Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Theoretical Physics and Astrophysics
Theoretical Physics and Astrophysics
Theoretical Physics and Astrophysics
Ebook771 pages9 hours

Theoretical Physics and Astrophysics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

The aim of this book is to present, on the one hand various topics in theoretical physics in depth - especially topics related to electrodynamics - and on the other hand to show how these topics find applications in various aspects of astrophysics. The first text on theoretical physics and astrophysical applications, it covers many recent advances including those in X-ray, -ray and radio-astronomy, with comprehensive coverage of the literature
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483293189
Theoretical Physics and Astrophysics

Read more from V. L. Ginzburg

Related to Theoretical Physics and Astrophysics

Related ebooks

Physics For You

View More

Related articles

Reviews for Theoretical Physics and Astrophysics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Theoretical Physics and Astrophysics - V. L. Ginzburg

    THEORETICAL PHYSICS AND ASTROPHYSICS

    V.L. GINZBURG

    P. N. Lebedev Physical Institute, Academy of Sciences of the USSR, Moscow

    Table of Contents

    Cover image

    Title page

    OTHER TITLES OF INTEREST

    Copyright

    PREFACE TO THE ENGLISH EDITION

    PREFACE TO THE RUSSIAN EDITION

    Chapter 1: THE HAMILTONIAN APPROACH TO ELECTRODYNAMICS

    Publisher Summary

    Chapter 2: RADIATION REACTION

    Publisher Summary

    Chapter 3: UNIFORMLY ACCELERATED CHARGE

    Publisher Summary

    Chapter 4: RADIATION OF A MOVING PARTICLE

    Publisher Summary

    Chapter 5: SYNCHROTRON RADIATION

    Publisher Summary

    Chapter 6: ELECTRODYNAMICS OF A CONTINUOUS MEDIUM

    Publisher Summary

    Chapter 7: CHERENKOV EFFECT, DOPPLER EFFECT, TRANSITION RADIATION

    Publisher Summary

    Chapter 8: ON SUPERLUMINAL RADIATION SOURCES

    Publisher Summary

    Chapter 9: REABSORPTION AND RADIATIVE TRANSFER

    Publisher Summary

    Chapter 10: ELECTRODYNAMICS OF MEDIA WITH SPATIAL DISPERSION

    Publisher Summary

    Chapter 11: DIELECTRIC PERMITTIVITY AND WAVE PROPAGATION IN A PLASMA

    Publisher Summary

    Chapter 12: THE ENERGY-MOMENTUM TENSOR IN MACROSCOPIC ELECTRODYNAMICS

    Publisher Summary

    Chapter 13: FLUCTUATIONS AND VAN DER WAALS FORCES

    Publisher Summary

    Chapter 14: SCATTERING OF WAVES IN A MEDIUM

    Publisher Summary

    Chapter 15: COSMIC RAY ASTROPHYSICS

    Publisher Summary

    Chapter 16: X-RAY ASTRONOMY

    Publisher Summary

    Chapter 17: GAMMA ASTRONOMY

    Publisher Summary

    REFERENCES

    INDEX

    OTHER TITLES IN THE SERIES IN NATURAL PHILOSOPHY

    OTHER TITLES OF INTEREST

    Books*

    CLARK & STEPHENSON: The Historical Supernovae

    ELGAROY: Solar Noise Storms

    HEY: The Radio Universe, 2nd edition

    LANDAU & LIFSHITZ: Electrodynamics of Continuous Media (Course of Theoretical Physics Volume 8)

    MEADOWS: Stellar Evolution, 2nd edition

    PACHOLCZYK: Radio Galaxies (Radiation Transfer, Dynamics, Stability and Evolution of a Synchrotron Plasmon)

    REDDISH: Stellar Formation

    SAHADE & WOOD: Interacting Binary Stars

    Journals**

    Chinese Astronomy

    Vistas in Astronomy


    *Available under the Pergamon textbook inspection copy scheme

    **Free specimen copy available on request

    Copyright

    Copyright © 1979 Pergamon Press Ltd.

    All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or otherwise, without permission in writing from the publishers.

    First edition 1979

    British Library Cataloguing in Publication Data

    Lazarevich Theoretical physics and astrophysics. - (International series in natural philosophy; vol. 99).

    1. Astrophysics 2. Physics

    I. Title II. Series

    523.01 QB461 78-41009

    ISBN 0-08-023067-9(hardcover)

    ISBN 0-08-023066-0(flexicover)

    In order to make this volume available as economically and as rapidly as possible the author’s typescript has been reproduced in its original form. This method unfortunately has its typographical limitations but it is hoped that they in no way distract the reader.

