297

of the complete wave vector range 0 ≤ |k| < ∞ in our turbulent path-
integral soluble model studies, we propose to consider a kind of generalized
Beltrami condition to overcome this possible drawback of our turbulence
modeling, namely:
rot v(r, t) = λ(r)v(r, t), (10.38)
where λ(r) is a positive function varying in the space and to be determined
from a phenomenological point of view. Note that the Fourier transformed
(wave vector) condition now takes the general form

|k| · |ṽ(k, t)| = d3 p|λ̃(p − q)| · |ṽ(p, t)| (10.39)
R3
which, by its turn, leads to the full range scale 0 < |k| < ∞ for the eddies
hydrodynamical motions under study. By supposing that the “vortical”
stirring equation (10.28) is a pure white noise process with strength D,

Ωext ext   3  
 (r, t)Ω (r, t ) = D · δ δ (r − r )δ(t − t ). (10.40)
It is a straightforward deduction by following our procedures as exposed
in the text to arrive at an analogous Gaussian path-integral for the gener-
alized Beltrami random hydrodynamical defined by Eq. (10.38). The gen-
eralized effective motion equation is given, in this new situation, by
   
∂ ∆λ ν ∂λ(r) ∂
−ν (r) − + νλ2 (r) δ ik
∂t λ λ(r) ∂xe ∂xe
 
∂λ(r)
+ ν εijk vk (r, t) = Ωext
i (r, t). (10.41)
∂xj
The Gaussian path-integral, thus, is exactly written as
 
3   +∞  ∞ 
3
Z[ji (r, t)] = F
D [vi (r, t)]exp i d r i
dt(j vi )(r, t)
i=1 ∞ 0
  +∞  ∞
1
× exp − d3 r dtvk (r, t)(M ∗ki · M is )vs (r, t) .
2D −∞ 0
(10.42)
Here, the differential operators entering in the kinetic term of the tur-
bulent path-integral are

∂ ν ∂λ(r) ∂ ν
M ∗ki = − + + · ∆λr
∂t λ(r) ∂xe ∂xe λ(r)

∆λ(r) ∂λ(r)
−ν + νλ2 (r) δ ki + νεkji (10.43)
λ(r) ∂xj
August 6, 2008 15:46 9in x 6in B-640 ch10

298 Lecture Notes in Applied Differential Equations of Mathematical Physics

and
 
∂ ν ∂λ(r) ∂ 2 ∆λ(r) ∂λ(r)
M is
= + − e
+ νλ (r) − ν δ is + νεijs .
∂t λ(r) ∂xe ∂x λ(r) ∂xj
(10.44)

It is worth pointing out that the exact evaluation of the variance in


Eq. (10.42) depends on the exact form of our rotation λ(r) defining the
Beltrami condition (10.38).
The vortex phase factor equation (10.34), takes now a form closely
related to the pure self-avoiding string theory in the case of a slowly varying
function |grad λ(r)| << λ(r) and λ(r) ∼ 1 (a very slowly r-varying function:
for instance as λ(r) = λ0 exp(−10−5 |r|2 ))
H   
i v(,t)d 1
e = exp − dA(x) · δ (3) (x − y) · dA(y)
ν S S
{S}
 
1
∼ exp − area(S) . (10.45)
ν
Now, if we follow Ref. 9 it is an easy task to deduce that the above
written time-fixed vortex phase factor satisfies the famous loop wave
equation for Abelian Q.C.D. at very low energy and a large number of
colors. It may be written in the geometrical (infinitely differentiable loops
(σ)) as the following:
δ  H  1 H
∂µx ei v(,t)d  = dyδ (3) (x − y)ei v(,t)d . (10.46)
δσµν (x) ν
The above obtained results rise hopes again that a string theory may be
relevant to understand turbulence modeled as an amalgamation of “rough”
roll up of random stirred fluid motions.

10.4. A Complex Trajectory Path-Integral Representation


for the Burger–Beltrami Fluid Flux

The Hopf wave equation for turbulence is a master functional compressing


the infinite hierarchy fluid velocity correlation functions in a single func-
tional differential equation.3
In this section, our aim is to a certain extent complete the previous path-
integral studies by presenting a complex trajectory path-integral represen-
tation for a reduced model simulating “Burger turbulence” by considering
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 299

directly the “experimental observable” N -point grid velocity observable as a


fundamental object of the proposed dynamically reduced Burger–Beltrami
turbulent flux model below defined.
Let us start with the dynamical equation defining our one-dimensional
Brownian-like fluid flux
 
∂v(x, t) ∂v
+ v (x, t) = −(νλ2 )v(x, t) + f (x, t),
∂t ∂x
(10.47)
v(x, 0) = g(x),

where we have replaced the usual fluid viscosity term νd2 v(x, t)/dx2 by
the pure damping term −νλ2 v(x, t) (the reader should compare our pro-
posed Brownian-like flux with that of the Navier–Stokes–Beltrami studied
in Sec. 10.3).
One of the most important observable objects in fluid turbulence is the
fixed velocities measurements at the grid points (x ) and at a common
observation time t
N 

δ(v(x ; t) − v ) (10.48)
=1

where the average   is defined by the random stirring satisfying the


gaussian statistics10

f (x, t)f (x , t ) = k(x − x )δ(t − t ). (10.49)

In momentum space, the observable equation (10.48) is given by the fol-


lowing (grid dependent) characteristic functional (the Hopf wave functional
restricted on the N -point grid)
  N 

ψ((x1 , . . . , xN ); (p1 , . . . , pN ); t) = exp i p v(x , t) . (10.50)
=1

The Hopf wave equation associated with our model equation (10.48) is
given in a closed form by applying straightforwardly the methods of Ref. 10

−i ψ((x1 , . . . , xN ); (p1 , . . . , pN ); t)
∂t
 
 N    
N 
∂ 1 ∂ ∂
= p − (νλ2 )p + (k(x − x )p p
 ∂p p ∂x ∂p 
=1 =1, =1

× ψ((x1 , . . . , xN ); (p1 , . . . , pN ); t) (10.51)


August 6, 2008 15:46 9in x 6in B-640 ch10

300 Lecture Notes in Applied Differential Equations of Mathematical Physics

Note that we must add the deterministic initial date condition to


Eq. (10.51)
 N 

ψ(x1 , . . . , xN ); (p1 , . . . , pN ); t → 0+ ) = exp i p g(x ) (10.52)
=1

(a)
Let us remark that in the physical grid on R3 = {xk ; a = 1, 2, 3; k =
1, . . . , N }, Eq. (10.51) naturally reads
∂ (a) (a) (a) (a)
−i ((x , . . . , xN ); (p1 , . . . , pN ); t)
∂t 1
3
   
N 
(a) ∂ 1 ∂ 2 (a) ∂
× p (a) (a) (a)
− νλ p (a)
=1 a=1 ∂p p ∂x ∂p
(a) (a) (a)
+ ψ((x1 , . . . , xN ); (p1 , . . . , pN ); t)
 
N
(a) (a) (a) (b)
+ Kab (x − x )p p 
=1, =1

(a) (a) (a) (a)


× ψ((x1 , . . . , xN ); (p1 , . . . pN ); t) (10.53)

Hereafter as said earlier we will present our study of Eq. (10.53) for the one-
dimensional case equation (10.51). By introducing the mixed coordinates
defined by the transformation law.

pj + xj = uj ; pj − xj = vj . (10.54)

The turbulent wave equation (10. 51) takes the more invariant form similar
to a many-particle Schrödinger equation in quantum mechanics.

