quantum aspects of light propagation

482

Upload: others

Post on 11-Sep-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Quantum Aspects of Light Propagation
Page 2: Quantum Aspects of Light Propagation

Quantum Aspects of Light Propagation

Page 3: Quantum Aspects of Light Propagation

Antonın Luks · Vlasta Perinova

Quantum Aspects of LightPropagation

1 3

Page 4: Quantum Aspects of Light Propagation

Antonın LuksJoint Laboratory of OpticsPalacky University and

Institute of Physics of the CzechAcademy of Sciences

772 07, OlomoucCzech [email protected]

Vlasta PerinovaJoint Laboratory of OpticsPalacky University and

Institute of Physics of the CzechAcademy of Sciences

772 07, OlomoucCzech [email protected]

Consulting Editor

D.R. VijKurukshetra UniversityE-5 University CampusKurukshetra 136119India

ISBN 978-0-387-85589-9 e-ISBN 978-0-387-85590-5DOI 10.1007/b101766Springer Dordrecht Heidelberg London New York

Library of Congress Control Number: 2009930842

c© Springer Science+Business Media, LLC 2009All rights reserved. This work may not be translated or copied in whole or in part without the writtenpermission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York,NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use inconnection with any form of information storage and retrieval, electronic adaptation, computersoftware, or by similar or dissimilar methodology now known or hereafter developed is forbidden.The use in this publication of trade names, trademarks, service marks, and similar terms, even ifthey are not identified as such, is not to be taken as an expression of opinion as to whether or notthey are subject to proprietary rights.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)

Page 5: Quantum Aspects of Light Propagation

Preface

Quantum descriptions of light propagation frequently exhibit a replacement of timeby propagation distance. It seems to be natural since a propagation lasts someamount of time. The primary intention was to inform more fundamentally inclined,open-minded readers on this approach by this book. We have included also spatio-temporal descriptions of the electromagnetic field in linear and nonlinear opticalmedia. We call some of these formalisms one dimensional (more exactly 1 + 1-dimensional), even though they comprise the time variable along with the positioncoordinate. These descriptions, however, are 3 + 1-dimensional in principle. Therapid development of applications of photonic band-gap structures and experimentson lasing in a disordered medium has directed us to pay attention even to thesetopics, which has influenced the style of the book, which becomes a very review ofthese streams.

This book has the following features. It reviews both macroscopic and micro-scopic theories of the electromagnetic field in dielectrics. It takes into account para-metric down-conversion experiments. It covers results on nonlinear optical couplers.It includes optical imaging with nonclassical light. It expounds basics of quasimodetheory. It respects success of the Green-function approach in describing opticalfield at dielectric devices, left-handed materials and the Casimir effect for somegeometries. It refers to quantization in waveguides, photonic crystals, disorderedmedia, and propagation in strongly scattering media, incoherent and coherent ran-dom lasers, and important problems in optical resonators including chaotic cavities.In our opinion it is appropriate to do something more than only formal comparisonof various approaches in the future, even though the reader will already have formedan idea of their scope.

The simplest approach with one variable (time or propagation distance) andwith several frequencies has proven its vitality in the development of the quan-tum information theory and the quantum computation. At present there exist evenbooks devoted to these fields: Alber, G., Beth, T., Horodecki, M., Horodecki, P.,Horodecki, R., Rotteler, M., Weinfurter, H., Werner, R., and Zeilinger, A. (2001),Quantum Information: An Introduction to Basic Theoretical Concepts and Experi-ments, Springer-Verlag, Berlin; Nielsen, Michael A. and Chuang, Isaac L. (2000),Quantum Computation and Quantum Information, Cambridge University Press,Cambridge.

v

Page 6: Quantum Aspects of Light Propagation

vi Preface

The fundamental problem of light propagation in dielectric media is connectedwith the role of nonclassical light in applications and has been pursued intensivelyin quantum optics since about 1984. In the present book we review spatio-temporaldescriptions of the electromagnetic field in linear and nonlinear dielectric mediaapplying macroscopic and microscopic theories. We mainly pay attention to canoni-cal quantum descriptions of light propagation in a nonlinear dispersionless dielectricmedium and linear and nonlinear dispersive dielectric media. These descriptions areregularly simplified by a transition to the one-dimensional propagation, which isillustrated also by descriptions of some optical processes.

Quantum theories of light propagation in optical media are generalized fromdielectric media to magnetodielectrics. Classical and nonclassical properties of radi-ation propagating through left-handed media will be presented. The theory is uti-lized for the quantum electrodynamical effects to be determined in periodic dielec-tric structures which are known to be a basis of new schemes for lasing and a controlof light field state. Quantum descriptions of random lasers are provided.

It is an interesting question, to what extent the topic of this book overlapswith the condensed-matter theory. Restricting ourselves to optical devices, we can-not exclude such overlap in principle, because many of them are made of condensedmatters. The condensed-matter theory, however, is devoted mainly to problems ofconductors and semi-conductors. Photonic crystals can be studied similarly as ordi-nary electronic crystals, even though for instance the conductivity is replaced by thetransmissivity. This does not mean any thematic overlap.

Texts on quantum optics have so far based the spatio-temporal description on thequantization of the electromagnetic field in a free space in the hope that differencesfrom the field in a medium are negligible or can be easily included in other ways.A rare exception was for instance the text Vogel, W. and Welsch, D.-G. (1994),Lectures on Quantum Optics, Akademie Verlag, Berlin, where a choice of a suitableapproach, albeit a selection of one of possibilities, was declared.

The book will be useful to research workers in the field of general optics,quantum optics and electronics, optoelectronics, and nonlinear optics, as well asto students of physics, optics, optoelectronics, photonics, and optical engineering.

Olomouc Vlasta PerinovaOlomouc Antonın Luks

Page 7: Quantum Aspects of Light Propagation

Acknowledgments

We have pleasure in thanking Dr. J. Perina, Jr., Ph.D., for communicating files tothe publisher, graphics, and word processing and Ing. J. Krepelka, Ph.D., for thecareful preparation of figures. This book has arisen under the financial support bythe Ministry of Education of the Czech Republic in the framework of the projectNo. 1M06002 “Optical structures, detection systems, and related technologies forfew-photon applications”.

vii

Page 8: Quantum Aspects of Light Propagation

Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Origin of Macroscopic Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72.1 Lossless Nonlinear Dielectric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72.2 Nondispersive Lossless Linear Dielectric . . . . . . . . . . . . . . . . . . . . . . 11

2.2.1 Quantization in Terms of a Dual Potential . . . . . . . . . . . . . . . 112.2.2 Momentum Operator as Translation Operator . . . . . . . . . . . . 132.2.3 Wave Functional Description of Gaussian States . . . . . . . . . 202.2.4 Source-Field Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.2.5 Continuum Frequency-Space Description . . . . . . . . . . . . . . . 31

2.3 Quantum Description of Experiments with Stationary Fields . . . . . . 362.3.1 Spatio-temporal Descriptions of Parametric

Down-Conversion Experiments . . . . . . . . . . . . . . . . . . . . . . . 382.3.2 From Coupled Quantum Harmonic Oscillators Back

to Interacting Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3 Macroscopic Theories and Their Applications . . . . . . . . . . . . . . . . . . . . . 853.1 Momentum-Operator Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

3.1.1 Temporal Modes and Their Application . . . . . . . . . . . . . . . . 863.1.2 Slowly Varying Amplitude Momentum Operator . . . . . . . . . 883.1.3 Space–Time Displacement Operators . . . . . . . . . . . . . . . . . . 1023.1.4 Generator of Spatial Progression . . . . . . . . . . . . . . . . . . . . . . 1043.1.5 Nonlinear Optical Couplers . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

3.2 Dispersive Nonlinear Dielectric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1173.2.1 Lagrangian of Narrow-Band Fields . . . . . . . . . . . . . . . . . . . . 1173.2.2 Propagation in One Dimension and Applications . . . . . . . . . 126

3.3 Modes of Universe and Paraxial QuantumPropagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1333.3.1 Quasimode Description of Spectrum of Squeezing . . . . . . . 1333.3.2 Steady-State Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1373.3.3 Approximation of Slowly Varying Envelope . . . . . . . . . . . . . 1433.3.4 Optical Imaging with Nonclassical Light . . . . . . . . . . . . . . . 152

ix

Page 9: Quantum Aspects of Light Propagation

x Contents

3.4 Optical Nonlinearity and Renormalization . . . . . . . . . . . . . . . . . . . . . 1733.5 Quasimode Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

3.5.1 Relation to Quantum Scattering Theory . . . . . . . . . . . . . . . . 1933.5.2 Mode Functions for Fabry–Perot Cavity . . . . . . . . . . . . . . . . 2003.5.3 Atom–Field Interaction Within Cavity . . . . . . . . . . . . . . . . . . 2073.5.4 Several Sets of Quasimodes . . . . . . . . . . . . . . . . . . . . . . . . . . 213

4 Microscopic Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2234.1 Method of Continua of Harmonic Oscillators . . . . . . . . . . . . . . . . . . . 223

4.1.1 Dispersive Lossy Homogeneous Linear Dielectric . . . . . . . . 2244.1.2 Correlation of Ground-State Fluctuations . . . . . . . . . . . . . . . 235

4.2 Green-Function Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2384.2.1 Dispersive Lossy Linear Inhomogeneous Dielectric . . . . . . 2394.2.2 Dispersive Lossy Nonlinear Inhomogeneous

Dielectric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2424.2.3 Elaboration of Linear Theory . . . . . . . . . . . . . . . . . . . . . . . . . 2454.2.4 Optical Field at Dielectric Devices . . . . . . . . . . . . . . . . . . . . . 2534.2.5 Modification of Spontaneous Emission by

Dielectric Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2584.2.6 Left-Handed Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2634.2.7 Application to Casimir Effect . . . . . . . . . . . . . . . . . . . . . . . . . 280

5 Microscopic Models as Related to Macroscopic Descriptions . . . . . . . . 3035.1 Quantum Optics in Oscillator Media . . . . . . . . . . . . . . . . . . . . . . . . . . 3035.2 Problem of Macroscopic Averages . . . . . . . . . . . . . . . . . . . . . . . . . . . 305

5.2.1 Conservative Oscillator Medium . . . . . . . . . . . . . . . . . . . . . . 3055.2.2 Kramers–Kronig Dielectric . . . . . . . . . . . . . . . . . . . . . . . . . . . 3115.2.3 Dissipative Oscillator Medium . . . . . . . . . . . . . . . . . . . . . . . . 312

5.3 Single-Photon Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317

6 Periodic and Disordered Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3216.1 Quantization in Periodic Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321

6.1.1 Classical Description of Electromagnetic Field . . . . . . . . . . 3226.1.2 Modal Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3236.1.3 Method of Coupled Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 3316.1.4 Normalized Modes of the Electromagnetic Field . . . . . . . . . 3346.1.5 Quantization in Linear Nonhomogeneous

Nonconducting Medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3426.2 Corrugated Waveguides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349

6.2.1 Lossless Propagation in a Waveguide Structure . . . . . . . . . . 3516.2.2 Coupled-Mode Theory Including Gain or Losses . . . . . . . . . 359

6.3 Photonic Crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3786.4 Quantization in Disordered Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393

6.4.1 Quantization in Chaotic Cavity . . . . . . . . . . . . . . . . . . . . . . . . 3946.4.2 Open Systems Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398

Page 10: Quantum Aspects of Light Propagation

Contents xi

6.4.3 Semiclassical Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4046.5 Propagation in Amplifying Random Media . . . . . . . . . . . . . . . . . . . . 408

6.5.1 Strongly Scattering Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4086.5.2 Incoherent and Coherent Random Lasers . . . . . . . . . . . . . . . 4146.5.3 Modal Decomposition in Optical Resonators . . . . . . . . . . . . 4446.5.4 Chaotic Resonators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445

7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

Page 11: Quantum Aspects of Light Propagation

Chapter 1Introduction

The importance of quantum optics has been recognized by both specialists and pub-lic since Roy J. Glauber was awarded the Nobel Prize in Physics 2005. Quantuminformatics is closely connected with this field. Ingenious, but simple solutions arepreferred to intricacies of the quantized field theory with the hope that experimentersrealize the simple proposals with appropriate means.

From the historical viewpoint, the problem of quantization of the electromag-netic field in vacuo was solved by Dirac (1927) long ago and the quantization ofa nonlinear theory is due to Born and Infeld (1934, 1935). With respect to thepropagation in linear dielectric media it is appropriate to refer first to Jauch andWatson (1948). A revived interest in this problem can be perceived since the 1990s.First it resembled some dissatisfaction with the situation following the advent oflaser in 1958. The new optical effects are analyzed both by the methods of non-linear optics which belong to classical physics and by those of quantum optics(Shen 1969). In quantum optics, the normal-mode expansion approach is used thatis well suited for systems in optical cavities, such as an optical parametric oscil-lator, but is not appropriate for open systems such as a parametric amplifier. Innonlinear optics (Bloembergen 1965, Shen 1984), the Maxwell equations com-pleted by the constitutive relations are solved with the method of slowly varyingenvelope approximation and the resultant equations are sometimes simplified onthe assumption of parametric approximation. It has become standard that the phe-nomenological Hamiltonians of quantum optics are frequently introduced withouta quantitative connection to the classical equations describing the nonlinear opticaleffects.

The quantization of the electromagnetic field in the presence of a dielectric ispossible. This can be done in two ways which are called the macroscopic andmicroscopic approaches. In the first, the macroscopic approach, the medium is com-pletely described by its linear and nonlinear susceptibilities. No matter degrees offreedom appear explicitly in this treatment. After a Lagrangian which produces themacroscopic Maxwell equations for the field in a nonlinear medium is found, thecanonical momenta and the Hamiltonian are derived. Quantization is accomplishedby imposing the standard equal-time commutation relations. In the second approach,the microscopic, a model for the medium is constructed and both the field and thematter degrees of freedom appear in the theory. Both are quantized. The result is a

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 1, C© Springer Science+Business Media, LLC 2009

1

Page 12: Quantum Aspects of Light Propagation

2 1 Introduction

theory of mixed matter-field (polariton) modes, which are coupled by a nonlinearinteraction.

Hillery and Mlodinow (1984) have used the electric displacement field as thecanonical variable for nonlinear quantization and they have explored the macro-scopic approach to the quantization of homogeneous nondispersive media. Theyhave pointed out that there is a difficulty in including the dispersion in the quantizedmacroscopic theory.

In the past, many authors that dealt with macroscopic quantum theories of lightpropagation wrote also on space displacements, shifts, and translations of the elec-tromagnetic field along with the time displacements, shifts, and translations, or sim-ply on the (time) evolution. Accordingly, they used the term “space evolution” inthe former case. In the following, we will use the term space progression insteadof space evolution. Abram (1987) intended to overcome the difficulties of the con-ventional quantum optics by reformulating its assumptions. He has based the for-malism on the momentum operator for the radiation field and investigated in thisway not only the spatial progression of the electromagnetic wave but also refractionand reflection. The importance of a proper space–time description of squeezing hasbeen recognized (Bialynicka-Birula and Bialynicki-Birula 1987). The problem ofa proper quantum mechanical description of the operation of optical devices hasbeen addressed (Knoll et al. 1986, 1987). Besides this, an attempt at a formula-tion of quantum theory of propagation of the optical wave in a lossless dispersivedielectric material has been made (Blow et al. 1990). The vacuum propagation andlow-order perturbation theory have sufficed for spatio-temporal descriptions of para-metric down-conversion experiments (Casado et al. 1997a, Casado et al. 1997b).The experiment on the “induced coherence without induced emission” has beendescribed on restriction to spatial behaviour of fields and the multimode descriptionhas been restored too (Perinova et al. 2003).

The applications have used the fact that the nonlinear processes of quantumoptics are described quantum optically in the parametric approximation with linearmathematical tools so that quantization procedures and solutions of the dynamicsneed not face immense difficulties as for the really nonlinear formalism (Huttneret al. 1990). The formalism of the macroscopic approach to the quantization hasbeen developed (Abram and Cohen 1991). The space–time displacement operatorshave been related to the elements of the energy–momentum tensor (Serulnik andBen-Aryeh 1991). The macroscopic quantization of the electromagnetic field wasapplied to inhomogeneous media (Glauber and Lewenstein 1991). The theoreticalmethods for investigating propagation in quantum optics, in which the momentumoperator is used along with the Hamiltonian, have been developed (Toren and Ben-Aryeh 1994). The excellent review of linear and nonlinear couplers (Perina, Jr. andPerina 2000), where the restriction to merely spatial behaviour of interesting opticalfields is accepted, has used a similar approach.

The optical solitons have been studied in the nonlinear optics and their quan-tum properties have been calculated using spatio-temporal descriptions by eruditeauthors. The dispersion has been treated on the assumption of a narrow-frequencyinterval (Drummond 1990). Drummond (1994) has presented a review of his theory

Page 13: Quantum Aspects of Light Propagation

1 Introduction 3

and its applications. In addition to previous work (Lang et al. 1973) devoted tothe concept of quasinormal modes, the modes of the universe have been used inthe treatment of the spectrum of squeezing (Gea-Banacloche et al. 1990a,b). Anoriginal approach to the description of a degenerate parametric amplifier (Deutschand Garrison 1991a) has been related to the theory of paraxial quantum propaga-tion (Deutsch and Garrison 1991b). One-dimensional description of beam propa-gation has been completed with transverse position coordinates. It has taken intoaccount the existence of small photodetectors or pixels (Kolobov 1999). Abram andCohen (1994) have developed a travelling-wave formulation of the theory of quan-tum optics and have applied it to quantum propagation of light in a Kerr medium.A quantum scattering theory approach to quantum-optical measurements has beenexpounded (Dalton et al. 1999a). In addition to (Lang et al. 1973) and along with anindependent work (Ho et al. 1998) devoted to the concept of quasinormal modes,quasimode theory of macroscopic canonical quantization has been invented andapplied (Dalton et al. 1999b,c). A macroscopic canonical quantization of the electro-magnetic field and radiating atom system involving classical, linear optical devices,based on expanding the vector potential in terms of quasimode functions has beencarried out (Dalton et al. 1999b). The relationship between the pure mode and quasi-mode annihilation and creation operators is determined (Dalton et al. 1999c). Aquantum theory of the lossless beam splitter is given in terms of the quasimodetheory of macroscopic canonical quantization. The input and output operators thatare related via scattering operator are directly linked to multi-time quantum corre-lation functions (Dalton et al. 1999d). Brown and Dalton (2001a) have generalizedthe quasimode theory of macroscopic quantization in quantum optics and cavityquantum electrodynamics developed by Dalton, Barnett, and Knight (1999a,b). Thisgeneralization admits the case where two or more quasipermittivities are introduced.The generalized form of quasimode theory has beeen applied to provide a fullyquantum-theoretical derivation of the laws of reflection and refraction at a boundary(Brown and Dalton 2001b).

Huttner and Barnett (1992a,b) have presented a fully canonical quantizationscheme for the electromagnetic field in dispersive and lossy linear dielectrics. Thisscheme is based on the Hopfield model of such a dielectric, where the matter isrepresented by a harmonic polarization field (Hopfield 1958). Following (Huttnerand Barnett 1992a,b), Gruner and Welsch (1995) have calculated the ground-statecorrelation of the quantum-mechanical fluctuations of the intensity. Gruner andWelsch (1996a) have realized the expansion of the field operators which is basedon the Green function of the classical Maxwell equations and preserves the equal-time canonical commutation relations of the field. They have found that the spatialprogression can be derived on the assumption of weak absorption. In (Schmidt et al.1998), the microscopic approach to the quantum theory of light propagation hasbeen extended to nonlinear media and the generalized nonlinear Schrodinger equa-tion well known from the description of quantum solitons has been derived for adielectric with a Kerr nonlinearity. Dung et al. (1998) have developed a quantiza-tion scheme for the electromagnetic field in a spatially varying three-dimensionallinear dielectric which causes both dispersion and absorption. In the case of a

Page 14: Quantum Aspects of Light Propagation

4 1 Introduction

homogeneous dielectric, the well-known Green function has been used and it hasbeen shown that the indicated quantization scheme exactly preserves the funda-mental equal-time commutation relations of quantum electrodynamics. The Greenfunction has also been used in the more complicated case of two dielectric bod-ies with a common planar interface. Spontaneous decay of an excited atom in thepresence of dispersing and absorbing bodies has been investigated using an exten-sion of this formalism (Dung et al. 2000). A microscopic theory of an optical fieldin a lossy linear optical medium has been developed (Knoll and Leonhardt 1992).Dutra and Furuya (1997) have considered a single-mode cavity filled with a mediumconsisting of two-level atoms that are approximated by harmonic oscillators. Theyhave shown that macroscopic averaging of the dynamical variables can lead to amacroscopic description. Dutra and Furuya (1998a,b) have observed that the (full)Huttner–Barnett model of a dielectric medium does not comprise all the dielectricpermittivities of the medium which can be expected from the classical electrody-namics, although the field theory in linear dielectrics should have such a property.Dalton et al. (1996) have dealt with the quantization of a field in dielectrics andhave applied it to the theory of atomic radiation in one-dimensional Fabry–Perotresonator.

Yablonovitch (1987) suggested that three-dimensional periodic dielectric struc-tures could have a photonic band gap in analogy to electronic band gaps in semicon-ductor crystals, namely, a band of frequencies for which an electromagnetic wavecannot propagate in any direction. This idea and its subsequent experimental proofin the macrowave domain have led to extensive activity aimed at the optimizationof photonic band-gap structures for the visible domain and the exploration of theirpotential applications (Journal of Modern Optics 1994, Journal of the Optical Soci-ety of America B 2002, etc.).

As soon as quantization in nonhomogeneous dielectric media is solved, not onlythe case of a finite dielectric medium is worth a treatment but also the case of aninfinite periodic medium (Caticha and Caticha 1992, Kweon and Lawandy 1995,Tip 1997). The idea of one-dimensional propagation may be compared with resultsconcerning a mirror waveguide. Nonlinear optics in a photonic band-gap structurehas been studied (Tricca et al. 2004, Perina, Jr. et al. 2004, 2005, Perina, Jr. et al.2007). Sakoda (2002) has formulated quantization of the electromagnetic field inphotonic crystals. Spontaneous parametric down conversion in a finite-length mul-tilayer structure has been considered (Centini et al. 2005, Perina, Jr. et al. 2006).

Both localization and laser theory, which were developed in the 1960s, werejointly applied in the study of random laser. They were used in strongly scatter-ing gain media. Lasing in disordered media has been a subject of intense theoret-ical and experimental studies. Random lasers have been classified into incoherentand coherent random lasers. Research works on both types of random lasers havebeen summarized in the monographic chapter (Cao 2003). In order to understandquantum-statistical properties of random lasers, quantum theory is needed. Standardquantum theory for lasers applies only to quasidiscrete modes and cannot accountfor lasing in the presence of overlapping modes. In a random medium, the characterof lasing modes depends on the amount of disorder. Weak disorder leads to a poor

Page 15: Quantum Aspects of Light Propagation

1 Introduction 5

confinement of light and to strongly overlapping modes. Statistics naturally belongsto the theory of amplifying random media (Beenakker 1998, Patra and Beenakker1999, 2000, Mishchenko et al. 2001), which is restricted to linear media.

Hackenbroich et al. (2002) have developed a quantization scheme for opticalresonators with overlapping (nonorthogonal) modes. Cheng and Siegman (2003)have derived a generalized formalism of radiation-field quantization, which neednot rely on a set of orthogonal eigenmodes. True eigenmodes of a system will benon-orthogonal and the method is intended for quantization of an open system, inwhich a gain or loss medium is involved.

We will use units following original papers and although the system of interna-tional (SI) units prevails, there are exceptions, so some of the relations (2.25)–(2.69)and (3.276)–(3.322) are in the Gaussian units, the relations (3.323)–(3.392) are inthe rationalized cgs units, and the relations (2.1)–(2.2), (3.14)–(3.108), (3.494)–(3.578), (3.109)–(3.125), and (2.15)–(2.24) are in the Heaviside–Lorentz units.

Page 16: Quantum Aspects of Light Propagation

Chapter 2Origin of Macroscopic Approach

With the birth of quantum optics in the 1960s it became clear that it would be easyto describe the interaction between the electromagnetic field and the matter in acavity even on elimination of matter degrees of freedom. A similar travelling-wavedescription for the electromagnetic field–matter interaction was considered to bepossible in terms of a virtual cavity and a momentum operator of the field.

This approach to quantization was rather distant from the quantum theory ofthe electromagnetic field. On a fundamental level the theory of the electromagneticfield in the free space does not differ from the theory of this field in the matter.Macroscopic approaches to quantization of the electromagnetic field are not funda-mental theories and modify the free-space electromagnetic-field theory. Especially,quantization of the field power has been assumed.

Although the virtual cavity has been beaten, the momentum operator has stillenabled one to study quantum aspects of nonlinear optical processes. Quantizationrestrictions of any kind such as the frequency dispersion of the refractive index wereapparent on published work.

Efforts emerged to formulate so simple a quantum theory of the electromagneticfield that it allows one to recognize the role of the momentum operator. Formalismswere presented which, to the contrary, did not consider the momentum operator.With the progress in (classical) optics interest in the quantization of the field powerin quantum optics has increased.

Not always is it necessary to utilize the formalism of the electromagnetic fieldin the matter. For description of experiments with correlated photons it suffices todescribe the electromagnetic field between optical devices and to know the input–output relations for the optical elements, both passive and active, with which theradiation is transformed.

2.1 Lossless Nonlinear Dielectric

An approach to the quantum theory of light propagation was considered stan-dard until the critique by Hillery and Mlodinow (1984) and is still. Concerningthis approach, let us consider papers by Shen (1967, 1969). Shen (1967) studied

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 2, C© Springer Science+Business Media, LLC 2009

7

Page 17: Quantum Aspects of Light Propagation

8 2 Origin of Macroscopic Approach

quantum statistics of nonlinear optics. He contributed to the contemporary research(Glauber 1965). Quantum theory of radiation had long been formulated (Heitler1954). For investigation of properties of a medium, incoherent scattering has been auseful tool. For nonlinear optics, coherent scattering has been interesting as well ormore.

Weak nonlinearity has a significant effect on light only after a longer interactiondistance. Light can cover a longer distance easily when contained in a cavity res-onator. Quantum statistics has been determined using descriptions suited to the caseof a cavity.

In principle, the same treatment can be applied to problems of light propagationin media (Shen 1967). But for coherent scattering, it becomes difficult. This caseshould be treated by the method of many-body transport theory (Ter Haar 1961). Inquantum optics, the cavity treatment of the problems of light propagation in mediaseems to be valid on the following assumption. Photon fields may be quantizedin a box of finite volume, which moves in the z direction with a light velocity c( c√

εin Shen 1969). One is advised to imagine a box of length cT , where T is the

counting time of photodetectors. A partial interaction of the light with the mediumcan be approximated with no interaction and a complete interaction, which lastsfor a time t . The finite medium can be extended to infinity. The resultant changeof statistical properties of fields in the box can now be calculated using the cavitytreatment (Shen 1967).

In nonlinear optics a number of classical descriptions have been developed bothas a cavity problem and as a steady-state propagation problem. Then a cavityproblem can be converted to a corresponding steady-state propagation problem byreplacing t by − z

c in the field amplitudes and the latter problem can be changed tothe former one by replacing z by −ct when ez is the direction of propagation. Itraises expectations that the same is true in the quantum treatment. Shen (1969) paysattention to replacing t by − z

√ε

c and to replacing z by ct√ε.

Here the dependence on the time seems to be more fundamental. It is evident thatone is interested in a conversion of a cavity problem to a corresponding steady-statepropagation problem. The operators will be space dependent (localized) insteadof time dependent. Transformations will be generated with a localized momentumoperator instead of the Hamiltonian operator.

On quantizing in a volume L3 and assuming that the field does not vary appre-ciably over a distance d large compared with the wavelength, and associating thediscrete values of the wave vector k with d (instead of L), the localized annihilationand creation operators bk(z) and b†

k(z) have been proposed. An appropriate compo-nent of the vector-potential operator has the expansion of the form

A(z, t) = c∑

k

√�

2ωkεk L3

{bk(z) exp[−i(ωk t − kz)]+ H.c.

}, (2.1)

where ωk is the frequency, � is the Planck constant divided by 2π , εk = ε(ωk) isthe value of the dielectric function ε at ωk , and H.c. denotes the term Hermitian

Page 18: Quantum Aspects of Light Propagation

2.1 Lossless Nonlinear Dielectric 9

conjugate to the previous one. The annihilation and creation operators bk(z) andb†

k(z), respectively, satisfy the equal-space commutation relation

[bk(z), b†

k ′(z)] = δkk ′ 1. (2.2)

The small variation of the field has been formulated as that of the normally ordered

moments⟨b†m

k (z)bnk (z)

⟩. It is also specified that k = 2πn

d , where n is an integer.

There is a difficulty. The above picture of a moving box requires a light velocityc independent of the frequency ωk . Shen (1969) utilizes the notation c√

εfor this

velocity. Here it is replaced by the phase velocity c√εk

, with c ≡ c0, the free-spacespeed of light. There is another difficulty in view of this picture that d has been usedinstead of cT . The localized photon-number operator is realized as a configuration-space photon-number operator (Mandel 1966)

n(z) = Ad

L3

k

b†k(z)bk(z), (2.3)

where A is the cross-sectional area of the beam.A Hamiltonian density H (z, t) is considered. The Hamiltonian is

H(t) = L2∫

LH (z′, t) dz′. (2.4)

A third difficulty is that the localized momentum operator is defined as H (z,t)c essen-

tially, not by using an integration with respect to time. It has been assumed thatk = 2πn

d , not that ωk = 2πnT . For free fields, the localized momentum operator is

P(z) =∑

k

�k[b†

k(z)bk(z)+ 12 1]. (2.5)

For interacting fields, the localized momentum operator has the form of a Hamilto-nian, but with bk(z) and b†

k(z) replacing ak(t) and a†k(t).

A momentum operator should have the form ezP(z), be a vector, but in fact onedoes not utilize this. The momentum operator generates translations

d

dzbk(z) = i

[bk(z), P(z)

]. (2.6)

The electric strength vector is derived from the vector potential according to therelation

E(z, t) = −1

c

∂tA(z, t). (2.7)

Page 19: Quantum Aspects of Light Propagation

10 2 Origin of Macroscopic Approach

We decompose this operator as

E(z, t) = E (+)(z, t)+ E (−)(z, t), (2.8)

where E (+)(z, t) (E (−)(z, t)) contains the functions exp(−iωk t)(exp(iωk t)). In Shen (1967) the opposite convention is used. Then

d

dzE (+)(z, t) = i

[E (+)(z, t), P(z)

]. (2.9)

Something is more suitable for propagation problem: We define all the quantities ata given plane z = z0 for all times and try to obtain the propagation towards z ≥ z0.According to equations (2.6) and (2.9), the unitary translation operator is

U (z, z0) = S exp

[i

∫ z

z0

P(z′) dz′]

, (2.10)

where S is the space-ordering operation. The space-ordered product has a similardefinition as the time-ordered product. Field operators at different spatial points z,z0 are connected by this unitary operator:

E(z, t) = U †(z, z0)E(z0, t)U (z, z0). (2.11)

There are indications that any “alternative” quantum theory is avoided. Such anindication is the fact that the localized momentum operator has been derived fromthe Hamiltonian density. With this in mind, we pass from the “spatial Heisenbergpicture” to a spatial Schrodinger picture. In the latter picture, a localized densitymatrix (statistical operator) progresses:

ρ(z) = U (z, 0)ρ(0)U †(z, 0). (2.12)

Here ρ(0) is a given statistical operator. Then the correlation function of fields atdifferent times is expressed in two forms:

⟨E (−)(z, t1) . . . E (−)(z, tn)E (+)(z, tn) . . . E (+)(z, t1)

= Tr{ρ(0)E (−)(z, t1) . . . E (−)(z, tn)E (+)(z, tn) . . . E (+)(z, t1)

}

= Tr{ρ(z)E (−)(0, t1) . . . E (−)(0, tn)E (+)(0, tn) . . . E (+)(0, t1)

}. (2.13)

The equation of motion for a statistical operator ρ(z) is

∂zρ(z) = i

[P(z), ρ(z)

]. (2.14)

With the help of these localized operators, the calculations for steady-state propa-gation in a medium become the same as the corresponding calculations for a cavity

Page 20: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 11

with t replaced by − zc (Shen 1967) and by z

√ε

c (Shen 1969). The problem of beamsplitting was mentioned. Essentially, the same proposal has been included in Shen(1969).

2.2 Nondispersive Lossless Linear Dielectric

The study of nonlinear optical phenomena and their inclusion in an effective non-linear theory of the electromagnetic field has utilized the asymmetry of most opticalmedia, which are nonlinear with respect to the electric-field, but linear relative tothe magnetic field. The canonical momentum should be the magnetic induction inplace of the more usual electric-field strength. Such a theory may not be capable ofdescribing the Bohm–Aharonov effect. Besides such a theory we expound a simplequantization connected to considerations of the role of the Poynting vector operatorand the momentum operator.

A description of the field distribution in space must be completed with a quantumstate of the field in quantum physics. A renewed interest in the spatio-temporaldescription leads to the study of the wave functional of the electromagnetic field des-pite the doubts of the pioneers of theoretical physics of the photonic wavefunction.

On neglecting dispersion and nonlinearity, a macroscopic theory of the quantizedelectromagnetic field in a medium can be very close to the usual theory of this fieldin free space. In contrast to this, solutions have been disseminated, which includethe dispersion and the nonlinearity at least approximately.

2.2.1 Quantization in Terms of a Dual Potential

According to a pioneering paper of Hillery and Mlodinow (1984), the standardmacroscopic quantum theory of electrodynamics in a nonlinear medium is due toShen (1967) and has been elaborated upon by Tucker and Walls (1969). Hillery andMlodinow (1984) have pointed out some problems with the standard theory, aboveall that it is not consistent with the macroscopic Maxwell equations.

One approach to the derivation of a macroscopic quantum theory would be tobegin from a quantum microscopic theory as explored in the linear case by Hopfield(1958). The other approach is to take the expression for the energy of the radiation innonlinear medium, which differs from the free-field Hamiltonian in part, and to keepinterpreting the electric-field (up to the sign) as the canonically conjugated variableto the vector potential. Then, this macroscopic classical theory is quantized. (Let usnote that it differs from Shen (1969)). The Hamiltonian formulation of the theoryconsists in the noncanonical Hamiltonian

Hnoncan = HEM + HInoncan, (2.15)

Page 21: Quantum Aspects of Light Propagation

12 2 Origin of Macroscopic Approach

where

HEM = 1

2

∫(E2 + B2) d3x, (2.16)

HInoncan = 1

2

∫E · P d3x, (2.17)

with E being the electric field strength operator and P being the polarization of themedium, and the Heaviside–Lorentz units having been used. The polarization is afunction of the electric field which may be written as a power series. This theory maybe called standard. It can easily be seen that, as an undesirable “quantum effect”,we obtain an improper expression for the time derivative of the magnetic-inductionfield B.

It is assumed that the medium is lossless, nondispersive, and homogeneous. ALagrangian is considered which gives proper equations of motion. The electric andmagnetic fields are expressed in terms of the vector potential A and the scalar poten-tial A0:

E = −∂A∂t

−∇A0, B = ∇ × A. (2.18)

The appropriate Lagrangian density depends on the first partial derivatives of thefour-vector A = (A0, A). The momentum canonical to A is Π = (Π0,ΠΠ), whereΠ0 = 0. The vanishing of Π0 indicates that the system is constrained. It has beenshown how to utilize the Dirac quantization procedure for constrained Hamiltoniansystems (Dirac 1964). It can be derived that the canonical momentum is Π = −D.The canonical Hamiltonian has the form

H = HEM + HI, (2.19)

where

HI =∫

E ·[

P−∫ 1

0P(λE) dλ

]d3x. (2.20)

In order to simplify the quantization of the macroscopic Maxwell theory, the dualpotential Λ has been introduced along with Λ and Λ0, which we call the dual vectorand scalar potentials. The relation (2.18) is replaced by

D = ∇ ×Λ, B = ∂Λ

∂t+ ∇Λ0. (2.21)

It can be shown that the canonical momentum is ΠΠΠ× = B. Upon expressingthe canonical Hamiltonian functional in terms of the electric-displacement and

Page 22: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 13

magnetic-induction fields, the results are the same:

H = H×. (2.22)

Then, the usual Hamiltonian theory for the electromagnetic field in a nonlineardielectric medium and the alternative have been quantized in the ordinary way. Wecan compare

[Ai (x, t), Π j (x′, t)

] = iδ⊥i j (x− x′)1, (2.23)

with

[Λi (x, t), Π×

j (x′, t)] = iδ⊥i j (x− x′)1. (2.24)

The transverse δ function has been used and made a reference to Bjorken andDrell (1965). Hillery and Mlodinow (1984) do not mention propagation except aparagraph on the interpretation problems, where they recommend to confine themedium to part of the quantization volume and to place the field source and thedetector outside of the medium, being aware that they require the consideration ofpropagation.

It is added that different diagonalizations indicated by the quadratic part ofthe total Hamiltonian generate different kinds of normal ordering. A doubt isexpressed that there is an appropriate kind and the microscopic approach is pro-pounded. Dispersion is also considered a reason for a microscopic theory to becontemplated.

2.2.2 Momentum Operator as Translation Operator

In the late 1980s, the problem of propagation did not seem to be typical of quantumoptics. Abram addressed the problem of light propagation through a linear nondis-persive lossless medium (Abram 1987). Although this model can be an appropriatelimit of the Huttner–Barnett model, we expound the main ideas of Abram (1987).Abram criticized the modal Hamiltonian formalism, especially the inclusion of thelinear polarization term in the Hamiltonian:

H ?= 1

V(E2 + H 2 + 4πχ E2) dV, (2.25)

where E (H ) is the magnitude of the electrical (magnetic) field strength, χ is the(linear) susceptibility of the material, and V is a quantization volume. This wouldlead to an incorrect result, mainly to a change of the frequencies of the modeswhich does not occur. He decided to extend the traditional theory of quantumoptics to describe propagation phenomena without invoking the modal Hamilto-nian. According to him one of the propagation phenomena, refraction, suggests

Page 23: Quantum Aspects of Light Propagation

14 2 Origin of Macroscopic Approach

the momentum as the concept appropriate for the description of these phenomena.Quantum mechanically, space and momentum are canonically conjugate variables.Let us remark that microscopic models demonstrate that a Hamiltonian includinglight–matter interaction can be considered. These are a good antidote against theidea that “space and momentum are canonically conjugate variables like time andenergy”.

Propagation of the electromagnetic field is described by the Maxwell equations:

∇ ×H = 1

c

∂D∂t

, (2.26)

∇ × E = −1

c

∂B∂t

, (2.27)

∇ · B = 0, (2.28)

∇ · D = 0, (2.29)

where D = E+4πP is the electric displacement, B is the magnetic induction, E andH are the electric and magnetic field strengths, respectively, P is the (linear and non-linear) polarization induced in the medium, and c is the speed of light. We assumethat there are no free charges or currents and that we are dealing with nonmagneticmaterials, so that B = H.

For simplicity, we shall consider only the case of plane waves propagating alongthe z-axis, with the electric field polarized along the x-axis and the magnetic fieldalong the y-axis. This reduces the Maxwell equations to scalar differential equa-tions, the directions of all vectors being implicit. We shall further assume that lightis propagating in a linear dielectric, where the induced polarization is at all timesproportional to the incident electric field:

P = χ E, (2.30)

where we assume the susceptibility of the material for simplicity to be a scalar(neglecting its tensorial properties), independent of frequency (no dispersion). It isconvenient to define also the dielectric function ε of the material

ε = 1+ 4πχ, (2.31)

and the refractive index n,

n = √ε. (2.32)

The change in the total energy which is given by the integrated energy flux (thePoynting vector) over the surface of a body or volume is proposed in Abram (1987)as the proper quantum-mechanical Hamiltonian. The change in the total momentumis given as the integrated flux of the Maxwell stress tensor. The momentum is treated

Page 24: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 15

on the same footing as the Hamiltonian. However, the enigma of the Hamiltonian(2.25) is solved. We may consider a square pulse which enters a dielectric. The totalenergy is conserved, but the energy density is increased by a factor of n, becausethe volume V reduces to V ′ = V

n . In volume V ′ the wavelengths of the modesbecome λ′ = λ

n , but the oscillator frequencies remain unchanged. It is interestingthat in the absence of reflection, the electric and magnetic fields of the transmitted(T ) waves in the dielectric are related to the corresponding incident (I ) fields in freespace by

ET = 1√n

EI , (2.33)

HT =√

nHI . (2.34)

This change in the energy density implies a similar increase for the total momentumof the pulse, the components of which are always proportional to the wave vectorsof the excited modes. In propagation along the z-axis the Maxwell stress tensor isreplaced by the energy density.