    Printed and bound at William Clowes & Sons Limited Beccles and London

    PREFACE TO THE ENGLISH EDITION

    As any author I am glad to see my book appear in English, a language accessible to all physicists nowadays. However, it is hardly necessary to wirte a special preface to say this or to mention that in the English edition a number of errors have been corrected. The idea that such a preface might be necessary arose when I received from Professor ter Haar the preliminary list of references for the English edition and noticed the large number of references to Ginzburg. Of course, the same references also appear in the Russian edition, but there they do not appear together in one lot, and are also not so obvious in the text. The reason is that in the Russian edition, as is the usage with us, references are indicated by a number, without mentioning the name of the authors. As a result my name never appears in the text, while in the English translation one meets it very often! Such a situation is unusual for a book meant for students or, at any rate, one which is not a monograph, and it may cause some unpleasant observations. I want therefore to add the following to what I said in the preface to the Russian edition. I am, in general, not a teacher and I do not like to lecture, especially not when the same material must be presented more than once. At the same time, for the first time lectures one needs a large amount of preparation. In such a situation and if I have the possibility to choose the subject of the lectures completely freely I have given some lectures just once; the present book was written in preparing such a course. When choosing the material and wishing to share with my audience or my readers what appeared to me to be interesting and important I widely used my own papers and review articles. Of course, the result was a very one-sided picture. However, as there was no pretension of priority claims to that I had chosen the most interesting problems, there was hardly any reason for reproaching me of immodesty. A different question is whether the book is useful and deserving attention, that that is, of course, up to the reader to decide.

    In conclusion I use this possibility to express my warmest thanks to Professor ter Haar for undertaking the arduous task of translation and also for a number of helpful remarks.

    Moscow, April, 1978

    V.L. Ginzburg

    PREFACE TO THE RUSSIAN EDITION

    There are many textbooks of theoretical physics among which the many-volume work of Landau and Lifshitz is the best known and most outstanding. It is impossible, however, to deal with all problems in such a course. Moreover, even the problems which are considered can usually not be looked at from different points of view. At the same time, depending on their peculiar abilities, the nature of their training, natural inclinations, and so on, different people often prefer different approaches, arguments, examples, and proofs.

    A natural possibility to satisfy an existing need is, clearly, to publish different textbooks and, in particular, different supplementary textbooks which are devoted to separate problems, aspects, and methods rather than to a systematic exposition of a topic. Such supplementary texts differ in principle from the systematic ones in that the choice of material to a large extent is not predetermined. One can say the same about the style and nature of the exposition whereas in a systematic textbook one must impose very rigid restrictions with respect to conciseness, content of technical methods, unified notation, and so on. The present book is just such a supplementary text which is devoted to a few problems in theoretical physics and astrophysics. From the table of contents it is immediately clear, but we may also state it that, in general, we are dealing with problems which are in one way or other connected with electrodynamics.

    In order not to violate this trend, even though it was not rigorously expressed, we left outside the limits of the book a number of problems of the general theory of relativity and of statistical physics which, in the opinion of the author, also should be the subject of supplementary courses of a similar kind.

    The basis of the exposition was a lecture course for students in the physics and astrophysics departments of the Moscow Physico–Technical Institute. These lectures were not meant to replace a systematic course and had just the character of ‘capita selecta’ taking into account the interests of the department and not least the interests and capabilities of the author. Of course, we are not saying that the problems which the author at a particular moment in the past was occupied with are more important or more interesting than many other ones. It is simply the case that merely presenting ‘more or less moved by the spirit’ material with which he is well familiar, the author may perhaps hope to supplement existing texts and monographs without checking whether he is rewriting or to some extent duplicating them.

    As to the nature of the exposition one should note that we are dealing here not with lectures which are written out but with a special text specially prepared for these lectures, in which rather often we have also included material which is not very suitable and in fact was not used for the lectures themselves (that is, for the oral presentation). In this respect the book is in style close to a monograph or a review article which finds reflection also in the rather large number of references to the literature. As amongst those there is a large number of references to work by the author, we emphasize that this, like the choice of material, is completely unconnected with any pretentions, but caused by the tendency, already mentioned, to touch only upon very familiar problems which were dealt with in detail in the papers referred to; moreover, a whole number of such papers were used directly in the text.