−i (ψ(u1 , . . . , uN ); (v1 , . . . vN ); t)
∂t
N    
1 ∂2 ∂2 2 ∂ ∂
= − − −
4 ∂ 2 u ∂ 2 v u + v ∂u ∂v
=1
 
νλ2 ∂ ∂
− (u + v ) + ψ((u1 , . . . , uN ); (v1 , . . . , vN ); t)
2 ∂u ∂v
 
 
1 
N
u − u + (v − v ) 
+ (u + v )(u + v )K
4 1
2
=1; =1

× ψ((u1 , . . . , uN ); (p1 , . . . , pN ); t) (10.55)


August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 301

and
 N    
 u + v u − v
ψ((u1 , . . . , uN ); (v1 , . . . vN ); 0) = exp i g
2 2
=1

(10.56)

The above written closed partial differential equation is the basic result of
this section. At this point we can implement perturbative calculations for
our turbulent wave equation by considering a physical slowly varying (even
function) correlation function of the form
 1/2
κ0 2 K(0)
K(x) ≈ K(0) − x ; |x|  ≡ L,
2 κ0
(10.57)
0; |x| L,

which by its turn leads to the harmonic and quartic anharmonic potential
which is written as

N
K(0)(u u + v v + u v + v u )
=1; =1

1
− κ0 (u + v )(u + v [u2 + u2 + v2 + v2 − 2u u − 2v v ]) .
8
(10.58)

In the important case of the single fluid velocity average, our turbulent wave
equation takes the following form, after making an analytic continuation
v → iv; namely

i ψ(u, v; t) = (L0 + L1 )ψ(u, v; t), (10.59)
∂t
with the initial condition
    
+ u + iv u − iv
ψ(u, v; t → 0 ) = exp i g . (10.60)
2 2
Here the kinetic and perturbation terms are
 
1 ∂2 ∂2 k(0) 2
L0 = − + 2 + (u − v 2 ),
4 ∂u2 ∂v 4
    (10.61)
2 ∂ 1 ∂ 2 ∂ 1 ∂
L1 = − − νλ (u + iv) + .
u + iv ∂u i ∂v ∂u i ∂v
August 6, 2008 15:46 9in x 6in B-640 ch10

302 Lecture Notes in Applied Differential Equations of Mathematical Physics

The harmonic oscillator propagator of the kinetic term equation (10.61) is


determined in a straightforward way and a Feynman diagramatic analysis
may be easily implemented for ν  1 by the same perturbative procedure
used in quantum mechanical problems. Similar remarks hold true in the
general case equation (10.55).
It is worth to point out that analogous results are easily obtained in the
physical case of turbulent Beltrami flux in the three-dimensional case.
Let us comment the case of general turbulent flux. In this case, although
being impossible to write a closed partial differential equation as we did in
Sec. 10.4, we can develop approximate schemes to solve the full functional
Hopf equation by approximating the fluid shear stress tensor by finite dif-
ferences, namely:
 
d2 v(xj ; t) ν
ν ≈ (−2v(xj , t) + v(xj+1 ; t) + v(xj−1 , t)) (10.62)
dx2j ∆

With the uniform grid spacing ∆ = |xj+1 − xj |.


Let us now write a trajectory functional integral representation for the
initial-value problem equation (10.55) after taking into account the analytic
continuation v → iv there.
As a first step to achieving our goal, we write the associated Green func-
tional of Eq. (10.55) in an operator form (the Feynman–Dirac propagation)
for the free case k ≡ 0

G[(u , v ); (u , v ; t)]


   N   
 1 2 ∂ 1 ∂

= (u , v ) exp it ∆(u ,v ) − −
4 u + iv ∂v i ∂v
=1
    
νλ2 ∂ 1 ∂   
− (u + iv ) + (u , v ) . (10.63)
2 ∂u i ∂v 

As in the usual Feynman analysis we write Eq. (10.63) as an infinite


product of short-time t-propagation and consider the standard short–time
expansion

(I) (I) (I−1) (I−1)


lim (u , v )| exp(isH)|(v , u 
s→0+
     
p2I + qI2 2  
() ()
= lim N N
d pI d qI exp is − (I) (I)
ipI − qI
s→0+ 4 u + iv
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 303

N 
νλ2 (I)     () () 
(I) () () ()
− (u + iv ) ipI + qI − exp ipI (uI − uI−1 )
2
=1
N 
  () 
× exp iqI (vI − vI−1
−1
) , (10.64)
=1

where H denotes the second-order differential operator inside the brackets


of Eq. (10.63).
If we substitute Eq. (10.64) into the short-time product expansion of
Eq. (10.63), namely

u , v |eitH |u , v 


M  +∞
       
I 
 t  I−1 I−1
= I
duI dvI (u , v ) exp i 
H (u , v ) (10.65)
−∞ M
I=1

and evaluate the (pI , qI )-momenta functional integrals (see Ref. 11 for a
detailed exposition), we get our searched trajectory path-integral represen-
tation for the Green-function of Eqs. (10.56)–(10.65) in the free case k ≡ 0

Ḡ [(u , v ); ((u , v ); t]


     2
i t d
= Ū (0)=u  v̄ (0)=v  exp dσ Ū (σ)
  4 0 ∂σ
Ū (t)=u v̄ (t)=v

N   
d () 2
−2 Ū (σ) − − (νλ2 )(Ū (σ) + iV̄ (σ))
∂σ Ū (σ) + iV̄ (σ)
=1

N  2 
2
+ − − (νλ2 )(Ū (σ) + iV̄ (σ))
=1
Ū  (σ) + i V̄ (σ)
   2 N  
i t ∂  ∂
× exp dσ V̄ (σ) − 2 V̄ (σ)
4 0 ∂σ ∂σ
=1
 2
2·i
× − + i(νλ2 )(Ū (σ) + iV̄ (σ))
U (σ) + iV̄ (σ)
N  
2i 2 ()
× − + i(νλ )(Ū (σ) + iVI (σ)) . (10.66)
=1
Ū (σ) + iV̄ (σ)

Note that the discrete index I = 1, . . . , M has become the continuous


time parameter σ ranging in [0, t].
August 6, 2008 15:46 9in x 6in B-640 ch10

304 Lecture Notes in Applied Differential Equations of Mathematical Physics

The general k ≡ 0 case is straightforwardly obtained from Eq. (10.66)


by only considering the additional weight
 
 
N
exp i  (Ū  (σ) + iV̄  (σ))(Ū  (σ) + iV̄  (σ))
 
=1, =1
 
(Ū  (σ) − Ū  (σ)) + i(V̄  (σ) − V̄  (σ))
×K . (10.67)
2

It is obvious from the above written N -body (complex valued!)


trajectory path-integrals representations that any analytical analysis will
be somewhat cumbersome. However, its numerical (Monte-Carlo and F.F.T
algorithms) studies may be useful to implement approximate evaluations
on applied problems.