When the propagation along the±z-axis in free space is considered with the elec-tric field polarized along the x-axis and the magnetic field along the y-axis (χ = 0,ε = 1), the electromagnetic vector-potential operator A ≡ A(z, t) is usually writtenas (� = 1)

A(z, t) = c∑

j

(2π

V ω j

) 12 (

a†j e

iω j t−ik j z + a j e−iω j t+ik j z

), (2.35)

where a†j , a j are the creation, annihilation operators, respectively, for a photon in

the j th mode of the wave vector k j (with k− j = −k j ) and the frequency ω j = c|k j |fulfilling the Bose commutation relations. To simplify the notation, we omit unitvectors. It is convenient to rearrange equation (2.35) in a manner that is familiar tosolid-state physicists,

A(z, t) = c∑

j

(2π

V ω j

) 12 (

a†j e

iω j t + a− j e−iω j t

)e−ik j z . (2.36)

The electric and magnetic field operators may be obtained as

E(z, t) = −1

c

∂tA(z, t) =

j

e j

= −i∑

j

(2πω j

V

) 12 (

b†j − b− j

), (2.37)

Page 25: Quantum Aspects of Light Propagation

16 2 Origin of Macroscopic Approach

and

H (z, t) = ∂

∂zA(z, t) =

j

h j

= −i∑

j

s j

(2πω j

V

) 12 (

b†j + b− j

), (2.38)

where s j ≡ sgn j and

b j = a j e−iω j t+ik j z . (2.39)

When products of these operators are encountered, we suppose that they are sym-metrized. The Hermiticity of the operators E ≡ E(z, t) and H ≡ H (z, t) can beverified using the relations

e†j = e− j , (2.40)

h†j = h− j . (2.41)

The energy density operator can be written as

u = 1

(E2 + H 2

)(2.42)

=∑

j

u j (2.43)

= 1

j

(e j e− j + h j h− j

)

= 1

2V

j

ω j

(b†

j b j + b†− j b− j + 1

)(2.44)

= 1

V

j

ω j

(b†

j b j + 1

21

). (2.45)

The energy fluxes due to the forward (backward) waves alone can be expresseduniquely:

u+ =∑

j(>0)

ω j

V

(b†

j b j + 1

21

), u− =

j(<0)

ω j

V

(b†

j b j + 1

21

). (2.46)

Page 26: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 17

The total momentum operator G is then

G = V

c(u+ − u−) =

j

k j

(b†

j b j + 1

21

). (2.47)

It is important to understand the relations (3.9) and (3.10) in Abram (1987) well.We interpret (3.9) concerning elementary quantum mechanics as

〈z| pz|ψ〉 = −i�∂

∂z〈z|ψ〉, (2.48)

where |z〉 are position coordinate states and |ψ〉 is an arbitrary pure state. The simi-larity with equation (3.10) from Abram (1987)

∂ Q

∂z= −i[G, Q], (2.49)

where Q is any operator, fades. We would prefer a definition of the operator Q. Letus consider

Q ≡ Q(z, t) = Q[E(z, t), H (z, t)], (2.50)

where Q[•, •] is a formal series in E and H . Since the differential operator ∂∂z is

just as differentiation as the superoperator −i[G, •], it suffices to verify the relation(2.49) for Q = E , H . It is true at least in the situations treated in Abram (1987).

Although the operators b j ≡ b j (z, t) are studied using (2.49), the Heisenbergequation of motion, and the initial condition

b j (0, 0) = a j (2.51)

as appropriate for any operator Q(z, t), we perceive that the operators do not obeyour definition of the operator Q.

We may calculate the Poynting vector operator as

S = c

4πE H = c

j

e j h− j (2.52)

=∑

j

s jcω j

V

(b†

j b j + 1

21

). (2.53)

The Poynting vector operators due to the forward (backward) waves alone can beexpressed uniquely:

S+ =∑

j(>0)

cω j

V

(b†

j b j + 1

21

), S− = −

j(<0)

cω j

V

(b†

j b j + 1

21

). (2.54)

Page 27: Quantum Aspects of Light Propagation

18 2 Origin of Macroscopic Approach

The total energy operator of the free field inside the volume of quantization is thus

H = U = V

c

(S+ − S−

) =∑

j

ω j

(b†

j b j + 1

21

). (2.55)

The investigation of the case χ �= 0, ε �= 1 does not lead to any new expansionsof the field operators E and H . The individual components of the rearranged elec-tric and magnetic field operators according to (2.37) and (2.38) satisfy a modifiedoperator algebra with respect to that of the harmonic oscillator:

[e j , el ] = [h j , hl ] = 0, (2.56)

[e j , hl] = s− j

(4πω j

V

)δ− j,l 1, (2.57)

where δ j,l is the Kronecker δ function. The knowledge of these commutators and ofthe generalized total momentum operator G, the derivation of (2.42) through (2.47),should have been generalized accordingly, e.g. the relation (2.42) becoming

u = 1

(ε E2 + H 2

)

= 1

j

(εe j e− j + h j h− j

), (2.58)

which enables us to derive the Maxwell equations both via the temporal derivativesand via the spatial derivatives.

The energy density operator (2.58) can be generalized. In the expansion (2.43)we can set u j = u j refr,

u j refr = ω j

2V

{b†

j b j + b− j b†− j − 2πχ

(b†

j − b− j

) (b†− j − b j

)}. (2.59)

The energy density operator urefr may be diagonalized through a Bogoliubov trans-formation. To this end we introduce an anti-Hermitian operator R of the form

R =∑

j

(b j b− j − b†

j b†− j

)(2.60)

and introduce the operators

B j = e−γ R b j eγ R = (cosh γ )b j − (sinh γ )b†

− j , (2.61)

where

γ = 1

4ln ε = 1

2ln n. (2.62)

Page 28: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 19

On substitution

b j = eγ R B j e−γ R = (cosh γ )B j + (sinh γ )B†

− j , (2.63)

the operator R takes the form

R =∑

j

(B j B− j − B†

j B†− j

), (2.64)

and the energy density operator has the diagonal form

u j refr = nω j

2V

(B†

j B j + B− j B†− j

). (2.65)

The momentum operator is then given by

Grefr = V

c(u+ − u−) =

j

K j

(B†

j B j + 1

21

), (2.66)

with K j = nk j and the Hamiltonian can be calculated as

Hrefr =∑

j

ω j

(B†

j B j + 1

21

). (2.67)

By inserting (2.63) into (2.37) and (2.38), respectively, we can obtain the electricand magnetic field operators inside the dielectric:

E(z, t) = −i∑

j

(2πω j

nV

) 12

(B†j − B− j ) (2.68)

and

H (z, t) = −i∑

j

s j

(2πnω j

V

) 12

(B†j + B− j ). (2.69)

Similarly as above, this relation can be interpreted as a result of the replacementb j �→ B j and a consequence of the quantized classical equations (2.33) and (2.34).

For normal incidence on a sharp vacuum–dielectric interface, both reflection anddiffraction occur. We will not treat this more general case according toAbram (1987).

Page 29: Quantum Aspects of Light Propagation

20 2 Origin of Macroscopic Approach

2.2.3 Wave Functional Description of Gaussian States

Białynicka-Birula and Białynicki-Birula (1987) have tried first to define the squeez-ing that is a generalization of the standard definition for one mode of radia-tion. This definition can be reformulated with respect to Białynicki-Birula (2000).The Riemann–Silberstein–Kramers complex vector has been introduced

F(r, t) = 1√2

[D(r, t)√

ε0+ i

B(r, t)√μ0

], (2.70)

where we have divided by√

ε0,√

μ0, as is appropriate with SI units. It has beenshown how the Green function method can be used for solving linear equations forthe field operator F(r, t). This approach allows that the medium under investigationis inhomogeneous and time dependent. It is not clear whether the complex vector(2.70) is then useful. It has been suggested that the periodicity of the electric per-mittivity tensor ε(r, t) or the magnetic permeability μ(r, t) can be important forthe generation of squeezed states. Only the dispersion of the medium has not beenconsidered. It has been derived that photon pair production is a necessary conditionfor squeezing.

It is tempting to generalize the concept of a Gaussian state of the finite-dimensional harmonic oscillator to the case of an infinite oscillator. Białynicka-Birula and Białynicki-Birula (1987) treat the time development of the Gaussianstates in the free-field case. There the Schrodinger picture is adopted and an ana-logue of the Schrodinger representation in quantum mechanics has been introduced.Let us recall the quadrature representation in quantum optics. This representation isa wave functional Ψ[A, t]. Let us observe that contrary to the operator A(r, t), theargument A(r) of the wave functional does not depend on t , but the wave functionaldoes depend on t . The Hamiltonian in this representation has the form

H = 1

2

∫ {−�

2

ε0

δ2

δA(r)2+ 1

μ0[∇ × A(r)]2

}d3r. (2.71)

In Białynicka-Birula and Białynicki-Birula (1987), the wave functional of the vac-uum state, i.e. the simplest Gaussian state of the electromagnetic field, can be found,as well as that of the “most general”. Thus, the exposition is confined to pureGaussian states while it is possible to generalize it also to mixed Gaussian statesof the electromagnetic field. The pure Gaussian state is determined by a complexmatrix kernel, i.e. by two real matrix kernels. It is shown that the expectation values〈B〉 = B and 〈D〉 = D (equivalently, 〈E〉 = E) evolve according to the free-fieldMaxwell equations and also the equations which the complex matrix kernel obeyscan be found there.

The whole electromagnetic field is treated as a huge infinite-dimensional har-monic oscillator. The wave function and the corresponding Wigner function becomethen functionals of the field variables. Mrowczynski and Muller (1994) have

Page 30: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 21

considered only the scalar field. Białynicki-Birula (2000) starts from the wave func-tional of the vacuum state (Misner et al. 1970)

Ψ0[A] = C exp

[− 1

4π2�

√ε0

μ0

∫ ∫B(r)

1

|r− r′|2 · B(r′) d3r d3r′]

(2.72)

and from the wave functional (we change A →−D)

Ψ0[−D] = C exp

[− 1

4π2�

√μ0

ε0

∫ ∫D(r)

1

|r− r′|2 · D(r′) d3r d3r′]

. (2.73)

The normalization constant C is an issue and it has not been completely solved inBiałynicki-Birula (2000). The analogy with the one-dimensional harmonic oscilla-tor leads to other notions. The Wigner functional of the electromagnetic field in theground state is

W0[A,−D] = exp{−2N [A,−D]}, (2.74)

where

N [A,−D] = 1

4π2�

∫ ∫ [√ε0

μ0B(r)

1

|r− r′|2 · B(r′)

+√

μ0

ε0D(r)

1

|r− r′|2 · D(r′)]

d3r d3r′. (2.75)

The expression (2.75) also plays the role of a norm for the photon wave function(Białynicki-Birula 1996a,b). The Wigner functional for the thermal state of the elec-tromagnetic field has been presented. This state is mixed and it even has infinitelymany photons in the whole field. In each of the subsequent cases, the wave func-tional and the Wigner functional have been introduced. The exception, the mixedstate, has no wave functional. Let us remark that for (the statistical operator of) sucha state the matrix element can be considered which is a functional of two arguments,A and A′.

In particular, the Wigner functional for the coherent state of the electromagneticfield |A,−D〉 has been presented, where A(r), D(r) are the vector potential and theelectric displacement vector, respectively, which characterize the state. The exposi-tion is related to the hot topic of the superpositions of coherent states of the electro-magnetic field. The exposition continues with the Wigner functionals for the statesof the electromagnetic field that describe a definite number of photons. An exampleof the functional for the one-photon state with the photon mode function f(r) hasbeen included.

The norm (2.75) has not been related to any inner product of the photon wavefunctions, but these notions are connected. In contrast to Białynicka-Birula and

Page 31: Quantum Aspects of Light Propagation

22 2 Origin of Macroscopic Approach

Białynicki-Birula (1987), we introduce quadrature operators as

X1[D] =∫

D(r, 0) · f(r) d3r, (2.76)

X2[B] =∫

B(r, 0) · g(r) d3r, (2.77)

where

f(r) = 1

4π2�

∫ √μ0

ε0

1

|r− r′|2 D(r′) d3r′, (2.78)

g(r) = 1

4π2�

∫ √ε0

μ0

1

|r− r′|2 B(r′) d3r′. (2.79)

The commutator of the X1 and X2 operators is

[X1[D], X2[B]] = i�∫

f(r) · [∇ × g(r)] d3r. (2.80)

Let us note that the right-hand sides of (2.78) and (2.79) comprise the operator |∇|−1

up to a certain factor (cf. Milburn et al. 1984). Without resorting to this notation, weobtain that

[X1[D], X2[B]] = i

4�

∫D(r1) ·A(r1) d3r11. (2.81)

We see easily that the usual commutator − 12 i1 is yielded by the field (D,B) (or

(A,−D)) with the property

∫[−D(r1)] ·A(r1) d3r1 = 2�. (2.82)

We have not deepened the contrast by introducing the notation X1[−D] and X2[A]on the left-hand sides of (2.76) and (2.77).

Białynicki-Birula (2000) presents the Wigner functional for the squeezed vacuumstate:

Wsq[A,−D] = exp

[−1

∫ ∫ (√ε0

μ0B ·KBB · B′

+√

μ0

ε0D ·KDD · D′ + B ·KBD · D′

)d3r d3r′

], (2.83)

where KBB, KDD, and KBD are real matrix kernels. The kernel KBD is not inde-pendent of KBB and KDD, but it must obey the condition that is reminiscent

Page 32: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 23

of the Schrodinger–Robertson uncertainty relation (Białynicki-Birula 1998). Theproblem of the time evolution is also discussed. It has been conceded that theWigner function is not a very powerful tool for making detailed calculations. Justas in the field theory, the symmetric ordering is vexed. Another open questionis how the projection of this Wigner functional onto Wigner functions of anyorthogonal (complete or incomplete) modal system looks out. It is appropriate tomention here work concerning the photon wave function (Inagaki 1998, Hawton1999, Kobe 1999), although it is relevant mainly to the electromagnetic field invacuo.

Using a straightforward procedure, Mendonca et al. (2000) have quantized thelinearized equations for an electromagnetic field in a plasma. They have determinedan effective mass for the transverse photons. An extension of the quantization pro-cedure leads to the definition of a photon charge operator. Zalesny (2001) has foundthat the influence of a medium on a photon can be described by some scalar andvector potentials. He has extended the concept of the vector potential to relativisticvelocities of the medium. He has derived formulae for the mass of photon in restingand moving dielectric and the velocity of the photon as a particle.

2.2.4 Source-Field Operator

Knoll et al. (1987) have compared the problem of quantum-mechanical treatmentof action of optical devices with the input–output formalism (Collett and Gardiner1984, Gardiner and Collett 1985, Yamamoto and Imoto 1986, Nilsson et al. 1986, cf.also Gea-Banacloche et al. 1990a,b). Apart from the fact that only a very particularsetup is considered in the input–output formalism, the theory does not take intoaccount the full space–time structure of the field.

Knoll et al. (1987) have elaborated on the approach developed on the basis ofquantum field theory and applied to the problem of spectral filtering of light (Knollet al. 1986). The only assumptions are that the interaction between sources andlight is linear in the vector potential and the optical system is lossless and that thecondition of sufficiently small dispersion is fulfilled. First, the classical Maxwellequations with sources and optical devices are formulated and solved by the proce-dure of mode expansion and the quantized version is derived. The classical Maxwellequations comprise the relative permittivity

ε(r) = n2(r), (2.84)

where n(r) is the space-dependent refractive index. The mode functions Aλ(r) areintroduced as the solutions of equation

∇ × (∇ × Aλ(r))− ε(r)ω2

λ

c2Aλ(r) = 0, (2.85)

Page 33: Quantum Aspects of Light Propagation

24 2 Origin of Macroscopic Approach

where ω2λ is the separation constant for each λ, from which the gauge condition can

be derived

∇ · [ε(r)Aλ(r)] = 0. (2.86)

It is assumed that these solutions are normalized and orthogonal in the sense ofequation

∫ε(r)Aλ(r) · Aλ′(r) d3r = δλλ′ 1. (2.87)

In terms of these functions, the vector potential can be decomposed. In the standardmanner the destruction and creation operators aλ and a†

λ are defined, which have theproperties

[aλ, a†λ′ ] = δλλ′ 1,

[aλ, aλ′ ] = 0 = [a†λ, a†

λ′ ]. (2.88)

On inserting the operators aλ and a†λ into the decomposition of the vector potential,

the operator of the vector potential A(r, t) is defined:

A(r, t) =∑

λ

Aλ(r)[aλ(t)+ a†

λ(t)]. (2.89)

The source quantities ra and pa are considered as the operators ra and pa , whichobey the standard commutation relations

[rka, pk ′a′] = i�δaa′δkk ′ 1,

[rka, rk ′a′] = 0 = [ pka, pk ′a′ ], (2.90)

and the commutation relations

[rka, aλ] = 0 = [rka, a†λ],

[ pka, aλ] = 0 = [ pka, a†λ]. (2.91)

The operator A(r, t) can be used for the derivation of the electric-field strengthoperator which is associated with the radiation field by the relation

E(r, t) = − ∂

∂tA(r, t) (2.92)

and for the derivation of the magnetic field strength operator

B(r, t) = ∇ × A(r, t). (2.93)

Page 34: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 25

Nevertheless, the mode functions are redefined so that they obey the normalizationcondition

∫ε(r)Aλ(r) · Aλ′ (r) d3r = �

2ε0ωλ

δλλ′ . (2.94)

The form of the normalization conditions (2.87) and (2.94) is tailored to real-mode functions and the necessity of modification of some fundamental relationsis commented on by Knoll et al. (1987). All of these field operators may be writtenin the form

F(r, t) =∑

λ

[Fλ(r)aλ(t)+ F∗λ(r)a†

λ(t)]. (2.95)

In dependence on the choice of the operator F(r, t), the functions Fλ(r) can bederived from the mode functions of the vector potential Aλ(r).

It is often convenient to decompose a given field operator F(r, t) into two partsby the relation

F(r, t) = F(+)(r, t)+ F(−)(r, t), (2.96)

where

F(+)(r, t) =∑

λ

Fλ(r)aλ(t), (2.97)

F(−)(r, t) = [F(+)(r, t)]†. (2.98)

Further, the Heisenberg equations of motion for the field operators are derived,so that the field operators can be expressed in terms of free-field and source-fieldoperators. It is typical of the approach of Knoll et al. (1987) that any field oper-ator F (+)

k is decomposed into a free-field operator and a source-field operator asfollows:

F (+)k (r, t) = F (+)

kfree(r, t)+ Fks(r, t), (2.99)

where

F (+)kfree(r, t) =

λ

Fkλ(r)aλfree(t), (2.100)

Fks(r, t) =∫ ∫

θ (t − t ′)Kkk ′(r, t ; r′, t ′) Jk ′(r′, t ′) d3r′ dt ′. (2.101)

Here vector components are labelled by the index k and repeated indices k ′ meansummation.

Page 35: Quantum Aspects of Light Propagation

26 2 Origin of Macroscopic Approach

Unfortunately, the operator aλfree(t) was not defined, so that we may only guessthat aλfree(t)|t=t0 = aλ(t0) for t = t0, and the dynamics for t ≥ t0 can be foundin Knoll et al. (1987). In equation (2.101), the kernel Kkk ′ is defined by therelation

Kkk ′(r, t ; r′, t ′) = − 1

i�

λ

Fkλ(r)A∗k ′λ(r′) exp[−iωλ(t − t ′)]. (2.102)

Inserting equation (2.99) yields the following representation of F (+)k :

F (+)k (r, t)

=∫ ∫

θ (t − t ′)Kkk ′(r, t ; r′, t ′) Jk ′(r′, t ′) d3r′ dt ′ + F (+)kfree(r, t). (2.103)

In particular, if F (+)k is identified with the vector potential A(+)

k , it holds that Fkλ =Akλ and the kernel Kkk ′ takes the form

Kkk ′(r, t ; r′, t ′) = − 1

i�

λ

Akλ(r)A∗k ′λ(r′) exp[−iωλ(t − t ′)]. (2.104)

Analogously, if one is interested in the electric-field strength of the radiation E (+)k ,

the appropriate form of the kernel Kkk ′ is

Kkk ′(r, t ; r′, t ′) = −1

λ

ωλ Akλ(r)A∗k ′λ(r′) exp[−iωλ(t − t ′)]. (2.105)

So the symmetry relations

K ∗kk ′(r, t ; r′, t ′) = ∓Kk ′k(r′, t ′; r, t) (2.106)

are valid for A(+)k and E (+)

k , respectively. The information on the action of the opti-cal instruments on the source field is contained in the space–time structure of thekernel Kkk ′ , which may be regarded as the apparatus function also used in classicaloptics.

Further, the commutation relations for various combinations of field operators atdifferent times are studied and relationships between field commutators and source-quantity commutators are derived. The following abbreviations of the notation areused:

x = {r, t} (2.107)

Page 36: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 27

and others, by which the superscripts +, − are introduced also for Jk(x) andKkk ′(x, x ′). With these generalizations, it holds that

F ( j)k (x) = F ( j)

kfree(x)+ F ( j)ks (x), j = +,−, (2.108)

F ( j)ks (x) =

∫θ (t − t ′)K ( j)

kk ′ (x, x ′) J ( j)k ′ (x ′) dx ′. (2.109)

When appropriate, the time ordering symbols T+ and T− are used. Let us considerany operator product A1(t1) A2(t2)... An(tn). The symbol T+ introduces the timeordering of the operators Ai (ti ) with the latest time to the far left:

T+ A1(t1) A2(t2)... An(tn)

= Ai1 (ti1 ) Ai2 (ti2 )... Ain (tin ) with ti1 > ti2 > · · · > tin , (2.110)

and the symbol T− introduces time ordering of the operators Ai (ti ) with the latesttime to the far right:

T− A1(t1) A2(t2)... An(tn)

= Ai1 (ti1 ) Ai2 (ti2 )... Ain (tin ) with ti1 < ti2 < · · · < tin . (2.111)

From (2.91) it follows that

[F ( j1)k1

(x1), F ( j2)k2

(x2)] = [F ( j1)k1free(x1), F ( j2)

k2free(x2)]

+ D( j1, j2)k1k2

(x1, x2)− D( j2, j1)k2k1

(x2, x1), (2.112)

where

D( j1, j2)k1k2

(x1, x2) = −∫ ∫

θ (t2 − t ′2)θ (t ′2 − t ′1)θ (t ′1 − t1)

⊗K ( j1)k1k ′1

(x1, x ′1)K ( j2)k2k ′2

(x2, x ′2)[ J ( j1)k ′1

(x ′1), J ( j2)k ′2

(x ′2)] dx ′1 dx ′2. (2.113)

From an inspection of equation (2.113), we readily learn that

D( j1, j2)k1k2

(x1, x2) = 0 if t1 > t2. (2.114)

The commutators in (2.112) are

[F ( j)k1free(x1), F ( j)

k2free(x2)] = 0, j = +,−, (2.115)

[F (+)k1free(x1), F (−)

k2free(x2)] = Fk1k2 (x1, x2)1, (2.116)

Page 37: Quantum Aspects of Light Propagation

28 2 Origin of Macroscopic Approach

where

Fk1k2 (x1, x2) =∑

λ

Fk1λ(r1)F∗k2λ

(r2) exp[−iωλ(t1 − t2)]. (2.117)

It would be interesting to find the particular forms of the commutators between Aand E or between A(+) and E (−).

Further, these commutation relations are used to express field correlation func-tions of free-field operators and source-field operators and to describe the effect ofthe optical system on the quantum properties of light fields.

The method of transformation of normal and time orderings is demonstrated forthe following important class of correlation functions:

G(m,n)k1...km+n

(x1, ..., xm+n)

=⟨⎡

⎣T−m∏

j=1

F (−)k j

(x j )

⎣T+m+n∏

j=m+1

F (+)k j

(x j )

⎦⟩

. (2.118)

This transformation is understood in the relation

T+ F (+)k1

(x1)F (+)k2

(x2)

= O+[

F (+)k1free(x1)+ F (+)

k1s (x1)] [

F (+)k2free(x2)+ F (+)

k2s (x2)]. (2.119)

In equation (2.119) and the following ones the ordering symbols O+ and O− areused. The symbol O+ introduces the following ordering of operators F (+)

ki s (xi ),

F (+)k j free(x j ):

(i) Ordering of the operators F (+)ki s (xi ), F (+)

k j free(x j ) with the operators F (+)k j free(x j ) to

the right of the operators F (+)ki s (xi ).

(ii) Substitution of equation (2.109) for the operators F (+)ki s (xi ) and T+ time ordering

of the source-quantity operators Jk ′i (x′i ) in the resulting source-quantity operator

products before performing the integrations with respect to t ′i .

The symbol O− introduces the following operator ordering in products of operatorsF (−)

ki s (xi ), F (−)k j free(x j ):

(i) Ordering of the operators F (−)ki s (xi ), F (−)

k j free(x j ) with the operators F (−)k j free(x j ) to

the left of the operators F (−)ki s (xi ).

(ii) Substitution of equation (2.109) for the operators F (−)ki s (xi ) and T− time ordering

of the source-quantity operators J †k ′i

(x ′i ) in the resulting source-quantity operator

products before performing the integrations with respect to t ′i .

Page 38: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 29

Equation (2.119) may now be generalized:

T+n∏

j=1

F (+)k j

(x j ) = O+n∏

j=1

[F (+)

k j free(x j )+ F (+)k j s (x j )

], (2.120)

T−n∏

j=1

F (−)k j

(x j ) = O−n∏

j=1

[F (−)

k j free(x j )+ F (−)k j s (x j )

]. (2.121)

Using relations (2.118), (2.120), and (2.121), we may represent the correlation func-tions as

G(m,n)k1...km+n

(x1, ..., xm+n)

=⟨⎧⎨

⎩O−m∏

j=1

[F (−)

k j free(x j )+ F (−)k j s (x j )

]⎫⎬

⊗⎧⎨

⎩O+m+n∏

j=m+1

[F (+)

k j free(x j )+ F (+)k j s (x j )

]⎫⎬

⟩. (2.122)

When at the points of observation the following conditions are fulfilled

⟨. . . F (+)

kfree

⟩ = 0 = ⟨F (−)kfree . . .

⟩, (2.123)

then the relation (2.122) can be simplified:

G(m,n)k1...km+n

(x1, ..., xm+n)

=⟨⎡

⎣O−m∏

j=1

F (−)k j s (x j )

⎣O+m+n∏

j=m+1

F (+)k j s (x j )

⎦⟩

. (2.124)

When written in more detail, into the relation (2.124), the complex kernels Kk j k ′j (x j ,

x ′j ) are introduced.It is noted that the effect of the beam splitter that is used for mixing of source

light with the reference beam in the case of homodyne detection is described by theassumption that the reference light beam is a free field. In Knoll et al. (1987) therelation (2.122) is specialized to a multimode coherent free field |{αλ}〉,

F (+)kfree(x)|{αλ}〉 = Fk(x)|{αλ}〉, (2.125)

Page 39: Quantum Aspects of Light Propagation

30 2 Origin of Macroscopic Approach

that is to say

G(m,n)k1...km+n

(x1, ..., xm+n)

=⟨⎧⎨

⎩O−m∏

j=1

[F∗

k j(x j )1+ F (−)

k j s (x j )]⎫⎬

⊗⎧⎨

⎩O+m+n∏

j=m+1

[Fk j (x j )1+ F (+)

k j s (x j )]⎫⎬

⟩. (2.126)

Finally, the theory is applied to the photocount statistics. Following Glauber’stheory of photon detection (Glauber 1965, Kelley and Kleiner 1964), the probabilityof observing precisely n events in a counting time interval [t, t+Δt) is given by therelation

pn(t,Δt) =⟨Ω

1

n!

[Γ(t,Δt)

]nexp

[−Γ(t,Δt)]⟩

, (2.127)

where

Γ(t,Δt) =∑

i

∫ t+Δt

t

∫ t+Δt

tS(t1 − t2)E (−)

k (ri , t1)E (+)k (ri , t2) dt1 dt2 (2.128)

may be interpreted as the operator of the integrated intensity. Here ri are positionvectors of the detector atoms and S(t) is a response function. Let us note that oneusually assumes that

S(t1 − t2) = ηδ(t1 − t2), (2.129)

with some η. In relation (2.127), the ordering symbol Ω introduces the followingoperator ordering:

(i) The normal ordering of the operators E (−)k (x), E (+)

k (x) with the operatorsE (−)

k (x) to the left of the operators E (+)k (x).

(ii) T+ ordering of the operators E (+)k (x) and T− ordering of the operators E (−)

k (x).

In analogy with (2.122), relation (2.127) becomes

pn(t,Δt) =⟨O 1

n!

[Γ(t,Δt)

]nexp[−Γ(t,Δt)]

⟩, (2.130)

Page 40: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 31

where the Ω ordering is simply replaced by the O ordering defined as follows:

(i) The normal ordering of the operators E (−)ks (x), E (−)

kfree(x), E (+)ks (x), E (+)

kfree(x) withthe operators E (−)

ks (x), E (−)kfree(x) to the left of the operators E (+)

ks (x), E (+)kfree(x).

(ii) O+ ordering of the operators E (+)ks (x), E (+)

kfree(x) and O− ordering of the opera-tors E (−)

ks (x), E (−)kfree(x).

The fulfilling of the conditions (2.123) causes a modification of relation (2.128) asfollows:

Γ(t,Δt) =∑

i

∫ t+Δt

t

∫ t+Δt

tS(t1 − t2)E (−)

ks (ri , t1)E (+)ks (ri , t2) dt1 dt2. (2.131)

In the case of mixing the source field light with a coherent free-field referencebeam, there is an analogy with the relation (2.126):

Γ(t,Δt) =∑

i

∫ t+Δt

t

∫ t+Δt

tS(t1 − t2)

×[E∗k (ri , t1)1+ E (−)

ks (ri , t1)] [

Ek(ri , t2)1+ E (+)ks (ri , t2)

]dt1 dt2. (2.132)

A generalization of the Wick theorem on transforming a time-ordered product ontoa sum of normally ordered terms was performed by Agarwal and Wolf (1970).

The quantum theory of the radiation field interacting with atomic sources inthe presence of a linear, dispersionless, and absorptionless dielectric with space-dependent refractive index has been applied to the description of the action of aresonator-like cavity with input–output coupling and filled with an active medium(Knoll and Welsch 1992).

2.2.5 Continuum Frequency-Space Description

Blow et al. (1990) have formulated the quantum theory of optical wave propagationwithout recourse to cavity quantization. This approach avoids the introduction ofa box-related mode spacing and enables one to use a continuum frequency-spacedescription. In this chapter and in that by Blow et al. (1991) a continuous-modequantum theory of electromagnetic field has been developed. As usual in the quan-tum field theory, the box-related modes are considered whose creation and destruc-tion operators satisfy the usual independent boson commutation relations:

[ai , a†j ] = δi j 1. (2.133)

Different modes of the cavity, labelled by i and j , have frequencies given by dif-ferent integer multiples of the mode spacing Δω. The mode spectrum becomes

Page 41: Quantum Aspects of Light Propagation

32 2 Origin of Macroscopic Approach

continuous as Δω → 0 and in this limit the transformation to continuous-modeoperators is convenient:

ai →√

Δω a(ω). (2.134)

A complete orthonormal set of functions was considered which may describe statesof finite energy. The set is numerable infinite and to each function in it a destructionoperator is assigned. Such operators have all the usual properties of the operators ofthe monochromatic mode.

Further specific states of the field have been treated such as coherent states, num-ber states, noise and squeezed states. With the use of noncontinuous operators, ageneralization of the single-mode normal ordering theorem was proved. Field quan-tization in a dielectric has been treated including the material dispersion and thetheory has been applied to the pulse propagation in an optical fibre. A comparisonwith results by Drummond (1990, 1994) would be in order.

Let us consider the fields in a lossless dielectric material with the real relativepermittivity ε(ω) and the refractive index n(ω) related by

ε(ω) = [n(ω)]2. (2.135)

Let us recall the definition of the phase velocity

vF(ω) = ω

k= c

n(ω)(2.136)

and that of the group velocity vG(ω)

1

vG(ω)= ∂k

∂ω= 1

c

∂ω[ωn(ω)]. (2.137)

The normalization of the field operators is fixed by requirement that the normallyordered total energy density operator U (z, t) has the diagonal form:

Hfree = A∫ ∞

−∞U (z, t) dz =

∫�ωa†(ω)a(ω) dω. (2.138)

The field operators are obtained in accordance with the relation

E (+)(z, t) = − ∂

∂tA(+)(z, t), B(+)(z, t) = ∂

∂zA(+)(z, t) (2.139)

and with the expansion of the vector-potential operator

A(+)(z, t) =∫ ∞

−∞

√�vG(ω)

4πε0cωn(ω)A

×∑

λ=1,2

ε(k, λ)a(k, λ) exp[−i(ωt − kz)] dk. (2.140)

Page 42: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 33

Noting that

dk = dω

vG(ω), a(k, λ) =

√vG(ω)a(ω), (2.141)

and taking the polarization to be parallel to the x-axis, it follows from (2.139) thatthe field operators are

E (+)(z, t) = i∫ √

�ω

4πε0cAn(ω)

× a(ω) exp

{−iω

[t − n(ω)z

c

]}dω (2.142)

and

B(+)(z, t) = i∫ √

�ωn(ω)

4πε0c3A

× a(ω) exp

{−iω

[t − n(ω)z

c

]}dω. (2.143)

Alternatively, the propagation constant can be expanded to the second order infrequency and a partial differential equation can be obtained (cf. Drummond 1990).Assuming a narrow bandwidth, the slowly varying field envelope can be representedby the operator a(z, t), which obeys the equation

i∂

∂za(z, t)+ k ′′

2

∂2

∂t2a(z, t) = 0, (2.144)

where k ′′ is the second derivative with respect to the frequency of the propagationconstant, evaluated at the central frequency. The equation has been simplified usingthe transformation of envelope into a frame moving with the group velocity. This isnecessary for the envelope to be slowly varying. In the classical nonlinear optics thestationary fields have also envelopes, but they seem to be defined otherwise.

The treatment of this problem in the noncontinuous basis proceeds from thereplacement

a(z, t) =∑

j

φ j (z, t)c j , (2.145)

where φ j (z, t) are a complete orthonormal set of functions on z and c j are destruc-tion operators obeying the usual commutation relations. The advantage of this treat-ment is that the functional dependence on z and t is contained in the c-numberfunctions rather than the operators a(z, t) as in the propagation equation (2.144)for example. It is not emphasized by Blow et al. (1990) that the solution of

Page 43: Quantum Aspects of Light Propagation

34 2 Origin of Macroscopic Approach

equation (2.144) preserves the equal-space, not equal-time, commutators. Similarly,the set of functions φ j (z, t) enjoys the orthonormality and completeness only as theequal-space, but not equal-time, properties. The propagation equation (2.144) nowyields the following equations for the noncontinuous basis functions:

i∂

∂zφ j (z, t)+ k ′′

2

∂2

∂t2φ j (z, t) = 0. (2.146)

Finally, the process of photodetection in free space is considered and the resultsapplied to homodyne detection with both local oscillator and signal fields pulsed.The results of sets of measurements in which the photocurrent is integrated overperiods T can be predicted by the use of an operator

M =∫ τ+T

τ

a†(t)a(t) dt. (2.147)

Here τ is the start time of the measurements, the detector is placed at z = 0, and

a(t) = 1√2π

∫a(ω) exp(−iωt) dω. (2.148)

Let us further consider a balanced homodyne detector in which the light beam understudy is superposed on a local oscillator by combining them at a 50:50 beam splitter.The measured quantity is the difference in the photocurrents of two detectors placedin the output arms of the beam splitter and it can be represented by the operator(Collett et al. 1987)

O = i∫ τ+T

τ

[a†(t)aL(t)− a†L(t)a(t)] dt, (2.149)

where a†L(t) and aL(t) are the continuum creation and destruction operators of the

local oscillator field and a†(t) and a(t) correspondingly for the signal field.For homodyne detection of pulsed signals it is advantageous to use a pulsed local

oscillator. The pulsed signal is described by the noncontinuous basis function φ0(t)and the local oscillator is described by a normalized function φL(t), the field of thelocal oscillator being in the coherent state |{αL(t)}〉, where

αL(t) =√

NL exp(iθL)φL(t), (2.150)

with NL the mean total number of photons in the pulse and θL the externally con-trolled local oscillator phase. Let us recall the definition of a coherent state:

|{α(t)}〉 = D({α(t)})|0〉, (2.151)

with

D({α(t)}) = exp{ ∫

[α(t)a†(t)− α∗(t)a(t)]}, (2.152)

Page 44: Quantum Aspects of Light Propagation

2.2 Nondispersive Lossless Linear Dielectric 35

which is close enough to that by Blow et al. (1990) except for the exchange of spacefor time. It is assumed that the signal field is described by a set of noncontinuousoperators di at the output of a nonlinear system and the signal field at the input tothe system is described by a similar set of operators ci . The action of the nonlinearsystem is defined by the relations

d0 = μc0 + νc†0,

di = ci , i > 0. (2.153)

In the relation (2.149) it is necessary to substitute

a(t) =∑

i

φi (t)di . (2.154)

In analogy, we consider

aL(t) =∑

i

φiL(t)ciL, (2.155)

where the subscript L only modifies the familiar meanings and φ0L(t) = φL(t).It is shown how the formulation of the quantum field theory is modified for the

one-dimensional optical system. The fields are defined in an infinite waveguideparallel to the z-axis, but of finite cross-sectional area A of the rectangular formwith sides parallel to the x- and y-axes. The x and y wave-vector components arethus restricted to discrete values and any three-dimensional integral over this spatialregion is converted according to

∫d3k → (2π )2

A∑

kx ,ky

∫dkz . (2.156)

On the assumption that the modes with kx �= 0 or ky �= 0 are vacuum ones, areduced Hilbert (namely Fock) space can be exploited. The summation in (2.156)can, therefore, be removed and putting kz = k, the other conversions are

δ(3)(k− k′) → A(2π )2

δ(k − k ′), (2.157)

a(k, λ) →√A

2πa(k, λ). (2.158)

The vector-potential operator has been modified for the dispersive lossless mediumand compared with Drummond (1990) and Loudon (1963), the positive-frequencypart is

Page 45: Quantum Aspects of Light Propagation

36 2 Origin of Macroscopic Approach

A(+)(r, t) =∫ √

�vG(ω)

16π3ε0cωn(ω)

×∑

λ=1,2

ε(k, λ)a(k, λ) exp[−i(ωt − k · r)] d3k, (2.159)

ω = c

n(ω)|k|. (2.160)

The expression (2.159) can be converted to the one-dimensional form easily (asindicated above, cf. (2.140)).

McDonald (2001) has considered a variation of the physical situation of “slowlight” to show that the group velocity can be negative at central frequency. A Gaus-sian pulse can emerge from the far side of a slab earlier than it hits the near sideand the pulse emission at the far side is accompanied by an antipulse emission, theantipulse propagating within the slab so as to annihilate the incident pulse at thenear side.

2.3 Quantum Description of Experiments with Stationary Fields

Burnham and Weinberg (1970) found that the measured value of the correlationtime between the two optical photons produced in a parametric process was verysmall, an effect of a practical interest. Laboratory techniques for doing experi-ments with single photons also have advanced. Since 1985, such photon pairshave become familiar for the study of nonclassical aspects of light (Horne et al.1990).

The process of optical parametric three-wave mixing in a second-order nonlinearmedium consists of the coherent interaction between pump, signal, and idler waves.This process may occur as frequency down conversions, specifically as an opticalparametric oscillation and an optical parametric amplification. In a travelling-wavesetting, the optical parametric generation is called a spontaneous parametric down-conversion. The photon pairs (biphotons) produced in parametric down-conversionare useful in experiments concerning fundamental questions of quantum theory. Thedescription of experiments has been facilitated by studies of Campos et al. (1990).The autocorrelation and cross-correlation properties of the signal and idler beamsproduced in the parametric down-conversion have been studied, e.g. in Joobeur et al.(1994). A unified treatment of the experiment on the interference of a “biphoton withitself” and of other three experiments has been provided by Casado et al. (1997a).A fourth-order interference has been obtained in the four cases, and the uniformityhas been achieved also by the use of the Wigner (or Weyl) representation of thefield operators. A similar treatment of the famous experiment and of another onehas been presented by Casado et al. (1997b). A second-order interference has been

Page 46: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 37

treated in the two cases and the stochastic properties of the pump beam have beenrespected.

The studies of the fundamentals of quantum mechanics underlie such interest-ing applications as quantum cryptography and quantum computing (Bowmeesteret al. 2000). The experiments have become very popular (Shih 2003). Designof experiments for undergraduate students has become feasible (Galvez et al.2005).

Here we return from the Wigner to the Hilbert-space formalism as in Perinovaand Luks (2003). First we consider the three-dimensional expansion of the operatorof a chosen component of the electric vector after Casado et al. (1997a). As inthe schematics of the experiments the field is restricted to paths leading to detec-tors, we introduce one-dimensional expansions of the electric-field operator. Weattempt to consider orthogonal modal functions, although we cannot define themeverywhere, but only on the paths. We are aware of the dangerous position, whereone cannot evaluate the orthogonality property for the lack of a complete definition.In this approach we do not start with the description of the process of parametricdown-conversion from a Hamiltonian, but with the response of the output fields of anonlinear crystal to the input fields (Casado et al. 1997a) when two paths cross sucha crystal. Such a response depends also on stochastic properties of the pump beam,which is assumed to be monochromatic however. The experiment on the interferenceof signal and idler photons (Ghosh et al. 1986) can do with the simple description,when the lack of the second-order interference is derived. In the use of two detectorswe consider four paths and modify (double) the description. Nevertheless, we do notreproduce the well-known result.

Similarly we proceed in the case of the experiment of Rarity and Tapster (1990),which was also used to test Bell’s inequality using phase and momentum. In con-trast, in the case of the experiment of Franson (1989) we are allowed to return to thesimple description, as essentially two paths are involved, even though the schematicis more complicated. This experiment was proposed in order to test a Bell inequalityfor energy and time.

Next we deal with induced coherence and indistinguishability in two-photoninterference (Zou et al. 1991). In this case the schematic comprises two nonlinearcrystals, the number of paths is greater, but since two paths belong to each crystal,the simple description is appropriate. The lack of induced emission made it a “mind-boggling” experiment (Greenberger et al. 1993), but the indistinguishability of thepaths along which the signal photon arrives at the detector (in fact, the biphotonarrives at the two detectors) is still held for the reason of interference. We mayrefer to Casado et al. (1997b), where stochastic properties of the pump beam aretaken into account. Two experiments are analysed: frustrated two-photon creationby interference, and induced coherence and indistinguishability. Coincidences arenot studied and a second-order interference has been obtained in the two cases.Last we mention the frustrated two-photon creation via interference (Herzog et al.1994) restricting ourselves to the second-order interference and the monochromaticpump.