    We note, finally, that the book is definitely not intended for people with a mathematical inclination — such as ‘pure’ theoretical physicists. The exceptionally large role played by mathematics in theoretical physics is completely unquestionable and natural, but aiming at mathematical generality and rigour is by far not always justified — one must pay for this. It is generally known, in particular, that most new physical results have been obtained by relatively simple means while the ‘mathematization’ occurred only in the later stages. At any rate, physics, and not mathematics is the main point of theoretical physics. An exposition of theoretical problem with a ‘general physics’ bias is at least as permissible as the nowadays more widely propagated tendency to mathematical perfection.

    One may hope that this book will turn out to be useful for graduate students and also for post-doctoral and research workers.

    In conclusion I use this opportunity to thank all who read the manuscript or parts of it and whose remarks contributed to improvements in the text.

    Moscow, July, 1974

    V.L. Ginzburg

    Chapter I

    THE HAMILTONIAN APPROACH TO ELECTRODYNAMICS

    Publisher Summary

    This chapter discusses the Hamiltonian approach to electrodynamics for the interpretation of a whole range of electrodynamical problems. The transition from classical to quantum mechanical electrodynamics in the Hamiltonian framework is completely analogous to the transition from classical, Newtonian mechanics to nonrelativistic quantum mechanics. The field of a uniformly moving charge is not at all necessarily a stationary one. The charge can already have moved for some time uniformly; however, the field entrained by it can still differ from the stationary field—which exists when the motion with a constant velocity has been going on for a sufficiently long time. For an electron that moves uniformly for t ≥ 0, there exists the difference between a free radiation field and the transverse entrained field. The actual construction of quantum electrodynamics is completely free from any assumption about the absence of charges and is in no way connected with any identification of a quantized transverse field with a free radiation field, that is, a collection of photons.

    The Hamiltonian method in classical electrodynamics in vacuo.

    Quantization. Photons and pseudophotons.

    Does a uniform moving electron radiate?

    We shall in what follows widely apply the so-called Hamiltonian method for the interpretation of a whole range of electrodynamical problems. When we use this method electrodynamics is formulated in a way which is strongly reminiscent of mechanics. The transition from classical to quantum mechanical electrodynamics is thus in the Hamiltonian framework completely analogous to the transition from classical, Newtonian mechanics to non-relativistic quantum mechanics. Nowadays much more sophisticated methods are predominant in quantum electrodynamics and in general in quantum field theory and there are strong arguments for using them. However, the use of the Hamiltonian method is still completely justified for the elucidation of a large number of physical aspects; this is, for instance, also done by Heitler (1947) in his book. Moreover, we shall in what follows apply the Hamiltonian method mainly to classical electrodynamics, both in vacuo and in a medium.

    Before introducing the Hamiltonian method we shall give the main equations and relations and we shall do that in considerable detail for future convenience.

    The usual form of the Maxwell equations in vacuo is:

    (1.1)

    Here H is the magnetic field strength, E the electrical field strength, ρ the charge density, v the velocity of the charges, and c the velocity of light in vacuo. We assume for the sake of simplicity that there is a single point charge e, at position ri(t), in the electromagnetic field. In that casethe charge density is given by a δ-function

    (1.2)

    It is well known that Eqs. (1.1) can be reduced to the equations for the electromagnetic potentials A and ϕ which are connected with the fields E and H sby the relations

    (1.3)

    The third and fourth of Eqs. (1.1) are automatically satisfied by virtue of (1.3), as can be verified by substitution.

    From the first and second of Eqs. (1.1) we get, using (1.3) and the identity

    (1.4)

    equations for the potentials A and ϕ :

    (1.5)

    The set of Eqs. (1.5) determines the potentials A and ϕ. The fields E and H can be found using Eqs. (1.3)

    It is well known that the vector potential A and the scalar potential ϕ are not uniquely determined. Indeed, we can change to new potentials:

    (1.6)

    where χ is an arbitrary function of the coordinates and the time. This is called a gauge transformation. One can easily show that the fields E and H do not change under a gauge transformation. They can be expressed in terms of A′ and ϕ′ just as well as in terms of A and ϕ; one can verify this by substituting (1.6) into (1.3).

    The fact that the definition of the potentials is not unique enables us to impose upon A and ϕ an additional condition. This condition can be chosen in such a way that the form of Eqs. (1.5)becomes as simple as possible.