References

1. R. Kraichnan and D. Montgomery, Rep. Prog. Phys. 43 (1980).


2. A. A. Migdal, Mod. Phys. Lett. A 6(11), 1023 (1991).
3. A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics (MIT. Press,
Cambridge, MA, 1971).
4. G. Rosen, J. Math. Phys. 12, 812 (1971).
5. L. C. L. Botelho, J. Phys. A: Math. Gen. 23, 1829 (1990).
6. W. G. Faris and G. J. Lasinio, J. Phys. A: Math. Gen. 15, 3025 (1982).
7. P. C. Martin, E. O. Siggia and H. A. Rose, Phys. Rev. A8, 423 (1973);
E. Medina, T. Hwa and M. Kardar, Phys. Rev. 39A (1989); V. Gurarie and
A. Migdal, Phys. Rev. E54, 4908 (1996); A. A. Migdal, Int. J. Mod. Phys.
A9, 1197 (1994).
8. L. C. L. Botelho, J Math. Phys. 42, 1689 (2001).
9. L. C. L. Botelho, J. Math. Phys. 30, 2160 (1989).
10. A. M. Polyakov, Phys. Rev. E 52, 4908 (1996); L. C. L. Botelho, Int. J. Mod.
Phys. B 13, 1663, (1999).
11. L. C. L. Botelho and R. Vilhena, Phys. Rev. E 49, 1003 (1994); L. C. L.
Botelho, J. Phys. A: Math. Gen. 34, L131–L137 (2001).

Appendix A. A Perturbative Solution of the Burgers


Equation Through the Banach Fixed
Point Theorem

A very important result in the theory of metric spaces is the famous


Banach–Picard fixed point theorem.
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 305

Theorem of Banach–Picard. Let T : (X, d) → (X, d) be a con-


tractive application in a given complete metric space (X, d) ≡ X. This
means that for any pair (x, y) ∈ X × X, we have the uniform bound
with L < 1

d(T x, T y) ≤ Ld(x, y). (A.1)

As a consequence, we have the following results:

(a) There is a unique solution of the functional abstract equation

T x̂ = x̂. (A.2)

(b) For any x̄ ∈ X, the sequence (xn ) defined by the recurrrence relationship

xn+1 = T n x̄ (A.3)

converges to the (unique) solution of Eq. (A.2).


(c) For each n, one has the mathematical control of the error from the
approximate sequence equation (A.3) to the real solution x̂ Eq. (A.2)

Ln−1
d(xn , x̂) ≤ d(x̄, T x̄). (A.4)
1−L

We left its (easy) detailed proof to our readers.

Another important result is the following theorem related to the con-


tinuous dependence of the solution equation (A.2) on parameters.

Theorem A.1. Let X be a complete metric space and A be a general


topological space. Let T : X × A → X be a family of locally contractive
applications. This means that for any λ ∈ A, there is a topological
neighbourhood of λ, namely Vλ ∈ A, and a constant Lλ < 1, such that
d(Tµ x, Tµ y) ≤ Lλ d(x, y), for µ ∈ Vλ . Then, we have that all fixed points of
Tλ are continuous functions of the parameter λ in A.

A curious interplay among compacity and “almost” contractive appli-


cations is the following lemma.

Lemma A.1. Let X be a compact metric space and T : X → X an “almost ”


contractive application d(T x, T y) < d(x, y) for x = y. (Example: T : [0, 1] →
August 6, 2008 15:46 9in x 6in B-640 ch10

306 Lecture Notes in Applied Differential Equations of Mathematical Physics

[0, 1]; T (x) = x − x2 ). Let us consider the following real function on X


ϕ: X → R
(A.5)
x → d(x, T x).
Now ϕ(x) has a minimum value of δ in X due to its compacity. On the
other side, if δ > 0, one can consider the restriction of ϕ to a new compact
domain set W = {x ∈ X | d(x, T x) ≤ δ2 }. One finds in W ⊂ X, another
different minimum of ϕ in X. So δ = 0, which means that there is a x̂ ∈ X,
such that
d(x̂, T x̂) = 0 ⇔ T x̂ = x̂. (A.6)
Let us examplify the Banach–Picard theorem by considering the fol-
lowing initial value problem for differential equations in Banach Space E,
where all initial conditions are contained in the Ball of radius R(||U0 || ≤ R)
and the nonlinearity is supposed to be Lipschitz in any bounded set in E
dU
= F (t, U (t)),
dt (A.7)
U (t0 ) = U0 .
We introduce the equivalent integral form of Eq. (A.7) in our proposed
complete metric space E(T,R) of continuous function taking values in E
(
E(T,R) = u ∈ C([0, T ], E); sup ||U (t)||E ≤ K(R);
0≤t≤T ) (A.8)
d(u, v) = sup {||u(t) − v(t)||E } .
0≤t≤T

Namely,
 t
U (t) = U0 + F (s, U (s))ds = F̂ (U )(t). (A.9)
0

Let us now choose the parameter T, R in the definition of Eq. (A.8)


in such a way that F̂ maps E(R,T,α) into E(R,T,α) and F̂ be a contractive
application there.
Now, we can see that
 t
d((F̂ U )(t), 0) ≤ ||U0 ||E + ||F (s, U (s))||E ds
0
 t
≤ ||U0 ||E + ||F (U ) − F (0) + F (0)||E (s)ds
0

≤ R + (||F (0)||E + 2L(KR ) · KR )t. (A.10)


August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 307

Here, for (x, y) ∈ E(T,R) we have the Lipschitz property


L(KR )
* +, -
||F (x, t) − F (y, t)||E ≤ sup {(R̄, t)} ·||x − y||E
0≤t≤T

≤ 2L(KR ) · K(R). (A.11)

Let us now choose the global time evolution T in such a way that

R + T (||F (0)||E + 2L(KR ) · KR ) ≤ K(R) (A.12a)

or
K(R) − R
T ≤ . (A.12b)
||F (0)||E + 2K(R)L(KR )
We have the following additional estimate:
 t
d(F̂ U, F̂ V ) = sup ds||F (s, U (s)) − F (s, V (s))||E
0≤t≤T 0

≤ T · L(KR ) · d(U, V ) . (A.13)

If one chooses the global time satisfying again the bound for any δ > 1.
1
T ≤ (A.14)
L(KR )(1 + δ)
we have a unique solution of Eq. (A.7) in E(T,R) . For instance, if F (0) = 0
and K(R) = R + b, the compatibility of Eqs. (A.12b) and (A.14) means the
existence of b > 0 such that b < 2R/(δ − 1).
Related to the material exposed in this chapter, let us apply the
Banach–Picard fix point theorem to a Burger equation with a viscosity
time-dependent parameter and zero initial conditions in R × R+
 t  +∞
1
U (x, t) = dτ dy 
ν 0 −∞
   