Page 47: Quantum Aspects of Light Propagation

38 2 Origin of Macroscopic Approach

2.3.1 Spatio-temporal Descriptions of ParametricDown-Conversion Experiments

In the Hilbert-space representation of the light field, the electric vector is representedas a sum of two mutually conjugate operators (Casado et al. 1997a)

E(r, t) = E(+)(r, t)+ E(−)(r, t), (2.161)

E(+)(r, t) = i∑

k,λ

√�ωk

2L3εk,λak,λ(t)eik·r, (2.162)

where L3 is the normalization volume, ak,λ(t) is the annihilation operator for aphoton whose wave vector is k and whose polarization vector is εk,λ, and ωk =c|k|. Equations (2.161) and (2.162) correspond to the Heisenberg picture, where alltime dependence of the averages comes from the creation and annihilation oper-ators a†

k,λ(t) and ak,λ(t). In this picture the state of the field is represented by atime-independent statistical operator ρ.

As we do not study experiments involving polarizing devices, we find it conve-nient to use a scalar approximation well known in classical optics. When Casadoet al. (1997a) use the subscripts on the (Wigner representations of) field opera-tors they indicate that the light beam contains frequencies within a range and that“transverse” components of wave vectors are limited by small upper values. Webelieve that such subscripts indicate which part of the field is considered.

The laser theory and, in general, the theory of resonators connect the quantumfield with the annihilation operators not via the complex exponentials, but via moregeneral modal functions, which are often related to the device. We suppose thatsuch an approach can be interesting also in our study, after we find modal functionsthat are connected to the linear devices used and to the mirrors. Obviously, the freeevolution of operators ak0 (zeroth-order solution) is transformed into a kind of lineardynamics of the “relevant” component E (+)

ss 0 (r, t) of the electric vector via the appro-priate modal functions, with ss being any subscript. This process can be formalizedby a quadratic Hamiltonian, which differs from the free-field Hamiltonian only bythe meaning of the creation and annihilation operators. We restrict ourselves to theoperator ρ that represents a vacuum state.

We suppose that one or two nonlinear crystals involved in the experiment aredescribed in terms of interaction Hamiltonians. The action of the scattering operatoron the initial field can be “guessed”. The interaction lasting only for a short time andbeing spatially confined to the medium suggests to us an appropriate modification ofthe linear dynamics. We modify also the notation for the resulting field by omittingthe initial subscript 0.

(i) The process of parametric down-conversion

We are going to study the process of parametric down-conversion of light inthe Hilbert-space representation. We refer to any of our figures for a sketch of the

Page 48: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 39

setup used for parametric down-conversion. A nonlinear crystal is pumped by a laserbeam V , producing a continuum of coloured cones around the axis defined by thepump. In experimental practice two narrow correlated beams, called “signal” Es and“idler” Ei, are selected by means of apertures, filters, or just the detectors. Let ustake the origin of the coordinate system, 0 ≡ 01, at the centre of the crystal. Wetreat the pump beam as an intense monochromatic wave represented, in the scalarapproximation, by

V (r, t) = V ei(k0·r−ω0t) + c.c., (2.163)

where V is a complex amplitude of a pump beam, ω0 is a frequency of the pumpbeam, k0 is an appropriate wave vector, and c.c. means the complex conjugate termto the previous one. In a product with the identity operator it may be added to theelectric-field operator.

Now, let us consider two narrow correlated beams, called signal and idler, withaverage frequencies ωs, ωi and wave vectors ks, ki, respectively, fulfilling the match-ing conditions

ωs + ωi = ω0, ks + ki = k0. (2.164)

The response E (+)s (r, t) and E (−)

i (r, t) of a nonlinear crystal to the input fieldsE (+)

0s (r, t) and E (−)0i (r, t) is as

E (+)s (r, t) = ( ˆ1+ g2|V |2 ˆJ )E (+)

0s (r, t)+ e−iω0t gV ˆG E (−)0i (r, t),

E (−)i (r, t) = eiω0t gV ˆG† E (+)

0s (r, t)+ ( ˆ1+ g2|V |2 ˆJ )E (−)0i (r, t), (2.165)

where g is an effective coupling constant, ˆ1 is the identity superoperator, ˆG andˆJ are antilinear and linear superoperators, respectively, which substitute expan-

sions in annihilation (creation) operators for annihilation (creation) operators ( ˆJ

yields an expansion in the annihilation operators in the first equation and ˆG yields

an expansion in the creation operators in the same equation), and ˆG† E (+)0s (r, t) ≡

[ ˆG E (−)0s (r, t)]†. The relation (2.165) can be modified (doubled) so that it relates out-

put fields E (+)s j

(r, t), E (−)i j

(r, t), j = 1, 2, to input fields E (+)0s j

(r, t), E (−)0i j

(r, t), j =1, 2. Since a pointlike crystal is considered (Casado et al. 1997b), it may be interest-ing to imagine Equations (2.165) at r = 0 without the subscript 0 on the right-handside. It can occur, but at the cost of other notation. The interaction does not changethe field just in front of the crystal, so we can interpret the initial field as the “in”resulting field. As it is almost at the centre of the crystal, it differs negligibly fromthe initial field just behind the crystal, which becomes the “out” resulting field.

In order to determine the detection probabilities in the Hilbert-space representa-tion, we adopt the correlation properties (Casado et al. 1997a). In such a work it hasproved convenient to substitute slowly varying amplitudes F (+)

J (r, t) [F (−)J (r, t)] for

Page 49: Quantum Aspects of Light Propagation

40 2 Origin of Macroscopic Approach

the amplitudes E (+)J (r, t) [E (−)

J (r, t)], the relation between them being

F (+)J (r, t) = eiωJ t E (+)

J (r, t), J = s, i. (2.166)

According to Casado et al. (1997a) it is essential to use the following relation,which is still an approximation:

F (+)(rb, t) = F (+)(

ra, t − rab

c

)exp

(iωa

rab

c

), (2.167)

where ωa is some frequency appropriate to a light beam and rab = ea ·(rb−ra), withea being the unit vector in the direction of propagation. Since the vectors whose dotproduct is taken are usually of the same direction, the magnitude of displacementvector may be evoked.

If we consider the signal beam emerging from the crystal at different times t andt ′, we can use the autocorrelations (Casado et al. 1997a):

〈F (−)J (r, t)F (+)

J (r, t ′)〉 = g2|V |2μJ (t ′ − t), J = s, i. (2.168)

The following autocorrelations, and their complex conjugates, vanish:

〈F (+)J (r, t)F (+)

J (r, t ′)〉 = 0, J = s, i. (2.169)

The relation holds at any point of the outgoing beam, most interestingly just behindthe crystal.

With respect to the cross correlation, we prefer to characterize the signal andidler field operators just behind the crystal at different times (Casado et al. 1997a):

〈F (+)s out(0, t)F (+)

i out(0, t ′)〉 = gV ν(t ′ − t). (2.170)

It is useful to know that

〈F (+)s (r, t)F (−)

i (r, t ′)〉 = 〈F (−)s (r, t)F (+)

i (r, t ′)〉 = 0. (2.171)

In the Hilbert-space formalism, the usual theory of detection (by photon absorp-tion) is based on the normal ordering. The joint detection rate is given by

Pab(ra, t ; rb, t ′) = K ′〈0|E (−)(ra, t)E (−)(rb, t ′)E (+)(rb, t ′)E (+)(ra, t)|0〉 (2.172)

in the Heisenberg picture, where K ′ = Ka Kb and Ka (Kb) is a constant related tothe efficiency of the detector and the energy of a single photon. The well-knownproperty of Gaussian random variables A, B, C , and D,

〈ABC D〉 = 〈AB〉〈C D〉 + 〈AC〉〈B D〉 + 〈AD〉〈BC〉, (2.173)

Page 50: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 41

applies not only in the Weyl (Casado et al. 1997a) but also in the normal orderingand entails that the joint detection rate is written in three terms. The first two termsare fourth order in g, while the last term is second order in g. We may discard thefirst two terms (Casado et al. 1997a) and finally obtain

Pab(ra, t ; rb, t ′) = K ′|〈E (+)(ra, t)E (+)(rb, t ′)〉|2. (2.174)

We will determine the visibility V ′ of the intensity interference:

V ′ = Rabmax − Rabmin

Rabmax + Rabmin, (2.175)

where

Rabmax =∫ w

2

− w2

Pabmax(τ ′) dτ ′, Rabmin =∫ w

2

− w2

Pabmin(τ ′) dτ ′, (2.176)

with w the coincidence window which we choose to be w = 13 × 10−9 s, definedin terms of the integral

M

(d

c,

h

c

)= K ′

∫ w2

− w2

∣∣∣∣ν(

τ ′ + d

c

)∣∣∣∣

∣∣∣∣ν(

τ ′ + h

c

)∣∣∣∣ dτ ′

= exp

[−σ 2

2

(h − d

c

)2]

× 1

2

{erf

[σ√

2

(d + h

c+ w

)]∓ erf

[∓ σ√

2

(w − d + h

c

)]}.

(2.177)

Here

erf(x) = 2√π

∫ x

0e−t2

dt, (2.178)

d, h are parameters of an experimental setup, c = 2.998× 108 ms−1 is the speed oflight, ν(τ ) is a Gaussian,

ν(τ ′) = |ν(τ ′)| =√

1

K ′

(2

π

)14 √

σe−σ 2τ ′2 , (2.179)

where σ = 1012 s−1. Let us remember that for σ−1 � w, we have erf(±∞) = ±1.Particularly,

M

(d

c,

d

c

)= 1

2

{erf

[σ√

2

(2d

c+ w

)]+ erf

[σ√

2

(w − 2d

c

)]},

M(0, 0) = erf

(σ√

2w

). (2.180)

Page 51: Quantum Aspects of Light Propagation

42 2 Origin of Macroscopic Approach

(ii) Experiment on the interference of signal and idler photons

Let us start with an experiment demonstrating the coherence properties of theparametric down-conversion photon pairs as proposed in Ghosh et al. (1986). It isassumed that ωs = ωi = ω0

2 and the signal and idler beams are directed to a screenby means of two mirrors. There is no second-order interference between the twobeams. When two detectors are put on the screen one can show a fourth-order, orintensity–intensity, interference. As seen from Fig. 2.1, the tracing of the beams isnot so evident, and we modify the well-known result (Casado et al. 1997a, Ghoshet al. 1986). A report on the experiment was brief (Ghosh and Mandel 1987).

Fig. 2.1 Experimental setupon interference on a screen

In what follows, we will specify modal functions and nonlinear dynamics of fieldoperators. We introduce the notation for the points of reflection 0Ms j

= (b, 0, zMs j),

0Mi j= (−b, 0, zMi j

), j = 1, 2, where

zMs j= bd

2b − x j, zMi j

= bd

2b + x j, j = 1, 2, (2.181)

with d being the distance from the centre of the crystal to the screen and b being thedistance from the axis of the pumping to the mirrors.

We consider the initial electric field in the form

E (+)0 (r, t) = V (+)(r, t)1+

2∑

j=1

[E (+)

s j 0(r, t)+ E (+)

i j 0(r, t)

], (2.182)

where r = (x, y, z), k = (kx , ky, kz), and

E (+)s j 0

(r, t) =∑

k∈[k]s j

vs j k(r)as j k0(t), (2.183)

E (+)i j 0

(r, t) =∑

k∈[k]i j

vi j k(r)ai j k0(t), (2.184)

Page 52: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 43

with the orthonormal systems of functions vs j k(r), k ∈ [k]s j , vi j k(r), k ∈ [k]i j ,j = 1, 2,

‖vs j k(r)‖2 =∫|vs j k(r)|2 d3r = �ωk, k ∈ [k]s j , (2.185)

‖vi j k(r)‖2 =∫|vi j k(r)|2 d3r = �ωk, k ∈ [k]i j . (2.186)

The [k]s j is a set of integer multiples of the vector 2πL es j , es j being a unit vector of

the signal beam at the origin. Similarly for [k]i j . A formal expression for vJk(r),J = s j , i j , j = 1, 2, is of the form

vJk(r) = i

√�ωk

ALexp(ik · r) for r ‖ k, z < zMJ , k ∈ [k]J ,

vJk(r) = −i

√�ωk

ALexp[ik′ · (r− 0′J )] for (r− 0′J ) ‖ k′,

z > zMJ , k ∈ [k]J , (2.187)

where J = s j , i j , j = 1, 2, A is the effective transverse area of the beam, 0′J =(2b, 0, 0) for J = s1, s2, 0′J = (−2b, 0, 0) for J = i1, i2, k′ = (−kx , ky, kz). Ina standard fashion, we associate the signal and idler modal functions (2.187) withfields we denote as E (+)

J0 (r, t), J = s j , i j , j = 1, 2.After switching on the nonlinear interaction, part of the field is not influenced:

E (+)s j

(r, t) = E (+)s j 0

(r, t), E (+)i j

(r, t) = E (+)i j 0

(r, t) for z < 0, (2.188)

whereas for z > 0 provided that g|V | � 1, the perturbative approximation of thesolution of the Heisenberg equations of motion that retains terms up to g2 can bewritten as

E (+)s j

(r, t) = E (+)s j 0

(r, t)+ e−iω0t gV ˆG j E (−)i j 0

(r, t)+ g2|V |2 ˆJj E (+)s j 0

(r, t), (2.189)

E (+)i j

(r, t) = E (+)i j 0

(r, t)+ e−iω0t gV ˆG j E (−)s j 0

(r, t)

+ g2|V |2 ˆJj E (+)i j 0

(r, t), j = 1, 2, (2.190)

where

E (−)s j 0

(r, t) = [E (+)s j 0

(r, t)]†, E (−)i j 0

(r, t) = [E (+)i j 0

(r, t)]†, (2.191)

and ˆG j and ˆJj are antilinear and linear superoperators, respectively, which substitute

the expansions in annihilation operators ( ˆJj yields an expansion in the annihilation

operators) for annihilation (creation) operators ( ˆG j yields an expansion in creationoperators). Compare Casado et al. (1997a), where G j and Jj are appropriate linear

Page 53: Quantum Aspects of Light Propagation

44 2 Origin of Macroscopic Approach

operators acting on functions of the argument r for complex-valued functions. The

superoperators ˆG j and ˆJj have the properties

ˆG j a j0k(t) =∑

[k′]s j

f (k, k′)u[Δt

2(ω0 − ωk − ωk′)

]

× exp [it(ω0 − ωk − ωk′)] a†j0k′ (t) for k ∈ [k]i j , (2.192)

ˆJj a j0k(t) =∑

[k′]i j

[k′′]s j

f (k, k′) f ∗(k′, k′′)

× u

[Δt

2(ωk′ + ωk′′ − ω0)

]u

[Δt

2(ωk′′ − ωk)

]exp [it(ωk′′ − ωk)]

× a j0k′′ (t) for k ∈ [k]s j , (2.193)

respectively, with

u(x) = sin x

xeix , (2.194)

a j0k(t) = a j0k(0)e−iωkt . (2.195)

Supposing that in the sense of classical nonlinear optics, f (k, k′) is a distributionwith a support determined by the condition

ω0 − ωk − ωk′ = 0, (2.196)

we easily obtain that

ˆG j a j0k(t) =∑

[k′]s j

f (k, k′)a†j0k′ (t) for k ∈ [k]i j , (2.197)

ˆJj a j0k(t) =∑

[k′′]s j

⎣∑

[k′]i j

f (k, k′) f ∗(k′, k′′)

⎦ a j0k′′ (t) for k ∈ [k]s j , (2.198)

which is a great unexpected simplification.Further we will express the intensity correlations that have been determined in

the experiment. Introducing

F (+)s j

(r, t) = exp (iωst)E (+)s j

(r, t), (2.199)

we express the field at a point r j , j = 1, 2, on the screen as

F (+)(r j , t) = F (+)s j out

(0, t − rs j

c

)exp

(iω0rs j

2c

)

+ F (+)i j out

(0, t − ri j

c

)exp

(iω0ri j

2c

), j = 1, 2, (2.200)

Page 54: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 45

where

rs j = rs(x j ) =√

z2Ms j

+ b2 +√

(d − zMs j)2 + (b − x j )2 , j = 1, 2,

ri j = ri(x j ) =√

z2Mi j

+ b2 +√

(d − zMi j)2 + (b + x j )2 , (2.201)

and the subscript out indicates that the field behind the nonlinear crystal is consid-ered. Hence, we can obtain the relations (2.209) below.

By taking into account the correlation relations

〈F (+)s1out(0, t)F (+)

i2out(0, t ′)〉 = 〈F (+)i1out(0, t ′)F (+)

s2out(0, t)〉

= gV ν(t ′ − t), (2.202)

we get (cf. Casado et al. 1997a)

〈F (+)(r1, t1)F (+)(r2, t2)〉

= gV

{ν(

t2 − t1 + rs1

c− ri2

c

)exp

[iω0

2c(ri2 + rs1 )

]

+ ν(

t1 − t2 + rs2

c− ri1

c

)exp

[iω0

2c(ri1 + rs2 )

] }. (2.203)

Assuming that the beams with different subscripts j are mutually uncorrelated, wefinally get (cf. Casado et al. 1997a)

P12(r1, t + τ1; r2, t + τ2) ≈ K ′g2|V |2

×{ ∣∣∣ν

(τ2 − τ1 + rs1

c− ri2

c

)∣∣∣2+∣∣∣ν(τ1 − τ2 + rs2

c− ri1

c

)∣∣∣2

+ 2Re

[ν(τ2 − τ1 + rs1

c− ri2

c

)ν∗(τ1 − τ2 + rs2

c− ri1

c

)

× exp[iω0

2c(ri2 + rs1 − ri1 − rs2 )

] ]}, (2.204)

where K ′ is a constant related to the efficiency of the detectors,

K ′ = K1 K2, K1 = 2η1

�ω0, K2 = 2η2

�ω0, (2.205)

ηJ , J = 1, 2, is the efficiency of the detector DJ .

Page 55: Quantum Aspects of Light Propagation

46 2 Origin of Macroscopic Approach

The visibility is expressed by the formula (2.175), where

Rsimax + Rsimin = 2g2|V |2

×[

M

(rs1 − ri2

c,

rs1 − ri2

c

)+ M

(ri1 − rs2

c,

ri1 − rs2

c

)], (2.206)

Rsimax − Rsimin = 4g2|V |2 M

(rs1 − ri2

c,

ri1 − rs2

c

). (2.207)

We may ask whether the visibility has its proper meaning for all x1, x2, whether itis associated only with the extremes of the detection rate, i.e. x1 and x2 for whichRsi = Rsimax or Rsi = Rsimin. By relation (2.204) and the choice (2.179) we areinterested in C = ±1, where

C = cos

[ω0

2

(ri2 + rs1

c− ri1 + rs2

c

)], (2.208)

with ω0 = 2πc351 × 109 Hz.

The original formulae for rs j , ri j (instead of the relation (2.201)) were as (Ghoshet al. 1986)

rs j = rs(x j ) =√

(2b − x j )2 + d2 , j = 1, 2,

ri j = ri(x j ) =√

(2b + x j )2 + d2 . (2.209)

In Fig. 2.2 a short period of the oscillations of the cosine is depicted after Ghoshet al. (1986). An equal phase is assumed on any of the lines y2 ≡ x2−x1 = constant.The short oscillation period corresponds to the change in the signed distance

Fig. 2.2 The dependence ofthe cosine C of the phase onthe position coordinates x1

and x2 for d = 1 and b = 0.1.The analysis of the setup inFig. 2.1 after Ghosh et al.(1986). The cosine C of thephase depends only on thesigned distance of thedetectors

Page 56: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 47

y2 ≡ x2 − x1. As obvious from Fig. 2.3, the complement of the visibility, 1 − V ′,depends only on the coordinate of the point amid the detectors D1 and D2. An equalvisibility V ′ is assumed on the lines y1 = 1

2 (x1 + x2) = constant. For y1 = 0.0015the visibility almost vanishes.

Fig. 2.3 The complement ofthe visibility, 1− V ′, versusthe position coordinates x1

and x2 for d = 1 and b = 0.1.The analysis of the setup inFigure 2.1 after Ghosh et al.(1986)

(iii) The experiment of Rarity and Tapster

Rarity and Tapster (1990) demonstrated a violation of Bell’s inequality usingphase and momentum of photon pairs instead of polarization as in previous exper-iments. They selected two signal beams of the same colour (the frequency ωs) andtwo idler beams also of the same colour (frequency ωi �= ωs). They directed one ofthe signal beams and one of the idler beams to a mirror M1 and another mirror M2

(see Fig. 2.4). On the paths to the mirror M2 they increased the phase of the signalby ϕs and that of the idler by ϕi. They coherently mixed the two signals and idlers.

Fig. 2.4 Experimental setupof Rarity and Tapster

Now we will determine nonlinear dynamics of field operators. We assume theelectric field in the form (2.161), (2.162), and (2.165), with another orthonormalsystem of functions vs j k(r), k ∈ [k]s j , vi j k(r), k ∈ [k]i j , j = 1, 2. The distinction isin the sets [k]s j and [k]i j , which are appropriate to the experimental setup.

Page 57: Quantum Aspects of Light Propagation

48 2 Origin of Macroscopic Approach

We will give a specification of vJk(r), J = s j , i j , j = 1, 2. We associate points0BSi , 0BSs with the beam splitters. We introduce the notation zBSi , zBSs such that

0BSi = (0, 0, zBSi ), 0BSs = (0, 0, zBSs ). (2.210)

We respect the phase shifters and the spots of reflection on the mirror M2 with thenotation zPSi , zMi2

, zPSs , zMs2. We have located the phase shifter for the idler beam,

the spot of reflection of the idler beam, the phase shifter for the signal beam, and thespot of reflection of the signal beam, respectively, at

(−b

zPSi

zMi2

, 0, zPSi

), (−b, 0, zMi2

),

(−b

zPSs

zMs2

, 0, zPSs

), (−b, 0, zMs2

). (2.211)

The origin 0 has its image 0′s2= 0′i2 = (−2b, 0, 0) in the mirror M2.

Then we assume that the mirror M1 is simple. We respect the spots of reflectionon the mirror M1 with the notation zMi1

, zMs1. We have located the spot of reflection

of the idler beam and the spot of reflection of the signal beam, respectively, at

(b, 0, zMi1), (b, 0, zMs1

). (2.212)

The origin 0 has its image 0′s1= 0′i1 = (2b, 0, 0) in the mirror M1, which is assumed

to be simple so far. We specify that

vJk(r) = i

√�ωk

ALexp(ik · r) for r ‖ k, z < zMJ , k ∈ [k]J , (2.213)

vJk(r) = −i

√�ωk

ALexp

[ik′ · (r− 0′J )+ ωJ1δx

c

]for (r− 0′J ) ‖ k′,

zMJ < z < zBSJ1, k ∈ [k]J , (2.214)

where δx is a path-length difference, J = s1, i1, J1 = s, i, k′ = (−kx , ky, kz),

vJk(r) = i

√�ωk

ALexp(ik · r) for r ‖ k, z < zPSJ1

, k ∈ [k]J , (2.215)

vJk(r) = i

√�ωk

ALexp[i(k · r+ ϕJ1 )] for r ‖ k,

zPSJ1< z < zMJ , k ∈ [k]J , (2.216)

vJk(r) = −i

√�ωk

ALexp{i[k′ · (r− 0′J )+ ϕJ1 ]} for (r− 0′J ) ‖ k′,

zMJ < z < zBSJ1, k ∈ [k]J , (2.217)

where J = s2, i2, J1 = s, i.

Page 58: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 49

Beam splitter BSs with the transmissivities ts, t′s and reflectivities rs, r′s: Inputvalues are

vs1kin(0BSs ) = −i

√�ωk

ALexp

[ik′ · (0BSs − 0′s1

)+ ωsδx

c

], k ∈ [k]s1 , (2.218)

vs2kin(0BSs ) = −i

√�ωk

ALexp{i[k′ · (0BSs − 0′s2

)+ ϕs]}, k ∈ [k]s2 . (2.219)

Under the assumption δx = 0 we can perform the exchange k ↔ k′ and write theoutput values for k ∈ [k]s1

vs1kout(0BSs ) = −i

√�ωk

ALexp{ik′ · (0BSs − 0′s1

)}ts

−i

√�ωk

ALexp{i[k · (0BSs − 0′s2

)+ ϕs]}r′s, (2.220)

vs2k′out(0BSs ) = −i

√�ωk

ALexp{ik′ · (0BSs − 0′s1

)}rs

−i

√�ωk

ALexp{i[k · (0BSs − 0′s2

)+ ϕs]}t′s. (2.221)

Performing the exchange k ↔ k′ in (2.221), we have for (r − 0′s1) ‖ k′,

z > zBSs , k ∈ [k]s1 ,

vs1k(r) = −i

√�ωk

ALexp[ik′ · (r− 0′s1

)]ts

− i

√�ωk

ALexp{i[k′ · (r− 0BSs )+ k · (0BSs − 0′s2

)+ ϕs]}r′s, (2.222)

and for (r− 0′s2) ‖ k′, z > zBSs , k ∈ [k]s2 ,

vs2k(r) = −i

√�ωk

ALexp{i[k′ · (r− 0BSs )+ k · (0BSs − 0′s1

)]}rs

−i

√�ωk

ALexp{i[k′ · (r− 0′s2

)+ ϕs]}t′s. (2.223)

Beam splitter BSi with the transmissivities ti, t′i and reflectivities ri, r′i can bedescribed analogously: In (2.222) and (2.223) we perform the replacement s ↔ i.

In a standard fashion, we associate the modal functions, e.g. (2.213), (2.214),(2.215), (2.216), (2.217), (2.222), and (2.223), with fields we denote E (+)

j0 (r, t), J =s1, s2, i1, i2. Nonlinear dynamics is described in the same way as for the experimenton the interference of signal and idler photons by relations (2.188), (2.189), (2.190),

Page 59: Quantum Aspects of Light Propagation

50 2 Origin of Macroscopic Approach

(2.192), and (2.193). It allows the fields with z < 0 to stay initial and those withz > 0 at least to obey the same rules we have used to calculate the modal functions.

We introduce the slowly varying field operators

F (+)J (r, t) = exp (iωJ1 t)E (+)

J (r, t), (2.224)

where J = s1, s2, J1 = s and J = i1, i2, J1 = i, for expressing the intensitycorrelations. The field operators at the signal and idler detectors placed at rs, ri,respectively, are

F (+)s2

(rs, t) = t′s F (+)s2out

(0, t − rs

c+ ϕs

ωs

)exp

(iωsrs

c+ iϕs

)

+ rs F (+)s1out

(0, t − rs

c

)exp

(iωsrs

c

), (2.225)

F (+)i2

(ri, t ′) = t′i F(+)i2out

(0, t ′ − ri

c+ ϕi

ωi

)exp

(iωiri

c+ iϕi

)

+ ri F(+)i1out

(0, t ′ − ri

c

)exp

(iωiri

c

), (2.226)

with 0 the centre of the coordinate system, rs and ri the path lengths of the lowersignal and idler beams, respectively, to the appropriate detector. In the experimentunder consideration both upper paths were modified by δx , since the upper and thelower mirrors were not at exactly the same distance from the pumping beam axis(Casado et al. 1997a) . The mirror above the pumping beam axis is not simple, but amirror assembly which enables one to change δx (Rarity and Tapster 1990). In thefollowing, we assume δx = 0.

We will take into account that the signal field F (+)s1

(r, t) is correlated with the

idler field F (+)i2

(r, t) and also F (+)s2

(r, t) is correlated with F (+)i1

(r, t), but F (+)s j

(r, t) is

uncorrelated with F (+)i j

(r, t), j = 1, 2, these pairs not fulfilling matching conditions.If we consider that the time intervals rs

c − ric − ϕs

ωs, rs

c − ric + ϕi

ωiare small in comparison

with the coherence time of signal and idler given by the function ν(τ ), we obtain

〈F (+)s2

(rs, t)F (+)i2

(ri, t ′)〉 ≈{rit

′s exp

[i(ωs

rs

c+ ωi

ri

c+ ϕs

)]

+ rst′i exp

[i(ωs

rs

c+ ωi

ri

c+ ϕi

)]}ν(t ′ − t). (2.227)

From this we have

Psi(rs, t + τ ; ri, t + τ ′) = K ′g2|V |2|ν(τ ′ − τ )|2

×[|rit

′s|2 + |rst

′i|2 + 2Re{r∗st′srit

′∗i exp[i(ϕs − ϕi)]}

]. (2.228)

Page 60: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 51

The visibility is expressed by the formula (2.175), where

Rsimax + Rsimin = 2g2|V |2 M (0, 0)(|t′sri|2 + |rst

′i|2), (2.229)

Rsimax − Rsimin = 8g2|V |2 M (0, 0) |t′sri||rst′i|. (2.230)

The dependence of the visibility on the beam splitters with the transmissivities|t′s| ∈ [0, 1], |t′i| ∈ [0, 1] and the reflectivities |rs| =

√1− |t′s|2, |ri|=

√1− |t′i|2

is plotted in Fig. 2.5.

Fig. 2.5 The visibility V ′versus moduli of theamplitude transmissivities t′s,t′i from the “lower side” ofthe beam splitters for signaland idler beams for δx = 0

Unfortunately, the quantity under consideration is not dependent on the charac-teristic of the nonlinear optical process. The surface plotted has a boundary condi-tion zero. It may be equal to unity in the sense of the equality |t′sri| = |t′irs|. Themaximum is attained on the line segment connecting the points |t′s| = 0, |t′i| = 0and |t′s| = 1, |t′i| = 1. The interference manifests itself as a cosine variation ofthe coincidence rate with ϕs − ϕi.

(iv) The experiment of Franson

Franson (1989) proposed a test of “Bell’s inequality for energy and time”. Hearranged two Mach–Zehnder interferometers and let a signal and an idler beameach pass through an interferometer (see Fig. 2.6). The experiment was originallyproposed for an atom and free-space propagation.

The coincidence detection shows a fourth-order interference as a cosine depen-dence on 1

c (ωsΔLs + ωiΔL i), where ΔLs (ΔL i) is the length difference betweenthe long (short) route of the signal (idler) beam through the corresponding inter-ferometer. In the past few years several groups have performed experiments of thattype. In Tapster et al. (1994) Franson’s experiment has been adapted to parametricdown-conversion and fibres.

Page 61: Quantum Aspects of Light Propagation

52 2 Origin of Macroscopic Approach

Fig. 2.6 Experimental setupof Franson’s type. Forsimplicity EiBS0 ≡ EiBS1s0 and

EsBS0 ≡ EsBS1i0

For the description of the nonlinear dynamics of field operators, we consider theinitial electric-field in the form

E (+)0 (r, t) = V (+)(r, t)1+ E (+)

s0 (r, t)+ E (+)iBS1s 0(r, t)

+E (+)i0 (r, t)+ E (+)

sBS1i 0(r, t), (2.231)

where

E (+)J0 (r, t) =

k∈[k]J

vJk(r)aJk0(t), J = s, i, iBS1s , sBS1i . (2.232)

The modal functions as restricted to linear segments are

vsk(r) = i

√�ωk

ALeik·r for r ‖ k, z < zBS1s , k ∈ [k]s, (2.233)

viBS1s k(r) = i

√�ωk

ALeik·r for (r− 0BS1s ) ‖ k, z > zBS1s , k ∈ [k]iBS1s

. (2.234)

Here 0BS1s is the centre of the beam splitter BS1s, zBS1s is the correspondingz-coordinate, and [k]iBS1s

is the set of wave vectors k of the beam correspondingto the unused input port of this beam splitter.

The modal functions at the output of this beam splitter are

vsk(r) = i

√�ωk

ALeik·rts + i

√�ωk

ALe

ik·(r−0′BS1siBS1s)rs

for r ‖ k, zBS1s < z < zBS2s, k ∈ [k]s, (2.235)

Page 62: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 53

viBS1s k(r) = i

√�ωk

ALeik·(r−0′BS1ss)rs + i

√�ωk

ALeik·rts

for (r− 0BS1s ) ‖ k, zM1s < z < zBS1s, k ∈ [k]iBS1s, (2.236)

where 0′BS1siBS1s, 0′BS1ss

are chosen such that

k · (r− 0′BS1siBS1s) = k · (r− 0BS1s )+ k′ · 0BS1s , k ∈ [k]s, (2.237)

k · (r− 0′BS1ss) = k · (r− 0BS1s )+ k′ · 0BS1s , k ∈ [k]iBS1s. (2.238)

Here zM1s is the z-coordinate of the centre of the signal mirror and zBS2s is thez-coordinate of 0BS2s , the centre of the beam splitter BS2s.

After the reflection from the first mirror, the modal function is

viBS1s k(r) = −i

√�ωk

ALeik′ ·(r−0′M1sBS1ss)rs

−i

√�ωk

ALeik′ ·(r−0′M1s

)ts for (r− 0M1s ) ‖ k′,

zM1s < z < zM2s, k ∈ [k]iBS1s, (2.239)

where zM2s is the z-coordinate of the centre of the second mirror and 0′M1sBS1ssand

0′M1sare chosen such that

k′ · (r− 0′M1sBS1ss) = k′ · (r− 0M1s )+ k · (0M1s − 0′BS1ss), k ∈ [k]iBS1s, (2.240)

k′ · (r− 0′M1s) = k′ · (r− 0M1s )+ k · 0M1s , k ∈ [k]iBS1s

. (2.241)

After the reflection from the second mirror, the modal function is

viBS1s k(r) = i

√�ωk

ALe−ik·(r−0′M2sM1sBS1ss)rs

+i

√�ωk

ALe−ik·(r−0′M2sM1s

)ts for (r− 0M2s ) ‖ −k,

zM2s < z < zBS2s , k ∈ [k]iBS1s, (2.242)

where 0′M2sM1sBS1ssand 0′M2sM1s

are chosen such that

−k · (r− 0′M2sM1sBS1ss) = −k · (r− 0M2s )

+k′ · (0M2s − 0′M1sBS1ss), k ∈ [k]iBS1s, (2.243)

−k · (r− 0′M2sM1s) = −k · (r− 0M2s )

+k′ · (0M2s − 0′M1s), k ∈ [k]iBS1s

. (2.244)

Page 63: Quantum Aspects of Light Propagation

54 2 Origin of Macroscopic Approach

The output modal functions for the beam to be detected are

vsk(r) = i

√�ωk

ALeik·rt2

s

+[

i

√�ωk

ALe

ik·(r−0′BS1siBS1s) + i

√�ωk

ALeik·(r−0′BS2sM2sM1s

)

]tsrs

+i

√�ωk

ALeik·(r−0′BS2sM2sM1sBS1ss)r2

s , for r ‖ k, z > zBS2s , k ∈ [k]s, (2.245)

where 0′BS2sM2sM1sand 0′BS2sM2sM1sBS1ss

are chosen such that

k · (r− 0′BS2sM2sM1s) = k · (r− 0BS2s )− k′ · (0BS2s − 0′M2sM1s

), k ∈ [k]s, (2.246)

k · (r− 0′BS2sM2sM1sBS1ss) = k · (r− 0BS2s )

−k′ · (0BS2s − 0′M2sM1sBS1ss), k ∈ [k]s. (2.247)

The output modal functions for the second beam are

viBS2s k(r) = i

√�ωk

ALe−ik·(r−0′BS2sBS1siBS1s

)r2

s

+[

i

√�ωk

ALe−ik·(r−0′BS2ss) + i

√�ωk

ALe−ik·(r−0′M2sM1sBS1s

)

]rsts

+i

√�ωk

ALe−ik·(r−0′M2sM1s

)t2s for (r− 0BS1s ) ‖ −k,

z > zBS2s , k ∈ [k]iBS1s, (2.248)

where 0′BS2sBS1siBS1sand 0′BS2ss

are chosen such that

−k · (r− 0′BS2sBS1siBS1s) = −k · (r− 0BS2s )

+k′ · (0BS2s − 0′BS1siBS1s), k ∈ [k]iBS1s

. (2.249)

−k · (r− 0′BS2ss) = −k · (r− 0BS2s )+ k′ · 0BS2s , k ∈ [k]iBS1s. (2.250)

In a standard fashion, we associate the modal functions which travel to the abovedetector, (2.233), (2.234), (2.235), (2.236), (2.239), (2.242), (2.245), and (2.248),with fields we denote as E (+)

s0 (r, t), E (+)iBS1s 0(r, t). Exchanging s ↔ i, we introduce

modal functions, which travel to the lower detector. We relate them with fields wedenote as E (+)

i0 (r, t), E (+)sBS1i 0

(r, t). Switching on the nonlinear interaction, we find thefield to obey the relations (2.189) and (2.190). Counter to propagation the field staysinitial and along with propagation it at least obeys the rules we have used to generatethe modal functions. For simplicity, it is assumed that tJ = t′J = rJ = r′J = 1√

2,

J = s, i.

Page 64: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 55

For the calculation of intensity correlations determined in the experiment, weintroduce the slowly varying field operators

F (+)s (r, t) = exp (iωst)E (+)

s (r, t),

F (+)i (r, t) = exp (iωit)E (+)

i (r, t). (2.251)

The field operators at the signal and idler detectors placed at rs, ri, respectively, are

F (+)s (rs, t)

= 1

2

{[F (+)

sout

(0, t − |rs − rBS2s |

c− Ls,long

c− |rBS1s |

c

)exp

(iωs|rBS1s |

c

)

− iF (+)iBS1s in

(0BS1s , t − |rs − rBS2s |

c− Ls,long

c

)]exp

(iωs

Ls,long

c

)

+[

F (+)sout

(0, t − |rs − rBS2s |

c− Ls,short

c− |rBS1s |

c

)exp

(iωs|rBS1s |

c

)

+ iF (+)iBS1s in

(0BS1s , t − |rs − rBS2s |

c− Ls,short

c

)]exp

(iωs

Ls,short

c

)}

× exp

(iωs|rs − rBS2s |

c

), (2.252)

F (+)i (ri, t) = F (+)

s (rs, t)∣∣∣s↔i

. (2.253)

Let us denote Ls,short (L i,short) a length of the short arm of the interferometer for thesignal (idler) beam. Supposing that ΔLs ≡ Ls,long−Ls,short (ΔL i ≡ L i,long−L i,short)is much greater than the coherence length of the signal (idler) in order to avoid thesecond-order interference, we get (Casado et al. 1997a)

〈F (+)s (rs, t + τ )F (+)

i (ri, t + τ ′)〉 = 1

4gV

×{ν(τ ′ − τ ) exp

[i

cωs(|rBS1s | + Ls,long + |rs − rBS2s |

)

+ i

cωi(|rBS1i | + L i,long + |ri − rBS2i |

) ]

+ ν

(τ ′ − τ + ΔL i −ΔLs

c

)exp

[i

cωs(|rBS1s | + Ls,short

+ |rs − rBS2s |)+ i

cωi(|rBS1i | + L i,short + |ri − rBS2i |

) ]}, (2.254)

Page 65: Quantum Aspects of Light Propagation

56 2 Origin of Macroscopic Approach

provided that |rBS1J | + L J,short + |rJ − rBS2J | is the same for J = s and J = i. Wefinally obtain

Psi(rs, t + τ ; ri, t + τ ′) = 1

16K ′g2|V |2

{|ν(τ ′ − τ )|2

+∣∣∣∣ν(

τ ′ − τ + ΔL i −ΔLs

c

)∣∣∣∣2

− 2Re

{ν(τ ′ − τ )ν∗

(τ ′ − τ + ΔL i −ΔLs

c

)

× exp

[i

c(ωsΔLs + ωiΔL i)

]}}. (2.255)

The visibility is given in (2.175), where

Rsimax + Rsimin

= 1

8g2|V |2

[M (0, 0)+ M

(ΔL i −ΔLs

c,ΔL i −ΔLs

c

)], (2.256)

Rsimax − Rsimin = 1

4g2|V |2 M

(0,

ΔL i −ΔLs

c

). (2.257)

The dependence of the visibility on the difference (ΔL i−ΔLs) is plotted in Fig. 2.7.The variation of the visibility is due to the function M

(dc , h

c

), but the function erf

does not contribute to it. The distance between the points of inflection is 2 cσ=

6×10−4 m. The interference manifests itself as a cosine variation of the coincidencerate with 1

c (ωsΔLs + ωiΔL i).

Fig. 2.7 The visibility V ′versus the difference(ΔL i −ΔLs) ∈[−10−3, 10−3] measured inmetres

Page 66: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 57

(v) Induced coherence and indistinguishability in two-photon interference

Zou et al. (1991) performed an experiment in which fourth-order interference isobserved in the superposition of signal photons from two coherently pumped para-metric down-conversion crystals, when the paths of the idler photons are aligned.The experimental setup is outlined in Fig. 2.8, in which two nonlinear crystals NL1and NL2 are optically pumped by two mutually coherent, classical pump waves ofcomplex amplitudes

Vj (r, t) = Vj exp[i(k0 · r− ω0t)], j = 1, 2. (2.258)

We assume that V1 = V2 exp(ik0 · 02) = V . On the contrary, there were simi-lar crystals in the experiment, but we consider more general ones. The parametricdown-conversion occurs at both crystals, each with the emission of a signal photonand an idler photon. We are interested in the joint detection rate of the detectors Ds

and Di when the trajectories of the two idlers i1, i2 are aligned and the path differ-ence between the two signals is varied slightly. Fourth-order interference disappearswhen the idlers are misaligned or separated by a beam stop.