    For instance, we can impose as such a condition the relation:

    (1.7)

    This is a relativistically invariant condition which is called the Lorentz condition, and the resulting gauge is called the Lorentz gauge. It can be written in the form

    (1.7a)

    where we have assumed summation over repeated indexes, as will be done everywhere in what follows. We use here and elsewhere in this book (apart from Chapter 12) the notation of Landau and Lifshitz (1975). We refer to that text for the definition of four-vectors, for the difference between covariant and components and for the summation convention.

    One sees easily that if condition (1.7) is satisfied, the Maxwell equations take the following form:

    (1.8)

    One should not think that the condition (1.7) and the set of Eqs. (1.8) determine A χ = 0. The fields E and H remain invariant under the transformation.

    Splitting the field into longitudinal and transverse components is important, especially in the Hamiltonian framework. We split the vectors E and H into components

    (1.9)

    where div Etr = 0 and, by virtue of (1.1), div Htr = div H= 0.

    We demand that the vector potential A describe only the transverse field; this means that we impose on it the condition

    (1.10)

    instead of the additional condition (1.7). Sometimes the potential which satisfies the condition (1.10) is denoted by Atr.

    If condition (1.10) is satisfied, the Eqs. (1.5) for A and ϕ take the form

    (1.11)

    (1.12)

    We see that we have obtained the ‘static’ Poisson equation for the potential ϕ.

    If ρ is the charge density (1.2) of a point charge, the solution is the well known one:

    (1.13)

    where ri(t) is the position of the charge at time t. The vector potential A now describes only the transverse field. The gauge (1.10) is called the Coulomb gauge† The potentials A and ϕ are here determined apart from a gauge function χ(r,t) which satisfies the condition ∇²χ = 0.

    We now evaluate the energy of the electromagnetic field,

    (1.14)

    We substitute here the expressions for the fields E and H in the form (1.9); it is clear that in the case of the Coulomb gauge (1.10) we have

    (1.15)

    Substituting (1.9) and (1.15) into (1.14) we get

    One shows easily that for a closed system, when the field ‘at infinity’ vanishes, the last integral equals zero. The total energy of the electromagnetic field is thus the sum of the energies of the transverse and of the longitudinal fields.

    If there are several point charges in the field, the energy of the longitudinal field is simply the energy of the Coulomb interaction between the charges, that is,

    (1.16)

    The self energy of the point charges is infinite and is here, of course, neglected. It is important to note that the longitudinal part of the electromagnetic field is not quantized. Only the transverse field is quantized (see Heitler 1947 and the discussion later in this chapter).

    e²/ro. The electrostatic (classical) electron radius, defined by the relation re = e²/mc², where e and m are the observed electron charge and mass, is equal to re = 2.8 × 10−13 cm. We shall not be concerned here with the problems connected with the electromagnetic mass of the electron – or of other particles – or whether it is a point charge, and so on.

    Proceeding now along the path which will lead to the Hamiltonian formalism for electrodynamics we expand the vector potential A of the transverse electromagnetic field in a Fourier series,

    (1.17)

    is a normalization factor. The polarization vector e; for the sake of simplicity we assume here and henceforth that the vectors eλ are real. In order that we may apply the expansion (1.17) we must imagine the electromagnetic field to be enclosed in a large, cubic ‘box’. One can verify that the dimensions of this ‘box’ do not enter in any of the expressions for physically observable quantities. We shall therefore everywhere put the size of this ‘box’ to be equal to unity:

    The vector potential A is a real quantity; it therefore follows from the expansion . As the field is transverse,(eλ·kλ) = 0, that is, the polarization vector of the harmonic with number λ of the potential is at right angles to the wavevector kλ of this harmonic. To each direction of kλ there correspond two vectors eλ. We should therefore introduce one more index which can take two values, or, in other words, distinguish the vectors eλ1 and eλ2. We shall not do this in what follows, in order to simplify the notation, but we shall sum over the polarizations, when necessary, in the final expressions; we shall assume in those cases that (eλ1 · eλ2) = 0.

    We can realize also a different expansion of the vector potential, namely,

    (1.18)

    where the index i can take on only the values 1 and 2, while

    (1.19)

    One sees easily that the functions Aλ1 and Aλ2 in which we have expanded in (1.18), are mutually orthogonal, that is,

    (1.20)

    where the integral is over the volume of the ‘box’.

    We assume the field to be enclosed in a ‘box’ with specularly reflecting walls; the components of the wavevector kλ must thus be integral multiples of the quantity 2π/L, where L is the linear size of the ‘box’, that is

    nx, ny, nz are here integers; the summation in (1.18) is over a hemisphere of the directions of kλ. Apparently, what we have just said contradicts the earlier statement that the size L of the ‘box’ is unimportant and can be put equal to 1. However, one can easily check that there is here no contradiction: for sufficiently large values of L this quantity does not occur in the final results.