∂ (x − y  )2 1 def
× −  exp − U (y  , τ )2 · τ (− 2 +β) ≡ F (u)
∂y 4τ
(A.15)

Let us introduce the complete metric space defined below

E(T,R) = {U (·, t) ∈ C([0, T ], R); sup (||U (·, t)||L∞ (R) ) ≤ R}. (A.16)
0≤t≤T
August 6, 2008 15:46 9in x 6in B-640 ch10

308 Lecture Notes in Applied Differential Equations of Mathematical Physics

Proceeding as in the previous example, we will choose T in such a way


that F (E(T,R) ) ⊂ E(T,R) . This means that (for β > − 21 )
sup ||(F U )(x, t)||L∞ (R)
0≤t≤T
  t  +∞
1  2
sup 
2
≤ dζ dy|e−(x−y) /4ζ |x − y|
ν 0≤t≤T  0 −∞ 4
 
ζβ  2 
× ||U (y, ζ)||L∞ 
ζ 1/2 ·ζ
 t  ∞ 
R2  
sup  dz · e−(z) /4ζ · z 
3 2
≤ dζζ (β− 2 )
ν 0≤t≤T 0 0
c
 t  * +, -
R 2  1 

≤ sup  dζζ (β− 2 
)
 dz̄e −z̄ 2 /4

ν 0≤t≤T 0 0

R2 c 1 1 ν(2β + 1)
≤ T β+ 2 < R ⇔ T β+ 2 < (A.17)
ν(β + 12 ) 2cR

Let us further make a choice of the global time T in such a way that F̂
is a contractive application in E(T,R) . In order to show such a result, let us
consider the estimate below
. /
sup ||(T U )(x, t) − (T V )(x, t)||L∞ (R)
0≤t≤T
  
 t  +∞ e−|x−y|2 · 2|x − y| · ζ β 
1  
≤ sup  dζ 

ν 0≤t≤T 0 −∞ 4·ζ 1/2 ·ζ 

× ||U 2 (·, t) − V 2 (·, t)||L∞ (R)


 t 
1  
≤ sup   dζζ β− 12
2 · c||U (·, t) − V (·, t)||L (R) 

2ν 0≤t≤T 0
≤2R
* +, -
× (||U (·, t)||L∞ (R) + ||V (·, t)||L∞ (R) )
1
 
2c T β+ 2
≤ R· sup ||U (·, t) − V (·, t)||L∞ (R) (A.18)
ν (β + 12 ) 0≤ζ≤T

So F̂ will be a contractive application in E(T,R) if


1
4cRT β+ 2 1 ν(2β + 1)
< 1 ⇔ T β+ 2 < . (A.19)
ν(2β + 1) 4cR
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 309

We have thus the existence and unicity for the Burger equation in the
intergral form (A.15) for global times satisfying Eq. (A.19).

Appendix B. Some Comments on the Support of Functional


Measures in Hilbert Space

Let us comment further on the application of the Minlos theorem in Hilbert


spaces. In this case one has a very simple proof which holds true in general
Banach spaces (E, || ||).
Let us thus, give a cylindrical measure d∞ µ(x) in the algebraic dual
E alg of a given Banach space E (Chap. 5).
Let us suppose either that the function ||x|| belongs to L1 (E alg , d∞ µ(x)).
Then the support of this cylindrical measures will be the Banach space E.
The proof is the following:
Let A be a subset of the vectorial space E alg (with the topology of
pontual convergence), such that A ⊂ E c (so ||x| = +∞). Let the sets be
0
Aµ = {x ∈ E alg | ||x|| ≥ n}. Then we have the set inclusion A ⊂ ∞ n=0 An ,
so its measure satisfies the estimates below:
µ(A) ≤ lim inf µ(An )
n

= lim inf µ{x ∈ E alg | ||x|| ≥ n}


n

1
≤ lim inf ||x||d∞ µ(x)
n n E alg
||x||L1 (E alg ,d∞
µ )
= lim inf = 0. (B.1)
n n
Leads us to the Minlos theorem that the support of the cylindrical
measure in E alg is reduced to the own Banach space E.
Note that by the Minkowisky inequality for general integrals, we have
that ||x||21 ∈ L1 (E alg , d∞ µ(x)). Now it is elementary evaluation to see that
if A−1 ∈ 1 (H), when E = H, a given Hilbert space, we have that

d∞ 2
A µ(x) · ||x|| = TrH (A
−1
) < ∞. (B.2)
Halg
This result produces another criterium for supp d∞
A µ = H (the Minlos
Theorem), when E = H is a Hilbert space.
It is easy to see that if

||x||d∞ µ(x) < ∞ (B.3)
H
August 6, 2008 15:46 9in x 6in B-640 ch10

310 Lecture Notes in Applied Differential Equations of Mathematical Physics

then the Fourier-transformed functional



Z(j) = ei(j,x)H d∞ µ(x) (B.4)
H

is continuous in the norm topology of H.


Otherwise, if Z(j) is not continuous in the origin 0 ∈ H (without loss
of generality), then there is a sequence {jn } ∈ H and δ > 0, such that
||jn || → 0 with

δ ≤ |Z(jn ) − 1| ≤ |ei(jn ,x)H − 1|d∞ µ(x)
H

≤ |(jn , x)|d∞ µ(x)
H
 
≤ ||jn || ||x||d∞ µ(x) → 0, (B.5)
H

a contradiction with δ > 0.


Finally, let us consider an elliptic operator B (with inverse) from
the Sobelev space H−2m (Ω) to H2m (Ω). Then by the criterium given by
Eq. (B.2) if
m m
TrL2 (Ω) [(I + ∆)+ 2 B −1 (I + ∆)+ 2 ] < ∞, (B.6)

we will have that the path-integral written below is well-defined for x ∈


H+2m (Ω) and j ∈ H−2m (Ω). Namely
  
1
exp − (j, B −1 j)L2 (Ω) = dB µ(x) exp(i(j, x)L2 (Ω) ). (B.7)
2 H+2m (Ω)

By the Sobolev theorem which means that the embeeded below is con-
tinuous (with Ω ⊆ Rν denoting a smooth domain), one can further reduce
the measure support to the Hölder α continuous function in Ω if 2m− ν2 > α.
Namely, we have a easy proof of the famous Wiener theorem on sample
continuity of certain path-integrals in Sobolev spaces

H2m (Ω) ⊂ C α (Ω) (B.8a)

The above Wiener theorem is fundamental in order to construct non-


trivial examples of mathematically rigorous Euclideans path-integrals in
spaces Rν of higher dimensionality, since it is a trivial consequence of
the Lebesgue theorem that positive continuous functions V (x) generate
August 6, 2008 15:46 9in x 6in B-640 ch10

Studies on Functional Integrals Representations 311

functionals integrable in {H2m (Ω), dB µ(ϕ)} of the form below



exp − V (ϕ(x))dx ∈ L1 (H2m (Ω), dB µ(ϕ)). (B.8b)