Fig. 2.8 Experimental setupon induced coherence withoutinduced emission. Forsimplicity, EsBS0 ≡ EsBSi0

In what follows, we will specify modal functions and the nonlinear dynamics offield operators. We consider the initial electric field in the form

E (+)0 (r, t) = V (+)(r, t)1+

2∑

j=1

E (+)s j 0

(r, t)+ E (+)i0 (r, t)+ E (+)

sBSi 0(r, t), (2.259)

where

E (+)J0 (r, t) =

k∈[k]J

vJk(r)aJk0(t), J = s1, s2, i, sBSi . (2.260)

The modal functions as restricted to linear segments are

vs1k(r) = i

√�ωk

ALeik·r for r ‖ k, z < zMs , k ∈ [k]s, (2.261)

vs2k(r) = i

√�ωk

ALeik·r for (r− 02) ‖ k, z < zBSs , k ∈ [k]s, (2.262)

Page 67: Quantum Aspects of Light Propagation

58 2 Origin of Macroscopic Approach

vs1k(r) = −i

√�ωk

ALeik′ ·(r−0′Ms

) for (r− 0Ms ) ‖ k′,

zMs < z < zBSs , k ∈ [k]s. (2.263)

Here 0′Msand k′ are used as in the definition of modal functions related to Fig. 2.1

and k′ has the same meaning. Here, as in the definitions related to Fig. 2.4, wespecify that 0′Ms

has been chosen so that

k′ · (r− 0′Ms) = k′ · (r− 0Ms )+ k · 0Ms , k ∈ [k]s. (2.264)

Similarly as in (2.222) and (2.223) for t = t′ = 1√2, r = r′ = i√

2, we have

vs1k(r) = −i

√�ωk

ALeik′ ·(r−0′Ms

) 1√2+ i

√�ωk

ALeik′ ·(r−0′BSs

) i√2

for (r− 0Ms ) ‖ k′, z > zBSs , k ∈ [k]s, (2.265)

vs2k(r) = −i

√�ωk

ALeik·(r−0′BSsMs

) i√2+ i

√�ωk

ALeik·r 1√

2for (r− 02) ‖ k, z > zBSs , k ∈ [k]s, (2.266)

where 0′BSsand 0′BSsMs

have been chosen so that, respectively,

k′ · (r− 0′BSs) = k′ · (r− 0BSs )+ k · 0BSs , k ∈ [k]s, (2.267)

k · (r− 0′BSsMs) = k · (r− 0BSs )+ k′ · 0′Ms

, k ∈ [k]s; (2.268)

vik(r) = i

√�ωk

ALeik·r for r ‖ k, z < zBSi , k ∈ [k]i, (2.269)

vsBSi k(r) = i

√�ωk

ALeik·r for (r− 0BSi ) ‖ k, z > zBSi , k ∈ [k]sBSi

, (2.270)

vik(r) = i

√�ωk

ALeik·rt+ i

√�ωk

ALeik·(r−0′BSisBSi

)r′

for r ‖ k, z > zBSi , k ∈ [k]i, (2.271)

vsBSi k(r) = i

√�ωk

ALeik·(r−0′BSii)r+ i

√�ωk

ALeik·rt′

for (r− 0BSi ) ‖ k, z < zBSi , k ∈ [k]sBSi, (2.272)

where k′ is defined relative to the beam splitter BSi and 0′BSiiand 0′BSisBSi

have beenchosen so that, respectively,

k · (r− 0′BSisBSi) = k · (r− 0BSi )+ k′ · 0BSii for k ∈ [k]i, (2.273)

k · (r− 0′BSii) = k · (r− 0BSi )+ k′ · 0BSi for k ∈ [k]sBSi. (2.274)

Page 68: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 59

The modal functions (2.269), (2.270), (2.271), and (2.272) travel to the lowerdetector. We relate them with fields we denote as Ei0(r, t), EsBSi 0

(r, t). Switching onthe first nonlinear interaction, we find the field to obey the relations (2.189), (2.190),and (2.191) for j = 1 with i1 → i. Counter to propagation the field remains initialand along with propagation it at least obeys the rules we have used to generate themodal functions.

Now we would like to interpret the subscript 0 not as the order of solution butas a number of the initial stage. Since the stage is followed by the first stage, wewould like to modify the relations (2.189), (2.190), and (2.191) so that they confessthe first-stage operators on the left-hand side, which would lead to the use of thesubscript 1.

Switching on the second nonlinear interaction, we find the field to obey therelations (2.189), (2.190), and (2.191) for j = 2 with i2 → i, but the first-stagefield operators to have been substituted for the operators on the right-hand side. The

action of the operators ˆG2 and ˆJ2 depends on the nonlinear crystal located at 02.It also depends on the pump beam at the same crystal. Counter to propagation thefield stays first stage and along with propagation it still at least obeys the rules togenerate the modal functions.

Further we will express the intensity correlations that have been determined inthe experiment. Again, we introduce the operators (2.224), where J = s1, s2, i,J1 = s, s, i. The field operators at the signal and idler detectors placed at rs, ri,respectively, are

F (+)s2

(rs, t) = 1√2

[− iF (+)

s1

(01, t − d

c

)exp

(iωs

d

c

)

+ F (+)s2

(02, t − h

c

)exp

(iωs

h

c

)], (2.275)

F (+)i (ri, t ′) = F (+)

i

(02, t ′ − l

c

)exp

(iωi

l

c

). (2.276)

We still assume different crystals and derive a slight generalization of the well-known experiment (Casado et al. 1997a). By taking into account the correlationrelations

〈F (+)s1

(01, t)F (+)i (02, t ′)〉 = tgV1ν1

(t ′ − t − f

c

)exp

(iωi

f

c

), (2.277)

〈F (+)i (02, t ′)F (+)

s2(02, t)〉 = gV2ν2(t ′ − t), (2.278)

Page 69: Quantum Aspects of Light Propagation

60 2 Origin of Macroscopic Approach

we get (Casado et al. 1997a)

〈F (+)s2

(rs, t)F (+)i (ri, t ′)〉 = gV√

2

×[− itν1

(τ ′ − τ − l

c− f

c+ d

c

)exp

{i

c[ωsd + ωi(l + f )]

}

+ ν2

(τ ′ − τ − l

c+ h

c

)exp

{i

c[ωsh + ωil]

}]. (2.279)

We finally obtain

Psi(rs, t + τ ; ri, t + τ ′) = 1

2K ′g2|V |2

×{ ∣∣∣∣tν1

(τ ′ − τ − l

c− f

c+ d

c

)∣∣∣∣2

+∣∣∣∣ν2

(τ ′ − τ − l

c+ h

c

)∣∣∣∣2

+ 2Im

[tν1

(τ ′ − τ − l

c− f

c+ d

c

)ν∗2

(τ ′ − τ − l

c+ h

c

)

× exp

{i

c[ωs(d − h)+ ωi f ]

}]}. (2.280)

We have hopefully corrected the factor, changed a sign with respect to the reflectionfrom the mirror, and changed signs of the argument of ν(τ ) relying on the identityν(τ ) = ν(−τ ), where ν1(τ ) = ν2(τ ) ≡ ν(τ ).

The visibility is expressed by the formula (2.175), where for d = l + f , l = h

Rsimax + Rsimin = g2|V |2 M (0, 0)(|t|2 + 1

), (2.281)

Rsimax − Rsimin = 2g2|V |2 M (0, 0) |t|. (2.282)

The maximum visibility is equal to unity and, in general, it depends on the transmis-sivity of the beam splitter, as can be seen from Fig. 2.9. The interference manifestsitself as a cosine variation of the coincidence rate with ω0

c f .

(vi) Frustrated two-photon creation via interference

Herzog et al. (1994) performed a simple experiment interpreted as showing inter-ference of two processes. They placed three mirrors in the three beams, laser, signal,and idler, that emerge from a nonlinear crystal, NL, and put a detector into thereflected idler beams (see Fig. 2.10). In the standard quantum interpretation a pairof correlated photons can be created either by the laser beam travelling from left toright or when the reflected laser beam travels from right to left. In both cases theidler photon may arrive at the detector. As the two possibilities are indistinguishable

Page 70: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 61

Fig. 2.9 The visibility V ′versus the modulus of theamplitude transmissivity|t| ∈ [0, 1] of the beamsplitter BSi. It is assumed thatd = l + f , l = h

Fig. 2.10 Experimental setupon frustrated two-photoncreation via interference

they interfere and the counting rate oscillates depending on the position of a chosenmirror.

Accordingly the description of the pump beam is given by

V (+)(r, t) = V ei(k0·r−ω0t) − V ei[−k0·(r−2l0e0)−ω0t]

= V ei(k0·r−ω0t) − V eiϕ0 ei(−k0·r−ω0t)

for x = 0, y = 0, z < l0, (2.283)

where e0 is the direction vector of the forward-propagating pump beam, ϕ0 =2|k0|l0 = 2ω0l0

c .The modal functions are

vsk(r) = i

√�ωk

2ALeik·r − i

√�ωk

2ALeiϕs e−ik·r

for r ‖ k, z < zMs , k ∈ [k]s, (2.284)

where ϕs = 2ωslsc , zMs = e0 ·0Ms . The modal functions vik(r) are expressed similarly.

Associating the modal functions with quantum fields, we must consider that

E (+)s0 (r, t) = E (+)

sF0 (r, t)+ E (+)sB0(r, t), (2.285)

Page 71: Quantum Aspects of Light Propagation

62 2 Origin of Macroscopic Approach

where

E (+)sF0 (r, t) = i

k∈[k]s

√�ωk

2ALeik·rask0(t),

E (+)sB0(r, t) = −i

k∈[k]s

√�ωk

2ALeiϕs e−ik·rask0(t) (2.286)

are the forward-propagating component and the backward-propagating component,respectively. The field operators E (+)

i0 (r, t), E (+)iF0 (r, t), and E (+)

iB0 (r, t) are expressedsimilarly. Nonlinear dynamics is described by the relations

E (+)sF (0− (r = 0)es, t) = E (+)

sF0 (0, t),

E (+)iF (0− (r = 0)ei, t) = E (+)

iF0 (0, t), (2.287)

where e j , j = s, i, is the direction vector of the forward-propagating signal, idlerbeam, respectively:

E (+)sF (0+ (r = 0)es, t) = (1+ g2|V |2 ˆJ )E (+)

sF0 (0, t)+ e−iω0t gV ˆG E (−)iF0 (0, t),

E (+)iF (0+ (r = 0)ei, t) = e−iω0t gV ˆG E (−)

sF0 (0, t)+ (1+ g2|V |2 ˆJ )E (+)iF0 (0, t),

(2.288)

E (+)sB (0+ (r = 0)es, t) = −E (+)

sF

(0+ (r = 0)es, t − 2ls

c

)

= −(1+ g2|V |2 ˆJ )E (+)sF0

(0, t − 2ls

c

)

− ei(ϕs+ϕi)e−iω0t gV ˆG E (−)iF0

(0, t − 2ls

c

), (2.289)

E (+)iB (0+ (r = 0)ei, t) = −E (+)

iF

(0+ (r = 0)ei, t − 2li

c

)

= −ei(ϕs+ϕi)e−iω0t gV ˆG E (−)sF0

(0, t − 2li

c

)

− (1+ g2|V |2 ˆJ )E (+)iF0

(0, t − 2li

c

), (2.290)

E (+)sBout(0, t) = (1+ g2|V |2 ˆJ )E (+)

sB (0+ (r = 0)es, t)

+ e−iω0t gV eiϕ0 ˆG E (−)iB (0+ (r = 0)ei, t),

E (+)iBout(0, t) = e−iω0t gV eiϕ0 ˆG E (−)

sB (0+ (r = 0)es, t)

+ (1+ g2|V |2 ˆJ )E (+)iB (0+ (r = 0)ei, t). (2.291)

Page 72: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 63

Further we will calculate the quantum mean intensities that have been determinedin the experiment. Introducing the slowly varying field operators

F (+)J (r, t) = eiωJ t E (+)

J (r, t),

F (+)JF (r, t) = eiωJ t E (+)

JF (r, t),

F (+)JB (r, t) = eiωJ t E (+)

JB (r, t), J = s, i, (2.292)

and F (−)J (r, t), F (−)

JF (r, t), F (−)JB (r, t), we express the field operators at the signal and

idler detectors placed at rs, ri, respectively, as

F (+)JB (rJ , t) = F (+)

JBout

(0, t − rJ

c

)exp

(iωJ rJ

c

), J = s, i. (2.293)

The quantum mean intensity or single photodetection rate in the detector Ds is

Ps(rs, t) = K⟨0∣∣∣ E (−)(rs, t)E (+)(rs, t)

∣∣∣ 0⟩

(2.294)

= K⟨0∣∣∣ F (−)

sB (rs, t)F (+)sB (rs, t)

∣∣∣ 0⟩, (2.295)

where K is a constant related to the efficiency of the detector and the energy of asingle photon. Considering the forward propagation, reflections, and the backwardpropagation, we obtain that

⟨F (−)

sB (rs, t)F (+)sB (rs, t)

⟩=⟨F (−)

sB

(0, t − rs

c

)F (+)

sB

(0, t − rs

c

)⟩

= 2g2|V |2[μs(0)+ μs

(2li

c− 2ls

c

)cos(ϕs + ϕi − ϕ0)

], (2.296)

where we have relied on the identity μs(τ ) = μs(−τ ). From this,

Ps(rs, t) = 2K g2|V |2[μs(0)+ μs

(2li

c− 2ls

c

)cos(ϕs + ϕi − ϕ0)

]. (2.297)

The photodetection rate in the detector Di is expressed similarly (Casado et al.1997b).

In conclusion, we have mostly dealt with the fourth-order interference in para-metric down-conversion experiments. The 1986, 1990, 1994 (adapted back to freespace), and 1991 experiments were chosen according to a review article of otherauthors. Coincidence measurements in the various setups are essentially (or suffi-ciently well) described in terms of the cross correlation between the signal and theidler.

We have “promoted” the schemes of the experiments, where only paths throughnonlinear and linear optical elements and the free space (with possible reflectionsfrom perfect mirrors) to detectors are drawn, to a reason of a certain neglect of the

Page 73: Quantum Aspects of Light Propagation

64 2 Origin of Macroscopic Approach

beams’ divergence. We have replaced the usual assumption that the electric field isexpanded in terms of an incomplete set of plane waves, which is relatively com-plete with respect to the expected direction of propagation, by the hypothesis thatthere exists an incomplete or relatively complete system of more complicated modalfunctions, which have still been specified only on the paths. We have performed con-ventional quantization by introducing annihilation operators in place of the classicalcomplex amplitudes of the modes. We tried to choose sufficiently realistic values ofthe parameters for all the four experiments and to find visibilities of the intensityinterference.

2.3.2 From Coupled Quantum Harmonic Oscillators Backto Interacting Fields

One of the interference experiments we have described in Section 2.3.1, the exper-iment of Zou et al. (1991) which has been analysed in Wang et al. (1991a,b), hasattracted much attention. The arrangement of two down-converters is pumped bymutually coherent beams and the two down-converters are connected by the idlerbeam. The spontaneous emission from the first nonlinear crystal in the idler servesas a stimulating idler input to the second nonlinear crystal that acts as an opticalamplifier. The interference of signal beams from both the crystals can be observed.A beam splitter placed between the two nonlinear crystals in the idler beam canchange the strength of their connection since it attenuates the emerging field.

The parametric down-conversion in the second nonlinear crystal is stimulated byidler photons when the idler field is strong “per frequency unit”. In this situation,the polychromatic theory yields results similar to those obtained by the monochro-matic treatment, i.e. about multiples of the latter. When the idler field is weak perfrequency unit, the second nonlinear crystal is proven to “ignore” the idler photons.The monochromatic description, even though completed by optimal scaling of itsresults, is far from being persuasive here. The assumption of the strong idler field isimplicit in work contributing to the monochromatic theory.

In Rehacek and Perina (1996) it has been shown that the distribution of photon-number sum in signal modes interpolates between a Bose–Einstein distribution and aconvolution of two Bose–Einstein distributions. The latter distribution occurs whenthe idler beam is blocked. In general, the photon-number sum is distributed as if itcorresponded to the number of signal degrees of freedom which varied between 1and 2. A nonclassical distribution of photon-number sums restricted to even sums ofphoton numbers cannot occur, because the correlation between the photon numbersof the two signal beams is not complete. It has been found that the distribution ofphase difference derived from the Q function narrows when the connection of boththe down-converters via the idler mode closes up.

The monochromatic treatment associates each travelling wave with a quantumharmonic oscillator. The simple formalism of several coupled harmonic oscillators isuseful for an analysis of the travelling-wave setup of interference experiment due to

Page 74: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 65

Zou et al. (1991) up to suppression of the induced emission. The more complicatedapproach used originally for the analysis seems to be unsuitable for treating thephenomenon of induced emission. We try to formalize here a comparison betweenthe two approaches.

When the induced emission occurs, it can be utilized. The phenomenon ofinduced emission makes the phase of an amplified field adopt the same phase as theincident locking field (Wang et al. 1991a, Wiseman and Mølmer 2000). The inducedemission can also be used in parametric down-conversion to lock the phase of theidler and, from this, that of the signal (since the phase sum of the signal and idler islocked to the pump phase). If the field used to lock the idler of one down-converter(crystal NL2 in Fig. 2.11) is itself the idler output of another down-converter (crystalNL1 in Fig. 2.11), the two signal fields will also be locked in phase. Thus, theywill have, in principle, perfect first-order coherence and so will interfere at a finalbeam splitter not included in Fig. 2.11. If there is no connection between the twodown-converters, and hence no induced emission, the two signals will be incoherentand there will be no interference.

Fig. 2.11 Scheme of twoparametric processes withaligned idler beams with thespatial Heisenberg picturemade explicit

Zou et al. (1991) and Wang et al. (1991a) had a negligible probability of bothcrystals producing a down-converted photon pair and used a quantum-mechanicalexplanation based on indistinguishability of paths to explain the interference theyobserved in the experiment. To the contrary, the interference was lost when onecould tell which crystal had emitted each signal photon. Using multimode analysisof the experiment, they derived that there could be no induced emission in theirexperiment. Nonetheless, they found that for perfect matching of idler modes, thesignal fields from NL1 and NL2 show perfect interference.

The multimode approach to the analysis used by Wang et al. (1991a) yields anexplanation involving a sufficient number of realistic parameters. Even though inthe foregoing section the formalism yielded results similar enough to those of Wanget al. (1991a), here we try to come near their analysis. However, there exists a sim-ple quantal description that may claim that it conforms to results of the multimodeanalysis. Such simple models have been published. Concerning this, we may referto Rehacek and Perina (1996), Wiseman and Mølmer (2000), and Perinova et al.(2000) and provide what is a continuation of Perinova et al. (2000).

(i) Formalism of several modes

We turn to the quantum analysis of the Zou–Wang–Mandel experiment. Theexperimental arrangement consists of two parametric down-conversion crystals with

Page 75: Quantum Aspects of Light Propagation

66 2 Origin of Macroscopic Approach

aligned idler beams, which are partially connected due to the presence of a beamsplitter in between, and is illustrated in Fig. 2.11.

Restricting ourselves to the quasimonochromatic light beams (or quasimonochro-matic components of these), we can describe the system by four modes, s1 (thesignal mode for crystal NL1), i (the idler modes, which are identified), s2 (the signalmode for crystal NL2), and 0 (the escape mode for the beam splitter) (Perinovaet al. 2000). We consider the input annihilation operators as1 (0), a0(1), as2 (2), ai(0),the output annihilation operators as1 (1), a0(2), as2 (3), ai(3), and the intermediateannihilation operators ai(1), ai(2). Here s1, s2 stand for the signal mode of crystal1 and that of crystal 2, respectively, i for the idler mode, and 0 for the “escape”mode of the beam splitter. To obtain four-mode unitary transformations between thestages 0, 1, 2, 3, we consider also the appendage input annihilation operators a0(0),as2 (0), as2 (1) and the appendage output annihilation operators as1 (2), as1 (3), a0(3).Of course, in the description of the dynamics below, we will have to be consistentwith the identities a0(0) = a0(1), as2 (0) = as2 (1) = as2 (2), as1 (1) = as1 (2) = as1 (3),a0(2) = a0(3). Let us consider, in the Hilbert space of these four modes, an arbitraryoperator M( j), j = 0, 1, 2, 3, a j th-stage operator. We will write the equation givingthe transformation of M( j) from its value M(0) before the interaction to its valueM(1) after the action of the first down converter, to its value M(2) after the action ofthe beam splitter, and to its value M(3) after the action of the second down converter.

The appropriate relations read

M( j + 1) = U †j+1( j)M( j)U j+1( j), j = 0, 1, 2. (2.298)

Having prescribed equations of motion of operators, we have adopted a spatial mod-ified Heisenberg picture of the dynamics. In the Heisenberg picture the input statedoes not change, while in the modified Heisenberg picture it changes like that ofthe free field. In our case of the “discrete” space (cf. j = 0, 1, 2, 3), the change ofthe state cannot be specified satisfactorily, but fortunately, we will not need it. In(2.298) U j+1( j) for j = 0, 2 describe the down conversion in the undepleted pumpapproximation. The crystals are assumed to be identical or distinct and pumps areassumed to be identical so that

U j+1( j) = exp{

iκ j2+1

[as j

2 +1( j)ai( j)+ H.c.

]}, j = 0, 2, (2.299)

where κ j2+1 = χ

c vp j2 +1

l j2 +1

, χ is the quadratic susceptibility of the matter of which

both the nonlinear crystals are made, vp j2 +1

are classical complex amplitudes of

pumping beams, c is the speed of light, l1 and l2 are the lengths of the first crystaland the second crystal, respectively. In between the down converters the idler fromcrystal NL1 is put through a beam splitter BS and becomes the idler for crystal NL2.This process is described by

U2(1) = exp{

i[ω0a†0(1)a0(1)+ ωia

†i (1)ai(1)+ (γ ∗a†

0(1)ai(1)+ H.c.)]}

,

(2.300)

Page 76: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 67

where

ω0

ωi

}= arg

(t+ t′

2

)∓ i

(t− t′)2

f (t,t′), (2.301)

γ = −ir f (t,t′), (2.302)

with

f (t,t′) =Arccos

∣∣∣t+t′

2

∣∣∣√

1− ∣∣t+t′2

∣∣2(t+ t′)∗

|t+ t′| . (2.303)

Here t and r are the transmission and reflection amplitude coefficients, respectively,for the idler mode and t′ and r′ are those for the “escape” mode. The modulus ofthe transmission amplitude coefficient |t| can vary between zero (where the secondemission is spontaneous) and unity (where the second down conversion is stimulatedin the highest degree).

Hence,

ai(2) = U †2 (1)ai(1)U2(1) = tai(1)+ r′a0(1),

a0(2) = U †2 (1)a0(1)U2(1) = rai(1)+ t′a0(1), (2.304)

and the unitarity of the transformation matrix implies that

|t|2 + |r|2 = 1, |r′|2 + |t′|2 = 1, tr′∗ + rt′∗ = 0. (2.305)

It is advantageous to assume that t′ = t∗, and from this r = −r′∗ (Rehacek andPerina 1996), and that Ret > 0. Then

f (t,t′) = Arccos(Ret)√1− (Ret)2

, (2.306)

ω0

ωi

}= ±(Imt) f (t,t′). (2.307)

On applying the relation (2.298) at the input ( j = 0) and at the stage 2 ( j = 2),we obtain that

as1 (1) = as1 (0) cosh(κ1)+ ia†i (0) sinh(κ1),

ai(1) = ia†s1

(0) sinh(κ1)+ ai(0) cosh(κ1), (2.308)

and

as2 (3) = as2 (2) cosh(κ2)+ ia†i (2) sinh(κ2),

ai(3) = ias2 (2) sinh(κ2)+ ai(2) cosh(κ2). (2.309)

Page 77: Quantum Aspects of Light Propagation

68 2 Origin of Macroscopic Approach

Using Equations (2.304), we easily obtain the following relations:

as1 (1) = cosh(κ1)as1 (0)+ i sinh(κ1)a†i (0), (2.310)

as2 (3) = t∗ sinh(κ1) sinh(κ2)as1 (0)+ it∗ cosh(κ1) sinh(κ2)a†i (0)

+ it∗ cosh(κ2)as2 (2)+ ir′∗ sinh(κ2)a†0(1). (2.311)

The statistical properties of the system in the Heisenberg picture can be obtainedwhen we take into account that the initial, in fact, “permanent” statistical operatorof the system is given as ρ ≡ ρ(0) and when we average

〈M( j)〉 = Tr{ρ M( j)}. (2.312)

Here, concretely, the statistical operator is a tensor product of separate vacuum sta-tistical operators

ρ =∏

j=s1,i,s2,0

|0〉 j j 〈0|. (2.313)

We may introduce also the abbreviations M ≡ M(0) and we consider the Schrodingerpicture, where the relation (2.298) is replaced by the evolution relations

ρ( j + 1) = U j+1ρ( j)U †j+1, j = 0, 1, 2, (2.314)

with U j ≡ U j (0) given in (2.299) and (2.300). The equivalence of both the picturescan be proved and the statistical properties can be expressed in similar terms as in(2.312)

〈M〉( j) = Tr{ρ( j)M} = 〈M( j)〉. (2.315)

Since all the initial fields are in the vacuum states, it is easy to obtain the expec-tation values

〈a†s1

(1)as1 (1)〉 = sinh2(κ1), (2.316)

〈a†s2

(3)as2 (3)〉 = sinh2(κ2)[1+ |t|2 sinh2(κ1)], (2.317)

〈a†s1

(1)as2 (3)〉 = t∗ sinh(κ1) cosh(κ1) sinh(κ2). (2.318)

We will show in the Heisenberg picture that the input–output relation is con-nected to the SU(2,2) group. In fact,

⎜⎜⎝

as1 (3)a†

i (3)as2 (3)a†

0(3)

⎟⎟⎠ =

⎜⎜⎝

ms1s1 ms1i ms1s2 ms10

m is1 m ii m is2 m i0

ms2s1 ms2i ms2s2 ms20

m0s1 m0i m0s2 m00

⎟⎟⎠

⎜⎜⎝

as1 (0)a†

i (0)as2 (0)a†

0(0)

⎟⎟⎠ , (2.319)

Page 78: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 69

where

ms1s1 = cosh(κ1), ms1i = i sinh(κ1), ms1s2 = ms10 = 0;

m is1 = −it∗ sinh(κ1) cosh(κ2), m ii = t∗ cosh(κ1) cosh(κ2),

m is2 = −i sinh(κ2), m i0 = r′∗ cosh(κ2); (2.320)

ms2s1 = t∗ sinh(κ1) sinh(κ2), ms2i = it∗ cosh(κ1) sinh(κ2),

ms2s2 = cosh(κ2), ms20 = ir′∗ sinh(κ2);

m0s1 = ir′ sinh(κ1), m0i = −r′ cosh(κ1), m0s2 = 0, m00 = t.

From the form of the relation (2.319) it is evident that the operator

N ( j) = ns1 ( j)+ ns2 ( j)− ni( j)− n0( j), j = 0, 3, (2.321)

is independent of j . This conservation law suggests the SU(2,2) group. The coeffi-cients of the transformation (2.319) verify the pseudoorthogonality relations

m js1 m∗ks1+ m js2 m∗

ks2− m j im

∗ki − m j0m∗

k0 = g jk, j, k = s1, i, s2, 0, (2.322)

where

g jk = g j jδ jk, gs1s1 = gs2s2 = 1, gii = g00 = −1. (2.323)

We observe that the antinormally ordered moments have the expression

〈a j (3)a†j (3)〉 = |m js1 |2 + |m js2 |2, j = s1, s2, (2.324)

and the normally ordered moments

〈a†j (3)a j (3)〉 = |m js1 |2 + |m js2 |2, j = i, 0. (2.325)

More generally,

〈as1 (3)a†s2

(3)〉 = ms1s1 m∗s2s1

+ ms1s2 m∗s2s2

,

〈as2 (3)a†s1

(3)〉 = 〈as1 (3)a†s2

(3)〉∗, (2.326)

〈a†i (3)a0(3)〉 = m is1 m∗

0s1+ m is2 m∗

0s2,

〈a†0(3)ai(3)〉 = 〈a†

i (3)a0(3)〉∗. (2.327)

Further nonvanishing moments are

〈a j (3)ak(3)〉 = m js1 m∗ks1+ m js2 m∗

ks2, j = s1, s2, k = i, 0, (2.328)

and

〈a†j (3)a†

k(3)〉 = 〈a j (3)ak(3)〉∗. (2.329)

Page 79: Quantum Aspects of Light Propagation

70 2 Origin of Macroscopic Approach

The rest second-order moments vanish:

〈a j (3)a†k(3)〉 = 〈a†

j (3)ak(3)〉 = 0, j = s1, s2, k = i, 0, (2.330)

〈a j (3)ak(3)〉 = 〈a†j (3)a†

k(3)〉 = 0, j = s1, s2, k = s1, s2,

and j = i, 0, k = i, 0. (2.331)

Quantum statistics of radiation in the process under study is that of a four-modeGaussian state, starting with the quantum characteristic function:

CS (βs1 , βs2 , βi, β0, 3)

= Tr{ρ(3)Ds1 (βs1 , 0)Ds2 (βs2 , 0)Di(βi, 0)D0(β0, 0)}= Tr{ρ Ds1 (βs1 , 3)Ds2 (βs2 , 3)Di(βi, 3)D0(β0, 3)}, (2.332)

where the displacement operators are given by

D j (β j , k) = exp[β j a†j (k)− β∗j a j (k)], j = s1, s2, i, 0, k = 0, 3. (2.333)

By the remark above, D j (β j ) ≡ D j (β j , 0). On substituting into the relation (2.332)according to (2.319), we obtain that

CS (βs1 , βs2 , βi, β0, 3)

= Tr{ρ(0)Ds1 (βs1 (3))Ds2 (βs2 (3))Di(βi(3))D0(β0(3))}, (2.334)

where

− β∗s1(3) = −β∗s1

ms1s1 − β∗s2ms2s1 + βim is1 + β0m0s1 ,

−β∗s2(3) = −β∗s1

ms1s2 − β∗s2ms2s2 + βim is2 + β0m0s2 ,

βi(3) = −β∗s1ms1i − β∗s2

ms2i + βim ii + β0m0i,

β0(3) = −β∗s1ms10 − β∗s2

ms20 + βim i0 + β0m00. (2.335)

From the known quantum characteristic function for the initial vacuum state

CS (βs1 , βs2 , βi, β0, 0) = exp

⎧⎨

⎩−1

2

j=s1,s2,i,0

|β j |2⎫⎬

⎭ , (2.336)

we derive that

CS (βs1 , βs2 , βi, β0, 3) = exp

{−

j=s1,s2,i,0

|β j |2 B jS

+⎡

⎣−βs1β∗s2

B∗s1s2− βiβ

∗0 B∗i0 +

j=s1,s2

k=i,0

β jβkC∗jk + c.c.

⎦}. (2.337)

Page 80: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 71

Here the coefficients B jS , B jk , C jk can be expressed in the form

B jS = 〈a j (3)a†j (3)〉 − 1

2, j = s1, s2,

B jS = 〈a†j (3)a j (3)〉 + 1

2, j = i, 0,

Bs1s2 = 〈as1 (3)a†s2

(3)〉, (2.338)

Bi0 = 〈ai(3)a†0(3)〉,

C jk = 〈a j (3)ak(3)〉, j = s1, s2, k = i, 0.

Taking into account that 〈a j (3)〉 = 0, j = s1, s2, i, 0, we see that we are consistentwith the more general notation

B jA = 〈Δa j (3)Δa†j (3)〉, j = s1, s2,

B jN = 〈Δa†j (3)Δa j (3)〉, j = i, 0, (2.339)

where Δa = a−〈a〉, and with the coefficients Bs1s2 , Bi0, C jk, j = s1, s2, k = i, 0,

after similar replacement.We confine ourselves to the study of the signal beams in what follows, which are

described by the reduced statistical operator

ρsignal(3) = TriTr0{ρ(3)}, (2.340)

where Tri and Tr0 are partial traces over the idler and escape modes, respectively.Quantum characteristic function in the state described by the statistical operator(2.340) can easily be obtained:

CS (βs1 , βs2 ) ≡ CS (βs1 , βs2 , 3) = CS (βs1 , βs2 , 0, 0, 3). (2.341)

In Perinova et al. (2003), the same function has been introduced as

CS (βs1 , βs2 ; 1, 3) = Tr{ρ(0)Ds1 (βs1 , 1)Ds2 (βs2 , 3)

}

= exp

⎣−∑

j=s1,s2

|β j |2 B jS +(−βs1β

∗s2

B∗s1s2+ c.c.

)⎤

⎦ . (2.342)

In the following we simplify the notation s1, s2 for the signal modes to 1, 2,respectively. From the characteristic function

Cs(β1, β2) = exp[ s

2(|β1|2 + |β2|2)

]CS (β1, β2), (2.343)

where s = 1, 0,−1 in the subscript and also s = N ,S,A denote the normal,symmetrical, and antinormal orderings of field operators, we can establish the Φs

Page 81: Quantum Aspects of Light Propagation

72 2 Origin of Macroscopic Approach

quasidistribution related to the respective ordering of field operators

Φs(α1, α2) = 1

π4

×∫

Cs(β1, β2) exp(α1β

∗1 − α∗1β1 + α2β

∗2 − α∗2β2

)d2β1 d2β2. (2.344)

After integrating, we obtain

Φs(α1, α2) = 1

π2 K12s

× exp

{1

K12s[−B2s |α1|2 − B1s |α2|2 + (B∗12α1α

∗2 + c.c.)]

}, (2.345)

where

B1A = cosh2(κ1), B1S = B1A − 1

2, B1N = B1A − 1,

B2A = cosh2(κ2)+ |t| sinh2(κ2) sinh2(κ1),

B2S = B2A − 1

2, B2N = B2A − 1,

B∗12 = t∗ sinh(κ1) cosh(κ1) sinh(κ2), (2.346)

and

K12s = B1s B2s − |B12|2. (2.347)

Especially, for s = −1 it holds that

ΦA(α1, α2) = 1

π2〈α1, α2|ρsignal(3)|α1, α2〉, (2.348)

where |α1, α2〉 is the two-mode coherent state, which yields the expansion

ΦA(α1, α2) = 1

π2exp(−|α1|2 − |α2|2)

∞∑

q=−∞

∞∑

m1=max(0,−q)

×∞∑

n2=max(0,−q)

ρ(m1 + q, m1, n2, n2 + q)α∗(m1+q)1 α

m11 α

∗n22 α

n2+q2√

(m1 + q)!m1!n2!(n2 + q)!(2.349)

for any ΦA quasidistribution that does not depend on α1α2, α∗1α∗2 . Here

ρ(n1, m1, n2, m2) = 〈n1, n2|ρsignal(3)|m1, m2〉 (2.350)

Page 82: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 73

are the usual matrix elements. Equating the expansion coefficients for (2.345) withs = −1 and those of (2.349), we arrive at the expression

ρ(m1 + q, m1, n2, n2 + q) =min(m1,n2)∑

p=max(0,−q)

√(m1 + q)!m1!n2!(n2 + q)!

(m1 − p)!(n2 − p)!p!(p + q)!

× (K12A − B2A)m1−p(K12A − B1A)n2−p B∗p12 B p+q

12

K m1+n2+q+112A

, (2.351)

while obviously

ρ(m1 + q1, m1, n2, n2 − q2) = 0 for q1 �= −q2. (2.352)

(ii) Photon-number statistics

Numbers of photons in signal modes complete the picture of the quantum cor-relation between these beams. The joint photon-number distribution p(n1, n2) canbe expressed in the concise form (Rehacek and Perina 1996). A substitution into(2.351) leads to slightly more complicated expression:

p(n1, n2) = ρ(n1, n1, n2, n2). (2.353)

This distribution can be seen in Fig. 2.12 for |t| = 1, B1N = B2N = 3, |B12| = 3.It differs from the product of pertinent marginal distributions by larger “diagonal”probabilities. To the contrary for |t| small, the joint photon-number distribution isapproximately the product of its marginal photon-number distributions.

Fig. 2.12 Jointphoton-number distributionfor |t| = 1; B1N = B2N = 3,n j ∈ [0, 10], j = 1, 2

As for the experimental arrangement under study, it depends on l1, t, l2, whereasthe numerical demonstration is restricted to the case when the length of the firstcrystal is kept fixed. Consequently, the mean photon number B1N in the first sig-nal mode is constant and this convenient behaviour is, for the sake of illustrations,

Page 83: Quantum Aspects of Light Propagation

74 2 Origin of Macroscopic Approach

extended also to the second one as a relationship between |t| and κ2,

sinh2(κ2) = B2N1+ |t|2 B1N

. (2.354)

The Glauber degree of coherence (Perina 1991) γ(2)12 is the complex-valued quan-

tum correlation measure related to the normal ordering:

γ(2)12 = B∗12√

B1N B2N. (2.355)

In the numerical illustration of the quantum correlation measures, we assume κ1 =κ2 = κ and find κ1 by the inversion of the formula

B1N = sinh2(κ1) = 〈a†s1

(1)as1 (1)〉 = n1(1). (2.356)

The limit case |t| = 1 is interesting, |γ (2)12 | = RN = 1, see Fig. 2.13. This maximum

correlation does not correspond to a weaker correlation between the signal photonnumbers. In the multimode analysis (Wang et al. 1991a) of the experiment (Zou et al.

Fig. 2.13 Quantumcorrelation measure RNversus the modulus of thetransmission amplitudecoefficient |t| ∈ [0, 1]; it isassumed thatn1(1) = 10−2, 1, 10, 100, 104

(the curves a, b, c, d, e,respectively)

1991), the visibility of the interference between the signal fields has been expressed,the interference manifests itself as oscillations in the counting rate Is (see (2.412)below) when propagation times of the idler beam from NL1 to NL2, of the firstsignal beam from NL1 to Ds, and of the second signal beam from NL2 to Ds areincremented by δτ0, δτ1, δτ2, respectively. Deriving a simplified visibility Vsimple forthe formalism of several modes provides

Vsimple = 2RN√

B1N B2NB1N + B2N

. (2.357)

As√

B1N B2N ≤ 12 (B1N+B2N ), the visibility cannot exceed the correlation measure

RN and the equality is attained for B1N = B2N . The maximum obtainable visibility

Page 84: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 75

between two fields in an experiment is given by the correlation measure RN , cf.Wiseman and Mølmer (2000).

On substituting (2.338) with (2.316), (2.317), and (2.318) into (2.355), we findthat

RN = |t| cosh(κ1)√1+ |t|2 sinh2(κ1)

. (2.358)

Noting that the idler beam, before it enters the beam splitter, has the same statisticsas the output signal 1, we can rewrite (2.358) in terms of the mean photon numbern1(1) = sinh2(κ1) as

RN = |t|√

1+ n1(1)

1+ |t|2n1(1). (2.359)

Wiseman and Mølmer (2000) considered the relevant limits in this form. The single-photon regime which is the regime of experiment and theory in Zou et al. (1991)and Wang et al. (1991a,b) occurs for n1(1) � 1. Up to the first order in the rescaledlengths κ1, κ2 of the crystals, we simply obtain RN = |t|. The probability of adown conversion at crystal NL1 over interaction time is less than or equal to n1(1),or it is small. The probability to have down conversions at both crystals over inter-

action time is less than or equal to⟨a†

s1 (1)as1 (1)a†s2 (3)as2 (3)

⟩= B1N B2N + |B12|2,

or it is negligible. The single-photon regime applies in the multimode analysis ofWang et al. (1991a), because each of the narrow-bandwidth signal modes (ks1 , ωs1 ),(ks2 , ωs2 ), with directions characterized by ks j and with the frequencies ωs j , j =1, 2, that form broad-band signal fields, receives only a small part of the pumpingphotons over interaction time. The same applies to idler modes and idler fields. Thesignal fields s1 and s2 from the two down converters are allowed to come togetherand interfere at the detector Ds.

In the spatial interaction picture, the state of the field produced by the crystals isgiven by

|ψ(3)〉 ≡ U3(0)U2(0)U1(0)|0〉s1,i,s2,0, (2.360)

where U j+1(0) for j = 1, 2, 3 are given by relations (2.299) and (2.300), wherethe annihilation operators as j

2 +1( j) → as j

2 +1(0), ai( j) → ai(0), a0( j) → a0(0).

We will drop the argument (0) at the annihilation operators in what follows. Here|0〉s1,i,s2,0 ≡ |0〉s1 |0〉i|0〉s2 |0〉0 and, in general,

|ns1 , ni, ns2 , n0〉 ≡ |ns1〉s1 |ni〉i|ns2〉s2 |n0〉0. (2.361)

In the Schrodinger picture, the operators do not change and in the interaction picture,which is the modified Schrodinger picture, they change like the Heisenberg picturefree-field operators. An analogue of relation (2.298) for a discrete space is not used

Page 85: Quantum Aspects of Light Propagation

76 2 Origin of Macroscopic Approach

in quantum optics (the free-field propagation is absorbed in the interaction). Fortu-nately, we will use just the interaction-picture annihilation operators. Expanding theoperators U j+1(0) according to κ j

2+1, j = 0, 2, we obtain that

|ψ(3)〉 � |0〉s1,i,s2,0 + iκ2|0, 1, 1, 0〉 + iκ1(t|1, 1, 0, 0〉 + r|1, 0, 0, 1〉), (2.362)

when κ j2+1 are small. For |t| = 1 we have a single-photon state in the idler mode

and in the collection of the signal modes. In general, one can infer a conversion atcrystal NL1 after a photocount in the escape mode. We introduce the probability ofthe detection

p1,0,0,1(3) = |κ1|2|r|2. (2.363)

Let us assume that one infers a conversion at crystal NL2 after no photocounts in theescape mode. We introduce the probability of a correct inference of the conversionat crystal NL2

p0,1,1,0(3) = |κ2|2 (2.364)

and that of such a wrong inference

p1,1,0,0(3) = |κ1|2|t|2. (2.365)

On a photocount in the escape mode it is certain that the conversion has happenedat NL1. On no photocounts in this mode, the posterior probabilities are

Prob(ns1= 0 ∩ ns2

= 1|ni = 1 ∩ n0 = 0) = |κ2|2|κ1|2|t|2 + |κ2|2 , (2.366)

Prob(ns1= 1 ∩ ns2

= 0|ni = 1 ∩ n0 = 0) = |κ1|2|t|2|κ1|2|t|2 + |κ2|2 . (2.367)

Here the underlining means a random variable. The counting rate registered byDs exhibits perfect interference when the idler fields are perfectly aligned. This maybe regarded as reflecting the intrinsic impossibility of knowing whether the detectedphoton comes from NL1 or NL2 (Wang et al. 1991a). The multiphoton conditionalstates can be found in Luis and Perina (1996a).