    It is clear from (1.18) that the transverse electromagnetic field is completely determined, if we give the set of quantities qλi (t). The quantities qλi(t) form a denumerable infinite set. The field is thus through (1.8) represented as a system of an infinite, though denumerable, number of degrees of freedom.

    Let us see how one can express the energy of the electromagnetic field in terms of the quantities qλi(t) which we can properly call the field coordinates. We are interested in the energy

    (1.21)

    If A is given in the form (1.18), we can use Eqs. (1.3) and (1.15) to determine the fields Etr and H; we can square them and substitute them into the integral (1.21) We then get

    (1.22)

    We have introduced here the notation

    (1.23)

    where the dot indicates differentiation with respect to the time. In deriving Eq. (1.22) we used the orthogonality condition (1.20)

    Each term in the sum (1.22) is the energy of a classical oscillator of frequency ωλ. Therefore, Eq. (1.22) is the sum of the energies of separate oscillators, which we call the field oscillators.

    If all qλi(t) in (1.18) are known, we can determine the energy of the transverse electromagnetic field. The problem is thus reduced to the determination of the qλi. (t).

    To find the equations for the qλi . (t) we substitute the expansion (1.18) into Eq. (1.12) for the transverse vector potential. Multiplying both sides of the resulting equation by Aλi and integrating over the volume of the ‘box’ we get the following equations for the qλi(t):

    (1.24)

    This is the equation for an oscillator when there is an exciting force present; for i = 1 we choose cos(kλ·r) and for i = 2: sin(kλ·r).

    We derived Eq. (1.24) in the assumption that there was a single point charge (electron with charge e; see (1.2)) moving with a velocity v(t) present in the field. The generalization to the case of several charges is obvious.

    All relations considered here can be written completely analogously to the Hamiltonian equations of classical mechanics:

    (1.25)

    (p,q) is the Hamiltonian function of the mechanical system, and q and p are, respectively, the generalized coordinates and momenta.

    (pλi, qλi) that we can obtain equations of motion such as (1.25) from it. It is clear that Eqs. (1.24) for the qλi. for the case of a free field – a field without charges – that is, the equations

    (1.26)

    can be written in Hamiltonian form, if

    (1.27)

    tr is the energy of the transverse electromagnetic field (1.22)

    Indeed, from (1.25) and (1.27) we get

    (1.28)

    which is the same as (1.26)

    The series (see (1.23)) and the field thus changes as cox{ωλ[t−(sλ·r)/c]} or sin{ω)λ[t−(sλ · r)/c]}, where sλ= k. When there are no charges the field thus consists of plane waves moving with the velocity of light; this result is, of course, already clear from the original equations.

    If the classical Hamiltonian of the electromagnetic field without charges in vacuo is the field energy, when there are charges present we must now take the interaction of them with the field into account. It is well known that in the non–relativistic case the energy of a charge in the field has the form (see, e.g., ter Haar, 1971)

    (1.29)

    The total Hamiltonian for the system of the electromagnetic field + charged particles is thus equal to the sum of expressions (1.27) and (1.29)

    (1.30)

    eis the sum of expressions such as (1.29); we shall sometimes in what follows drop the index i of ri).

    Using this Hamiltonian (in actual fact we have spoken so far all the time about a classical Hamiltonian) and (1.25) we get the following set of equations

    This set can be reduced to Eq. (1.24) Indeed, as the quantity r =/∂p = [p−(e Awith respect to r we find

    (1.31)

    We thus get from Eq. (1.30) for the Hamiltonian both the equation of motion for the field oscillators and the equation of motion for the charged particles.

    We have here considered everything in the non-relativistic approximationparticles, where one uses the Dirac equation which does not have such an obvious classical analogue, this similarity is lost to a certain degree.