As a last important remark on cylindrical measures in separable Hilbert


spaces, let us point at to our reader that the support of such above mea-
sures is always a σ-compact set in the norm topology of H. In order to see
such result let us consider a given dense set of H, namely {xk }k∈I + . Let
{δk }k∈I + be a given sequence of positive real numbers with δk → 0. Let
2∞
{εn } be another sequence of positive real numbers such that n=1 εn < ε.
3∞
Now, it is straightforward to see that H ⊂ h=1 B(xk , δk ) ⊂ H and thus
3n
lim sup µ{ k=1 B(xk , δk )} = µ(H) = 1. As a consequence, for each n, there
3kn
is a kn , such that µ k=1 B(xk , δk ) ≥ 1 − ε.
0∞ 4 3kn 5
Now, the sets Kµ = n=1 k=1 B(xk , δk ) are closed and totally
bounded, so they are compact sets in H with µ(H) ≥ 1 − ε. Let us now
choose ε = n1 and the associated compact sets {Kn,µ }. Let us further con-
3n
sider the compact sets K̂n,µ = =1 K,µ . We have that K̂n,µ ⊆ K̂n+1,µ ,
3∞
for any n and lim sup µ(K̂n,µ ) = 1. So, supp dµ = n=1 K̂n,µ , a σ-compact
set of H.
We consider now an enumerable family of cylindrical measures {dµn }
in H satisfying the chain inclusion relationship for any n ∈ I +
supp dµn ⊆ supp dµn+1 .
(n)
Now it is straightforward to see that the compact sets {K̂n }, where
3∞ (m) 3∞ (n)
supp dµm = n=1 K̂n , is such that supp{dµm } ⊆ n=1 K̂n , for any
m ∈ I +.
Let us consider the family of functionals induced by the restriction of
(n)
this sequence of measures in any compact K̂n . Namely

µn → Ln(n) (f ) = f (x) · dµp (x). (B.8c)
(n)
K̂n
(n) 3∞ (n)
Here f ∈ Cb (K̂n ). Note that all the above functionals in n=1 Cb (K̂n )
are bounded by B.1. By the Alaoglu–Bourbaki theorem a compact set in
3∞ (n) ∗
the weak star topology of n=1 Cb (K̂n ) is formed, so there is a sub-
sequence (or better the whole sequence) converging to a unique cylindrical
measure µ̄(x). Namely,
 
lim f (x)dµn (x) = f (x)dµ̄(x) (B.8d)
n→∞ H H
3∞ (n)
for any f ∈ n=1 Cb (K̂n ).
August 6, 2008 15:46 9in x 6in B-640 ch11

Chapter 11

The Atiyah–Singer Index Theorem:


A Heat Kernel (PDE’s) Proof∗

Dirac-type operators in Riemannian manifolds are mathematical objects of


cornerstone importance in differential geometry1–3 and with a growing use-
fulness in quantum physics.4,5 Unfortunately, many studies related to the
theory of Dirac operators are very sophisticated for applied scientists due
to the use of sophisticated and cumbersome methods of algebraic topology
and bundles manifolds.1
In this chapter, we intend to present in a relatively simple way and
based in the elementary aspects of the Seeley theory of pseudo-differential
operators, one of the most celebrated differential topological theorem: The
Atiyah–Singer index theorem which (in one of its “palatable” version) says
roughly that the trace of the evolution operator associated to a certain
class of Dirac operators defined in two-dimensional orientable compact
Riemannian manifolds possess a manifold topological index — its Euler–
Poincaré genus.
Let us thus start with A denoting an elliptic differential operator of
second order acting in the space Cc∞ (R2 )q×q formed by all these functions
infinitely differentiable with compact support in R2 and taking values in
Mq×q (C) (the vector space of the complex matrices q × q)

A= aα (x)Dxα , (11.1)
|α|≤2

where α = (α1 , α2 ) are non-negative integers associated with the basic self-
adjoint differential operators
 α1  α2
1 ∂ 1 ∂
Dxα = , (11.2)
i ∂x1 i ∂x2
with Aα (x) ∈ Cc∞ (R2 )q×q .
∗ Author’s original results.

312
August 6, 2008 15:46 9in x 6in B-640 ch11

The Atiyah–Singer Index Theorem 313

Let us now consider the (contractive) semigroup generated by the


operator A through the spectral calculus for the operators in Banach spaces

1
exp(−tA) = dλe−tλ (λ11 ε×ε − A)−1 , (11.3)
2πi Cb

where C is a given closed curve containing in its interior the spectrum of


A : σ(A) which is supposed to be in the semiline R+ .
According to Seeley,1,2 let us consider the operational symbol of the
operator (λ11q×q − A) which is defined by the relationship below:

σ(A − λ11)(ξ)def = e−ixξ (A − λ11)eixξ


 

= Aα (x)(ξ1 )α1 (ξ2 )α2  − λ11q×q
|α|≤2

= Aj (x, ξ, λ), (11.4)
|j|=0

where for 0 ≤ j < 2,



Aj (x, ξ, λ) = Aα (x)(ξ1 )α1 (ξ2 )α2 , (11.5a)
|α|=α1 +α2 =j

and for j = 2,
 

A2 (x, ξ, λ) = −λ11 q×q +  Aα (x)(ξ1 )α1 (ξ2 )α2  . (11.5b)
|α|=α1 +α2 =2

Let us now consider the usual Fourier transforms in L1 (R2 ) ∩ L2 (R2 )


with its operational rule
1
F(ξ) = √ d2 xeiξx f (x) (11.6a)
2π R2

1
(Af )(x) = d2 ξeixξ [σ(A)(x, ξ)]F (ξ). (11.6b)
(2π)2 R2

At this point, it is of crucial importance to note that the functions aj (x, ξ, λ)


are all homogeneous √ functions of degree j when considered as functions of
the variable ξ and λ:

aj (x, cξ, c2 λ) = (c j )aj (x, ξ, λ). (11.7)


August 6, 2008 15:46 9in x 6in B-640 ch11

314 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now it is an important technical theorem of Seeley that the Green function


of the operator (A − λ11) has a symbol which is given by a series of smooth
functions {C−2−j (x, ξ)}j=0 in the Seeley topology


σ((A − λ11)−1 ) = C−2−j (x, ξ). (11.8)
j=0

As a consequence of the above formulae, we have the following exact for-


mulae for the inverse of any elliptic inversible operator A

(A−1 f )(x) = dt(e−tA f )(x)
0

   
1 C−2−j (x, ξ, λ)
= d2 ξd2 xd2 yeiξ(x−y) dλ f (y) .
j=0
2πi C b
C λ
(11.9)
A crucial observation can be made at this point. One can introduce a
noncommutative multiplicative operation among the Seeley symbols asso-
ciated to elliptic operators A and B acting on the domain Cc∞ (R2 )q×q
namely,
def
σ(A) ◦ σ(B) ≡ σ(A ◦ B), (11.10)

where A ◦ B is the operator composition. Explicitly, we have that



σ(A) ◦ σ(B) = [Dξα σ(A)(x, ξ)][iDxα σ(B)(x, ξ)]/α!. (11.11)
|α|=α

Here

α! = α1 ! · α2 !. (11.12)