Let us consider the annihilation operators as j2 +1

, j = 0, 2. The action of the two

operators on the state |ψ(3)〉 is asymptotically for small κ j2+1 expressed as

as1 |ψ(3)〉 � iκ1(t|0, 1, 0, 0〉 + r|0, 0, 0, 1〉), (2.368)

as2 |ψ(3)〉 � iκ2|0, 1, 0, 0〉. (2.369)

Hence

〈a†s1

as2〉 � κ1κ2t∗, (2.370)

〈a†s1

as1〉 � κ21 , 〈a†

s2as2〉 � κ2

2 . (2.371)

Page 86: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 77

It can be verified that these relations hold up to the second order in κ1, κ2. We canfind that γ

(2)12 = t∗ in the limit of small κ j

2+1, j = 0, 2. Obviously, the equationholds up to the zeroth order, but it can be verified that it is valid up to the first orderin κ1, κ2.

The opposite regime is that where n1(1) � 1. Here there are many photonson average in all of the down-converted beams. That is, the phase of the stage-1or stage-2 idler mode should lock the phase of the output signal-2 mode for anynonzero transmission amplitude coefficient t.

(iii) Multimode formalism

In Wang et al. (1991a) the pump beams at each crystal are represented by com-plex analytic signals V1(t) and V2(t) such that |Vj (t)|2 is in units of photons persecond ( j = 1, 2). The multimode formalism enables one to respect that the twocrystals are centred at 01 and 02. The multimode formalism views the electric fieldsas temporal-interaction-picture operators

E (+)m (r, t) =

√δω

ωm

am(ωm) exp[i(km · r− ωmt)], m = s1, i, s2, 0, (2.372)

where δω is the mode spacing and am(ωm) is the photon annihilation operatorfor narrow-bandwidth signal (m = s j , j = 1, 2), idler (m = i), and “escape”(m = 0) modes (km, ωm) at the frequency ωm . The Hilbert space for the mul-timode analysis is a tensor product of those Hilbert spaces of separate modes,whose vacuum states may be designated as |0〉m(ωm), m = s1, i, s2, 0. We con-sider photon-flux amplitude operators E (+)

s j(t), j = 1, 2, at the appropriate detector,

E (+)s j

(t)≡Es j (rs, t) = E (+)s j

(0 j , t − τ j ), where τ j , j = 1, 2, is the propagation timeof s j from NL j to Ds.

In order to compare the sophisticated multimode formalism with the simple for-malism of several modes, we must present another appropriate description of thedynamics of the same down-conversion experiment. We adopt the temporal interac-tion picture and combine it with the spatial interaction picture. In this case, the stateproduced by the crystal is given by

|ψ(3, t)〉 ≡ U3(0, t)U2(0, t)U1(0, t)|0〉s1,i,s2,0. (2.373)

Here |0〉s1,i,s2,0 is the vacuum state of all the narrow-bandwidth signal, idler, and“escape” modes (ks j , ωs j ), j = 1, 2, (ki, ωi), (k0, ω0), respectively. U j+1(0, t), j =0, 2, are unitary operators:

U j+1(0, t) = limt0→−∞ U j+1(0, t, t0), (2.374)

where U j+1(0, t, t0) are unitary operators that obey the initial condition

U j+1(0, t, t0)∣∣t=t0

= 1, (2.375)

Page 87: Quantum Aspects of Light Propagation

78 2 Origin of Macroscopic Approach

U j+1(0, t, t0) for j = 0, 2 describe the parametric down conversion and areexpressed, indirectly, in terms of E (+)

s j2 +1

(r, t), E (+)i (r, t); U2(0, t) describes the beam

splitter and is expressed as

U2(0, t) = exp{

i∑

ω′′

[ω0a†

0(ω′′)a0(ω′′)+ ωia†i (ω′′)ai(ω

′′)

+(γ ∗a†

0(ω′′)ai(ω′′)+ H.c.

) ]}. (2.376)

In this point we differ from the paper by Wang et al. (1991a), who used the initialcondition at t ′ = t − t1 and they did not write down the decomposition into stages.Let T denote the time ordering. We will introduce the unitary operator

U3(0, t, t − t1) = T exp

{∫ t

t−t1

[ν2V2(0)δω

ω′

ω′′φ(ω0 − ω′′, ω′′)

×e−i(k′s2+k′′)·02 e−i(ω0−ω′−ω′′)t ′ a†

s2(ω′)a†

i (ω′′)− H.c.]

dt ′}, (2.377)

where ν j , j = 1, 2, is a constant such that |ν j |2 gives the fraction of incident pumpphotons that is spontaneously down converted in the steady state, ω0 is the frequencyof the monochromatic pump beam, k′s2

(k′′) is a wave vector that is determined bythe frequency ω′s2

(ω′′) and the direction of the second signal beam (the idler beam).To the first order in the processes the unitary operator may be expressed as

U3(0, t, t − t1) = 1+{ν2V2(0)δω

ω′

ω′′φ(ω0 − ω′′, ω′′)

× e−i(k′s2+k′′)·02

sin[

12 (ω0 − ω′ − ω′′)t1

]

12 (ω0 − ω′ − ω′′)

× exp

[−i(ω0 − ω′ − ω′′)

(t − t1

2

)]a†

s2(ω′)a†

i (ω′′)− H.c.

}.

(2.378)

From this we obtain the vector

|ψ(3, t, t − t1)〉 = U3(0, t, t − t1)|ψ(2, t, t − t1)〉= |ψ(2, t, t − t1)〉 + ν2V2(0)δω

ω′

ω′′φ(ω0 − ω′′, ω′′)

× e−i(k′s2+k′′)·02

sin[

12 (ω0 − ω′ − ω′′)t1

]

12 (ω0 − ω′ − ω′′)

× exp

[−i(ω0 − ω′ − ω′′)

(t − t1

2

)]|ω′〉s2 |ω′′〉i|0〉s1,0,

(2.379)

Page 88: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 79

where |ω′〉s2 and |ω′′〉i are frequency eigenstates of the second signal and the idlerbeam, respectively, |0〉s1,0 is the vacuum state of the first signal and escape modes,and

|ψ(2, t, t − t1)〉 = U2(0, t)U1(0, t, t − t1)|0〉s1,i,s2,0

= U2(0, t)U1(0, t, t − t1)U †2 (0, t)|0〉s1,i,s2,0. (2.380)

We transform the vector |ψ(3, t, t − t1)〉 to a vector

Es2 (t)|ψ(3, t, t − t1)〉 = ν2V2(0)

√δω

2πδω∑

ω′

ω′′φ(ω0 − ω′′, ω′′)

× e−ik′′ ·02sin[

12 (ω0 − ω′ − ω′′)t1

]

12 (ω0 − ω′ − ω′′)

× exp

[−i(ω0 − ω′ − ω′′)

(τ2 − t1

2

)]exp[−i(ω0 − ω′′)(t − τ2)]

× |ω′′〉i|0〉s1,s2,0 (2.381)

� ν2V2(0)

√δω

ω′′φ(ω0 − ω′′, ω′′)

× e−ik′′ ·02

∫ ∞

−∞

sin[

12 (ω0 − ω′ − ω′′)t1

]

12 (ω0 − ω′ − ω′′)

× exp

[−i(ω0 − ω′ − ω′′)

(τ2 − t1

2

)]dω′ exp[−i(ω0 − ω′′)(t − τ2)]

× |ω′′〉i|0〉s1,s2,0 (2.382)

= ν2V2(t − τ2)|1(02, t − τ2)〉i|0〉s1,s2,0, (2.383)

where the single-photon state of the idler beam

|1(r, t)〉i =√

2πδω∑

ω′′φ(ω0 − ω′′, ω′′)

×e−i(k′′ ·r−ω′′t)|ω′′〉i, (2.384)

φ(ω′′, ω′′) is connected with spectral functions φ j (ω′, ω′′; ω) characterizing the sig-nal and idler fields at any crystal NL j ,

φ(ω, ω) = φ1(ω, ω) = φ2(ω, ω),

φ j (ω′′, ω′′) = φ j (ω

′′, ω′′; ω0), j = 1, 2. (2.385)

The frequency eigenstates are single-photon states:

|ω′′〉m = a†m(ω′′)|0〉m, (2.386)

Page 89: Quantum Aspects of Light Propagation

80 2 Origin of Macroscopic Approach

where

|0〉m =⊗

ω′′|0〉m(ω′′), m = i, 0. (2.387)

Unfortunately, the nonvanishing result is obtained only for

0 < τ2 < t1. (2.388)

To resolve this, Wang et al. (1991a) let t1 →∞. Should t1 mean the interaction time,it is better to change the integration limits, namely not to consider the integrationinterval [t − t1, t], but, for instance,

[t − (K2 + 1)t1, t − K2t1], (2.389)

for

K2t1 < τ2 < (K2 + 1)t1 (2.390)

to hold.We further calculate

Es1 (t)|ψ(3, t, t − t1)〉 = Es1 (t)|ψ(2, t, t − t1)〉. (2.391)

We obtain the appropriate component of the vector |ψ(2, t, t − t1)〉 by the action ofthe unitary operator

U2(0, t)U1(0, t, t − t1)U †2 (0, t) = T exp

{∫ t

t−t1

[ν1V1(0)δω

×∑

ω′

ω′′φ(ω0 − ω′′, ω′′)e−i(k′s1

+k′′)·01 e−i(ω0−ω′−ω′′)t ′

× a†s1

(ω′)[t∗ai(ω′′)+ r∗a0(ω′′)]† − H.c.

]dt ′}. (2.392)

The calculation proceeds similarly as in the case of NL2, but we replace a†i (ω′′) by

ta†i (ω′′)+ra†

0(ω′′), |ω′′〉i by t|ω′′〉i+r|ω′′〉0 and we change all the other subscriptsthat underlie to a change, so that

Es1 (t)|ψ(3, t, t − t1)〉 = ν1V1(01, t − τ1)[t|1(01, t − τ1)〉i|0〉s1,s2,0

+ r|1(01, t − τ1)〉0|0〉s1,i,s2

], (2.393)

where |0〉s1,s2,0 ≡ |ψvac〉s1,s2,0, |0〉s1,i,s2 ≡ |ψvac〉s1,i,s2 stand for vacuum states,|1(r, t)〉i is defined in (2.384), and |1(r, t)〉0 stands for single-photon state of the“escape” beam

|1(r, t)〉0 ≡√

2πδω∑

ω′′φ(ω0 − ω′′, ω′′) exp[−i(k′′ · r− ω′′t)]|ω′′〉0. (2.394)

Page 90: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 81

Here a nonvanishing result is obtained only for

0 < τ1 < t1. (2.395)

Considering a change of the integration limits as above, we see that, to the first orderof the processes, no difficulties arise if we change the limits independently of NL2.We do not consider the integration interval [t − t1, t], but, for instance,

[t − (K1 + 1)t1, t − K1t1], (2.396)

for

K1t1 < τ1 < (K1 + 1)t1 (2.397)

to hold.In other words, the relations (2.383) and (2.393) can be generalized to provide

Acτ1 [E (+)s1

(t)|ψ(3, t)〉] = Es1 (t)|ψ(3, t − K1t1, t − (K1 + 1)t1)〉, (2.398)

Ac(τ0+τ2)[E (+)s2

(t)|ψ(3, t)〉] = Es2 (t)|ψ(3, t − K2t1, t − (K2 + 1)t1)〉, (2.399)

where τ0 is the propagation time of the idler from NL1 to NL2. A f denotes anattenuation of the field down to the vacuum state outside an interaction lengthcentred at the distance f from NL1 in the direction of propagation of the beam.The operator A f compensates for the difference we have caused with the initialcondition at t ′ = t0 → −∞ instead of the Wang–Zou–Mandel shortening of theintegration interval. The operator A f is not unitary and is even “slightly” nonlinear.Its consideration depends on a neglect of the coherence length in comparison withthe interaction length. Using such an operator we can describe, where (within whichinteraction length) the single-photon states are localized at the time t ,

Acτ [|1(01, t − τ )〉i|0〉s1,s2,0] = |1(01, t − τ )〉i|0〉s1,s2,0, (2.400)

Acτ [|1(01, t − τ )〉0|0〉s1,i,s2 ] = |1(01, t − τ )〉0|0〉s1,i,s2 . (2.401)

The angular brackets will mean averages and, when operators are involved, thebrackets are supposed to average in the state |ψ(3, t)〉,

〈M〉 = 〈ψ(3, t)|M |ψ(3, t)〉, (2.402)

with M being an operator. When the operator is situated inside an interaction lengthcentred in the propagation distance f from NL1, it also holds that

〈M〉 = A f [〈ψ(3, t)|]M A f [|ψ(3, t)〉]. (2.403)

Hence, one may omit the unusual notation when no ambiguities arise.

Page 91: Quantum Aspects of Light Propagation

82 2 Origin of Macroscopic Approach

Letting ωs and ωi denote the centre frequency of the signal beam and the idlerbeam, respectively, we have ωs + ωi = ω0. Introducing the normalized correlationfunction μ(τ ) of the down-converted light

i〈1(r1, t − τ1)|1(r2, t − τ2)〉i = μ(τ0 + τ2 − τ1) exp[−iωi(τ0 + τ2 − τ1)], (2.404)

where

e−iωiτμ(τ ) = 2π

∫ ω0

0|φ(ω, ω)|2e−iωτ dω, (2.405)

we obtain that the relations (2.370) and (2.371) ought to read

〈E (−)s1

(t)E (+)s2

(t)〉 � ν∗1ν2t∗〈V ∗

1 (t − τ1)V2(t − τ2)〉×μ(τ0 + τ2 − τ1) exp[−iωi(τ0 + τ2 − τ1)], (2.406)

〈E (−)s1

(t)E (+)s1

(t)〉 � |ν1|2〈|V1(t − τ1)|2〉,〈E (−)

s2(t)E (+)

s2(t)〉 � |ν2|2〈|V2(t − τ2)|2〉, (2.407)

where we introduce E (−)s j

(t) ≡[

E (+)s j

(t)]†

. Hence the modulus of the normalized

correlation function is

|〈E (−)s1

(t)E (+)s2

(t)〉|√〈E (−)

s1 (t)E (+)s1 (t)〉〈E (−)

s2 (t)E (+)s2 (t)〉

= 〈V ∗1 (t − τ1)V2(t − τ2)〉√

〈|V1(t − τ1)|2〉〈|V2(t − τ2)|2〉|μ(τ0 + τ2 − τ1)||t|. (2.408)

The maximum value is equal to |t|, which is predicted also by equation (2.359).A linear dependence of visibility on |t|, as seen convincingly in the original work(Zou et al. 1991, Wang et al. 1991a), is the true signature of induced coherencewithout induced emission.

As concerns E (+)s j

(t)|ψ(t)〉, j = 1, 2, they are not explicitly presented in Wanget al. (1991a), but they may be derived. It emerges that the parameters of the beamsplitter do not enter the relation for E (+)

s1(t)|ψ(t)〉. On the contrary, E (+)

s1(t)|ψ(t)〉 in

Wang et al. (1991a) comprises the parameters t∗, r′∗. Nevertheless, the statisticalproperties in Perinova et al. (2003) coincide with those in Wang et al. (1991a),because the differences under discussion resemble distinct, yet equivalent pictures.Especially, considering the photon-flux amplitude operators E (+)

s (t) at the detectorDs with a quantum efficiency ηs (Wang et al. 1991a),

E (+)s (t) = 1√

2

[iE (+)

s1(t)+ E (+)

s2(t)], (2.409)

substituting into the formula for the average rate of photon counting

Is = ηs〈ψ(t)|E (−)s (t)E (+)

s (t)|ψ(t)〉, (2.410)

Page 92: Quantum Aspects of Light Propagation

2.3 Quantum Description of Experiments with Stationary Fields 83

where

E (−)s (t) = [E (+)

s (t)]†, (2.411)

and taking into account the orthogonality of single-photon states |1(r1, t)〉i and|1(r1, t)〉0 uniquely results in the relation

Is = 1

2ηs{|ν1|2〈|V1(t − τ1)|2〉 + |ν2|2〈|V2(t − τ2)|2〉

+[−iν∗1ν2〈V ∗1 (t − τ1)V2(t − τ2)〉t∗μ(τ0 + τ2 − τ1)e−iωi(τ0+τ2−τ1) + c.c.]}. (2.412)

Perinova et al. (2000) have studied quantum statistics of radiation in signal modesof the two-mode parametric processes with aligned idler beams. They have foundthat the signal beams are in the correlated chaotic state. The strength of correla-tion depends on the degree to which the paths of the idler beams are superposedand aligned. They have compared different measures of correlation, especiallythe entropic or information-based measure with the modulus of the usual degreeof coherence in the dependence on absolute value of the transmission amplitudecoefficient of the beam splitter inserted as an attenuator of the perfect alignment.Some other measures have been introduced taking into account the symmetrical andantinormal orderings of field operators. In contrast to the normal ordering, theseorderings do not indicate the maximum correlation for the perfect alignment. Thesituation with the photon numbers in the signal modes, whose correlation is notmaximum for the perfect alignment, serves as motivation for such a more generalconsideration. The theory of canonical correlation has been applied to the quasidis-tribution of complex amplitudes related to the symmetrical ordering of field oper-ators. They have taken into account that the quantum correlation has a significanteffect on the photon-number sum, photon-number difference, and quantum phase-difference statistics. Essentially, it concerned the variances of number sum and num-ber difference and the dispersions of quantum phase differences according to vari-ous definitions. A comparison of distributions of quantum phase difference derivedfrom the phase-space distributions has shown that the phase-difference uncertaintyincreases from the normal ordering, through the symmetrical and antinormal order-ings, whereas the system of canonical phase related to the antinormal ordering ofexponential phase operators ranges between the symmetrical and antinormal order-ings, but by no means exactly. The paper (Perinova et al. 2000) reveals that thecorrelated chaotic state is the mixed partial phase-difference state. In addition tothe marginal distributions, the joint number-sum and phase-difference distributionhas been considered, but for the canonical quantum phase difference and the Luis–Sanchez-Soto phase difference only. The quasidistribution of number difference andphase difference has been defined with the properties that the marginal distributionof the phase difference is the canonical one. They have addressed the number sumand the quantum phase difference as simultaneously measurable observables andthe number difference and the quantum phase difference as canonically conjugateobservables.

Page 93: Quantum Aspects of Light Propagation

84 2 Origin of Macroscopic Approach

Perinova et al. (2003) have compared the simple formalism of several coupledharmonic oscillators with multimode formalism in the analysis of an interferenceexperiment. On focusing on several modes they have been able to study phase prop-erties of “correlated chaotic beams”. Then they have assumed the single-photonregime as also previous authors did. They have indicated that, assuming severalcoupled harmonic oscillators, the previous authors did not try to include time delaysbetween optical elements into the analysis. Perinova et al. (2003) have also formallyexpressed, for instance, that one works with a single-photon state of some signalmodes in the several-mode formalism whenever one describes the experiment witha superposition of single-photon states of modes that form the signal beam.

The utility of a simple single-mode theory has been clarified in the case wheresingle spatial mode filters and narrow-band optical filters are used to filter the outputstate of parametric down-conversion Li et al. 2005).

Perina and Krepelka (2005) have derived joint photon-number distributions insignal and idler modes and have illustrated related concepts taking into accountexperimental data. Perina and Krepelka (2006) have provided the generalization ofthis description to stimulated parametric down-conversion. Perina et al. (2007) havereported on a measurement of the joint signal–idler photoelectron distribution oftwin beams. Parameters of the previously published model (Perina and Krepelka2005) have been estimated. The specific result that the joint signal–idler quasidis-tribution of integrated intensities can be approximated by a well-behaved functioneven in the case where the quasidistribution is not an ordinary function has beencomprised. Perina (2008) has shown that a nonlinear planar waveguide pumped bya beam orthogonal to its surface may serve as a versatile source of photon pairs.He considers the pump-pulse duration, pump-beam transverse width, and angulardecomposition of the pump-beam frequency and their effect on characteristics ofa photon pair, such as the spectral widths of signal and idler fields, the pair timeduration, and the degree of entanglement between the two fields.

Page 94: Quantum Aspects of Light Propagation

Chapter 3Macroscopic Theories and Their Applications

There were several attempts at a justification of the momentum-operator approach.It is appropriate that they include quantization of the electromagnetic field at least inthe one-dimensional case. A complete analysis could be provided only for the para-metric processes, in which the momentum operator is effectively quadratic. It hasbeen noted that the nonlinearity of the process may lead up to a need of a renormal-ization. Nevertheless, there is a modicum of papers on this theme in quantum optics.

A general approach to quantization of the electromagnetic field in a nonlinearmedium enables one to compare properties of the momentum operator with thoseof the space–time displacement operator. We present applications of the traditionalapproach in quantum optics.

The spatio-temporal approach has been developed with respect to quantum soli-tons. An attempt has been made to take into account the frequency dispersion of amedium at least up to inclusion of the group velocity and to preserve the traditionaldescriptions of nonlinear processes by introducing narrow-band fields.

A mention of the quasimode description of the spectrum of squeezing will berestricted to an analysis of coupling of the cavity modes and propagating modes.The paraxial propagation of a light beam with the parabolic approximation and theasymptotic expansion of the beam has been completed with a quantized version. Asan example the nonlinear process has been presented, whose description includesthe renormalization.

In optical imaging with nonclassical light one wants to investigate the quantumfluctuations of light at different spatial points in the plane perpendicular to the prop-agation direction of the light beam. In such spatial points very small photodetec-tors or pixels may be located. Finally the application of one of the macroscopicapproaches has led to the description of several linear optical devices and to thestudy of radiating atoms in a linear medium, which is a recurrent theme by the way.

3.1 Momentum-Operator Approach

Several papers devoted to macroscopic approaches to quantization of the elec-tromagnetic field advocate the momentum operator. In general, such an operator

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 3, C© Springer Science+Business Media, LLC 2009

85

Page 95: Quantum Aspects of Light Propagation

86 3 Macroscopic Theories and Their Applications

should be one of the space–time displacement operators. To the contrary, almostparadoxical properties of these operators have been derived. The exposition isconcluded with applications of the traditional approach to the nonlinear opticalcouplers.

3.1.1 Temporal Modes and Their Application

Huttner et al. (1990) have developed a formalism that describes in a fully quantum-mechanical way the propagation of light in a linear and nonlinear lossless dispersivemedium. At first, they assume a similar situation as Abram (1987), i.e. they consideronly the one-dimensional case restricting themselves even to fields propagating inthe +z-direction only. They take for granted that in quantum field theory there is agenerator for spatial progression, i.e. that relation (2.49) holds for any operator. Theyremark that the change in the quantization volume pointed out by Abram (1987) isnot defined when the medium is dispersive, i.e. when the refractive index dependson the frequency, but they develop Abram’s idea of the use of the energy flux notbeing dependent on the medium (cf. Caves and Crouch 1987). In their opinion,the classical analysis of nonlinear optical processes shows that in order to obtainsimple equations of propagation it is useful to introduce photon-flux amplitude, i.e.,a quantity whose square is proportional to the photon flux.

At present we hesitate to accept the consequences of their approach (cf. howeverBen-Aryeh et al. 1992). Specifying the state at a given point (e.g. z = 0) and withina time period T cannot substitute specifying the state at an initial time (t = 0)and within a quantization length L . Temporal modes of discrete frequencies ωm ,where ωm = 2mπ

T , cannot substitute the spatial modes. The equal-space commutationrelations

[a(z, ωi ), a†(z, ω j )] = δi j 1 (3.1)

cannot substitute the usual equal-time commutation relations.In comparison with Abram (1987), we see the following changes. In Huttner et al.

(1990), the MKSA (SI) system of units is used. Instead of immediately reducing theunsymmetrical Maxwell stress tensor to a single component, the momentum densityis first reduced. The normal ordering is used where necessary. The notation ceasesto express the dependence on both z and t and states the dependence on z only. “Thegeneralization” of the relation for the momentum-flux operator

g(z, t) = c[D(−)(z, t)B(+)(z, t)+ H.c.], (3.2)

where H.c. means the Hermitian conjugate to the previous term, to the form

g(z, t) = [D(−)(z, t)E (+)(z, t)+ H.c.] (3.3)

is not founded well. Its integration over T gives the momentum operator G(z),

G(z) ≡∫ t0+T

t0

g(z, t) dt. (3.4)

Page 96: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 87

In the case of a linear dielectric medium in contrast with Abram (1987), theelectric-field operator E(z, t) is dependent on the refractive index n(ωm)

E(z, t) =∑

m

√�ωm

2ε0cT n(ωm)

[a(z, ωm)e−iωm t + H.c.

], (3.5)

while in Abram (1987) the operator is independent of the medium. In Abram (1987)there is not pure Heisenberg picture, so that the equivalence of the two theories (forn independent of ω) is not excluded. From relation (3.3), the momentum operator isobtained

G lin(z) =∑

m

(�km)a†(z, ωm)a(z, ωm), (3.6)

where km = n(ωm )ωm

c is the wave vector in the (linear) medium. The equal-space com-mutation relations are conserved. For such a medium, the equal-time commutationrelations can be derived

[A(z, t),−D(z′, t)

] = i�δ(z − z′)1. (3.7)

Attempting at the quantization in a nonlinear medium, Huttner et al. (1990)have concentrated on the propagation of light in a multimode degenerate paramet-ric amplifier. The postulated relation (3.3) then leads to the nonlinear part of themomentum-flux operator

gnonlin(z, t) = χ (2){E (+)(z, t)

[E (−)(z, t)

]2 + H.c.}

, (3.8)

where E (+)(z, t) = |E |e−i(ωpt−kpz) is the positive-frequency part of the pump field,with the pump frequency ωp. From relations (3.4) and (3.8), the momentum operatoris obtained

Gnonlin(z) = �

4

m

λ(εm)[a†(z, ω0 + εm)a†(z, ω0 − εm)eikpz + H.c.

], (3.9)

where

εm ≡ ωm − ω0, ω0 = ωp

2(3.10)

and

λ(εm) ≡ χ (2)|E |ε0c

√ω0 + εm

n(ω0 + εm)

ω0 − εm

n(ω0 − εm)(3.11)

is the coupling constant between the different modes.

Page 97: Quantum Aspects of Light Propagation

88 3 Macroscopic Theories and Their Applications

It is assumed that the phase-matching condition at ω0, n(ωp)= n(ω0), is satisfied.It is found that the phase-mismatch Δk(εm) is proportional to ε2

m . As far as |Δk(εm)|< λ(εm), the Bogoliubov transformation for squeezing emerges and amplifyingbehaviour can be recognized. In the case |Δk(εm)| > λ(εm), the evolution is notessentially different from that in a linear medium, the squeezing effect is band lim-ited. For the equality |Δk(εm)| = λ(εm), the amplifying is present, but the increaseis only linear not exponential.

For the nonlinear medium, the equal-time commutation relations are

[A(−)(z, t),−D(+)(z′, t)

] ≈ i�

2δ(z − z′)1 (3.12)

and relation (3.7) can be recovered only approximately. In relation to the experi-ment, a standard two-port homodyne detection scheme is assumed, where the lightis mixed at a beam splitter with a strong local oscillator ε(z, t) of the frequencyω0. For the correlation function gS(τ ) of the photocurrent difference and its Fouriertransform

y(η) =∫

gS(τ )e−iητ dτ, (3.13)

we refer to Huttner et al. (1990). It has been shown that the values of y(η) can beminimized uniformly enough by an adequate choice of the local oscillator phase.The result is comparable with Crouch (1988), where the usual interpretation ofhomodyne detection in terms of the field quadratures is used.

3.1.2 Slowly Varying Amplitude Momentum Operator

Nevertheless, there is a class of problems, for which the modal approach is veryconvenient. It is the cavity quantum electrodynamics. Let us mention its use inthe development of the input–output formalism for nonlinear interactions in cav-ity (Yurke 1984, 1985, Collett and Gardiner 1984, Gardiner and Collett 1985,Carmichael 1987). The modal approach can describe many of the features oftravelling-wave phenomena, but, in principal, it mixes effects related to spatial pro-gression of the beam with the spectral manifestations of the nonlinearity. For exam-ple, for the case of the travelling-wave parametric generation (Tucker and Walls1969), a wave vector mismatch appears as energy (frequency) nonconservation.Several authors have tried to return the quantum-mechanical propagation to directspace. One technique (Drummond and Carter 1987) involves the partition of the boxof quantization into finite cells. Another technique considers the spatial progressionof the temporal Fourier components of the local electric field (Yurke et al. 1987,Caves and Crouch 1987). The propagation of light in a magnetic (dielectric) mediumis not considered in quantum optics.

We proceed with the field inside an effective (linear or nonlinear) medium and thedirect-space formulation of the theory of quantum optics as presented by Abram andCohen (1991). It is an alternative of the conventional reciprocal space approach to

Page 98: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 89

quantum optics. Their approach relies on the electromagnetic momentum operatoras well as on the Hamiltonian and is restricted to the dispersionless lossless nonmag-netic dielectric medium. They have derived an operatorial wave equation that relatesthe temporal evolution of an electromagnetic pulse to its spatial progression. Thetheory is applied to squeezed light generation by the parametric down-conversionof a short laser pulse as an illustration. This approach does not use the conventionalmodal description of the field.

The appeal of the classical theory of optics may consist in its considering materialas a continuous dielectric characterized by a set of phenomenological constants.In classical nonlinear optics the slowly varying amplitude approximation of theelectromagnetic wave equation has arisen. An important simplification of quantumoptics results when the microscopic description of the material is replaced by amacroscopic description, in terms of an effective linear or nonlinear polarization.In spite of the phenomenological treatment of the medium, such an effective theorystill permits a quantum-mechanical description of the field (Jauch and Watson 1948,Shen 1967, Glauber and Lewenstein 1989, Glauber and Lewenstein 1991, Hilleryand Mlodinow 1984, Drummond and Carter 1987).

In propagation problems, the interactions undergone by a short pulse of light areexamined. Abram and Cohen (1991) simplify the geometry for the electromagneticfield so that the electric field E is polarized along the x-axis, the magnetic field Balong the y-axis, while propagation occurs along the z-axis. They use the Heaviside–Lorentz units and take � = c = 1. In this simple geometry the Maxwell equationsreduce to two scalar differential equations

∂ E

∂z= −∂ B

∂t, (3.14)

∂ B

∂z= −∂ D

∂t, (3.15)

where the electric displacement field D is defined by

D = E + P, (3.16)

with P being the polarization of the medium, which can be expressed as a converg-ing power series in the electric field E ,

P = χ (1) E + χ (2) E2 + · · · + χ (n) En + · · · , (3.17)

where χ (n) is the nth-order susceptibility of the medium. The dispersion cannot betaken into account rigorously within a quantum-mechanical theory based on theeffective (macroscopic) Hamiltonian formulation (Hillery and Mlodinow 1984), butit can be introduced phenomenologically (Drummond and Carter 1987). To imposethe canonical structure on the field, they introduce the vector potential A and adoptthe Coulomb gauge in which the scalar potential vanishes and A is transverse. In theassumed geometry, the vector potential is polarized along the x-axis and is related

Page 99: Quantum Aspects of Light Propagation

90 3 Macroscopic Theories and Their Applications

to the electric and magnetic fields by

E = −∂ A

∂t(3.18)

and

B = ∂ A

∂z. (3.19)

The effective Lagrangian density has been chosen (Hillery and Mlodinow 1984,Drummond and Carter 1987),

L = 1

2(E2 − B2)+ 1

2χ (1) E2 + 1

3χ (2) E3 + 1

4χ (3) E4 + · · · , (3.20)

which is known to be the most general density dependent only on the electric fieldand having the gauge invariance.

Let us note that the theory with the effective Lagrangian density (3.20) is notrenormalizable (Power and Zienau 1959, Woolley 1971, Babiker and Loudon 1983,Cohen-Tannoudji et al. 1989). The canonically conjugated momentum of A withrespect to the Lagrangian density (3.20) is the electric displacement

Π = ∂L∂ A

= −D. (3.21)

The Lagrangian density is then transformed to some components of the energy–momentum tensor of the electromagnetic field inside a nonlinear medium Θμν ,namely the energy density

Θt t = Π∂ A

∂t− L (3.22)

= 1

2(B2 + A2)+ 1

2χ (1) E2 + 2

3χ (2) E3 + 3

4χ (3) E4 + · · · , (3.23)

and the momentum density

Θt z = −Π∂ A

∂z= DB. (3.24)

In setting up the Hamiltonian functional, the electric field E is to be expressed interms of the electric displacement, which is the canonically conjugated momentumof A according to relation (3.21). It is assumed that

E = β (1) D + β(2) D2 + β(3) D3 + · · · , (3.25)

where the β coefficients may be expressed in terms of the susceptibilities χ (n)

through definition (3.16) and relation (3.17) (Hillery and Mlodinow 1984).

Page 100: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 91

The Hamiltonian functional is then written as

H =∫

VΘt t d3r

=∫

V

(1

2B2 + 1

2β(1) D2 + 1

3β(2) D3 + 1

4β(3) D4 + . . .

)d3r, (3.26)

while the momentum has the form

G =∫

VΘt z d3r =

VB D d3r, (3.27)

where the integration is over the cavity obeying periodic boundary conditions andlower and upper limits are supposed to converge to −∞ and ∞, respectively.

The field can now be quantized by replacing each field variable by the corre-sponding operator and by replacing the Poisson bracket between the displacementD and the vector potential A by the equal-time commutator

[D(r, t), A(r′, t)] = iδT (r− r′)1, (3.28)

exactly by its (−i) multiple, where the transverse δT (r− r′) reduces to the ordinaryδ function and where the three-dimensional position vector r can be replaced by thecoordinate z.

The vector potential A replaced by the operator A does not appear explicitly inthe Hamiltonian (3.26) and momentum (3.27) operators, but rather in terms of itsspatial derivative B. Taking the curl according to r′ of both the sides of the canonicalcommutation relation (3.28) and using relation (3.19) in the simple geometry, weobtain that

[D(z, t), B(z′, t)] = −iδ′(z − z′)1, (3.29)

where

δ′(z − z′) = d

dzδ(z − z′) (3.30)

is the derivative of the δ function.Not caring for divergencies, Abram and Cohen (1991) consider any product of

noncommuting operators that appear in an expression as fully symmetrized, i.e.including all possible permutations of the individual field operators, such as

B D2 �→ B D2 + D B D + D2 B

3. (3.31)

In contrast, more recently they carried out the renormalization, i.e. the normal order-ing and an elimination of divergencies.

Page 101: Quantum Aspects of Light Propagation

92 3 Macroscopic Theories and Their Applications

The description of propagative optical phenomena has been discussed within theframework of a direct-space formulation of quantum optics and the operatorial (orbetter commutator) equivalent of the Maxwell equations and the electromagneticwave equation (Abram and Cohen 1991). It is emphasized that in the Hamiltonianformulation of mechanics the time variable plays a particular role.

The integrals in (3.26) and (3.27) and the equal-time commutator (3.28) corre-spond to the requirement that the field be specified over all space at one instant oftime (e.g. at t = 0).

Abram and Cohen (1991) use the Kubo (1962) notation for the commutator ormore exactly for a corresponding superoperator. The superoperator assigns opera-tors to operators. Respecting this, the Heisenberg equation can be written as follows

∂ Q

∂t= i[H , Q] = iH× Q, (3.32)

where Q is any field operator and the superscript × denotes the superoperator,namely the commutation of the operator it loads with another operator which fol-lows. Equation (3.32) has the solution

Q(t) = exp(it H×)Q(0) (3.33)

= eiH t Q(0)e−iH t . (3.34)

The Heisenberg-like equation involving the momentum can be considered

∂ Q

∂z= −iG× Q. (3.35)

This equation has the solution

Q(z) = exp[−i(z − z0)G×]Q(z0). (3.36)

Apart from the obvious similarity of equations (3.32) and (3.35), there is also adifference. The Hamiltonian of the electromagnetic field relates the desired spatialdistribution of the field at the instant t + dt to its spatial distribution at t , but themomentum operator G relates the translated and nontranslated fields only (at thesame instant of time).

In analogy with the classical equations (3.14), (3.15) rather than with the classicalequations

∂(β (1) D + β(2) D2 + β(3) D3 + · · · )∂z

= −∂ B

∂t,

(3.37)∂ B

∂z= −∂ D

∂t,

Page 102: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 93

two commutator equations can be derived

G× E = H× B, (3.38)

G× B = H× D. (3.39)

On assuming that the medium is homogeneous, so that the Hamiltonian and momen-tum operators commute with each other, that is

G× H = 0, (3.40)

Equations (3.38) and (3.39) may be combined into the commutator equivalent of theelectromagnetic wave equation

G×G× E = H× H× D. (3.41)

Let us note again that it is an analogue of the wave equation

∂2 E

∂z2= ∂2 D

∂t2, (3.42)

not of the more complicated equation

∂2(β(1) D + β(2) D2 + β(3) D3 + · · · )∂z2

= ∂2 D

∂t2. (3.43)

In Abram and Cohen (1991) the direct-space description of propagation (i.e.without resorting to a modal decomposition of a propagating pulse) is illustratedby examining the propagation of light (of the short light pulse) through a linearmedium and through a vacuum–dielectric interface.

For a linear medium the commutator wave equation (3.41) reduces to

(G×G× − ε H× H×)E = 0, (3.44)

where ε = 1 + χ (1) is the dielectric function of the medium. It is also convenient todefine v = 1√

ε, the velocity of an electromagnetic wave in a refractive medium. In

the following exposition, the convention c = 1 will be dissolved or not used. Waveequation (3.44) enables one to rewrite equations (4.2a) and (4.2b) of Abram andCohen (1991) in the form

E(z, t) = cosh(−ivt G×)E(z, 0)− iv sinh(−ivt G×)B(z, 0), (3.45)

B(z, t) = − i

vsinh(−ivt G×)E(z, 0)+ cosh(−ivt G×)B(z, 0). (3.46)

Equations (3.45) and (3.46) indicate that the linear combination

W+v (z, t) = E(z, t)+ v B(z, t) (3.47)

Page 103: Quantum Aspects of Light Propagation

94 3 Macroscopic Theories and Their Applications

evolves in time as

W+v (z, t) = exp(ivt G×)W+

v (z, 0) = W+v (z − vt, 0). (3.48)

Similarly, the linear combination

W−v (z, t) = E(z, t)− v B(z, t) (3.49)

evolves as

W−v (z, t) = exp(−ivt G×)W−

v (z, 0) = W−v (z + vt, 0). (3.50)

To examine the problem of the interface, we now consider two half-spaces suchthat the z ∈ (−∞, 0) half-space is empty, while the z ∈ (0,+∞) half-space consistsof a transparent linear dielectric. We then consider three waves, incident, W+

c (z, t),reflected, W−

c (z, t), and transmitted, W+v (ζ, t), and the relations they obey.

Abram and Cohen derive the commutator equivalent of the slowly varying ampli-tude wave equation, on which the classical theory of nonlinear optics is based(Abram and Cohen 1991). Not even in classical optics, the problem of propagationof a short pulse in a nonlinear medium can be solved in the general case. In classicalnonlinear optics, the assumption of a weak nonlinearity makes the slowly vary-ing amplitude (SVA) approximation of the electromagnetic wave equation possible(Shen 1984). In Abram and Cohen (1991) further a perturbative treatment of thetime evolution of the field in a nonlinear medium is examined that corresponds topropagation within the slowly varying amplitude approximation. For simplicity, asingle nonlinear susceptibility χ (n) is considered.

In the perturbative treatment of the nonlinear propagation it is assumed that theoptical nonlinearity of the medium is absent at t =−∞ and turned on adiabatically.In the absence of the nonlinearity, the electric and magnetic fields in the medium,E0 and B0, as well as the displacement field D0,

D0 = ε E0 (3.51)

propagate under the energy operator

H0 = 1

2

∫ (B2

0 +1

εD2

0

)d3r (3.52)

and the momentum operator

G0 =∫

B0 D0 d3r, (3.53)

which are of zeroth order in the nonlinear susceptibility χ (n). Following the stan-dard perturbation theory (Itzykson and Zuber 1980), the exact field operators in the

Page 104: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 95

nonlinear medium, D and B, can be derived from the zeroth-order fields by theunitary transformation

D(z, t) = U−1(t)D0(z, t)U (t) (3.54)

and

B(z, t) = U−1(t)B0(z, t)U (t). (3.55)

Here U (t) is the unitary operator, which is the solution to the differential equation

∂tU (t) = −iλ ˆH1(t)U (t), (3.56)

with ˆH1(t) the nonlinear interaction part of the Hamiltonian,

ˆH1(t) ≡ 1

n + 1

∫β(n) Dn+1

0 d3r = − 1

n + 1

∫χ (n) En+1

0 d3r (3.57)

or

ˆH1(τ ) = exp(iτ H×0 )H1, (3.58)

and obeys the initial condition

U (t)∣∣∣t→−∞

= 1. (3.59)

The Hamiltonian H1 is first order in the nonlinear susceptibility χ (n), which isexpressed also by λ, a dimensionless parameter, which has been introduced for thebookkeeping of this and higher powers of χ (n). The exact Hamiltonian (3.26) canthen be expressed perturbatively up to the first order in λ as

H = H0 + λH1S + O(λ2), (3.60)

where H1S is the “diagonal part” of H1, S stands for stationary, which commuteswith the linear Hamiltonian H0,

H1S ≡ H (n)1S = − χ (n)

n + 1

∫Sn+1 d3r, (3.61)

with

Sn = 2−n+1[ n

2 ]∑

m=0

n!

(n − 2m)!(2m)!ε−m B2m

0 En−2m0 , (3.62)

Page 105: Quantum Aspects of Light Propagation

96 3 Macroscopic Theories and Their Applications

which leads to the decoupling of opposite-going fields. In the context of (3.61) and(3.62), a connection with the modal approach has been mentioned in Abram andCohen (1991). The notion of the diagonal part belongs to the perturbation theorywhich was treated in Sczaniecki (1983).