    The transition from classical electrodynamics in a Hamiltonian form to quantum electrodynamics is accomplished in exactly the same way as the change from classical non–relativistic mechanics to quantum mechanics. In fact, the classical Hamiltonian for a particle with momentum p and position r,

    (1.32)

    is replaced by the Hamiltonian operator, or simply, the Hamiltonian

    (1.33)

    is the operator of the particle momentum which satisfies the commutation relations (r≡{xj}, j = 1, 2, 3; note that there is no summation over j here)

    (1.34)

    and is given by the expression

    (1.35)

    If the particle is in an electromagnetic field, p in (1.32) is replaced by p −(eAin (1.33) by

    The state of the system is determined by the wavefunction ψ(r,t); the change of the wavefunction with time is described by the Schrödinger equation

    (1.36)

    The wavefunctions of stationary states have the form

    (1.37)

    where the ψn(r) are independent of t (n is the number of the stationary state, or its quantum number). The square of the modulus of the wavefunction of a stationary state, that is, the probability that a particle will be observed in a given point in space, is independent of the time. Substituting) we have

    (1.38)

    Let us consider a one-dimensional harmonic oscillator with unit mass. It is well known that the Hamiltonian of such an oscillator has the form

    (1.39)

    is the coordinate and ω0 the angular frequency of the oscillator. The energy of the nth stationary state is equal to

    (1.39a)

    and the wavefunction has the form

    (1.39b)

    Hn (x) is a Hermite polynomial, and Cn a normalization factor. In particular,

    (1.39c)

    The matrix elements of the coordinate, qnn′ and of the momentum, pnn′, corresponding to transitions from a state with quantum number n to a state with quantum number n vanish unless n′ ≠ n ± 1, and when n′ = n ± 1 equal

    (1.40)

    We mentioned earlier that the transition from classical to quantum electrodynamics is achieved in the same way as in mechanics. The Hamiltonian of a system consisting of a particle and the field,

    (1.41)

    e is the Hamiltonian for a particle in the field, which we considered earlier (see (1.29)), is replaced by the Hamiltonian operator

    (1.42)

    where

    (1.43)

    are, as in mechanics, equal to

    (1.44)

    and satisfy the commutation relations

    (1.45)

    ; can be neglected, the wavefunction ψ(q,t) which describes the state of the field (q stands for the whole set of field coordinates qλj) will satisfy the equation

    (1.46)

    The wavefunction ψn(q) of a stationary state then satisfies the equation

    (1.47)

    One can easily check that En has the following form

    (1.48)

    , which is constant, does not occur in the final result. Secondly, the transition from the classical to the quantummechanical equations is not unique. One can find a method of quantizing the field equations in which this term is absent. Indeed, we shall start from the classical Hamiltonian for a single oscillator in the form

    , we find

    this result comes about as the operators p and q do not commute. Thus, if we apply this to the field, we are led to the expression

    (1.49)

    for the energy of a stationary state. The wavefunction of the field has the form of a product of the wavefunctions of the separate oscillators, that is,

    (1.50)

    It is clear from . Such pseudo-photons, which are also called virtual photons, occur in the intermediate states in perturbation theory calculations(vide infraωλ and their momentum pkwhich is valid for photons with a given momentum. We shall see below that for the transverse field carried along with a moving charge we have ω = (k.v) where v ω they refer to pseudo–photons and one sometimes says that they form the ‘coat’ of the moving charge.

    We emphasize that we do not at all wish to insist on the advisability of introducing the pseudo–photon concept, although we use this term† The importance for us lies in the explication of the fact that the transverse field is in general not a collection of photons. This will be discussed below.

    The pseudo-photons which occur in ω/c) (k/k) even for a radiation field. This is connected with the fact that above we used an expansion in standing waves (see (1.18) and (1.19)). Standing waves are not eigenfunctions of the momentum operator and their quantization leads to photons (we are thinking now of a pure, free, radiation field) with a vanishing momentum††

    ω/c) (k/k) we write the vector potential as a sum of travelling waves:

    (1.52)

    where

    (1.53)

    while the summation in (1.52) ism in contrast to (1.17), over half-a-sphere, that is, over half of all kλ-directions.

    For the Hamiltonian of the transverse part of the field whave then

    (1.54)

    (see also (1.26) and (1.28)) and we have

    (1.55)

    are not canonically conjugate variables, as the equations of motion in those variables do not have the form

    (1.56)

    We therefore introduce new variables:

    (1.57)

    We then have

    (1.58)

    Quantizing we get clearly Eq. (1.49) for the energy, and evaluating the momentum of the field we find

    (1.59)

    that is, the momentum of a single photon is, indeed, equal to

    (1.60)

    The field and the quanta corresponding to it are characterized not only by their energy and momentum, but also by their angular momentum. The angular momentum of the electromagnetic field is classically defined as follows:

    (1.61)