From the obvious relationship

(A − λ11) ◦ (A − λ11)−1 = 1, (11.13)

we have that

σ(A − λ11) ◦ σ((A − λ11)−1 ) = 1. (11.14)

As a consequence it yields the result


  
1   ∞ 
(Dξα σ(A − λ11)(x, ξ))iDxα  C−2−j (x, ξ) = 1, (11.15)
 α! 
|α|≤2 j=0
August 6, 2008 15:46 9in x 6in B-640 ch11

The Atiyah–Singer Index Theorem 315

or in an equivalent way
 
1   

[(Dξα σ(A − λ11)(x, ξ))iDξα (C−2−j (x, ξ))] = 1. (11.16)
α!  
|α|≤2
j=0

The above equations can be solved by introducing the scaled variables


(t > 0)

ξ = tξ  ; λ1/2 = t(λ )1/2 , (11.17)

and
1
C−2−j (x, tξ  , ((tλ ) 2 )2 ) = t−(2+j) C−2−j (x, ξ  , λ ). (11.18)

After substituting Eq. (11.18) into Eq. (11.16) and by comparing the
resulting power series in the scale factor 1/t, one obtains the famous Seeley
recurrence equations which for t → 1, give us explicitly expressions of all
Seeley coefficients {C(−2+j) (x, ξ)} of the Green function of the operator A
(note that they are elements of Mq×q (C))

C−2 (x, ξ) = (a2 (x, ξ))−1 , (11.19)

0 = a2 (x, ξ)C−2−j (x, ξ)


 
1   
 (Dξα ak (x, ξ))(iDxα C−2− (x, ξ))
+
α!  . (11.20)
<j
k−|α|−2−=−j

For the Laplacian-like elliptic operator below,


 
∂2 ∂2
A = − g11 (x1 , x2 ) 2 + g22 (x1 , x2 ) 2 1 q×q
∂x1 ∂x2
∂ ∂
+ (A1 (x1 , x2 ))q×q + (A2 (x1 , x2 ))q×q + (A0 (x, x2 ))q×q ,
∂x1 ∂x2
(11.21)

where {g11 (x), g22 (x); x ∈ R2 } are positive definite functions in Cc∞ (R2 )q×q ,
we can evaluate exactly the Seeley relationship (Eq. (11.20)).
August 6, 2008 15:46 9in x 6in B-640 ch11

316 Lecture Notes in Applied Differential Equations of Mathematical Physics

Now, it is a tedious evaluation to see that (important exercise left to


our readers) we have the result

a2 (x, ξ, λ) = (g11 (x)ξ12 + g22 (x)ξ22 − λ11), (11.22a)


a1 (x, ξ, λ) = −iA1 (x)ξ1 − iA2 (x)ξ2 , (11.22b)
a0 (x, ξ, λ) = −A0 (x), (11.22c)
C−2 (x, ξ, λ) = (g11 (x)ξ12 + g22 (x)ξ22 − λ11)−1 , (11.22d)
2
c−3 (x, ξ, λ) = i(A1 (x)ξ1 + A2 (x)ξ2 )(C−2 (x, ξ, λ))
   
∂ 2 ∂
− 2ig11 (x)ξ1 g11 (x)ξ1 + g22 (x) ξ22
∂x1 ∂x1
× (C−2 (x, ξ, λ))3
   
∂ ∂
− 2ig22 ξ2 g11 (x) (ξ1 )2 + g22 (x) (ξ2 )2
∂x2 ∂x2
× (C−2 (x, ξ, λ))3 . (11.22e)

We now analyze the symbolic–operational Seeley expansion for the evo-


lution semigroup Eq. (11.3)

Tr(Cc∞ (R2 ))q×q [exp(−tA)]


∞  2
1
= d2 xσ(exp(−tA))(x, ξ)
j=0
2π R2

∞  2 −∞
1 1
= d(is)eist d2 xd3 xC−2−j (x, ξ, −is)
j=0
2π 2πi +∞ R2 ×R2


 +∞   
1 1 2 2 is −is
= d ξ d x e C−2−j x, ξ, ds
j=0
t (2π)3 R2 R2 −∞ t
 
∞
1 1 +∞
2+ξ

=− d2 xd2 ξ eis (t)( 2 ) C−2−j (x, t1/2 ξ, −is)
 t (2π)3 R2 ×R2 −∞ 
j=0


 +∞ 
(ξ−2)
= t 2 (2π)3 d2 ξ d2 x dsC−2−j (x, ξ, −is) . (11.23)
j=0 R2 R2 −∞
August 6, 2008 15:46 9in x 6in B-640 ch11

The Atiyah–Singer Index Theorem 317

After substituting Eq. (11.22) (together with the term C−4 (x, ξ, λ) not
written here), we have the Seeley short-time expansion for the Heat–Kernel
evolution operator associated with our Laplacean A given by Eq. (11.21):

TrCc∞ (R2 )q×q [exp(−tA)]


 
t→0+ 1 2 √
∼ d x ( g11 g22 ) (x) 1 q×q
4πt R2
  
1 2 √ 1
+ d x g11 g22 − R(g) 1 q×q
4π R2 6
   
1 1 ∂ √ ∂ √
− ×√ ( g11 g22 · A1 ) + ( g11 g22 · A1 ) (x)
2 g11 g22 ∂x1 ∂x1
1 
− g̃11 (A1 )2 + g̃33 (A2 )2 + A0 (x) + O(t). (11.24)
4
Here R(g) is the curvature scalar associated with the metric tensor ds2 =
g11 (dx1 )2 + g22 (dx2 )2 . The inverse metric is denoted by {g̃11 , g̃22 }.
Let us give a proof of the famous Atiyah–Singer index theorem in the
context of the heat kernel PDE’s techniques.
Let us thus consider complex coordinates in a given (bounded) open
subset W ⊂ R2 , supposed to be a chart of a given Riemann surface M:
∂ ∂ ∂
z = x1 + ix2 ; = −i ,
∂z ∂x1 ∂x2
(11.25)
∂ ∂ ∂
z̄ = x1 − ix2 ; = +i .
∂ z̄ ∂x1 ∂x2

For each integer j, let us introduce Hilbert spaces H1 and H̄j defined as5


 vectorial set of all complex valued functions which under the


 action of a complex coordinate transformation z = z(w) of W ,
Hj =

 has the tensorial like transformation law



f (z, z̄) = (∂w/∂z)−j f˜(w, w̄).
(11.26)

We now introduce a Hilbertian structure in Hj by the following inner


product:

(g, f )Hj = dzdz̄(ρ(z, z̄)j+1 f (z, z̄)g(z, z̄), (11.27)


R2
August 6, 2008 15:46 9in x 6in B-640 ch11

318 Lecture Notes in Applied Differential Equations of Mathematical Physics

where ρ(z, z̄) denotes a real continuous function in W and of the prove-
nance of a conformal metric structure in a given Riemann surface with local
chart W

ds2 = ρ(z, z̄)dz ∧ dz̄. (11.28)

The Hilbert space H̄j is defined in an analogous way to the definition


Eq. (11.26), but with the dual tensor transformation law

f (z, z̄) = ((∂w/∂z))−j f˜(w, w̄). (11.29)

As an exercise for our readers we have that the above inner products are
invariant under the action of the conformal (complex) coordinate transfor-
mations of the open set W .
Let us now introduce the following self-adjoint operators in L2 (W̄ ):

Lj : Hj −→ H̄−(j+1)
(11.30)
f −→ ρj ∂z̄ f

and its adjoint operator

L∗j : H̄−(j+1) −→ H( j
(11.31)
f −→ ρ−(j+1) ∂z f.