According to (3.33), the time evolution of the displacement operator D can beexplained and cast in the form

D(z, t) ≈ exp(itλH×1S)D0(z, t)+ λD1(z, t), (3.63)

where λ ≡ 1 and

D1(z, t) =[

i∫ t

−∞ˆH1(τ ) dτ

]×D0(z, t) (3.64)

is the first-order correction to the displacement field. The action of the superoperatoron D0 in relation (3.63) can be compared with the multiplication of the fast-varying(“carrier”) wave by a slowly varying envelope function.

On introducing the nonlinear polarization

PNL = −εβ(n) Dn0 = χ (n) En

0 , (3.65)

it can be shown that the exact commutator wave equation (3.41) can be written upto order λ0 as

(G×0 G×

0 − ε H×0 H×

0 )D0 = 0 (3.66)

and that to order λ1 as

2ε H×0 H×

1S D0 = −ε H×0 H×

0 D1 + G×0 G×

0 D1 − G×0 G×

0 PNL. (3.67)

The nonlinear polarization PNL consists of two parts, PW and its complement, andPW obeys the zeroth-order wave equation

(G×0 G×

0 − ε H×0 H×

0 )PW = 0, (3.68)

namely

PW ≡ P (n)W = χ (n) Sn. (3.69)

This partition again eliminates all terms that couple opposite-going waves in PNL.Relying on the relation

0 = −ε H×0 H×

0 D1 + G×0 G×

0 D1 − G×0 G×

0 (PNL − PW ), (3.70)

we can derive the commutator equation

2ε H×0 H×

1S D0 = −G×0 G×

0 PW , (3.71)

Page 106: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 97

which has been compared to the classical slowly varying amplitude (SVA) waveequation, which is written as Abram and Cohen (1991)

2ik∂

∂zE = ∂2

∂t2PW , (3.72)

or, more often, in terms of the temporal Fourier components of E and PW as

∂zE(ω) = iω

2√

εPW (ω), (3.73)

where E is the envelope function of the electric field. Also concerning PW , theconnection to the modal approach has been shown in Abram and Cohen (1991).

The commutator equivalent of the slowly varying amplitude wave equation willbe applied to the quantum-mechanical treatment of the propagation in a nonlinearmedium. Let us consider equation (3.71) whose right-hand side does not containD0 in contrast to the left-hand side. This problem can be remedied by defining aneffective “SVA” momentum operator such that it obeys

G×SVA D0 = 1

2G×

0 PW . (3.74)

The solution GSVA is the stationary part of the effective “interaction” momentumoperator G1,

G1 = 1

2

∫B0 PNL d3r, (3.75)

namely

GSVA ≡ G(n)SVA =

χ (n)

n + 1

∫Rn+1 d3r, (3.76)

with

Rn = 2−n+1[ n

2 ]−1∑

m=0

n!

(n − 2m − 1)!(2m + 1)!ε−m B2m+1

0 En−2m−10 . (3.77)

Also, in the context of (3.76) and (3.77) for GSVA, the connection to the modalapproach can be shown. With definition (3.74), the commutator wave equation(3.71) can be written as

(G×SVAG×

0 + ε H×1S H×

0 )D0 = 0. (3.78)

In this form, the commutator SVA equation relates directly the slow componentof the temporal evolution of a short pulse of the displacement field D0 to the long-scale modulation of its spatial progression.

Page 107: Quantum Aspects of Light Propagation

98 3 Macroscopic Theories and Their Applications

In order to clarify the role of Equation (3.78), the forward (+) and backward (−)polarization waves are defined in analogy with (3.47) and (3.49),

V± = D ±√ε B, (3.79)

which in the absence of the nonlinearity have the form

V±0 = εW±

v , (3.80)

where in accordance with the perturbation theory the forward and backward elec-tromagnetic waves are defined as

W±v = E0 ± v B0. (3.81)

Relation (3.63) now becomes

V+(z, t) ≈ exp(it H×1S)εW+

v (z, t)+ V+1 (z, t), (3.82)

V−(z, t) ≈ exp(it H×1S)εW−

v (z, t)+ V−1 (z, t), (3.83)

where V±1 are the first-order corrections to V given by equations analogous to (3.64).

Equation (3.78) simplifies to

√ε H×

1S W+v = −G×

SVAW+v , (3.84)√

ε H×1S W−

v = G×SVAW−

v , (3.85)

for the forward-going and backward-going waves, respectively. These equationsprovide a simple rule for converting the temporal evolution of the modulation enve-lope to the spatial progression, i.e. relations (3.82) and (3.83) can be written asfollows

V+(z, t) = ε exp(−ivt G×SVA)W+

v (z − vt, 0)+ V+1 (z, t), (3.86)

V−(z, t) = ε exp(ivt G×SVA)W−

v (z + vt, 0)+ V−1 (z, t). (3.87)

In most practical situations, the first-order terms V±1 may be neglected and W±

v canbe introduced also on the left-hand sides of equations (3.86) and (3.87) using (3.80).Nevertheless, V±

1 play an important role in that they incorporate the coupling to thewave going in the opposite direction and do give rise to the nonlinear reflection.

As an illustration of the above quantum treatment, the travelling-wave generationof squeezed light by the parametric down-conversion of a short pulse is examined.

For the case of a classical pump, this problem was treated through a modal anal-ysis by Tucker and Walls (1969). More recently, Yurke et al. (1987) and Cavesand Crouch (1987) treated this problem by using spatial differential equations forappropriately defined creation and annihilation operators.

Page 108: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 99

The full modulated wave W can be written in terms of a carrier wave W asfollows

W (z, t) = exp(−izG×SVA)W (z, t), (3.88)

where the subscript v and the superscript + have been omitted. For a medium thatexhibits a second-order nonlinearity, the right-hand side of relation (3.84) can beexpressed as

− G×SVAW = −1

2

χ (2)

√ε

H×0 W 2. (3.89)

It is convenient to separate the field into its positive- and negative-frequency parts:

W = 2(E (+) + E (−)

), (3.90)

where the factor of 2 arises, because it is not in definition (3.47). Similarly, the mod-

ulated wave solution W can be separated into its positive- and negative-frequencyparts

W = 2

(E

(+) + E(−))

. (3.91)

In the first-order perturbative treatment, the optical frequencies retain the same signand relation (3.88) can be modified to

E(+)

(z, t) = exp(−izG×SVA)E (+)(z, t), (3.92)

with a similar equation holding for the negative-frequency part.In Abram and Cohen (1991), equations similar to familiar classical first-order

differential equations have been derived (see Shen 1984). The two fields involved inparametric down-conversion are introduced: the pump field with the central pumpfrequency ωP and the signal field that oscillates at approximately ωS , ωS = ωP

2 . In

fact, we introduce also the notation E (±)P , E (±)

S , E(+)

P , E(+)

S , and we modify relation(3.88) to

E(+)

P (z, t) = exp(−izG×SVA)E (+)

P (z, t),

E(+)

S (z, t) = exp(−izG×SVA)E (+)

S (z, t), (3.93)

and similar equations for the negative-frequency parts. The pump field consists ofa short pulse whose duration TP is much longer than the optical period 2π

ωP. On this

assumption, relation (3.89) for the signal field becomes

G×SVA E (+)

S = −κ E (+)P E (−)

S , (3.94)

G×SVA E (−)

S = κ E (−)P E (+)

S , (3.95)

Page 109: Quantum Aspects of Light Propagation

100 3 Macroscopic Theories and Their Applications

where κ = χ (2)ωS√ε

,

G×SVA E (+)

P = −κ E (+)S E (+)

S (3.96)

and

G×SVA E (−)

P = κ E (−)S E (−)

S . (3.97)

Let us observe that on the substitution Q = E (+)S , E (−)

S , E (+)P , E (−)

P into relation(3.35) and comparison with (3.94), (3.95), (3.96), and (3.97), we have an analogueof the classical description of the spatial progression. Within the undepleted pumpassumption, the solution of (3.93) becomes

E(+)

S (z, t) = coshN

[κz√

IP (z, t)

]E (+)

S (z, t)

+ i

{sinh

[κz√

IP (z, t)

]E (+)

P (z, t)√IP (z, t)

}

N

E (−)S (z, t), (3.98)

where

IP (z, t) = E (−)P (z, t)E (+)

P (z, t) (3.99)

is essentially the intensity operator for the pump field and the subscript N denotesthe normal ordering of the operators E (±)

P , which means that E (−)P stands to the left

from E (+)P .

Deviating slightly from Abram and Cohen (1991), we formulate the interactionpicture as follows:

|(P + S)(t)〉 = U (t)|P(t)〉, (3.100)

where |P(t)〉 is the state of the field (pump and signal) in the remote past,

|P(t)〉 = |P(t)〉P ⊗ |0〉S. (3.101)

We are now in a position to describe a travelling-wave experiment of paramet-ric down-conversion. In such an experiment, a pump pulse expressed by the state|P(t)〉P initially traverses a nonlinear crystal that extends from z = 0 to L andgenerates a signal pulse in the course of its propagation. Using the previous approx-imations, we rewrite (3.100) as

|(P + S)(t)〉 = exp(iGSVAvt)|P(t)〉. (3.102)

Page 110: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 101

The dependence on z which has been introduced by the substitution t = zv

in theprevious exposition is not explicit here. The interaction picture is not needed forcalculation of expectation values of the observables.

For example, a measurement of the intensity profile of the signal pulse can beexpressed by the equal-time function IS(t)

IS(t) = 〈P(t)|E (−)

S (L , t)E(+)

S (L , t)|P(t)〉. (3.103)

Since E(+)

S (L , t) = E(+)

S (L − vt, 0), relation (3.103) after a simplification yields

IS(t) = P

⟨P

(t − L

v

) ∣∣∣∣∣sinh2N

[κL

IP

(0, t − L

v

)]∣∣∣∣∣ P

(t − L

v

)⟩

P

× S〈0|E (+)S E (−)

S |0〉S, (3.104)

with IP(0, t − L

v

) = IP (L − vt, 0); relation (3.104) is simplified by omitting theterms without the antinormal ordering of signal field operators. This can be consid-ered as legitimate, because the vacuum expectation value of the operator productS〈0|E (+)

S E (−)S |0〉S diverges. There exist results for the two-time correlation function

g(1)S (t2, t1) and for the nth-order photon-coincidence rate for the signal pulse g(n)

S .The numerical results have been obtained for a laser pulse that has an amplitude

profile AP (t ′) at z = 0, but have been restricted to the intensity profile measured atthe exit of the crystal z = L . Let us assume that |P(t ′)〉 is a coherent state with theproperty

E (+)P (0, t ′)|P(t ′)〉 = AP

(t ′)

UP(t ′) |P(t ′)〉, (3.105)

where

AP(t ′) ≥ 0,

∣∣UP(t ′)∣∣ = 1. (3.106)

Hence,

P

⟨P

(t − L

v

) ∣∣∣∣ IP

(0, t − L

v

)∣∣∣∣ P

(t − L

v

)⟩

P

=[

AP

(t − L

v

)]2

, (3.107)

so that (after a renormalization)

IS(L , t) = sinh2

[κL AP

(t − L

v

)]. (3.108)

Page 111: Quantum Aspects of Light Propagation

102 3 Macroscopic Theories and Their Applications

3.1.3 Space–Time Displacement Operators

Serulnik and Ben-Aryeh (1991) have discussed a general problem of the electro-magnetic wave propagation through nonlinear nondispersive media. They have usedthe four-dimensional formalism of the field theory in order to develop an extensionof the formalism introduced by Hillery and Mlodinow (1984). The complicationsfollowing from the common definitions for the vector and scalar potentials are indi-cated. It is shown that the scalar potential can be neglected only by using alternativedefinitions.

First, it is shown that the conventional approach that uses the standard potentialsA and V is not appropriate for treating the general case of nonlinear polarizationwhen ∇ · P �= 0, since for such cases V does not vanish. As a solution to thisproblem it is proposed to use vector potential ψ ,

D = −∇ × ψ, (3.109)

which fulfils the relation ∇ · D = 0. This choice enables Serulnik and Ben-Aryeh(1991) to work in the new Coulomb gauge, where ∇ · ψ = 0, so that from thecondition ∇ · B = 0 it follows that the dual scalar potential ξ obeys the equation

∇2ξ = 0. (3.110)

It is then consistent to assume ξ = 0 everywhere in a nonlinear medium and thedual scalar potential need not be taken into account.

The Lagrangian and Hamiltonian densities are derived from the Maxwell equa-tions by using nonconventional definitions for the scalar and vector potentials. Thegeneral form of the energy–momentum tensor is derived and explicit expressionfor its elements is given. The relation between this tensor and the space–timedescription of propagation is analysed. Further the quantization is performed andthe properties of space–time displacement operators are presented. The space–timeis described by a Lie transform (Steinberg 1985). The displacement operators areobtained from the energy–momentum tensor developed by Serulnik and Ben-Aryeh(1991) with an alternative definition for the vector potential. It has been possible toobtain explicit expressions for all the elements of the energy–momentum tensor andto discuss their physical meaning.

In the following we will show that the relationship between the energy–momentumtensor and the space–time description of propagation is different from that derivedby Serulnik and Ben-Aryeh (1991). Let us restrict ourselves to the usually treatedone-dimensional case, where only the fields E1, D1, B2, and Λ2 are significant,where we use Λ2 = −ψ2 according to Drummond (1990, 1994). The arguments ofthese fields are x3 and ct . The corresponding quantum fields obey the commutationrelation

[Λ2(x3, ct), B2(y3, ct)] = i�cδ(x3 − y3)1, (3.111)

[D1(x3, ct), B2(y3, ct)] = −i�cδ′(x3 − y3)1. (3.112)

Page 112: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 103

Considering for this case the Hamiltonian density

H = 1

2

⎜⎜⎜⎝

ζ︷︸︸︷1

εD2

1 + B22

⎟⎟⎟⎠ , (3.113)

where the right-hand side is symmetrically ordered (cf. Abram and Cohen 1991),we obtain the equations of motion in the Heisenberg picture as follows

∂Λ2

∂t= − i

[Λ2,

∫H dx3

]= cB2, (3.114)

∂ B2

∂t= − i

[B2,

∫H dx3

]= −c

∂(

D1)

∂x3. (3.115)

Relation (3.114) explains the role of the dual vector potential and relation (3.115) isessentially the second of the Maxwell evolution equations. We could obtain the firstof them as the equation of motion for the quantum field D1.

It is a question whether the tensor element

T 03 = 1

c(D1 B2)S (3.116)

is a correct quantum density for generation of the displacement as Serulnik andBen-Aryeh (1991) indicate. The presumable equations of the spatial progression are

∂Λ2

∂x3= i

�c

[Λ2,

∫(D1 B2)S dx3

]= −D1, (3.117)

∂ B2

∂x3= i

�c

[B2,

∫(D1 B2)S dx3

]= ∂ B2

∂x3. (3.118)

Relation (3.117) expresses the role of the dual vector potential and relation (3.118)is a mere tautology. The same is obtained for the quantum field D1. This failureof the application of the ordinary presentations of quantum field theory has beenpublished by Ben-Aryeh and Serulnik (1991).

Considering, in contrast, the equal-space commutators

[Λ2(x3, ct), D1(x3, ct ′)] = 0, (3.119)

[Λ2(x3, ct), B2(x3, ct ′)] = i�cδ(ct − ct ′)1, (3.120)

[B2(x3, ct), B2(x3, ct ′)] = i�cδ′(ct − ct ′)1, (3.121)

[D1(x3, ct), B2(x3, ct ′)] = 0, (3.122)

[D1(x3, ct), D1(x3, ct ′)] = i�cεδ′(ct − ct ′)1, (3.123)

Page 113: Quantum Aspects of Light Propagation

104 3 Macroscopic Theories and Their Applications

we obtain peculiar equation of the spatial progression

∂Λ2

∂x3= i

[Λ2,

∫(D1 B2)S dt

]= −D1, (3.124)

∂ D1

∂x3= i

[D1,

∫(D1 B2)S dt

]= −∂ B2

∂t. (3.125)

Relation (3.124) expresses the role of the dual vector potential and relation (3.125) isessentially the second of the Maxwell evolution equations. We could obtain the firstof them as the equation of the spatial progression for the quantum field B2. Sincewe must often guess the commutators never known before, the above example is awarning against excessive trust in the spatial progression technique.

For a medium with nonlinear polarization, the global nature of creation andannihilation operators is lost. By consistently following this idea, Serulnik andBen-Aryeh (1991) have introduced the shift operators which, by their definition,are based on the energy–momentum tensor. They have followed in their treatmentPeierls’ solution of the problem of momentum conservation in matter (Peierls 1976,1985) by which the atoms or the bulk matter is considered to be at rest while theelectromagnetic field is propagating. They show that it is always possible to relatethe external field in front of the medium to that behind it by the use of the shift oper-ators, that is by a Lie transformation. As we can see from relations (3.114), (3.115),(3.124), and (3.125), there exists no space–time description which in addition to theso-called time displacement operators suggests the use of their space-displacementanalogues.

Leonhardt (2000) has determined an energy–momentum tensor of the electro-magnetic fields in quantum dielectrics. The tensor is Abraham’s plus the energy–momentum of the medium characterized by a dielectric pressure and enthalpydensity (Abraham 1909). While the consistency of this picture with the theory ofdielectrics has been demonstrated, a direct derivation from the first principles hasbeen announced only.

The theory of the radiation pressure on dielectric surfaces (Loudon 2002) acceptsthe expression for the momentum density of an electromagnetic wave in a transpar-ent material medium due to Abraham (1909). The force of the radiation pressure isobtained similarly using the Lorentz force density operator as using the momentumdensity according to Abraham (1909, 1910). A colloquium has been devoted to themomentum of an electromagnetic wave in dielectric media (Pfeifer et al. 2007).

3.1.4 Generator of Spatial Progression

Theoretical methods for treating propagation in quantum optics have been devel-oped in which the momentum operator is used in addition to the Hamiltonian. Asuccessful quantum-mechanical analysis has been given for various physical sys-tems which include amplification and coupling between electromagnetic modes

Page 114: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 105

(Toren and Ben-Aryeh 1994). Distributed feedback lasers have been described, butthe overarching generalization of both successful analyses has not been developed.The authors have drawn attention to the distributed feedback lasers (Yariv and Yeh1984, Yariv 1989), in which contradirectional beams are amplified by an activemedium and are coupled by a small periodic perturbation of a refractive index.

The energy and momentum properties of the electromagnetic field can be des-cribed, in a four-dimensional form, by the energy–momentum tensor T jk , wherej,k = 0,1,2,3 (Roman 1969),

T jk =

⎢⎢⎣

H gx gy gz

Sx σxx σxy σxz

Sy σyx σyy σyz

Sz σzx σzy σzz

⎥⎥⎦ . (3.126)

The tensor element T 00 represents the energy density. The density of the vectorialmomentum (proportional to D × B) is represented by the vector (gx ,gy,gz). Letus take further, for example, line 3. The tensor element T 30 is the component ofthe Poynting vector standing for the flux of energy in the z-direction. The vector(σzx ,σzy,σzz) refers to a flux of momentum in the propagation direction of z.

In the conventional approach (Roman 1969), the four-vector p0k is defined as

p0k =∫ ∫ ∫

T 0k(x, y, z, t) dx dy dz. (3.127)

The energy p00 is used as the Hamiltonian for the description of time evolution,but the momentum component p03 rather merely translates in the z-direction. Ben-Aryeh and Serulnik (1991) have shown that for the description of the spatial pro-gression, the four-vector p3k

p3k =∫ ∫ ∫

T 3k(x, y, z, t)c dt dx dy (3.128)

can be used. The momentum component in the z-direction p33 can be used as thegenerator of the spatial progression, but the energy p30 rather merely causes trans-lation of the field in time.

In Toren and Ben-Aryeh (1994), problems of propagation are treated by expand-ing the field operators in terms of mode operators associated with definite frequen-cies. Starting with Caves and Crouch (1987), the approach has been associated withthe conservation of commutation relations for creation and annihilation operators,which are space dependent (cf. Huttner et al. 1990). Imoto (1989) has developedthe basic equation of motion by using a modified procedure of canonical quantiza-tion in which time and space coordinates are interchanged in comparison with theconventional procedure. Ben-Aryeh et al. (1992) did the same by using a slightlydifferent notation. Toren and Ben-Aryeh (1994) dissociate themselves from suchan approach, but they are not very explicit about the point that the use of the

Page 115: Quantum Aspects of Light Propagation

106 3 Macroscopic Theories and Their Applications

integrals (3.128) is compatible with the canonical quantization in which the timecoordinate plays the usual role.

Linear amplification is treated by the use of momentum for space-dependentamplification. Travelling-wave attenuators and amplifiers can be treated as continu-ous limits of an array of beam splitters (Jeffers et al. 1993, Ban 1994). According toToren and Ben-Aryeh (1994), the propagating modes are coupled to a momentumreservoir. The Hamiltonian of this system is given by the relation

H = �

(ωa†a −

j

ω jresb†j b j

)(3.129)

and the total momentum operator is

G = �

[βa†a −

j

β jresb†j b j +

j

(κ j a†b†

j + κ∗j ab j )], (3.130)

where the subscript res stands for the reservoir and κ j are appropriate couplingconstants, a and b j represent (in the zeroth-order) modes which are propagating inthe positive direction of the z-axis.

The equations of motion obtained from the momentum operator (3.130) are

da

dz= i

�[a, G] = iβa + i

j

κ j b†j , (3.131)

db†j

dz= i

�[b†

j , G] = −iκ∗j a + iβ jresb†j . (3.132)

By using the spatial Wigner–Weisskopf approximation, the Heisenberg–Langevinequations can be obtained

da

dz=[

i(β −Δβ)+ 1

]a + L†, (3.133)

where

Δβ = −V.p.∫ ∞

−∞

|κ(βres)|2ρ(βres)

βres − βdβres, (3.134)

γ = {2π |κ(βres)|2 ρ(βres)} |βres=β, (3.135)

with ρ(βres) the density function of the reservoir wave propagation constants β jres,and

L† =∑

j

iκ j b†j e

iβ jresz . (3.136)

Page 116: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 107

The codirectional coupling is analysed. It is assumed that two modes are propa-gating in the same direction and they are coupled by a periodic change in the refrac-tive index. For a classical description, we refer to Yariv and Yeh (1984) and Yariv(1989). The Hamiltonian is given by

H0 = H = �ω[a†1a1 + a†

2a2], (3.137)

where the classical relation ω1 = ω2 = ω has been used. The total momentum oper-ator is

G = �[β1a†1a1 + β2a†

2a2 + κ a1a†2 + κ∗a†

1a2], (3.138)

where β1 and β2 are components of the wave vectors of the two modes in the prop-agation direction of z and

κ = κ exp

(− im2π

Λz

), (3.139)

with κ a coupling constant, m an integer, and Λ the “wavelength” of the spatialperiodic change in the index of refraction (a perturbation in the dielectric constant).In this connection papers Perinova et al. (1991) and Ben-Aryeh et al. (1992) arecriticized that they do not take account of the spatial dependence (3.139).

The equations of motion obtained from the momentum operator (3.138) are

da1

dz= i

�[a1, G] = iβ1a1 + iκ∗a2, (3.140)

da2

dz= i

�[a2, G] = iκ a1 + iβ2a2. (3.141)

We define slowly varying operators of the form

A1(z) ≡ a1(z)e−iβ1z , A2(z) ≡ a2(z)e−iβ2z . (3.142)

Substituting the operators (3.142) into equations (3.140), (3.141), we get

d A1

dz= iκ∗ A2e−iΔβ z, (3.143)

d A2

dz= iκ A1eiΔβ z, (3.144)

where

Δβ ≡ β1 − β2 − m2π

Λ(3.145)

Page 117: Quantum Aspects of Light Propagation

108 3 Macroscopic Theories and Their Applications

is the mismatch. A “field” mismatch may be cancelled by a medium component.For the input–output relations we refer to Yariv and Yeh (1984), Yariv (1989), andPerinova et al. (1991).

In Perinova et al. (1991), m = 0 and 2δ = −Δβ has been introduced in anapplication to the codirectional coupler. In general, the solution to equations (3.143),(3.144) coincides with the classical solution, A j ↔ A∗j , j = 1, 2, where A1, A2 arethe amplitudes of the waves propagating in the +z-direction.

The counterdirectional coupling is analysed in Toren and Ben-Aryeh (1994). Thetotal Hamiltonian is given by

H0 ≡ H = �ω[a†1a1 − a†

2a2]. (3.146)

The momentum operator is

G = �[β1a†1a1 − β2a2a†

2 + κ a1a†2 + κ∗a†

1a2]. (3.147)

It is reasonable that the Hamiltonian and the zeroth-order operator are related,respectively, to the flux of energy and that of the component of momentum in thez-direction. Compared to Toren and Ben-Aryeh (1994), the operators a2 and a†

2 have

been exchanged. They criticize our assumption [a2, a†2]

?= −1, which we obtainedin Perinova et al. (1991) from the usual equal-space commutator [a2, a†

2]= 1 by thisinterchange. It is tempting to have the same alternation between the opposite-goingmodes as can be seen in comparison of (3.524) with (3.526) (cf. Abram and Cohen1994). The equations of motion obtained from operator (3.147) are given by

da1

dz= i

�[a1, G

∣∣a2↔a†

2] = iβ1a1 + iκ∗a2, (3.148)

da2

dz= i

�[a2, G

∣∣a2↔a†

2] = −iκ a1 + iβ2a2. (3.149)

Using the slowly varying operators (3.142), in contrast with Toren and Ben-Aryeh(1994), we obtain that

d A1

dz= iκ∗ A2e−iΔβ z ,

d A2

dz= −iκ A1eiΔβ z, (3.150)

where Δβ has been defined by equation (3.145).In Perinova et al. (1991) still 2δ = −Δβ holds in an application to the counter-

directional coupler. The solution to equations (3.150) coincides with the classicalsolution, A j ↔ A∗j , j = 1, 2, where A1, A2 are the amplitudes of the waves prop-agating in the +z- and −z-directions. In Yariv and Yeh (1984), the solution to thecorresponding classical equations has been obtained for the boundary conditions

Page 118: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 109

A1(z)|z=0 = A1(0), A2(z)|z=L = A2(L). First, however, one obtains the solution forthe usual condition at z = 0.

In Perinova et al. (1991), the output operators have been obtained in terms ofthe input ones. While we simply determine the operators A1(L), A2(0) from twoequations for the operators A1(0), A2(0), A1(L), A2(L), we observe that in this

procedure the equal-space commutator [ A2, A†2]

?= −1 must depend on both z andL in a complicated manner, and simplifies to [ A2, A†

2] = 1. Since the commutatorscorrespond to the Poisson brackets, much is illustrated by the appropriate classicaltheory (Luis and Perina 1996b). One must be aware of the fact that in formulatingthe theory, Luis and Perina (1996b) have avoided the above considerations on thez coordinate and the generator of spatial progression and they used in the bulk oftheir paper the usual time dependence and the Hamiltonian function. Although stillobscure in the case of commutators, the situation is clear in the classical case, whenthe input–output transformation is characterized by the usual Poisson brackets andthe solution for the usual boundary conditions at z = 0 requires the noncanonicaltransformation α2 ↔ α∗2 , with the complex amplitude α2. The richness of theirtheory is due to nonlinearities, whereas it is shown that in the quantum case only apoor linear theory is possible.

The difficulty lies in the formulation of an appropriate dynamical operator.Tarasov (2001) has defined a map of a dynamical nonlinear operator into a dynam-ical superoperator. He had in mind quantum dynamics of non-Hamiltonian and dis-sipative systems.

A quantum-mechanical treatment of distributed feedback laser using the momen-tum operator in addition to the Hamiltonian is developed in Toren and Ben-Aryeh(1994). The authors start from the classical description based on two coupledequations

dA1

dz= 1

2γ A1 − iκ A2eiΔβ z,

dA2

dz= iκ∗A1e−iΔβ z − 1

2γ A2, (3.151)

where A1, A2 are the amplitudes of the waves propagating in the +z- and −z-directions, respectively, κ is the coupling constant, γ the amplification constant,and Δβ is given by (3.145), with β1 = β, β2 = −β. The solution of the classicalequations is well known (Yariv and Yeh 1984, Yariv 1989) and it shows that on acondition the amplification becomes extremely large. The classical theory does notinclude the quantum noise which follows from the amplification process. Unfortu-nately, Toren and Ben-Aryeh (1994) did not develop the overarching generalizationof the analysis of amplification and that of the counterdirectional coupling. To thebest of our knowledge, such a quantum-mechanical theory is not in hand.

The treatment of parametric down-conversion and parametric up-conversion byDechoum et al. (2000) is interesting with its use of the Wigner representation ofoptical fields, but it starts just from the Maxwell equations for the field opera-tors for the lossless neutral nonlinear dielectric medium. With the use of common

Page 119: Quantum Aspects of Light Propagation

110 3 Macroscopic Theories and Their Applications

approximation of treating the laser pump as classical, classical equations of nonlin-ear optics are obtained.

3.1.5 Nonlinear Optical Couplers

Optical couplers employing evanescent waves play an important role in optics, opto-electronics, and photonics where they may be conveniently used in the switching oflight beams. They also provide a means for controlling light by light. Amplitudeand intensity behaviour of linear couplers has been investigated extensively (Yarivand Yeh 1984, Solimeno et al. 1986, Saleh and Teich 1991). Substantial progressin controlling light beams has been achieved after nonlinear waveguides with bothlinear and nonlinear coupling have been taken into account (Finlayson et al. 1988,Townsend et al. 1989, Leutheuser et al. 1990, Weinert-Raczka and Lederer 1993,Assanto et al. 1994, Hatami-Hanza and Chu 1995, Hansen 1995, Weinert-Raczka1996).

This gave new possibilities of fast all-optical switching, including digital switch-ing, and reduction of switching power. New ways of controlling optical beams innonlinear couplers have been invented.

Nonlinear waveguide materials used in composing nonlinear couplers providenew possibilities in constructing switching and memory elements for all-opticaldevices. These elements are necessary for further development of optical process-ing and computing. Classically, all-optical devices are analysed from the viewpointof their amplitude or intensity dependences. However, they can be treated fully inquantum theory. Noise of light beams in nonlinear couplers is naturally included inthis quantum treatment.

Perina, Jr., and Perina (2000) in Section 2 indicate a consequential use of themomentum operator we have mentioned in Section 3.1.4. In nonlinear quantumsystems where both directions of propagation are present this formalism confrontsdifficulties. The generator of spatial progression is related to commutators whichdo not lead to proper input (and output) commutators. The two ways of introducingphoton annihilation and creation operators obeying boson commutation relations arenot consistent mutually. Working with operators is not secure.

The authors work with the two algebras transparently. The commutators whichare preserved in spatial progression are exploited only for the derivation of evolutionequations for operators. The proper input commutators are used to the other goals.

In order to determine quantum-statistical properties of light beams, solutions ofnonlinear operator equations have to be found first. One can apply a short-lengthapproximation. Or pump modes can be assumed to be in strong coherent states anda parametric approximation can be used.

The short-length approximation is used as a tool in the treatment of nonlinearquantum systems. Calculations with operators are safe when one direction of prop-agation is present, because the two algebras coincide. If both directions of propaga-tion are present it is easy to use one of the two algebras in principle (cf. Perinovaet al. 1991). Paradoxes may occur.

Page 120: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 111

For instance, the operator equations of spatial evolution differ from the Heisen-berg equations in interaction picture only by one or a few changes of sign. But thesechanges entail, in the formalism of input commutators, that operator products on theright-hand sides of the equations may have an incidental order. Elas, this problem isnot present in the formalism of the generator of spatial progression. We suppose itshould be better to consider boundary-value problems for equations for c-numbersin the systems where both directions of propagation underlie to description and toquantize at the end.

The parametric approximation is used as another instrument. It leads to linearevolution equations of operators. Also here a description of a quantum system isspecific in which two directions of propagation are present. But linear equationscomprise no products on the right-hand sides.

A number of quantum descriptions are related to two modes and can be specifiedas two linear equations (and their conjugates) for the operators Aa , Ab, A†

a , A†b. The

right-hand sides of these equations are often independent of time. We may let λ1,2,3,4

denote the eigenvalues of the matrix of the right-hand sides. Introducing operatorsAc, Ad as appropriate Bogoliubov transforms of the operators Aa and Ab, one canexpress any quantum description (a regularity is assumed) in one of the six normalforms (Williamson 1936)

d Ac

dt= a A†

c,d Ad

dt= bA†

d (3.152)

for λ1,2,3,4 = ±a,±b;

d Ac

dt= a A†

c + bAd ,d Ad

dt= −bAc + a A†

d (3.153)

for λ1,2,3,4 = ±a ± ib;

d Ac

dt= −iρa Ac,

d Ad

dt= −iσbAd , ρ, σ = ±1, (3.154)

for λ1,2,3,4 = ±ia,±ib;

d Ac

dt= a A†

c,d Ad

dt= −iρbAd , ρ = ±1, (3.155)

for λ1,2,3,4 = ±a,±ib;

d Ac

dt= a A†

c +1

2( A†

d − Ad ),d Ad

dt= 1

2( A†

c + Ac)+ a A†d (3.156)

for λ1,2,3,4 = ±a,±a;

d Ac

dt= i

ρ

2( A†

c − Ac)− a Ad ,d Ad

dt= a Ac + i

ρ

2( A†

d − Ad ), ρ = ±1 (3.157)

Page 121: Quantum Aspects of Light Propagation

112 3 Macroscopic Theories and Their Applications

for λ1,2,3,4 = ±ia,±ia. The formulae for Ac and Ad are complicated and they arenot presented here.

Perina and Perina, Jr. (1995a) have studied a codirectional coupler composedof one linear and one nonlinear waveguide. Second-subharmonic mode (b1) andpump mode (b2) nonlinearly interact in the waveguide b. The second-subharmonicmode b1 also interacts linearly with mode a in the waveguide a. The correspondingmomentum operator in interaction pictures is written in the form

G int = �

[−κab1 Aa A†

b1− Γb A2

b1A†

b2exp(iΔkbz)+ H.c.

], (3.158)

where κab1 denotes the linear coupling constant of modes a and b1 and Γb is the non-linear coupling constant between modes b1 and b2. The nonlinear phase mismatchΔkb is defined as Δkb = 2kb1 − kb2 and kb1 (kb2 ) means the wave vector of mode b1

(b2). In (3.158), Aa , Ab1 , and Ab2 stand for optical-field operators of modes a, b1,and b2 in interaction pictures. The conservation law

A†a(z) Aa(z)+ A†

b1(z) Ab1 (z)+ 2 A†

b2(z) Ab2 (z) = const. (3.159)

is fulfilled by the solution of Heisenberg equations in the interaction picture.Perina (1995a,b) and Perina and Bajer (1995) have analysed squeezing of the

light in a short-length approximation. The assumption of a strong coherent field inmode b2 with the amplitude ξb2 leads to the linearization of the operator equationsof motion. The analysis leads to at least one positive eigenvalue. Amplification mayoccur dependent on the initial conditions. In the case where |Γbξb| < |κab1 |, oscil-lations also occur in the spatial development of quantities characterizing the fields.Results on squeezing of the light have been obtained (Perina and Perina, Jr. 1995a).

The assumption of a strong coherent field in mode b2 and the linearization arerelated here to three types of behaviour, which can be distinguished also using nor-mal forms (3.152), (3.153), and (3.156).

Perina and Perina, Jr. (1995b) have treated a contradirectional coupler composedagain of one linear and one nonlinear waveguide. Mode a propagates against modesb1 and b2. The appropriate conservation law reads as

− A†a(z) Aa(z)+ A†

b1(z) Ab1 (z)+ 2 A†

b2(z) Ab2 (z) = const, z = 0, L; (3.160)

i.e.

A†a(0) Aa(0)+ A†

b1(L) Ab1 (L)+ 2 A†

b2(L) Ab2 (L)

= A†a(L) Aa(L)+ A†

b1(0) Ab1 (0)+ 2 A†

b2(0) Ab2 (0). (3.161)

A phase matching (Δkb = 0) is assumed.A formulation of short-length approximation seems to be obvious, but it uses

equal-space products of field operators. We can return to the boundary-value prob-lem for classical equations which have the same solutions in the case of

Page 122: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 113

contradirectional propagation as the initial-value problem for codirectional couplerup to the second order in L provided we write Aa(L) in place of Aa(0). Quantizationup to the second order in L can be done in the case of the contradirectional propaga-tion by using the quantum input–output relations of the codirectional coupler whereAa(L) is written in place of Aa(0).

The assumption of a strong coherent field in mode b2 leads to linear operatorequations of motion as in the case of the codirectional coupler. Introducing sa = ±1,sa = 1 when mode a propagates along with modes b1 and b2 and sa = −1 when itpropagates counter to the latter, we can write the eigenvalues as

λ1,2,3,4 = ±(|Γbξb2 | ±

√|Γbξb2 |2 − sa|κab1 |

). (3.162)

In the case of the contradirectional coupler oscillations cannot occur. Results onsqueezing of the light have been obtained (Perina and Perina, Jr. 1995b,c).

Let us note that one cannot assess input–output relations so easily in this caseusing only the eigenvalues. In the case of the codirectional coupler the input–outputrelations are just the solutions of the initial-value problem and their dependence onexp(λ1,2,3,4 L) is linear. In the case of the contradirectional coupler the input–outputrelations depend on exp(λ1,2,3,4 L) in a nonlinear way.

Perina and Bajer (1995) have studied also a codirectional coupler with fourmodes. A mode of frequency ω1 (b1) and a mode of frequency ω2 (b2) nonlinearlyinteract in the waveguide b. The pump mode b1 of frequency ω1 interacts linearlywith mode a1 and the mode b2 of frequency ω2 is coupled linearly with mode c(ω2 = 2ω1 holds). The momentum operator (3.158) is modified to the form

G int = �

(−κab1 Aa A†

b1− κcb2 Ac A†

b2+ Γb A2

b1A†

b2+ H.c.

), (3.163)

where κcb2 is the linear coupling constant of modes c and b2, κab1 and Γb have theoriginal meaning. The conservation law

A†a(z) Aa(z)+ A†

b1(z) Ab1 (z)+ 2 A†

b2(z) Ab2 (z)+ 2 A†

c(z) Ac(z) = const. (3.164)

is obeyed by the solutions of Heisenberg equations in the interaction picture. Perinaand Bajer (1995) have treated squeezing of the light in the short-length approxima-tion also in this case.

Mista, Jr. and Perina (1997) have investigated a codirectional coupler with fivemodes. Second-subharmonic modes (a1, c1) and pump modes (a2, c2) nonlinearlyinteract in the respective parts (a, c). The second-subharmonic mode a1 also inter-acts linearly with mode c1 via mode b in the part b. On assuming linear and non-linear phase matching, Perina, Jr. and Perina (2000) describe the coupler with thefollowing momentum operator:

G int = �(Γa A2

a1A†

a2+ Γc A2

c1A†

c2+ κa1b A†

a1Ab + κbc1 A†

c1Ab + H.c.

). (3.165)

Page 123: Quantum Aspects of Light Propagation

114 3 Macroscopic Theories and Their Applications

The solutions of Heisenberg equations in the interaction picture obey the conserva-tion law

A†a1

(z) Aa1 (z)+ 2 A†a2

(z) Aa2 (z)

+ A†b(z) Ab(z)+ A†

c1(z) Ac1 (z)+ 2 A†

c2(z) Ac2 (z) = const. (3.166)

In a short-length approximation results on squeezing of the light have been obtained.Perina and Perina, Jr. (1996) have studied a codirectional coupler composed of

two nonlinear waveguides. While they have used a parametric approximation fromthe very beginning (Perina, Jr. and Perina 2000), here we present a generalization ofthe momentum operator (3.158)

G int = �[Γa A2

a1A†

a2exp(iΔkaz)+ Γb A2

b1A†

b2exp(iΔkbz)

+ κab Aa1 A†b1+ H.c.

], (3.167)

where a1 ≡ a, κab ≡ κab1 , and Γa is the nonlinear coupling constant between modesa1 and a2. The nonlinear phase mismatch Δka is defined as Δka = 2ka1 − ka2 andka1 (ka2 ) means the wave vector of mode a1 (a2). The parametric approximationhas consisted in replacements Aa2 → ξa2 exp(iΔlz), Ab2 → ξb2 exp(−iΔlz), whereΔl = 1

2 (kb2 − ka2 ), on assuming also some linear coupling between modes a2 andb2. Korolkova and Perina (1997a) have obtained and discussed solutions on somesimplifying assumptions. Karpati et al. (2000) studied all-optical switching in thissystem. Perina and Perina, Jr. (1996) and Korolkova and Perina (1997a) have con-sidered a contradirectional coupler as well.

Janszky et al. (1995) were first to investigate a coupler composed of two nonlin-ear waveguides with nondegenerate optical parametric processes. Pump (aP), signal(aS), and idler (aI) modes in one waveguide are assumed to interact linearly withtheir counterparts (bP, bS, bI) in the other waveguide. The coupler is described bythe following momentum operator

G = �

{ ∑

i=aP,aS,aI,bP,bS,bI

ki a†i ai + [ΓaaaP a†

aSa†

aI+ ΓbabP a†

bSa†

bI+ H.c.]

+ [κPaaP a†bP+ κSaaS a†

bS+ κIaaI a

†bI+ H.c.]

}, (3.168)

where ki denotes the wave vector of the i th mode along z-axis, Γa (Γb) is the non-linear coupling constant of modes aP, aS, aI (bP, bS, bI) in waveguide a (b), and κP,κS, and κI stand for the linear coupling constants between the two pump, the twosignal, and the two idler modes.

Herec (1999) has used a short-length approximation to solve the Heisenbergequations in the interaction picture and he has obtained results on squeezing of thelight. Mista, Jr. (1999) has assumed strong coherent field in pump modes, κP = 0,and phase matching, discerned three (of five) regimes in spatial evolution of thecoupler, and obtained various results on the squeezing.