    An unbounded plane wave does not have a non-vanishing angular momentum along the direction of the wavevector kλ, as the Poynting vector S = c[EH]/4π is in the direction of kλ in such a wave. However, for a wave in a cylindrical waveguide with perfectly conducting walls, for instance, the value of Memz ≡ Mz, where the z-axis is along the axis of the waveguide, may be different from zero. To be concrete, for monochromatic circularly polarized waves in a waveguide we have (see Heitler, 1947 and the references cited there)

    (1.62)

    ωλ nλ and , that is, the spin of the photons is equal to unity. It is important to consider the problem of the angular momentum of the electromagnetic field in general and of the radiation field in particular when one expands the field in spherical waves (such waves are emitted by electrical and magnetic multipoles, including dipoles). We refer to the books by Heitler (1947) and by Berestetskii, Lifshitz and Pitaevskii (1971) for a detailed discussion of the angular momentum of the radiation field.

    Let us now once again consider the complete system consisting of field and charge. In the non–relativistic approximation we have for the Hamiltonian of the system:

    (1.63)

    and the equation for the wavefunction takes its usual form:

    (1.64)

    In order to be able to apply perturbation theory for solving this problem we write the Hamiltonian in the following form

    (1.65)

    is considered to be a perturbation.

    , as it is just the former expression which is Hermitean. However, in the case of a transverse field, when div A = 0, the terms with (pA) and (Ap) are equal to one another.

    The effects which are connected with the interaction of light with electrons are proportional to the ‘electromagnetic interaction constant’ which is the so-called ‘fine structure constant’

    (1.66)

    1, the interaction of an electron with the electromagnetic field is weak in a well understood sense†. We shall therefore assume that the wavefunction of a stationary state must differ little from the solution of the equation

    (1.67)

    which can be found, at least, in a number of cases. In particular, when there is no external electromagnetic field

    (1.68)

    That the difference between the exact wavefunction and the solution of Eq. (1.67) is small is, for instance, clear from considering electrons in excited levels in the hydrogen atom. The lifetime of an electron in an excited level is of the order of 10−9s and the time for completing one orbit is of the order of 10−15 s. If we use the uncertainty relation for the energy,

    (1.69)

    c is much less than unity. If such a constant were of the order of unity – as is the case for the interaction between a nucleon and the meson field – the width of a level would be of the order of the distance between the levels themselves and it would be totally impossible to speak, in general, of a ‘quasi–stationary’ motion†.

    As the functions ψn0 from (1.67) form a complete set we can write Ψ in Eq. (1.64) in the form

    (1.70)

    Substituting , integrating over the whole of space, and using the orthonomality of the functions ψn0 we get

    (1.71)

    Let us assume that at t = 0 we have bk = 1 and bn≠k = 0. If we assume that the bn with n ≠ k are small at all times and drop higher–order terms, we then have

    (1.72)

    whence we easily get

    (1.73)

    If in the first order of perturbation theory |bn(t)|² = 0, we can use a similar procedure to find the next approximation. For instance, the matrix element is in second approximation given by

    (1.74)

    Equation (1.73) determines the probability for a transition to only a single final state with energy En0. We are usually interested in the transition to any of all possible states, that is, in the integral

    (1.75)

    In this equation ρ(En0)dEn0 is the number of final states – which are assumed to be ‘densely’ distributed – in the energy range from En0 to En0 +dEn0. As t tends to infinity, the integral (1.75) equals (see Eq. (1.84) or for details, Heitler’s book (1947)),

    (1.76)

    and the probability for a transition per unit time is thus given by the formula

    (1.77)

    Note that the transition occurs only if there are states En0 which are arbitrarily close to Ek0; this was reflected in one must use Eqs. (1.65), and also (1.18), (1.19), and (1.40) and understand by q the operators qλi (for details see Heitler’s book (1947)).

    Using perturbation theory in such a simple or in a somewhat more complicated form enables us to find the answers to a whole set of problems in radiation theory (Heitler, 1947; Berestetskii, Lifshitz and Pitaevskii, 1971). However, a wider application of perturbation theory encounters considerable difficulties which is formally reflected in the appearance of divergent (infinite) expressions. The appearance of divergent expressions is connected with the assumption that the electron is a point particle, with the fact that the field has an infinite number of degrees of freedom, and so on. Some of those difficulties are not caused by the quantization and are classical in nature. It is sufficient to remember that the electrostatic energy of a point charge is infinite. Even in classical electrodynamics one had learned to avoid such difficulties. In particular, one used for this the mass ‘renormalization’ method†. In quantum electrodynamics one also renormalizes the charge of the particle and the situation altogether becomes more complicated. The study of the corresponding field of problems was for a long time in the centre of the attention of theoretical physics. As a result a great deal of progress was made and the infinities in quantum electrodynamics are practically made ‘harmless’. A framework was developed in which one is able to find the answers to problems which may arise and, in particular, in which one can take extremely minute radiative effects into account (Heitler, 1947; Berestetskii, Lifshitz and Pitaevskii, 1971).