We consider the further positive definite (inversible) self-adjoint oper-


ators

Lj : L∗j Lj : Hj −→ Hj , (11.32a)
L∗j : Lj L+
j : H̄−(j+1) −→ H̄−(j+1) . (11.32b)

Note the explicit expression:

Lj L∗j = (−ρ−1 ∂z̄ ∂z ) + ((j + 1)ρ−1 (∂z̄ (gρ∂z ))), (11.33a)


L∗j Lj = (−ρ −1
∂z ∂z̄ ) − (jρ −1
(∂z gρ∂z̄ )). (11.33b)

By using now the Seeley asymptotic expansion in Cc∞ (R2 ), we have the
following t → 0+ expansions for the heat kernels associated with the Dirac
operator Eq. (11.33)5 :

lim TrCc∞ (R2 )2×2 (exp(−tLj L∗j ))


t→0+
August 6, 2008 15:46 9in x 6in B-640 ch11

The Atiyah–Singer Index Theorem 319

  
dz ∧ dz̄ ρ(z, z̄) (1 + 3j)
= − (∆gρ)(z, z̄) + O(t) (11.34a)
W 2i 2πt 12π
lim TrCc∞ (R2 )2×2 (exp(−tL∗j Lj ))
t→0+
 
dz ∧ dz̄ ρ(z, z̄) (2 + 3j)
= + (∆gρ)(z, z̄) + O(t). (11.34b)
W 2i 2πt 12π

Now the famous topological heat kernel index can be obtained easily
through the Atiyah–Singer definition:

index(Lj ) = lim [TrCc∞ (R2 )2×2 [exp(−tLj L∗j ) − exp(−tL∗j Lj )]]


j→0+
 
(1 + 2j) dz ∧ dz̄ 1
= ρ(z, z̄) + ∆gρ(z, z̄)
4π W 2πi ρ(z, z̄)
 
1 + 2j
= χ(M). (11.35)

Here, the curvature of the Riemann surface M with metric ds2 = ρ(z, z̄)dz
∧ dz̄ is given by
 
1
R(z, z̄) = ∆gρ (z, z̄), (11.36)
ρ

which is related to the topological invariant of the Euler–Poincaré genus of


M through the Gauss theorem

ρ(z, z̄)R(z, z̄) = 2π(2 − 2g). (11.37)


M

References

1. N. Berline, E. Getzler and M. Vergne, Heat Kernels and Dirac Operators,


Vol. 298 (Springer-Verlag); M. F. Atiyah, R. Bott and V. K. Patodi, On
the heat equation and the index theorem, Invent. Math. 19, 279–330 (1973);
Errata, Invent. Math. 28, 277–280 (1975); M. F. Atiyah and I. M. Singer,
Dirac operators coupled to vector potentials, Proc. Mat. Acad. Sci. USA 81,
2597–2600 (1984).
2. P. B. Gilkey, Invariance, the Heat Equation and the Atiayah-Singer Index
Theorem (Publish or Perish, Washington, 1984).
3. J. M. Bismut, The infinitesimal Lefschetz formulas: a heat equation proof,
J. Funct. Anal. 62, 435–457 (1985).
4. A. S. Schwartz, Commun. Math. Phys. 64, 233 (1979).
August 6, 2008 15:46 9in x 6in B-640 ch11

320 Lecture Notes in Applied Differential Equations of Mathematical Physics

5. B. Durhuus, P. Olesen and J. L. Petersen, Polyakov’s quantized string with


boundary terms, Nucl. Phys. B 198, 157–188 (1982).
6. R. S. Hamilton, The Ricci flux on surfaces, Contemp. Math. 71, 237–262
(1988); B. Osgood, R. Phillips and P. Sarnak, Extremals of determinants of
Laplacians, J. Funct. Anal. 80, 148–211 (1988); L. C. L. Botelho and M. A. R.
Monteiro, Fermionic determinant for massless QCD, Phys. Rev. 30D, 2242–
2243 (1984); A. Kokotov and D. Korotkin, Normalized-Ricci flow on Riemann
surfaces and determinant of Laplacian, Lett. Math. Phys. 71, 241–242 (2005).
7. L. C. L. Botelho, Path integral Bosonization for the thirring model on a
Riemann surace, Eur. Phys. Lett. 11, 313–318 (1990); L. Alvarez-Guamé, G.
Moore and G. Vafa, Commun. Math. Phys. 106, 1–35 (1980); M. F. Atiyah
and I. M. Singer, Dirac operators coupled to vector potentials, Proc. Math.
Acad. Sci., USA 81, 2597–2600 (1984).
8. V. N. Romanov and A. S. Schwarz, Anomalies and elliptic operators, Teo-
reticheskaya, Matematicheskaya Fizika 41(2), 190–204 (1979); M. Nakahara,
Geometry, Topology and Physics, Graduate Student Series in Physics (IOP
Publishing, London, 1990); L. C. L. Botelho, Covariant functional diffusion
equation for Polyakov’s Bosonic String, Phys. Rev. D 40, 660–665 (1989);
D’Hoker and D. Phong, The geometry of string perturbation theory, Rev.
Mod. Phys. 60, 917 (1988).

Appendix A. Normalized Ricci Fluxes in Closed Riemann


Surfaces and the Dirac Operator in the
Presence of an Abelian Gauge Connection

In this appendix we wish to expose a new result of ours on the Riemannian


geometry of Riemann surfaces M. Let us consider the nonlinear parabolic
PDE’s equation governing the flux dynamics of a given metric hµν (x) in
M, namely,


hµν = (R0 − R)hµν , (A.1)
∂t

where R is a scalar of curvature and R0 its mean value.