Page 124: Quantum Aspects of Light Propagation

3.1 Momentum-Operator Approach 115

Optical fibres and certain organic polymers with high third-order nonlinearitiesmay be used for the construction of couplers based on the Kerr effect. The nonlineardirectional couplers are interesting as they exchange energy periodically betweenthe guides like linear ones for low total intensities while they trap the energy in theguide into which it has been launched initially for high intensities (Jensen 1982).A coupler with two nonlinear waveguides (denoted as a and b) with Kerr nonlin-earities has been considered by Chefles and Barnett (1996). It is described by themomentum operator G int (ka = kb is assumed) (Korolkova and Perina 1997b),

G int = �

[g A†2

a A2a + g A†2

b A2b + gab A†

a Aa A†b Ab

+(κab Aa A†

b + H.c.) ]

. (3.169)

Here g means a Kerr nonlinear coupling coefficient which is the same in both waveg-uides. The real constant gab describes nonlinear coupling of the modes and the realconstant κab characterizes linear coupling of the modes. The operator A†

a Aa+ A†b Ab

is a constant of motion.Korolkova and Perina (1997b) have assumed that Aa and Ab are fast-oscillating

operators due to the linear coupling and they have introduced the operators

Ba = 1√2

[ Aa exp(−iκabz)+ Ab exp(iκabz)],

Bb = 1√2

[ Aa exp(−iκabz)− Ab exp(iκabz)]. (3.170)

On an approximation a solution has been obtained. Then the operators B†a Ba and

B†b Bb are conserved along z. We note that the solution is exact for gab = 2g.

In the interaction picture a numerical solution of the Schrodinger equation mayuse invariant subspaces defined as eigenspaces of the constant of motion (Cheflesand Barnett 1996) and it is exact for any initial state from a finite direct sum of suchsubspaces. This way, Fiurasek et al. (1999a,b) have obtained interesting results onassuming also initial coherent states.

The behaviour of the coupler may exhibit a bifurcation in dependence on theparameter

η = 1

2κab|2g − gab| (|Aa|2 + |Ab|2) (3.171)

in the classical regime. The threshold is at η = 1. The behaviour is more compli-cated in the quantum regime. An optimum energy exchange between modes a andb occurs if the difference of the phases of the complex amplitudes of modes a andb equals π

2 or −π2 .

The character of the evolution of mean photon numbers in the regions of revivalscan be controlled by the z-dependent linear coupling constant κab(z) (Korolkova

Page 125: Quantum Aspects of Light Propagation

116 3 Macroscopic Theories and Their Applications

and Perina 1997c). Switching of energy between the waveguides is achieved bya suitable profile of the coupling functions. Squeezing in a given waveguide alsois preserved in such a way. In the classical regime, a nonlinear optical switchingmatrix has been considered (Liu et al. 2003).

Perina, Jr. and Perina (1997) have paid attention to the couplers which are basedon Raman and Brillouin scattering. Fiurasek and Perina (1998, 1999, 2000a,b)have continued this work. A codirectional coupler composed of two waveguidesis described with the momentum operator G:

G = �

j=a,b

[ ∑

l=L,S,A,V

k jl a†jl a jl +

(g jAa jLa jVa†

jA + g jSa jLa†jVa†

jS + H.c.) ]

+ �

[κSaaS a†

bS+ κAaaA a†

bA+ H.c.

], (3.172)

where k jl are wave vectors of modes l, l = L (laser), S (Stokes), A (anti-Stokes),and V (phonon) in waveguides j , j = a, b. Here g jS (g jA) describes the Stokes(anti-Stokes) nonlinear coupling in waveguide j , κS (κA) is the linear couplingconstant between the Stokes (anti-Stokes) modes in different waveguides. Vectorscharacterizing phase mismatches are defined as follows: Δk jS = k jL − k jV − k jS ,Δk jA = k jL + k jV − k jA , ΔkS = kaS − kbS , and ΔkA = kaA − kbA . A parametricapproximation consists in assuming strong coherent states in pump modes aL andbL.

Fiurasek and Perina (1999) have used another approximation in solving Heisen-berg equations in the interaction picture. The method utilized is based on linearoperator corrections to a classical solution.

Fiurasek and Perina (2000a) have considered a Raman coupler with broadphonon spectra. They have described the phonon systems of the waveguides withmultimode boson fields. Then these fields have been eliminated from the descriptionof the coupler using the Wigner–Weisskopf approximation (Perina 1981a,b).

Mogilevtsev et al. (1997) have treated one central waveguide (a) which interactslinearly with a greater number of mutually noninteracting waveguides (b j ) in itssurroundings. It is described by the following momentum operator G

G = �

⎣kaa†aaa +

N∑

j=1

kb j a†b j

ab j +N∑

j=1

gab j (ab j a†a + a†

b jaa)

⎦ , (3.173)

where ka (kb j ) is the wave vector of mode a (b j ), gab j is the linear coupling constantbetween modes a and b j , and N denotes the number of surrounding waveguides. Ifthe central waveguide a contains a second-order nonlinear medium, the momentumoperator Gn is

Gn = G + �

2[ξa2 exp(2ikaz)a†2

a + H.c.], (3.174)

Page 126: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 117

where ξa2 stands for the amplitude of the pump field in the central waveguide.The behaviour of the linear and nonlinear couplers agrees with the idea that thesurrounding waveguides form a reservoir. A reservoir spectrum has been consideredwhich has a gap.

Mogilevtsev et al. (1996) have considered a coupler composed of one waveguidewith χ (2) medium and the other one with χ (3) medium. Only a nonlinear coupling ispresent. The interaction momentum operator G int is

G int = �

[ga( A†2

a + A2a)+ gb A†2

b A2b + gab A†

a Aa A†b Ab

], (3.175)

where ga describes the process of second-subharmonic generation and includes thecoherent pump amplitude, gb stands for the Kerr constant, and gab means the non-linear coupling constant between modes a and b.

Assuming the vacuum state in mode a and an incident pure state in mode b, onecan distinguish two regimes essentially. The third type of behaviour could resultusing the superposition principle. When an initial Fock state with Nb photons inmode b is assumed, the solution for the operator Ab(z) oscillates in z for 2ga <

Nbgab, whereas it has an exponential character for 2ga > Nbgab. When an initialcoherent state with the amplitude ξb in mode b is assumed, the mean number ofphotons in mode a oscillates in z and the exponential terms can be neglected for2ga � |ξb|2gab. It increases exponentially in z and the oscillating terms do notcontribute significantly for 2ga � |ξb|2gab.

3.2 Dispersive Nonlinear Dielectric

The spatio-temporal quantum description has been adopted in optics in spite of itscomplexity due to quantum solitons. As known, the existence of the optical fieldsthat do not change during propagation is conditioned by the frequency dispersionand the nonlinearity of the medium. A macroscopic quantization was to take intoaccount both the properties. The nonlinearity would appear as an interaction ofnarrow-band fields.

3.2.1 Lagrangian of Narrow-Band Fields

Drummond (1990) has presented a technique of canonical quantization in a gen-eral dispersive nonlinear dielectric medium. Contrary to Abram and Cohen (1991),Drummond creates an arbitrary number of slightly varied copies of the vacuumelectromagnetic field for the nonlinear dielectric medium, essentially the numberrequired by the classical slowly varying amplitude approximation. But Abram andCohen (1991) work with a single field. The paradox of the validity of both theapproaches can be resolved only by a detailed microscopic theory.

Drummond (1990) generalizes the treatment of a linear homogeneous disper-sive medium (Schubert and Wilhelmi 1986). Till 1990, papers by Knoll (1987),Białynicka-Birula and Białynicki-Birula (1987), and Glauber and Lewenstein (1989)

Page 127: Quantum Aspects of Light Propagation

118 3 Macroscopic Theories and Their Applications

could be referred to as devoted to the theory of inhomogeneous nondispersive lineardielectric. Hillery and Mlodinow (1984) were attractive with their use of the idea dueto Born and Infeld (1934) for the quantization of homogeneous nonlinear nondis-persive medium.

The macroscopic quantization is a route to the simplest quantum theory compati-ble with known dielectric properties unlike the microscopic derivation of the nonlin-ear quantum theory of electromagnetic propagation in a real dielectric. Drummond(1994) compares the quantum theory obtained via macroscopic quantization withthe traditional quantum-field theory. He concedes that most model quantum fieldtheories prove to be either tractable, but unphysical, or physical, but intractable. Thetractable model quantum field theory ceases to be unphysical when it is tested exper-imentally in quantum optics. An excellent example of this is the fibre optical soli-tons whose quantization is given in detail in Drummond (1994). In agreement withtheoretical predictions (Carter et al. 1987, Drummond and Carter 1987, Drummondet al. 1989, Shelby et al. 1990, Lai and Haus 1989, Haus and Lai 1990), experiments(Rosenbluh and Shelby 1991) led to evidence of quantum solitons. More recentexperiments (Friberg et al. 1992) demonstrate that solitons can be considered to benonlinear bound states of a quantum field. In addition to the quadrature squeezing in(Rosenbluh and Shelby 1991), quantum properties of soliton collisions were mea-sured (Watanabe et al. 1989, Haus et al. 1989). Similar nonlinearities are encoun-tered in photonic band-gap theory (Yablonovitch and Gmitter 1987), microcavityquantum electrodynamics (Hinds 1990), pulsed squeezing (Slusher et al. 1987), andquantum chaos (Toda et al. 1989).

In description of a nonlinear dielectric medium, tensorial notation is used whichwill occur also elsewhere in this book. Let u, v, w, . . . be vectors. On using atensorial product, a tensor of rank 2, e.g. uv, is formed, a tensor of rank 3, e.g.uvw, is constructed, etc. The scalar product denoted by the dot · is generalized to acontraction, i.e. a simple sum after the tensors are replaced by their components andproducts of corresponding components are formed. The correspondence of compo-nents is achieved by using the same notation for the last subscript of the tensor to theleft as for the first subscript of the tensor to the right. Also the pieces of notation : and... mean contractions, i.e. a double sum and a triple sum of products of correspondingcomponents.

It is advantageous to begin with the treatment of a classical dielectric introduc-ing the nonlinear response function in terms of the electric displacement field D.Contrary to the usual description (Bloembergen 1965), which uses the dielectricpermittivity tensors, the inverse expansion is necessary here. For simplicity, thedielectric of interest is regarded as having uniform linear magnetic susceptibility.The charges are assumed to occur only in the induced dipoles of polarization. Thefield equations are therefore

∇ × E(x, t) = −∂B(x, t)

∂t,

∇ ×H(x, t) = ∂D(x, t)

∂t,

Page 128: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 119

∇ · D(x, t) = 0,

∇ · B(x, t) = 0, (3.176)

where

D(x, t) = ε0E(x, t)+ P(x, t),

B(x, t) = μH(x, t). (3.177)

Here

P(x, t) =∫ ∞

0χ (x, τ ) · E(x, t − τ ) dτ

+∫ ∞

0

∫ ∞

0χ (2)(x, τ1, τ2) : E(x, t − τ1)E(x, t − τ2) dτ1 dτ2

+∫ ∞

0

∫ ∞

0

∫ ∞

0χ (3)(x, τ1, τ2, τ3)

...E(x, t − τ1)E(x, t − τ2)

× E(x, t − τ3) dτ1 dτ2 dτ3

+ · · · , (3.178)

where the tensor of rank 2, in general, the (n + 1)th-rank susceptibility tensor read

χ (x, τ ) = 1

∫χ (x, ω)e−iωτ dω,

χ (n)(x, τ1, ..., τn) =(

1

)n

×∫

...

∫χ (n)(x, ω1, ..., ωn)e−i(ω1τ1+···+ωnτn ) dτ1... dτn,

(3.179)

respectively. After adding the vacuum electric displacement field ε0E(x, t) to bothsides of (3.178), we express the electric vector in the form

E(x, t) =∫ ∞

0ζ (x, τ ) · D(x, t − τ ) dτ

+∫ ∞

0

∫ ∞

0ζ (2)(x, τ1, τ2) : D(x, t − τ1)D(x, t − τ2) dτ1 dτ2

+∫ ∞

0

∫ ∞

0

∫ ∞

0ζ (3)(x, τ1, τ2, τ3)

...D(x, t − τ1)D(x, t − τ2)

× D(x, t − τ3) dτ1 dτ2 dτ3

+ · · · , (3.180)

Page 129: Quantum Aspects of Light Propagation

120 3 Macroscopic Theories and Their Applications

where

ζ (x, τ ) = 1

∫ζ (x, ω)e−iωτ dω,

ζ (n)(x, τ1, ..., τn) =(

1

)n

×∫

...

∫ζ

(n)(x, ω1, ..., ωn)e−i(ω1τ1+...+ωnτn ) dτ1... dτn (3.181)

and the tensors on the right-hand sides of (3.181) can be expressed from the equa-tions

ε(x, ω) · ζ (x, ω) = 1,

ε(x, ω1 + ω2) · ζ (2)(x, ω1, ω2)+ χ (2)(x, ω1, ω2) : ζ (x, ω1)ζ (x, ω2) = 0(2),

ε(x, ω1 + ω2 + ω3) · ζ (3)(x, ω1, ω2, ω3)

+2χ (2)(x, ω1, ω2 + ω3) : ζ (x, ω1)ζ (2)(x, ω2, ω3)

+χ (3)(x, ω1, ω2, ω3)...ζ (x, ω1)ζ (x, ω2)ζ (x, ω3) = 0(3),

... , (3.182)

with ε(x, ω) = ε01+ χ (x, ω) the usual frequency-dependent tensor of permittivity.In particular,

ζ (x, ω) = [ε(x, ω)]−1. (3.183)

Here 1 and 0(n) are the second-rank unit tensor and the (n + 1)th-rank zero tensor,respectively. Introducing the Fourier components of the electric strength field andelectric displacement field, respectively,

E(x, ω) =∫

E(x, t)eiωt dt,

D(x, ω) =∫

D(x, t)eiωt dt, (3.184)

and performing the Fourier transform of both sides of (3.180), we obtain that

E(x, ω) = ζ (x, ω) · D(x, ω)

+∫

ζ(2)(x, ω1, ω − ω1) : D(x, ω1)D(x, ω − ω1) dω1

+∫ ∫

ζ(3)(x, ω1, ω2, ω − ω1 − ω2)

...D(x, ω1)D(x, ω2)

× D(x, ω − ω1 − ω2) dω1 dω2

+ · · · . (3.185)

Page 130: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 121

The inverse relation of D(x, ω) comprises χ (x, ω)≡ χ (1)(x,−ω; ω), χ (2)(x, ω1, ω−ω1) ≡ χ (2)(x,−ω; ω1, ω − ω1),... . A similar extension of notation is conceivablealso in tensors ζ (x, ω), ζ

(2)(x, ω1, ω − ω1),... .Let us note that the similar relation for P(x, ω) in Perina (1991) comprises sums

instead of the integrals. Relation (3.185) can resemble relation (2.4) in Drummond(1990) on the condition that the integrals will be replaced by the sums. Such achange does not affect only the meaning of the tensors χ (n) and ζ

(n) but also (andabove all) the physical unit of their measurement.

We will treat the time-averaged linear dispersive energy 〈H〉 for a classicalmonochromatic field at nonzero frequency ω. For a permittivity ε(x, ω), this can bewritten in terms of a complex amplitude E (Bloembergen 1965, Landau and Lifshitz1960, Bleany and Bleany 1985),

〈H〉 =∫

V

{E∗(x) · ∂

∂ω[ωε(x, ω)] · E(x)+ 1

2μ〈B(x, t) · B(x, t)〉

}d3x, (3.186)

where the angular brackets mean the time average and

E(x, t) = 2 Re{E(x)e−iωt

}. (3.187)

It is important to distinguish the monochromatic case from the case of quasi-monochromatic fields. In the more general case, the displacement D is expanded interms of a series of complex (envelope) functions, each of which has a restrictedbandwidth. The relevant nonzero central frequencies are then ω−N , ..., ωN , thus

D(x, t) =N∑

ν=−N

Dν(x, t), (3.188)

where

D−ν = (Dν)∗ (3.189)

and in the monochromatic case

Dν(x, t) = Dν(x)e−iων t . (3.190)

Here, our notation slightly differs from that in Drummond (1990). Again, theelectric-field vector can be expanded as

E(x, t) =N∑

ν=−N

Eν(x, t), (3.191)

where

E−ν = (Eν)∗ (3.192)

Page 131: Quantum Aspects of Light Propagation

122 3 Macroscopic Theories and Their Applications

and in the monochromatic case

Eν(x, t) = Eν(x)e−iων t . (3.193)

In the case of quasimonochromatic fields, relations (3.190) and (3.193) should bereplaced by

Dν(x, t) = 1

∫ ων+δ

ων−δ

D(x, ω)e−iωt dω,

Eν(x, t) = 1

∫ ων+δ

ων−δ

E(x, ω)e−iωt dω. (3.194)

Bloembergen (1965) presented the relation (3.186) as sufficiently accurate for sucha case. If relation (3.186) is exact for monochromatic fields, it must be modified fora quasimonochromatic field as follows:

⟨H (t ′)

⟩(t) =

∫ {1

2

1∑

ν=−1

E−ν(x, t) · ∂

∂ων[ωνε(x, ων)] · Eν(x, t)

+ 1

2μ〈B(x, t ′) · B(x, t ′)〉(t)

}d3x. (3.195)

By modifying the summation, we obtain the energy integral in terms of the electricdisplacement fields

⟨H (t ′)

⟩(t) =

∫ {1

2

N∑

ν=−N

D−ν(x, t) · [ζ (x, ων)− ων ∂

∂ωνζ′(x, ων)] · Dν(x, t)

+ 1

2μ〈B(x, t ′) · B(x, t ′)〉(t)

}d3x. (3.196)

To achieve a completeness, we supplement relations (3.188) and (3.191) with theexpansion of the magnetic induction field

B(x, t) =N∑

ν=−N

Bν(x, t), (3.197)

where

B−ν = (Bν)∗, (3.198)

Bν(x, t) = 1

∫ ων+δ

ων−δ

B(x, ω)e−iωt dω. (3.199)

Next, ζ (x, ω) can be approximated near ω = ων by a quadratic Taylor polynomial,

ζ (x, ω) ≈ ζ ν(x)+ ωζ′ν(x)+ 1

2ω2ζ

′′ν(x) ≡ ζ ν(x, ω), (3.200)

Page 132: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 123

so that

ζ (x, ω)− ω∂

∂ωζ (x, ω) ≈ ζ ν(x)− 1

2ω2ζ

′′ν(x). (3.201)

For brevity, the prime stands for the partial derivative ∂∂ω

. Moreover, the Taylorpolynomial is not in a standard form, which comprises the brackets (ω − ων) and(ω − ων)2. Further explanations can be found in Drummond (1990) if necessary.

Using the notation D ≡ ∂∂t D, we rewrite relation (3.196) in the form

⟨H (t ′)

⟩(t) = 1

2

N∑

ν=−N

∫ [D−ν(x, t) · ζ ν(x) · Dν(x, t)

− 1

2D−ν(x, t) · ζ ′′ν(x) · Dν(x, t)+ 1

μB−ν(x, t) · Bν(x, t)

]d3x.

(3.202)

Here we deviate slightly from Drummond (1990). Drummond speaks of time aver-ages and he indicates the time average on the left-hand side and partially on theright-hand side in (3.196), but he does not remove the time dependence from theright-hand side.

A canonical theory of linear dielectric will be obtained using the causal localLagrangian. Drummond (1990) considers a Lagrangian L(Λ−N , ..., ΛN ), which isa functional of (components of) the dual vector potential. This is defined as Λ,

D(x, t) = ∇ ×Λ(x, t),

B(x, t) = μΛ(x, t). (3.203)

We introduce also

Λ(x, ω) =∫

Λ(x, t)eiωt dt, (3.204)

Λν(x, t) = 1

∫ ων+δ

ων−δ

Λ(x, ω)e−iωt dω. (3.205)

Each quasimonochromatic field obeys the Maxwell equations

∇ × Eν(x, t) = −Bν(x, t),

∇ ×Hν(x, t) = Dν(x, t),

∇ · Dν(x, t) = 0,

∇ · Bν(x, t) = 0, (3.206)

where

Eν(x, t) = ζ ν(x) · Dν(x, t)+ iζ ′ν(x) · Dν(x, t)− 1

2ζ′′ν(x) · Dν(x, t),

Hν(x, t) = 1

μBν(x, t). (3.207)

Page 133: Quantum Aspects of Light Propagation

124 3 Macroscopic Theories and Their Applications

The components of the dual vector potential fulfil linear wave equations.On the basis of (3.202) we can infer the Hamiltonian function of the form

H = H0 = 1

2

∫ N∑

ν=−N

{[∇ ×Λ−ν(x, t)] · ζ ν(x) · [∇ ×Λν(x, t)]

− 1

2[∇ × Λ−ν(x, t)] · ζ ′′ν(x) · [∇ × Λν(x, t)]

+ μΛ−ν(x, t) · Λν(x, t)

}d3x. (3.208)

In order to quantize the theory, a canonical Lagrangian must be found that corre-sponds to (3.208) while generating the Maxwell equations (3.206) as Hamilton’sequations. It is next necessary to derive a Lagrangian whose Lagrange’s variationalequations correspond to obvious wave equations and whose Hamiltonian corre-sponds to (3.208). Since Λν can be specified to be transverse fields, the variationscan also be restricted to be transverse. The use of restricted variations can be realizedusing transverse functional derivatives (Power and Zienau 1959, Healey 1982).

Drummond (1994) derived the Lagrangian using the method of indeterminatecoefficients in the form

L = L0 = 1

2

∫ N∑

ν=−N

{μΛ−ν(x, t)Λν(x, t)

− [∇ ×Λ−ν(x, t)] · ζ ν(x) · [∇ ×Λν(x, t)]

− i[∇ ×Λ−ν(x, t)] · ζ ′ν(x) · [∇ × Λν(x, t)]

− 1

2[∇ × Λ−ν(x, t)] · ζ ′′ν(x) · [∇ × Λν(x, t)]

}d3x. (3.209)

The canonical momenta are

Πν(x, t) = B−ν(x, t)− 1

2∇ × [iζ ′ν(x) · D−ν(x, t)+ ζ ′′νx) · D−ν(x, t)

], (3.210)

where we introduce for brevity the fields (3.203) again. We can rewrite also theLagrangian of Drummond in the form

L = L0 = 1

2

∫ N∑

ν=−N

{1

μB−ν(x, t) · Bν(x, t)

− D−ν(x, t) · ζ ν(x) · Dν(x, t)− iD−ν(x, t) · ζ ′ν(x) · Dν(x, t)

− 1

2D−ν(x, t) · ζ ′′ν(x) · Dν(x, t)

}d3x. (3.211)

Page 134: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 125

On the contrary, the Legendre transformation, i.e. a substitution of Λν in Hamilto-nian (3.208) with an expression in Πν and Λν , was not performed in Drummond(1990). A reason is that each Λν is to be found from (3.210) considered as a par-tial differential equation. Also this theory simplifies a great deal if the plane waveone-dimensional propagation is considered.

The local Lagrangian method is used as the foundation of a nonlinear canonicalLagrangian and Hamiltonian. The objective is the total Lagrangian and Hamiltonianof the form

L = L0 −∫

U N(x, t) d3x,

H = H0 +∫

U N(x, t) d3x, (3.212)

where U N(x, t) is a nonlinear energy density

U N(x, t) = 1

3

N∑

ν1=−N

N∑

ν2=−N

N∑

ν3=−N

Dν1 (x, t)

· ζ (2)(x, ων2 ,−ων1 − ων2 ) : Dν2 (x, t)Dν3 (x, t)δ−ων1 ,ων2+ων3

+ 1

4

N∑

ν1=−N

N∑

ν2=−N

N∑

ν3=−N

N∑

ν4=−N

Dν1 (x, t)

· ζ (3)(x, ων2 , ων3 ,−ων1 − ων2 − ων3 )...Dν2 (x, t)Dν3 (x, t)Dν4 (x, t)

× δ−ων1 ,ων2+ων3+ων4

+ · · · . (3.213)

In order to give an example, a one-dimensional case is treated and the nonlinearrefractive index as the lowest nonlinearity of most universal interest. For N = 1,

U N(x, t) = 1

(3)|D1(x, t)|4. (3.214)

Drummond (1990) has presented the quantization of the nonlinear mediumusing a treatment of modes defined relative to the new Lagrangian. The canonicalmomenta have the form (3.210) also in the nonlinear case. In the correspondingquantum theory, the field operators Λν and Πν are introduced, which obey thetransverse commutation relations of the form

[Λνi (x, t), Πμ

j (x′, t)] = i�δ⊥i j (x− x′)δμν 1. (3.215)

Since these operators are not Hermitian, it is also interesting to note that

Λ−ν = (Λν)†, Π−μ = (Πμ)†. (3.216)

Page 135: Quantum Aspects of Light Propagation

126 3 Macroscopic Theories and Their Applications

This entails that Λνi , (Πν

j )† commute. Then, a set of Fourier transform fields is

defined and the annihilation operators aνk and bν

k are introduced. The operators aνk

correspond to the normal modes while bνk generate additional necessarily vacuum

modes. This feature of the theory is due to the dependence of the Hamiltonian(3.208) on both the real and imaginary parts of the components Λν(x, t) (Λν(x, t)).Conversely, the right-hand side of relation (3.202) can be completed with termswhich make up the Hamiltonian dependent solely on 2 Re{Λν(x, t)}.

3.2.2 Propagation in One Dimension and Applications

Drummond (1994) discusses in detail a simplified model of a one-dimensionaldielectric, where ζ (x, ω) = ζ (ω)1, ζ

(n)(x, ω1, . . . , ωn) = ζ (n)(ω1, . . . , ωn)1(n), with1(n) the (n + 1)th-rank unit tensor for n odd and ζ

(n)(x, ω1, . . . , ωn) = 0(n) for neven, n ≤ 3.

Instead of the time average of energy (3.186), (3.195), Drummond (1994) haspresented the total energy in the length L ,

W (t) =∫ L

0

[1

2μH 2(x, t)+

∫ t

t0

E(x, τ )D(x, τ ) dτ

]dx . (3.217)

Here H (x, t) = 1μ

B(x, t) is the magnetic strength vector. In part of the exposition,single polarization components are considered only. The Hillery–Mlodinow theorywhich does not take account of dispersion (Ho and Kumar 1993) has the electric-field commutation relation with the magnetic field modified from its free-field value.Drummond (1994) points out that the solution of this commutator problem is theinclusion of the important physical property of a real dielectric. Traditionally, thedescription of the nonlinear medium assumes that the dispersion terms are negligi-ble. Neglecting the unphysical modes, the dual vector potential has the expansion

Λ1(x, t) =∑

k

√�

∂ω∂k

2Lkζ (ωk)akeikx , (3.218)

where ak , a†k ′ have the standard commutators

[ak, a†k ′ ] = δk,k ′ 1, (3.219)

and ωk are solutions to the equations

ωk = k

√ζ (ωk)

μ. (3.220)

This enables one to write

H0 =∑

k

�ωk a†k ak (3.221)

Page 136: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 127

and (reintroducing D1)

H =∑

k

�ωk a†k ak +

∫U N(D1) dx . (3.222)

When there is a nonlinear refractive index or ζ (3) term, the free particles interactvia the Hamiltonian nonlinearity. It is this coupling that leads to soliton formation.It is also possible to involve other types of nonlinearity, such as ζ (2) terms, that leadto second harmonic and parametric interactions.

With respect to practical applications, it is necessary to define photon-density andphoton-flux amplitude fields. The photon-density amplitude field reminds us of theso-called detection operator (Mandel 1964, Mandel and Wolf 1995). A polariton-density amplitude field is simply defined as

Ψ(x, t) =√

1

L

k

ei[(k−k1)x+ω1t]ak, (3.223)

where k1 = k(ω1) is the centre wave number for the first envelope field. This fieldhas an equal-time commutator of the form

[Ψ(x1, t), Ψ†(x2, t)] = δ(x1 − x2)1, (3.224)

where δ is a version (L-periodic) of the usual Dirac delta function

δ(x1 − x2) ≡ 1

L

Δk

eiΔk(x1−x2), (3.225)

where the range of Δk is equal to that of k−k1. The total polariton number operatoris

N =∫

Ψ†(x, t)Ψ(x, t) dx . (3.226)

A polariton-flux amplitude can also be approximately expressed as

Φ(x, t) =√

v

L

k

ei[(k−k1)x+ω1t]ak, (3.227)

where v is the central group velocity at the carrier frequency ω1. This flux has anequal-time commutator of the form

[Φ(x1, t), Φ†(x2, t)] = vδ(x1 − x2)1. (3.228)

Page 137: Quantum Aspects of Light Propagation

128 3 Macroscopic Theories and Their Applications

Operationally, 〈Φ†(x, t)Φ(x, t)〉 is the photon-flux expectation value in units of pho-tons/second.

A common choice is to define the dimensionless field φ(x, t) by the scaling

φ(x, t) = Ψ(x, t)

√vt0n

, (3.229)

where n is the photon-number scale and t0 is a timescale, defined so that the expec-tation value 〈φ†(x, t)φ (x, t)〉 is appropriate for the system. This scaling transforma-tion is accompanied by a change to a comoving coordinate frame. The first choiceof an altered space variable gives

ξv =xv− t

t0, τ = vt

x0. (3.230)

Here x0 is a spatial length scale introduced to scale the interaction times. An alter-native moving frame transformation is

ξ = x

x0, τv =

t − xv

t0. (3.231)

The quantization technique developed by Drummond (1990) was applied to thecase of a single-mode optical fibre (Drummond 1994). On simplified assumptions,the nonlinear Hamiltonian is (cf. (3.214))

H =∫

�ω(k)a†(k)a(k) dk + 1

4ζ (3)

∫D4(x) d3x. (3.232)

Here ω(k) are the angular frequencies of modes with wave vectors k describingthe linear photon or polariton excitations in the fibre including dispersion. a(k) arecorresponding annihilation operators defined so that, at equal times,

[a(k ′), a†(k)] = δ(k − k ′)1. (3.233)

In terms of the waveguide, the electric displacement field D(x) is expressed as

D(x) = i∫ √

�ε(k)kv(k)

4πa(k)u(k, r)eikx dk + H.c., (3.234)

where x = (x, r) and

∫|u(k, r)|2 d2r = 1. (3.235)

Here v(k) is the group velocity and ε(k) is the dielectric permittivity. The mode func-tion u(k, r) is included here to show how the simplified one-dimensional quantum

Page 138: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 129

theory relates to vector mode theory. When the interaction Hamiltonian describingthe evolution of the polariton field Ψ(x, t) in the slowly varying envelope androtating-wave approximations is considered, the coupling constant χe is introduced

χe ≡ 3�[χ (3)(ω1)]2[v(k1)]2

4ε(k1)c2

∫|u(r)|4 d2r. (3.236)

After taking the free evolution into account, the following Heisenberg equation ofmotion for the field operator propagating in the +x-direction can be found

[v

∂x+ ∂

∂t

]Ψ(x, t) =

[iω′′

2

∂2

∂x2+ iχeΨ

†(x, t)Ψ(x, t)

]Ψ(x, t), (3.237)

where v = v(k1) = ∂ω∂k

∣∣k=k1 , ω′′ = ∂2ω

∂k2

∣∣∣k=k1

. In a comoving reference frame, this

reduces to the usual quantum nonlinear Schrodinger equation

i∂

∂tψ1(xv, t) =

[−ω′′

2

∂2

∂x2v

− χeψ†1(xv, t)ψ1(xv, t)

]ψ1(xv, t), (3.238)

where ψ1(xv, t) = Ψ(xv + vt, t). In the case of anomalous dispersion which occursat wavelength longer than 1.5 μm, allowing solitons to form, the second derivativeω′′ can be expressed as

ω′′ = �

m, (3.239)

where m = �

ω′′ is an effective mass of a particle. Similarly, the nonlinear term χe

describes an interaction potential

V (xv, x ′v) = −χeδ(xv − x ′v). (3.240)

This interaction potential is attractive when χe is positive as it is in most Kerr media.It is known that this potential has bound states and is one of the simplest exactlysoluble known quantum field theories. The repulsive and attractive cases were inves-tigated by Yang (1967, 1968). This theory is one-dimensional and tractable anddoes not need renormalization, while two- and three-dimensional versions do needrenormalization.

In calculations, it is preferable to substitute flux amplitude operator

Ψ(x, t) =√

1

vΦ(x, t) (3.241)

into equation (3.237). Drummond (1994) associates an idea of the spatial pro-gression with the flux amplitude operator. Upon modifying the time variable, heobtains an “unusual form” of the quantum nonlinear Schrodinger equation, whichhe reduces to a more usual form again. Since the operators there have their standard

Page 139: Quantum Aspects of Light Propagation

130 3 Macroscopic Theories and Their Applications

meaning, they must have equal-time commutators. In contrast, the resulting equa-tion (Drummond 1994) appears as the quantum nonlinear Schrodinger equationwith time and space interchanged. Such an interpretation means that the operatorshave equal-space commutators. The problem is whether these commutators are welldefined.

An important physical effect in propagation is that from molecular excitations.For this reason, the nonlinear Schrodinger equation requires corrections due torefractive-index fluctuations for pulses longer than about 1 ps, especially when highenough intensities are present, and fails for pulse duration much shorter than this.

The treatment of the quantum theory can start from the classical theory developedby Gordon (1986). The Raman interaction energy of a fibre is known to be [Carterand Drummond (1991)]

WR =∑

j

ζ Rj

...D(x j )D(x j )δx j . (3.242)

Here D(x j ) is the electric displacement at the j th mean atomic location x j , δx j isthe atomic displacement operator, and ζ R

j is a Raman coupling tensor.In order this interaction to be quantized, the existence of a corresponding set

of phonon operators must be taken into account. The Raman effect can be includedmacroscopically through a continuum Hamiltonian term coupling photons to phononsof the form (Drummond and Hardman 1993)

HR = �

∫ ∞

−∞

∫ ∞

0Ψ†(z, t)Ψ(z, t)r (z, ω)[ A(z, ω, t)+ A†(z, ω, t)] dω dz

+ �

∫ ∞

−∞

∫ ∞

0ω A†(z, ω, t) A(z, ω, t) dω dz, (3.243)

where

[ A(z, ω, t), A†(z′, ω′, t)] = δ(z − z′)δ(ω − ω′)1, (3.244)

and r (z, ω) is a macroscopic frequency-dependent coupling which can be assumedto be independent of z. Here the Raman excitations are treated as an inhomoge-neously broadened continuum of modes localized at each longitudinal location z.The corresponding coupled set of nonlinear operator equations is

[v

∂z+ ∂

∂t

]Ψ(z, t) =

[iω′′

2

∂2

∂z2+ iχeΨ

†(z, t)Ψ(z, t)

]Ψ(z, t)

− i

[∫ ∞

0r (ω)[ A(z, ω, t)+ A†(z, ω, t)] dω

]Ψ(z, t)

(3.245)

and

∂tA(z, ω, t) = −i A(z, ω, t)− ir (ω)Ψ†(z, t)Ψ(z, t). (3.246)

Page 140: Quantum Aspects of Light Propagation

3.2 Dispersive Nonlinear Dielectric 131

The phonon operators do not have white-noise behaviour, but a coloured noise prop-erty.

In practical terms, the known exact solutions (Yang 1968) of the quantum non-linear Schrodinger equation can be hardly utilized at typical photon numbers of 109.It is often more useful to employ phase-space distributions or operator distributionssuch as the Wigner representation (Wigner 1932) and the Glauber–Sudarshan P-representation (Glauber 1963, Sudarshan 1963). In the review (Drummond 1994),the generalized P-representation is mentioned and the positive P-representation isused. Using this method, the operator equations are transformed to complex Itostochastic equations which involve only c-number commuting variables. In otherwords, an operator equation can be transformed to an equivalent pair of c-numberstochastic equations for Ψ(z, t) and A(z, ω, t). On the transformation (3.231), equa-tions (3.245) and (3.246) are ready for the positive P-representation. Thus, equiv-alent stochastic differential equations are obtainable. Substituting the integratedphonon variables into the equations for the photon field gives the following equationfor a new function φ(ζ, τv):

∂ζφ(ζ, τv) ≈ [i f φ†(ζ, τv)φ(ζ, τv)± i

2

∂2

∂τ 2v

φ(ζ, τv)

+ i∫ τv

−∞h(τv − τ ′v)φ†(ζ, τ ′v)φ(ζ, τ ′v) dτ ′v

+√

i f

nΓ(ζ, τv)+ iΓR(ζ, τv)]φ(ζ, τv). (3.247)

There is a corresponding Hermitian conjugate equation for φ† obtained by makingthe substitutions φ → φ†, i → −i, Γ → Γ†, ΓR → Γ

†R, with Γ, Γ†, ΓR, Γ

†R being

independent noise terms. In relation (3.247),

h(τv) = 2∫ ∞

0r2

t0

)sin(ντv)

χ, n = |k ′′|v2

χ t0, (3.248)

ζ = z

z0, f = χe

χ, z0 = t2

0

|k ′′| , k ′′ = −ω′′

v3. (3.249)

The last terms appearing in Equation (3.247) are stochastic functions. Γ representsthe quantum noise of a field introduced by the electronic nonlinearity and ΓR is thethermal noise due to the phonon coupling. In numerical simulation, the use of anenlarged nonclassical phase space can increase computation times. For this reason,the Wigner function defined on a classical phase space is useful. The Wigner func-tion does not have an exact stochastic equation. This is because there are third-orderderivative terms in the Fokker–Planck equation for the Wigner function that haveno stochastic equivalent. In sufficiently intense fields, the disagreeable terms can beneglected. The Wigner function represents symmetrically ordered operators which

Page 141: Quantum Aspects of Light Propagation

132 3 Macroscopic Theories and Their Applications

have a diverging vacuum noise term, because a cutoff is not considered or is takento be infinity.

As the χ (3) nonlinearity in silica optical fibres is low, Drummond and He (1997)have proposed the investigation of a quantum soliton, which occurs in parametricwaveguides. Such an object consists of a superposition of a second-harmonic photonwith a localized pair of subharmonic photons. The system is analogous to the quarkmodel of the meson.

On the simplifying assumption that the medium is homogeneous and isotropic,Milonni (1995) has considered the classical expression for the field energy and liftedthe restriction to the magnetic susceptibility independent on frequency (cf. (3.177)).Then he has applied the results to treat basic emission and absorption processesfor atoms in dispersive dielectric host media. Using this simple approach to quan-tization, Milonni and Maclay (2003) have shown how radiative recoil, the Dopplereffect, and spontaneous and stimulated radiation rates are set up when the radiatoris embedded in a host medium having a negative index of refraction.

Matsko and Kozlov (2000) have presented an approach which absorbs the resultsof two previous studies (Drummond and Carter 1987, Haus and Lai 1990). The twotheories have been shown to provide similar outcomes of a homodyne measurement.It has been concluded that both equal-time and equal-space commutation relationsare valid for the quantum soliton description. In Matsko and Kozlov (2000) the workwith physical units could be amended.

Korolkova et al. (2001) have studied a quantum soliton in a Kerr medium. Theyhave simplified, implicitly, the classical propagation equation for the slowly vary-ing electric-field envelope by introducing a new time measurement in dependenceon a position. In changing to dimensionless variables, they make the new time a“position” and the position a “time” variable and then get a classical nonlinearSchrodinger equation.

Raymond Ooi and Scully (2007) have studied three-level extended medium,which is utilized as an amplifier. They begin with a single three-level cascade atomand with a χ (2) crystal, which is described by coupled parametric amplifier equations(Boyd 2003, Yariv 1989, Shen 1984). Further they present the theory and the resultsof the three-level cascade scheme. They compare this model with the simple one.They calculate cross correlation of the idler E1(z, t) and the signal E2(z, t),

G(2)21 (τ ) ≡ 〈E†

1(z, t)E†2(z, t + τ )E2(z, t + τ )E1(z, t)〉. (3.250)

The neglect of the Langevin noise seems to be admissible especially for large detun-ing of the pump. They calculate also reverse correlation,

G(2)12 (τ ) ≡ 〈E†

2(z, t)E†1(z, t + τ )E1(z, t + τ )E2(z, t)〉. (3.251)

They complete the observation of antibunching and oscillations in the reverse cor-relation with an interesting physics of the three-level atomic system.

Page 142: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 133

3.3 Modes of Universe and Paraxial QuantumPropagation

The laser physics and the optical engineering are typical of their calculation meth-ods and the effort for their improvement is apparent. The laser cavity is coupledto the outer space and the mode coupling can be investigated in detail. The parax-ial description of light propagation can be quantized. Detectors of radiation with aspatial resolution motivate the inclusion of the optical imaging in quantum optics.

3.3.1 Quasimode Description of Spectrum of Squeezing

Toward the end of the 1980s, it had become clear that the use of squeezed states(Walls 1983, Loudon and Knight 1987) in the interferometry can lead to theenhancement of signal-to-noise ratios. Milburn and Walls (1981) have shown thatthe cavity of a degenerate parametric oscillator admits only a 50% amount ofsqueezing (in the steady state). Yurke was first to realize that the pessimistic conclu-sions do not hold as the noise reduction in the transmitted field can be quite differentfrom that in the intracavity field (Yurke 1984). As a first step one had to relate thefield operators inside and outside the cavity. Whereas it was obvious that the fieldoperators inside the cavity remain the usual quantum-mechanical annihilation oper-ators of one or a small number of harmonic oscillators, the connection of the fieldoperators outside the cavity with the “Langevin-noise operators” was establishedas late as 1980s by Collett and Gardiner (1984), Gardiner and Collett (1985), andCarmichael (1987). These authors have cleared up the relation of this subtle propertyof squeezed light and its generation with the concept of light propagation. Not onlythe interpretation but also the derivation of the Langevin-like “noise” terms waspresented by Lang and Scully (1973), after they introduced and studied the “modesof the universe” (Lang 1973, Ujihara 1975, 1976, 1977). It is in order to mentiona book of Scully and Zubairy (1997), where the results of Gea-Banacloche et al.(1990a) are expounded or formulated as exercises.