    The present text is not concerned with any of these problems, although the material discussed a moment ago may, we hope, be useful for understanding the physical basis of quantum electrodynamics. It was only important for us in the plan for our further exposition to formulate the Hamiltonian method in classical electrodynamics and to get acquainted with the most elementary aspects of quantum electrodynamics.

    It is interesting that the Hamiltonian method has in the past hardly ever been applied in classical electrodynamics; the method became popular only when one changed to quantum electrodynamics. However, as so often happens, there was a ‘feedback’ afterwards. To be precise, it became clear that the Hamiltonian method is very convenient also for a number of classical problems, especially when there is a medium present (see Chapters 6 and 7). Recently, when many problems turned out to be already solved, when there appeared new and more complicated problems and, when moreover, a number of powerful mathematical methods, such as the diagram technique or the Green function method, were developed and started to be widely applied, the Hamiltonian method dropped out of sight both in the quantum and in the classical radiation theory. We are, however, convinced that the Hamiltonian method nevertheless retains the advantage of clarity, simplicity, and rather large universality which makes its exposition and use very expedient, at least, for pedagogical purposes.

    As an illustration we use the Hamiltonian formalism to discuss the problem of the radiation by an oscillator – a harmonically oscillating charge. To determine the field we must find the quantities qλi in the expansion (1.18), while the equations of motion for these quantities have the form (1.24)

    (1.24)

    where r = r(t) is the position vector of the emitting charge e; in the case of an oscillator we have

    (1.78)

    The argument (kλ·r(t)) in (1.24) is small, if

    (1.79)

    where λ0 is the wavelength of the emitted radiation. Let us accept condition c, while λ0 = 2πc/ω0. In that case (kλ · r) in (1.24) is much less than unity and we are justified in putting cos(kλ · r) = 1, sin(kλ · r) = 0. Therefore qλ2 = 0 and for qλ1 we get the equation

    (1.80)

    , when t = 0 has the form

    (1.81)

    Having obtained the qλ1 we have thereby completely determined the electromagnetic field and we can now evaluate all other quantities. Let us, for instance, find the energy emitted by the charge (oscillator) per unit time. To do this we must clearly use tr/dt. That will just be the energy emitted by the oscillator per unit time.

    tr we get the expression

    (1.82)

    tr with time; the other terms which are omitted in (1.82) do not contribute to the expression for the energy emitted by the oscillator per unit time for large t (we assume here that t is large).

    For the evaluation of the sum in (1.82) it is convenient to change from a summation to an integration. We must then first multiply expression (1.82) by the number of field oscillators with frequencies between ω and ω + dω; this number is equal to

    (1.83a)

    where dΩ is an element of solid angle.

    The transition from a sum to an integration is thus reduced to the substitution

    (1.83b)

    is connected with the fact that we integrate over all directions, rather than over a hemisphere of k-directions.

    We can easily integrate expressions (1.82) over ω if we use the following relation which is valid for large values of t :

    (1.84)

    As a result of all these simple calculations we get an expression for the energy emitted by the oscillator per unit time into a solid angle dΩ:

    (1.85)

    where θ is the angle between the direction of the oscillations a0 and the wavevector k0, where k0 = ω0/c.

    The emission which we have just determined – that is, that part of the field which increases proportional to the time t – arises when the frequency of the ‘force’ which occurs on the right–hand side of Eq. (1.24) is equal to the eigenfrequency of the field oscillator ωλ = ckλ. In this respect the harmonically oscillating charge is completely typical even though in the dipole approximation (condition (1.79)) which we considered there is emission at only one frequency, ω0. We note that in the quantum theory of radiation the situation is completely analogous in the perturbation theory framework (compare (1.82) and (1.73); for details see Heitler, 1947).

    It is convenient at this stage to elucidate a few important aspects which usually are kept out of sight by discussing the somewhat rhetorical question: can a uniformly moving electron radiate?

    The standard, one might say automatic, answer to that question is negative. In actual

    Enjoying the preview?
    Page 1 of 1