An important result in this subject of Ricci fluxes is the famous theorem
of Osgood–Phillips–Sarnak, which shows that in the class of all metric
structures with a fixed conformal structure, the determinant of the Laplace
operator takes its maximal value in the metric of constant curvature.6 Let
us present a generalization of such result for a Dirac operator in the presence
of an Abelian gauge connection (with a fixed spin structure).
August 6, 2008 15:46 9in x 6in B-640 ch11

The Atiyah–Singer Index Theorem 321

Theorem A.1. The determinant of the Dirac operator in the presence of


a gauge connection in a compact orientable Riemann manifold (equivalent
to a complex curve = Riemann surface) has a monotonic grown under the
Ricci flux.
Proof. Let us consider a Dirac operator with a spin structure (v i , ui ) in
the presence of a U (1) Abelian gauge connection which in the physicists
tensor notation reads as7
 
 a µ 1 ab
D(A, ĥ) = iγ êa ∂µ + Wµ,ab (ê)ε γ5 + Aµ . (A.2)
8

The matrices γ µ = êµa γa are the (Euclidean) Dirac matrices associated with
the metric Riemann surface structure ĥµν = êµa êνa . Note that ĥµν can always
be written in the canonical conformal form (ζ ∈ M)

1 µν
ĥµν (ζ) = h̃ (ζ), (A.3)
ρ(ζ)

where h̃µν is the element representative of ĥµν in the Teichmiller modulo


space of M. We also note that the Abelian gauge connection (the Hodge
Abelian connection) has the usual decomposition7

εµν
Aµ = − √ ∂µ φ + AH
µ (A.4)
h
with

∇µ (Aµ − AH
µ ) ≡ 0. (A.5)

It is useful to remark that the effects of the nontrivial topology of the genus
of Riemann surface M reflects itself in the Hodge harmonic term AH µ of the
U (1)-Connection through the Abelian differential forms (and its respective
complex conjugates)7 
 g 

AH
µ = 2h (p αµ + g βµ ) , (A.6a)
=1

αiµ = −Ω̄ik (Ω − Ω̄)−1 f


ks wµ + c.c, (A.6b)
βµi = (Ω − Ω̄)−1 j
ij wµ + c.c. (A.6c)
August 6, 2008 15:46 9in x 6in B-640 ch11

322 Lecture Notes in Applied Differential Equations of Mathematical Physics

The Riemann surface matrix period Ω is defined by the usual homo-


logical relationships (Abel integrals)

αi = δij ; β j = Ωij , (A.7)


aj bi

where {ai } and {bj } are the canonical homological cycles of M.


Let us now consider the  variation of the (functional) determinant of
the Dirac operator det1/2 (DD∗ ) = det(D) in relation to an infinitesimal
variation of the metric with a fixed conformal class8
  
∂ det(D(A, ĥ))
g 
∂t area(M) det(D(A = 0, h̃))

1 dz ∧ dz̄ zz̄
= [h̃ (∂z ϕ∂z̄ ϕ)(z, z̄)]
π M 2i
1 dz ∧ dz̄ zz̄
+ h̃ [(gρ)t (gρ)zz̄ ]. (A.8)
12π M 2i
Note that this metric variation is evaluated for the normalized Ricci flux
below

gρ ≡ [gρ]t = R0 − R, (A.9a)
∂t
2π(2 − g)
R0 = , (A.9b)
area(M)
1
R = − ∂zz̄ (gρ). (A.9c)
ρ
As a consequence we have the positivity of Eq. (A.8):
  
∂ det D(A, ĥ)
g  ≥ 0. (A.10)
∂t area(M) det(D(A = 0, h̃))
At this point we remark that Eq. (A.10) vanishes solely for all those metric
of constant curvature, which at the asymptotic limit t → ∞ leads us to the
usual result (R → R0 )
√ t→∞
(R0 − R)R · h dz ∧ dz̄ −→ 0,
M

since for t → ∞ all (smooth-C ∞ ) metrics on M are attracted to the


metric of constant curvature under the action of Ricci flux with a fixed area

( M ĥ dz∧dz̄
2i = area(M)).
August 6, 2008 15:47 9in x 6in B-640 index

Index

Algorithm, 146 Feynman, 99, 214, 275


Anomalous, 96 Field theory, 112, 129
Arzelá, 224 Fluid, 96
Ascoli, 224 Fourier, 29
Asymptotic completeness, 39, 48 Fredholm, 56
Aubin, 87, 213, 223 Frequency, 254, 274
Friedrichs, 18
Banach, 10, 20, 87 Fubbini, 229
Beltrami, 11 Functional, 81, 108, 179
Borel, 177
Bosonization, 187, 202 Galerking, 210, 216
Gärding, 64, 130, 236
Cameron, 218 Gauss, 4, 84, 217
Cauchy, 1, 25, 271 Gelfand, 223, 272
Chilov, 28 Generating function, 111
Classical mechanics, 231 Graham, 16
Conductivity, 76, 80 Green, 3
Continuity, 116, 228, 230 Green function, 3, 8, 29, 67, 275
Cook, 40
Haar, 157
Decay, 70, 80 Halmos, 43
Decomposition, 46, 118, 122 Hamell, 113, 146
Determinants, 150, 160 Hamiltonian, 231, 234, 240
Diffusion, 84, 94, 209, 216 Hilbert, 15
Dirac, 204 Hilbert space, 58, 62, 78, 84, 114, 255
Dirichlet, 1, 12 Hille, 3
Hyperbolic damping, 91
Einstein, 31
Elasticity, 70, 71, 72 Jauch, 40
Energy, 58, 62, 70, 231
Enss, 42 Kac, 221
Ergodic theorem, 231, 241, 257 Kakutani, 169

323
August 6, 2008 15:47 9in x 6in B-640 index

324 Lecture Notes in Applied Differential Equations of Mathematical Physics

Kato, 65, 99, 251 Probability, 81, 82, 256, 257


Kirchoff, 23 Prokhorov, 250
Kirchoff potentials, 23 Propagation, 84, 120
Klein Gordon operator, 73
Kolmogorov, 80 Radon, 28
Kuroda, 40 Rellich, 64, 99, 224
Renormalization, 114, 124, 127
Laplace, 1 Riemann, 10
Laplace Operator, 1 Riemann conjecture, 282
Lax, 18 Riesz, 18, 113, 185
Lebesgue, 181
Legendre, 2 Sazonov, 180
Leibnitz, 25 Scattering operators, 39, 68
Light Deflection, 31 Schimidt, 16, 181
Lion, 94, 224 Schwartz, 28, 83, 116, 185
Liouville, 104, 270 Seeley, 160
Lipschitz, 20 Shanon, 259
Loop Space, 193 Sobolev, 130, 134
Stochastic processes, 61, 95, 254, 255
Markov, 113 Stone, 45, 145
Measure, 82, 96, 98, 114, 169, 177, 178 Synge, 34
Mercer, 10
Milgram, 18 Topology, 109, 111, 147
Minlos, 114, 121 Trotter, 65
Möller, 37 Turbulence, 84, 131, 254
Motion, 240, 247
Wave propagation, 58, 84
Newmann, 1 Weak problem, 183
Newton, 3 Weierstrass, 5, 45
Non-linear, 73, 94, 209 Wide sense, 255
Nyquist, 259 Wiener, 83, 273

Path-integrals, 110, 121, 124, 130, 146 Yoshida, 63


Poincaré, 8, 65, 130
Poisson, 1, 131 Zorn, 70

You might also like