The modes of universe are discussed, which include the interior of the imperfectcavity of interest, and are used to define the intracavity quasimode, the incidentexternal field mode, and the output field mode. The mutual coupling of these modesemerges naturally in this formalism. Following Lang (1973), the one-sided emptycavity is described also by the relation

E (+)cav (t) = r E (+)

cav

(t − 2l

c

)+ t E (+)

in (t), (3.252)

where l is the cavity length, r is a real amplitude reflection coefficient, and t isa respective transmission coefficient, t = √

1− r2. Here E (+)in (t) is a positive-

frequency part of the input field and it fulfils the commutation relation

[E (+)in (t), E (−)

in (s)] = K δ(t − s)1, (3.253)

Page 143: Quantum Aspects of Light Propagation

134 3 Macroscopic Theories and Their Applications

where E (−)in (t) = [E (+)

in (t)]† and

K = �Ω

2ε0cA , (3.254)

with Ω being a quasimode frequency. For the full Fox–Li quasimode (Fox andLi 1961, Barnett and Radmore 1988), a single-mode annihilation operator a(t) isdefined

a(t) =√

2l

K cE (+)

cav (t)eiΩt . (3.255)

It is convenient to use the slowly varying amplitudes E (+)in (t), E (+)

out (t), and E (+)cav (t)

for the input, output, and cavity fields related to the cavity frequency Ω, respectively,

E (+)in (t) = E (+)

in (t)e−iΩt , (3.256)

E (+)out (t) = E (+)

out (t)e−iΩt , (3.257)

E (+)cav (t) = E (+)

cav (t)e−iΩt . (3.258)

We propose to consider the definition

a(t) =√

c

K 2l

∫ t

t− 2lc

E (+)cav (t ′) dt ′ (3.259)

instead of (3.255), which is in a better agreement with the quantum field theory.To approve this change we denote the rightward and leftward travelling positive-frequency parts as E (+)

> (z, t) and E (+)< (z, t) so that E (+)

in (t) ≡ E (+)> (−0, t), E (+)

out (t)≡ E (+)

< (−0, t), E (+)cav (t) ≡ E (+)

> (+0, t). The factor at the delta function in the equal-time commutator of the field is K c and from this we can calculate the equal-spacecommutator (t > 0, s > 0 without loss of generality)

[E (+)in (t), E (−)

in (s)] = [E (+)> (−0, t), E (−)

> (−0, s)]

= [E (+)> (−ct, 0), E (−)

> (−cs, 0)] (3.260)

= K cδ(−ct + cs)1 = K δ(t − s)1. (3.261)

Hence, definition (3.259) can be rewritten in the form,

a(t) =√

1

K c2l

∫ l

0

[E (+)

> (z, t)+ E (+)< (z, t)

]dz, (3.262)

where (cf. (3.258)) E (+)> (z, t) and E (+)

< (z, t) mean the slowly varying smooth ampli-tude with the property

E (+)> (z, t) = E (+)

> (z, t)e−ik0z+iΩt , k0 = Ω

c, (3.263)

E (+)< (z, t) = E (+)

< (z, t)eik0z+iΩt . (3.264)

Page 144: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 135

Performing the time integration on both sides of relation (3.252), we obtain that

a(t) = r a

(t − 2l

c

)+ t b(t), (3.265)

where

b(t) =√

c

K 2l

∫ t

t− 2lc

E (+)in (t ′) dt ′. (3.266)

Recalling the space integration, we see that the annihilation operator a(t) corre-sponds to the same quasimode in all times, whereas b(t) is appropriate to manydistinct modes.

In the situation when it holds that

r a

(t − 2l

c

)≈(

1− 1

2t2

)a(t)− 2l

c

d

dta(t), (3.267)

in the short cavity round-trip time limit we get

d

dta(t) = −Γa(t)+

√2Γ

√c

2lb(t), (3.268)

where

Γ = c

2l

1

2t2. (3.269)

Noting that

√c

2lb(t) =

√1

K

c

2l

∫ t

t− 2lc

E (+)in (t ′) dt ′

=√

1

KE (+)

in (t), (3.270)

where we replaced the average over the short-time interval by the value of thefunction (at the upper limit), we rewrite equation (3.268) as a quantum Langevinequation

d

dta(t) = −Γa(t)+

√2Γ

√1

KE (+)

in (t). (3.271)

Further, Gea-Banacloche et al. (1990a) define, for arbitrary measurement times,the spectrum of squeezing of the output field via the quadrature variances. Theypresent a microscopic effective Hamiltonian model of balanced homodyne detection.

Page 145: Quantum Aspects of Light Propagation

136 3 Macroscopic Theories and Their Applications

They refer to the fundamental papers (Collett and Gardiner 1984, Gardiner and Col-lett 1985, Caves and Schumaker 1985, Yurke 1985), where this concept of spectralsqueezing was originally treated. As an approximation, the operator is introduced

ˆAout(δω) = N√T

∫ T

0E (+)

out (t)eiδωt dt, (3.272)

where

N =√

1

K. (3.273)

As shown also by Yurke (1985) and Carmichael (1987), with a balanced homodynedetector one measures the combinations

Eoutθ (t) = 1

2

[eiθE (+)

out (t)+ e−iθE (−)out (t)

], (3.274)

and from this the natural generalization of the single-mode quadrature concept is

ˆAoutθ (δω) = 1

2

[eiθ ˆAout(δω)+ e−iθ ˆA†

out(−δω)]. (3.275)

We may wonder why a non-Hermitian operator is taken for such a generalization ofthe Hermitian operator. Finally, the connection between single quasimode squeezingand spectral squeezing is explored and the difference in the noise reduction insideand outside the cavity is clarified in a way that lends itself to a simple visualization.

Gea-Banacloche et al. (1990b) have first analysed measurements of small phaseor frequency changes for an ordinary laser and calculated the extra cavity phasenoise for a phase-locked laser. These analyses are based on the mean values andthe normally ordered variances of quantum operators for which classical Langevinequations may be written down. The classical Langevin formalism is further replacedby the alternative Fokker–Planck formalism for the calculation of the spectrum ofsqueezing. This general Fokker–Planck formalism was applied to the two-photoncorrelated-spontaneous-emission laser. It has been shown that without one-photonresonance and initial atomic coherences involving the middle level, the maximumsqueezing of the ultracavity mode is 50% while the detected field can be almost per-fectly squeezed. Almost the exact reverse holds, however, if one-photon resonanceand initial atomic coherences involving the middle level are present. In particular,the intracavity field may be perfectly squeezed while the outside field is not onlyunsqueezed but has, in fact, increased noise in the conjugate quadrature. Finally, theeffect of finite measurement time on the quadrature variances is briefly analysed.

Dutra and Nienhuis (2000) have unified the concept of normal modes used inquantum optics and that of Fox–Li modes from semiclassical laser physics. Theirone-dimensional theory solves the problem of how to describe the quantized radia-tion field in a leaky cavity using Fox–Li modes. In this theory, unlike conventional

Page 146: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 137

models, system and reservoir operators no longer commute with each other, as aconsequence of natural cavity modes having been used. Aiello (2000) has derivedsimple relations for an electromagnetic field inside and outside an optical cavity,limiting himself to one- and two-photon states of the field. He has expressed input–output relations using a nonunitary transformation between intracavity and outputoperators.

Brown and Dalton (2002) have considered three-dimensional unstable opticalsystems. They have defined non-Hermitian modes and their adjoints in both thecavity and external regions. A number of concepts and properties resulting from thestandard canonical quantization procedure have been suited to the non-Hermitianmodes by the exact transformation method. The results are applied to the sponta-neous decay of a two-level atom inside an unstable cavity.

3.3.2 Steady-State Propagation

In Deutsch and Garrison (1991a) it is assumed that in the case of amplifier, one isusually interested in the spatial dependence of temporally steady-state fields. It is noattempt at a reformulation of one-dimensional propagation, cf. Abram and Cohen(1991), where the temporal evolution by the Hamiltonian is supplemented by thespatial progression with the momentum operator. An alternative proposal is madethat the quantum-mechanical equivalent of the classical steady-state condition is thedescription of the system by a stationary state of a suitable Hamiltonian. There isa formal resemblance to a nonrelativistic many-body theory for a complex scalarfield (Deutsch and Garrison 1991b), which helps determine the Hamiltonian. In thistheory a non-Hermitian envelope-field operator Ψ(z, t), with the property

[Ψ(z, t), Ψ†(z′, t)] = δ(z − z′)1, (3.276)

is introduced. In the application to the optical field, the vector potential (or electricfield in the lowest order) corresponding to a carrier plane wave of a given polariza-tion e is expressed as follows:

E (+)ω (z, t) = e

√2π�ω

An2(ω)Ψ(z, t) exp[i(kz − ωt)], (3.277)

where A is the beam area and n(ω) is a dispersive index of refraction.In contrast to Deutsch and Garrison (1991a), we will make a simplification, i.e.

we will not consider a carrier wave Hamiltonian. For the single wave interactingnonlinearly with matter, the total Hamiltonian can be written as

H = Henv + Hint, (3.278)

Page 147: Quantum Aspects of Light Propagation

138 3 Macroscopic Theories and Their Applications

Henv = − i�c

2n(ω)

×∫ (

Ψ†(z, t)∂Ψ(z, t)

∂z− ∂Ψ†(z, t)

∂zΨ(z, t)

)dz, (3.279)

where Henv is the Hamiltonian governing the free progression of the envelope andHint is a general interaction Hamiltonian. In fact, the generality will not be exer-cised and we will treat only the vacuum input and the case of degenerate parametricamplifier.

In the standard Heisenberg picture, the equation of motion for the envelope-fieldoperator reads

∂Ψ(z, t)

∂t= − i

�[Ψ(z, t), H ], (3.280)

or for a linear medium

∂Ψ(z, t)

∂t= − c

n(ω)

∂Ψ(z, t)

∂z. (3.281)

The solution is

Ψ(z, t) = Ψ

(z − ct

n(ω), 0

). (3.282)

In the standard Schrodinger picture, the state |Φ〉 evolves by the Schrodingerequation

∂|Φ〉∂t

= − i

�(Henv + Hint)|Φ〉. (3.283)

The introducing of the carrier wave Hamiltonian has revoked the considering ofthe Schrodinger picture (Deutsch and Garrison 1991b), along with the envelope pic-ture which we have confined ourselves to. Relation (3.277) is the positive-frequencycomponent of the electric-field in the envelope picture similarly as relation (4.5b) inCaves and Schumaker (1985) is this component in the interaction picture. The enve-lope picture is essentially the modulation picture in Caves and Schumaker (1985).

For application under consideration there will be exact frequency matchingbetween the carrier frequencies of the various waves which interact so that theHamiltonian in equation (3.283) will be independent of time, thus the steady state(ss) solutions are identified with the stationary solutions to equation (3.283):

(Henv + Hint

) |Φ〉ss = λ|Φ〉ss. (3.284)

For the stationary solutions, the label (ss) will be omitted.

Page 148: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 139

(i) In the case

Hint = 0, (3.285)

i.e. in the case of vacuum propagation, the stationary solutions are the translation-invariant states. To have a unitary representation of the translation, we may considereither the limits−∞,∞ in the integral on the left-hand side in (3.279) or the spatialperiodicity. We prefer the latter possibility.

As an example, we can consider a coherent state corresponding to a constantone-photon wave function. We define a functional displacement operator

D[α] ≡ exp

(∫[α(z)Ψ†(z)− α∗(z)Ψ(z)] dz

)(3.286)

= exp(ρa† − ρa), (3.287)

where

ρ =√∫

|α(z)|2 dz, (3.288)

a ≡ a

ρ

]≡ 1

ρ

∫α∗(z)Ψ(z) dz. (3.289)

The coherent state is defined as the displaced vacuum

|{α}〉 ≡ D[α]|0〉. (3.290)

Since

[a, a†] = 1, (3.291)

relation (3.286) can be rewritten in the normal ordering form

D[α] = exp

(−1

2

∫|α(z)|2 dz

)

× exp

(∫α(z)Ψ†(z) dz

)exp

(−∫

α∗(z)Ψ(z) dz

)(3.292)

and the corresponding one-photon state is

∣∣∣∣1,

ρ

}⟩≡ 1

ρ

∫α(z)Ψ†(z)|0〉 dz. (3.293)

On substituting |Φ〉 = |{α}〉 into relation (3.284) and applying then the operatorD†[α] from left to both sides, we derive that α(z) and λ should make the vacuum

Page 149: Quantum Aspects of Light Propagation

140 3 Macroscopic Theories and Their Applications

state the eigenstate of the operator

D†[α]Henv D[α] = − i�

2

c

n(ω)

∫ { [Ψ†(z)+ α∗(z)1

] ∂[Ψ(z)+ α(z)1

]

∂z

− ∂[Ψ†(z)+ α∗(z)1

]

∂z

[Ψ(z)+ α(z)1

] }dz. (3.294)

The eigenvalue is λ again. On equating

D†[α]Henv D[α]|0〉 = −i�c

n(ω)

∫Ψ†(z)

dα(z)

dz|0〉 dz

− i�c

n(ω)

∫α∗(z)

dα(z)

dzdz|0〉 (3.295)

with λ|0〉, we see that

dα(z)

dz= 0 (3.296)

and hence

λ = 0. (3.297)

Instead of a translation-invariant wave function, we may try one that is an eigen-function of the translation operator. When the boson number operator commuteswith the operator

A = Henv + Hint, (3.298)

on the left-hand side of (3.284), this problem can be generalized by the insertion ofthe number operator N ,

N =∫

Ψ†(z)Ψ(z) dz, (3.299)

behind λ on the right-hand side. The idea takes into account that the wave functionof any number of particles should be the eigenfunction of the operator A

exp(i A)|Φ〉 = exp(iλN )|Φ〉 (3.300)

or

exp(i( A − λN ))|Φ〉 = |Φ〉. (3.301)

Condition (3.301) is equivalent to

( A − λN )|Φ〉 = 0 (3.302)

or to the relation with the insertion. On substituting again Φ into the new relation,and applying the operator D†[α], we obtain the right-hand side in the form

Page 150: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 141

λ

∫ [Ψ†(z)+ α∗(z)1

] [Ψ(z)+ α(z)1

]dz. (3.303)

Condition (3.296) is thus generalized,

− i�c

n(ω)

dα(z)

dz= λα(z). (3.304)

The solution to equation (3.304) reads

α(z) = α(0) exp

[iλn(ω)

�cz

], (3.305)

where λ is any real number. Expression (3.305) for the complex amplitude is suffi-cient for the fulfilment of the relation A|Φ〉 = λN |Φ〉.

(ii) In the case of the degenerate parametric amplifier, the interaction Hamiltoniancan be written (Hillery and Mlodinow 1984) as follows:

Hint = −1

2

∫ ∫ ∫ {χ (2)(z)E∗p (z) exp[−i(kpz − ωpt)][E (+)

ω (z, t)]2

+ H.c.}

dx dy dz, (3.306)

where ωp is the pump frequency, ωp = 2ω, χ (2)(z) is the second-order susceptibilitycoupling the pump to the degenerate signal and idler fields and Ep(z) is the pumpamplitude. Substituting for E (+)

ω (z, t) from relation (3.277) gives the interactionHamiltonian in the envelope picture

Hint = i�

2

c

n(ω)

∫[κ∗(z)Ψ2(z)− κ(z)Ψ†2(z)] dz, (3.307)

with

κ(z) = 1

2g0(z) exp[iφ(z)], (3.308)

g0(z) = 4πω

n(ω)c|χ (2)(z)||Ep(z)|, (3.309)

φ(z) = −π

2+Δk z + β(z). (3.310)

Here g0(z) is the standard power gain coupling constant (Yariv 1985), Δk = 2k−kp

is the phase mismatch at the degenerate frequency, and β(z) is the remaining phaseoriginating from the product χ (2)(z)E∗p (z).

Page 151: Quantum Aspects of Light Propagation

142 3 Macroscopic Theories and Their Applications

To solve the time-independent Schrodinger equation, Deutsch and Garrison(1991a) assume that the eigenstate is a squeezed vacuum state corresponding toa two-photon wave function. They define a functional squeezing operator

S[ξ ] ≡ exp

{1

2

∫[ξ (z)Ψ†2(z)− ξ ∗(z)Ψ2(z)] dz

}, (3.311)

with z-dependent squeezing parameter ξ (z) = −r (z) exp[iθ (z)]. The squeezed vac-uum is defined as

|0〉{ξ} ≡ S[ξ ]|0〉. (3.312)

Similarly as in case (i), ξ (z) and λ should be solutions of the equation

S†[ξ ](Henv + Hint)S[ξ ]|0〉 = λ|0〉. (3.313)

Applying the operator Ψ(z) to both the sides of (3.313) and taking into account that

λΨ(z)|0〉 = 0 = S†[ξ ](Henv + Hint)S[ξ ]Ψ(z)|0〉, (3.314)

we rewrite the eigenvalue problem in the λ-independent form

[S†[ξ ](Henv + Hint)S[ξ ], Ψ(z)

] |0〉{ξ} = 0 = ˆC |0〉, (3.315)

where the commutator ˆC is

ˆC = i�c

n(ω)

{− exp[iθ (z)]

(exp[−iθ (z)] cosh[r (z)]

d

dz

(exp[iθ (z)] sinh[r (z)]

)

− d

dz

(cosh[r (z)]

)sinh[r (z)]

+ κ∗(z) exp[iθ (z)] sinh2[r (z)]− κ(z) exp[−iθ (z)] cosh2[r (z)])Ψ†(z)

+(

cosh[r (z)]d

dz

(cosh[r (z)]

)

− exp[iθ (z)] sinh[r (z)]d

dz

(exp[−iθ (z)] sinh[r (z)]

)

+ κ∗(z) exp[iθ (z)] sinh[r (z)] cosh[r (z)]

− κ(z) exp[−iθ (z)] sinh[r (z)] cosh[r (z)])Ψ(z)+ ∂

∂zΨ(z)

}. (3.316)

Page 152: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 143

The eigenvalue condition requires that the real and imaginary parts of the coefficientat Ψ† vanish yielding the desired propagation equations

dr

dz= 1

2g0 cos(θ − φ), (3.317)

dz= −g0 coth(2r ) sin(θ − φ). (3.318)

On introducing the complex amplitude

ζ (z) = − exp[iθ (z)] tanh[r (z)], (3.319)

we can write equations (3.317) and (3.318) in the compact form

d(−ζ )

dz= κ − κ∗ζ 2, (3.320)

which may be useful for guessing the boundary condition

r (z) |z=0 = 0, θ(z) |z=0 = φ(0). (3.321)

When β(z) = 0 and the phase difference θ (z)− φ(0) is small, the squeezing param-eter r (z) integrates values of experimental parameter 1

2 g0(z′), z′ ∈ [0, z], whenmoreover g0(∞) > |Δk|, the squeezing parameter θ (z) converges to the function

of the experimental parameters{φ(z)− arcsin

[Δk

g0(∞)

]}.

A direct solution of (3.313) requires that the real and imaginary parts of thecoefficient at Ψ†(z) vanish yielding the propagation equations (3.317) and (3.318)again. The presence of the singular operator Ψ(z)Ψ†(z) indicates that λ has no finitevalue in general. Resorting to the partition of the field into finite elements of lengthΔz in each of which we can define local field operators, we find that λ = O

(1

Δz

),

λ � − 1

Δz

�c

n(ω)

∫sin(θ − φ)

sinh3 r

cosh rdz. (3.322)

3.3.3 Approximation of Slowly Varying Envelope

The macroscopic approach to the quantum propagation aims at a quantum version ofthe slowly varying envelope approximation. Such an envelope implies that the waveis paraxial and monochromatic. The problem of quantum propagation of paraxialfields was considered first by Graham and Haken (1968). The revived interest isindicated by Kennedy and Wright (1988). Deutsch and Garrison (1991b) begin withgeneralizing the results of Lax et al. (1974), which develop the classical theory of astrictly monochromatic wave in an inhomogeneous nonlinear (perhaps amplifying)medium. The generalization is made only to a quasimonochromatic wave and the

Page 153: Quantum Aspects of Light Propagation

144 3 Macroscopic Theories and Their Applications

quantum theory is presented in the simplest system of codirectional propagationconsidering only the free-field dynamics.

In the Coulomb gauge, the positive-frequency component of the vector potentialsatisfies the free-field wave equation

∇2A(+)(x, t)− 1

c2

∂2

∂t2A(+)(x, t) = 0. (3.323)

The approximation of the slowly varying envelope is introduced by expressingA(+)(x, t) as envelope modulating a carrier plane wave propagating in the z-directionwith the wave number k0 and the frequency ω0 = ck0,

A(+)(x, t) = A0Ψ(x, t) exp[i(k0z − ω0t)]. (3.324)

Here, Ψ(x, t) is a vector-valued function, henceforth referred to as an envelope field,and A0 is a normalization constant, which we will specify before relation (3.333).The initial positive-frequency component can be expressed as follows:

A(+)(x, t = 0) = 1

(2π )32

∫ √�c

2|k|∑

λ=1,2

eλ(k)Fλ(k)eik.x d3k. (3.325)

Here eλ(k) are the orthogonal polarization unit vectors and the reduced Planckconstant is introduced in view of the possible later quantization. Fλ(k) are thusmomentum-space wave functions.

The intuitive notion of a paraxial field is that it is composed of rays making smallangles with the main propagation axis. In other words, a paraxial wave function{Fλ(k)} is concentrated in a small neighbourhood k0 = k0e3 of the wave vector ofthe carrier wave. We define fλ(q) by the relation

fλ(q) = Fλ(q+ k0), (3.326)

where q is the relative wave vector. Let us observe that q= (qT, qz), where qT is thetransverse part of q, q = qT + qze3.

In contrast to Deutsch and Garrison (1991b), we stress that we express the con-centration in a small neighbourhood of q0 = 0, by letting the wave function { fλ(q)}depend on a small positive parameter θ , fλ(q) ≡ fλ(q, θ ). Let us assume that

fλ(qT, qz, θ ) =√V f λ

(qT

θk0,

qz

θ2k0, θ

), (3.327)

where

V = 1

θ4k30

, (3.328)

Page 154: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 145

and we have introduced the notational convention that an overbar indicates a dimen-sionless function of the scaled variables (and perhaps θ ). This relates to defining adimensionless “momentum” vector η = (ηT, ηz), where

ηT =qT

θk0, (3.329)

ηz = qz

θ2k0. (3.330)

The functions of interest are those that have a convergent power-series expan-sion in θ ,

f λ(η, θ ) =∞∑

n=0

θn f(n)λ (η). (3.331)

In contrast to Deutsch and Garrison (1991b), we note that

f λ(η, θ ) = f(0)λ (η). (3.332)

Relations (3.329) and (3.332) lead to the wave function being θ -dependent, a differ-ence from Deutsch and Garrison (1991b).

Substituting integral (3.325) for the envelope field defined by (3.324) at t = 0,with the momentum-space wave function given by equation (3.326), and choosing

A0 =√

�c2k0

, we find that

Ψ(x, t = 0, θ ) = 1

(2π )32

×∫ √

k0

|q+ k0|∑

λ=1,2

eλ(q+ k0) fλ(q, θ )eiq.x d3q. (3.333)

Here, the parameter θ has been introduced, which is not present in integral (3.325),where Fλ(k) ≡ Fλ(k, θ ), A(+)(x, t) ≡ A(+)(x, t, θ ).

In Deutsch and Garrison (1991b), the integro-differential form of the wave equa-tion for A(+)(x, t, θ ) is investigated,

i∂

∂tA(+)(x, t, θ ) = c(−∇2)

12 A(+)(x, t, θ ), (3.334)

where (−∇2)12 is an integral operator defined by

(−∇2)12 F(x) = 1

(2π )32

∫|k|F(k)eik.x d3k, (3.335)

with F(k) being the Fourier transform. Let us note that ∇ = (∇T, ∂∂z ). Substituting

from (3.324) into (3.334) gives

i∂

∂tΨ(x, t, θ ) = (cΩ− ω0)Ψ(x, t, θ ), (3.336)

Page 155: Quantum Aspects of Light Propagation

146 3 Macroscopic Theories and Their Applications

where

Ω ≡ Ω

(∇T,

∂z

)=(

k20 − 2ik0

∂z− ∇2

T −∂2

∂z2

) 12

. (3.337)

The scaled configuration-space variables ξ = (ξT, ζ ) are

ξT = θk0xT, ζ = θ2k0z, (3.338)

and the dimensionless time variable

τ = θ2ω0t. (3.339)

After expressing the envelope field in the form

Ψ(x, t, θ ) ≡ Ψ(xT, z, t, θ ) = 1√V

Ψ(θk0xT, θ2k0z, θ2ω0t, θ ), (3.340)

we can rewrite relation (3.336) as follows:

i∂

∂τΨ(ξ , τ, θ ) = H(θ )Ψ(ξ , τ, θ ), (3.341)

where

H(θ ) = 1

θ2[ Ω(θ )− 1 ], (3.342)

Ω(θ ) =Ω(θk0∇T, θ2k0

∂∂ζ

)

k0. (3.343)

This provides the possibility of expanding the differential operator H(θ ),

H(θ ) =∞∑

n=0

θnH(n), (3.344)

where the differential operators H(n)are just defined by the formal expression.

The dimensionless amplitude Ψ(ξ , τ, θ ) has the expansion

Ψ(ξ , τ, θ ) =∞∑

n=0

θnΨ(n)

(ξ , τ ). (3.345)

It is evident that the terms satisfy the equations

i∂

∂τΨ

(n)(ξ , τ ) =

n∑

m=0

H(n−m)Ψ

(m)(ξ , τ ), n = 0, 1, 2, . . . . (3.346)

Page 156: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 147

In Deutsch and Garrison (1991b), the discussion of the classical equation ofmotion is completed by considering the initial-value problem. We rewrite equa-tion (3.333) as

Ψ(x, θ ) = 1

(2π )32

∫ ∑

λ=1,2

Kλ(q) fλ(q, θ )eiq·x d3q, (3.347)

where the function Kλ(q) is defined by

Kλ(q) =√

k0

|q+ k0| eλ(q), (3.348)

where

eλ(q) = eλ(q+ k0). (3.349)

Reexpressing (3.347) in terms of the scaled variables gives

Ψ(ξ , τ = 0; θ ) = 1

(2π )32

∫ ∑

λ=1,2

Kλ(η, θ ) f λ(η)eiη·ξ d3η. (3.350)

The scaled kernel function is

Kλ(η, θ ) = eλ(η, θ )√w(η, θ )

, (3.351)

where

w(η, θ ) =√

1+ θ2(2ηz + η2T )+ θ4η2

z . (3.352)

Considering the expansion

Kλ(η, θ ) =∞∑

n=0

θnK(n)λ (η), (3.353)

we obtain the initial expansion of the envelope fields

Ψ(n)

(ξ ) = 1

(2π )32

∫ ∑

λ=1,2

K(n)λ (η) f λ(η)eiη·ξ d3η. (3.354)

Deutsch and Garrison (1991b) claim that the preceding arguments can be usedto identify the subspace of the photon Fock space consisting of the paraxial states

Page 157: Quantum Aspects of Light Propagation

148 3 Macroscopic Theories and Their Applications

of the field. They resorted to the space S, the infinitely differentiable functions thatdecrease, when |η| → ∞, faster than any power of |η|−1.

Let us recall the standard plane wave creation and annihilation operators a†λ(k),

aλ(k), and the wave packet creation and annihilation operators

a†[F ] =∫ ∑

λ=1,2

Fλ(k)a†λ(k) d3k (3.355)

and its conjugate. We now define

c†[ f ] =∫ ∑

λ=1,2

c†λ(q) fλ(q) d3q, (3.356)

where

c†λ(q) = a†λ(q+ k0) (3.357)

are the creation operators corresponding to the envelope field.Before we generalize the definition of a state with exactly one photon present,

| f ; 1〉〉 = c†[ f ]|0〉, (3.358)

where f is a normalized one-photon wave function, we return to the original oper-ators. Let us proceed with the generalized commutation relations

[a[F ], a†[G]

] = (F ,G)1 =∫ ∑

λ=1,2

F∗λ (k)Gλ(k) d3k 1. (3.359)

If F is normalized, then

[a[F ], a†[F ]

] = 1. (3.360)

Let us observe that

a†2[F ] =∫ ∫ ∑

λ1

λ2

Fλ1λ2 (k1, k2)a†λ1

(k1)a†λ2

(k2) d3k1 d3k2, (3.361)

where

Fλ1λ2 (k1, k2) = Fλ1 (k1)Fλ2 (k2) (3.362)

exemplifies a normalized symmetric function. We are led to the definition of thestate with exactly two photons present

|F ; 2〉〉 = 1√2

∫ ∫ ∑

λ1

λ2

Fλ1λ2 (k1, k2)a†λ1

(k1)a†λ2

(k2)|0〉 d3k1 d3k2, (3.363)

Page 158: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 149

where F is a normalized symmetric two-photon wave function. In general,

|F ; m〉〉 = 1√m!

∫. . .

∫ ∑

λ1

. . .∑

λm

Fλ1...λm (k1, . . . , km)

× a†λ1

(k1) . . . a†λm

(km)|0〉 d3k1 . . . d3km, m ≥ 1, (3.364)

where F is this time a normalized symmetric wave function of m photons and

|F ; 0〉〉 = F |0〉, (3.365)

with F a complex unit. This almost completes the definition of the Fock space, sinceany element of this space has the form

|Φ〉 =∞∑

m=0

|F (m); m〉〉, (3.366)

where F (m) are (unnormalized) symmetric m-photon wave functions, m ≥ 1, andF (0) is a complex number. The pure state |Φ〉 is normalized if and only if the func-tions F (m) are jointly normalized by

∞∑

m=0

(F (m),F (m)) = 1. (3.367)

Similarly,

| f ; m〉 = 1√m!

∫. . .

∫ ∑

λ1

. . .∑

λm

fλ1...λm (q1, . . . , qm)

× c†λ1(q1) . . . c†λm

(qm)|0〉 d3q1 . . . d3qm, m ≥ 1, (3.368)

where

fλ1...λm (q1, . . . , qm) = Fλ1...λm (q1 + k0, . . . , qm + k0) (3.369)

and

| f ; 0〉 = f |0〉, (3.370)

where f = F , and any element of the Fock space has the form

|Φ〉 =∞∑

m=0

| f (m); m〉, (3.371)

Page 159: Quantum Aspects of Light Propagation

150 3 Macroscopic Theories and Their Applications

where

f (m)λ1...λm

(q1, . . . , qm) = F (m)λ1...λm

(q1 + k0, . . . , qm + k0) (3.372)

and f (0) = F (0).For the subsequent analysis, a unitary operator T (θ ) is of interest such that

T (θ )| f ′; m〉 = | f (θ ); m〉, (3.373)

where (cf. (3.327))

fλ1...λm (qT1, qz1, . . . , qTm, qzm, θ ) = V m2

× f ′λ1...λm

(qT1

θ,

qz1

θ2, . . . ,

qTm

θ,

qzm

θ2

).

(3.374)

In the description of the dynamics using the Schrodinger picture, the paraxialapproximation means mainly the evolution of the initial state |Φ(t, θ )〉,

∂t|Φ(t, θ )〉 = − i

�H |Φ(t, θ )〉, (3.375)

where

|Φ(t, θ )〉|t=0 = T (θ )|Φ′(t = 0)〉. (3.376)

Defining the state

|Φ′(t, θ )〉 = T †(θ )|Φ(t, θ )〉, (3.377)

for all times, we can rewrite (3.375) in the form

∂t|Φ′(t, θ )〉 = − i

�H ′(θ )|Φ′(t, θ )〉, (3.378)

where

H ′(θ ) = T †(θ )H T (θ ). (3.379)

Using the expansion

H ′(θ ) =∞∑

m=0

θm H ′(m), (3.380)

we may expand equation (3.378) into the coupled equations for the coefficients ofthe series

|Φ′(t, θ )〉 =∞∑

m=0

θm |Φ′(m)(t)〉. (3.381)

Page 160: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 151

Describing the dynamics in the Heisenberg picture, we should generalize relations(3.379) and (3.380) to an arbitrary operator M(t),

M ′(t, θ ) = T †(θ )M(t)T (θ ), (3.382)

M ′(t, θ ) =∞∑

m=0

θm M ′(m)(t). (3.383)

We can then rewrite the equation of motion

∂tM(t) = − i

�[M(t), H (t)] (3.384)

in the form

∂tM ′(t, θ ) = − i

�[M ′(t, θ ), H ′(t, θ )]. (3.385)

We may expand this equation into coupled equations similarly as (3.378). Since

T (1) = 1, M ′(t, 1) = M(t), (3.386)

relation (3.383) simplifies for θ = 1, or any operator M(t) can be expressed as

M(t) =∞∑

m=0

M (m)(t). (3.387)

In both the pictures, definition (3.339) can be used whenever it is advantageous.According to Deutsch and Garrison (1991b), we introduce the operator Ψ(x) by

relation (3.333), with Ψ(x, t = 0, θ ) �→ Ψ(x) on the left-hand side and fλ(q, θ ) �→cλ(q) on the right-hand side. This operator can be expanded as

Ψ(x) =∞∑

n=0

Ψ(n)(x), (3.388)

where

Ψ(n)(x) = 1

(2π )32

∫ ∑

λ=1,2

K(n)λ (q)cλ(q)eiq·x d3q. (3.389)

Using this expansion, we can compute the commutators between fields of differentorders that are

[Ψ(n)i (x), Ψ(m)†

j (x′)] = 1

(2π )32

∫ ∑

λ=1,2

K (n)λi (q)K (m)

λ j (q)eiq·(x−x′) d3q. (3.390)

Page 161: Quantum Aspects of Light Propagation

152 3 Macroscopic Theories and Their Applications

As expected, the nth-order commutator can be expressed as

[Ψi (x), Ψ†j (x

′)](n) =n∑

m=0

[Ψ(n−m)i (x), Ψ(m)†

j (x′)]. (3.391)

The equal-time commutation relations are preserved by the dynamics in each orderof the approximation scheme,

∂t[Ψi (x, t), Ψ†

j (x′, t)](n) = 0. (3.392)

In the zeroth order, the theory yields a quantized analogue of the classical parax-ial wave equation, and formally resembles a nonrelativistic many-particle theory.This formalism is applied to show that Mandel’s local-photon-number operator andGlauber’s photon-counting operator reduce, in the zeroth order, to the same truenumber operator. In addition, it is shown that the O(θ2)-difference between themvanishes for experiments described by stationary coherent states.

A nonperturbative quantization of a paraxial electromagnetic field has beenachieved by forcing the plane waves involved in the expression for the vector-potential operator to obey paraxial wave equations at the time origin (Aiello andWoerdman 2005).

3.3.4 Optical Imaging with Nonclassical Light

In optical imaging with nonclassical light or quantum imaging it is important toknow how quantum entanglement properties of light beams in the spatial domain canbe exploited in order to improve the quality of processing of images and of parallelsignals (Gatti 2003). In this section, we first expound some general concepts asspatially multimode squeezing and spatial entanglement, and describe some opticaldevices that are able to generate light beams with these properties (Kolobov 1999).Then we provide some references to interesting approaches in this field.

Kolobov (1999) enriches exposition of the usual quantum optics by new facts.The time moments are completed with space points ρ. It is connected with exis-tence of very small photodetectors or pixels. The observed quantity is the surfacephotocurrent density operator, an Hermitian operator, i(ρ, t).

Let the photodetection plane be located at the point with longitudinal coordinatez normal to the z-axis. Let E (+)(z, ρ, t) mean the positive-frequency operator of theelectric field of a quasiplane and a quasimonochromatic wave travelling in the +z-direction, where ρ is the position vector in the transverse plane of the wave. Thisoperator can be written in terms of space- and time-dependent photon annihilationand creation operators a(z, ρ, t) and a†(z, ρ, t) as

E (+)(z, ρ, t) = i

√�ω0

2ε0cexp[i(k0z − ω0t)]a(z, ρ, t). (3.393)

Page 162: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 153

Here ω0 is the carrier frequency of the wave and k0 is its wave number. But a(z, ρ, t)and a†(z, ρ, t) are not the standard-modal annihilation and creation operators. Theyobey the commutation relations

[a(z, ρ, t), a†(z, ρ ′, t ′)] = δ(ρ − ρ ′)δ(t − t ′)1,

[a(z, ρ, t), a(z, ρ ′, t ′)] = 0 (3.394)

and are normalized so that the mean value 〈a†(z, ρ, t)a(z, ρ, t)〉 determines themean photon-flux density in photons per cm2 per second at point ρ and time t .

The quantum theory of photodetection provides the following expressions forthe mean value of the photocurrent density operator 〈i(ρ, t)〉, and its space–timecorrelation function 〈 1

2 {δi(ρ, t), δi(ρ′, t ′)}+〉:

〈i(ρ, t)〉 = η〈 I (ρ, t)〉, (3.395)⟨

1

2{δi(ρ, t), δi(ρ′, t ′)}+

⟩= 〈i(ρ, t)〉δ(ρ − ρ ′)δ(t − t ′)

+ η2〈: I (ρ, t) I (ρ ′, t ′) :〉 − 〈i(ρ, t)〉〈i(ρ ′, t ′)〉. (3.396)

Here I (ρ, t) = a†(z, ρ, t)a(z, ρ, t) is the photon-flux density operator.The second contribution to the correlation function of the photocurrent density

operator is proportional to the normal- and time-ordered space–time intensity cor-relation function

G(2)(ρ, t ; ρ ′, t ′) = 〈: I (ρ, t) I (ρ ′, t ′) :〉. (3.397)

This correlation function is proportional to the probability of detecting a photon attime t ′ and at the spatial point ρ ′ under the condition that the previous detectionhappened at time t and point ρ. When the intensity of light is stationary in time anduniform in the transverse area of the light beam, this correlation function dependsonly on the time difference τ = t ′ − t and the spatial difference ξ = ρ ′ −ρ betweentwo points, G(2)(ρ, t ; ρ ′, t ′) = G(2)(ξ , τ ).

One can define the degree of second-order spatio-temporal coherence as

g(2)(ξ , τ ) = 〈: I (ρ, t) I (ρ + ξ , t + τ ) :〉〈 I (ρ, t)〉2 . (3.398)

If the correlation function g(2)(ξ , τ ) has its maximum at ξ = 0 and at τ = 0,g(2)(0, 0) > g(2)(ξ , τ ), it is natural to speak of the bunching in space–time. Analo-gously, if g(2)(0, 0) < g(2)(ξ , τ ), one may speak of the antibunching in space–time.

The antibunching in space–time is a purely quantum-mechanical phenomenon.Indeed, it follows from the Schwarz inequality that the correlation function g(2)(ξ , τ )of a classical electromagnetic field stationary in time and uniform in space mustsatisfy

g(2)(0, 0) ≥ g(2)(ξ , τ ) (3.399)

Page 163: Quantum Aspects of Light Propagation

154 3 Macroscopic Theories and Their Applications

for arbitrary ξ and τ . Since the antibunching in space–time means an exactly oppo-site inequality, it cannot be explained within the framework of semiclassical theory,i.e. when the light field is treated as a c-number.

The photocurrent noise spectrum is defined as a Fourier transform of the pho-tocurrent correlation function. As follows from relation (3.396), for a light fieldstationary in time and uniform in the transverse plane, the correlation function ofthe photocurrent density operator 〈 1

2 {δi(ρ, t), δi(ρ′, t ′)}+〉 depends only on the timedifference τ and the spatial difference ξ . The noise spectrum of the photocurrentdensity operator is the spatio-temporal Fourier transform of this correlation func-tion,

(δi)2(q,Ω) =∫ ∫ ⟨

1

2{δi(0, 0), δi(ρ, t)}+

× exp[i(Ωt − q · ρ)] dt d2ρ. (3.400)

Using the photodetection formula (3.396), we can write the noise spectrum

(δi)2(q,Ω) as follows

(δi)2(q,Ω) = 〈i(ρ, t)〉 + G(2)(q,Ω)− 〈i(ρ, t)〉2δ(Ω)δ(q). (3.401)

Here the first contribution comes from the shot-noise term in relation (3.396), thesecond one is the spatio-temporal Fourier transform of the intensity correlation func-tion,

G(2)(q,Ω) =∫ ∫

exp[i(Ωt − q · ρ)]G(2)(ρ, t) dt d2ρ, (3.402)

and the last one from the space–time-independent product of two mean photocurrentdensities. One can show that in semiclassical theory the sum of the second and thirdcontributions is always nonnegative. Therefore the semiclassical minimum value ofthe photocurrent density noise is given by the shot noise in space–time,

(δi)2(q,Ω) = 〈i(ρ, t)〉. (3.403)

This formula is a generalization of the standard quantum limit for a single-modefield described by the photon annihilation and creation operators a(t) and a†(t),respectively,

(δi)2(Ω) =∫ ⟨

1

2{δi(0), δi(t)}+

⟩exp(iΩt) dt = 〈i(t)〉, (3.404)

where i(t) = a†(t)a(t), {·, ·}+ indicates an anticommutator, from the temporaldomain into the space–time one. In quantum theory the sum of the second andthird terms in relation (3.401) can be negative and compensate partially or evencompletely for the shot-noise contribution for some frequencies Ω and spatial fre-quencies q.

Page 164: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 155

An opinion of many workers in quantum optics is expressed in Kolobov (1999).The difficulties associated with the quantum-mechanical description of field propa-gation in free space or a nonlinear medium lie in the usual procedure of field quan-tization. Evolution of the quantized field due, for instance, to the interaction withan atomic medium is described in terms of the Heisenberg equations for annihila-tion and creation operators, i.e. as purely temporal evolution. Such a descriptionof field dynamics is not well suited to the problem of field propagation in freespace or a medium. At the start of such a study, it would be more appropriate tohave a quantum-mechanical analogue of the classical wave-optical propagation anddiffraction theory. Such a description for transparent nonlinear media when the fieldinteraction with atoms is described in terms of an effective Hamiltonian is muchappreciated. But the simpler question of quantized field propagation in free space isconsidered first.

Let E (+)(r, t), where r = (x, y, z) is the spatial coordinate, be the positive-frequency operator of the electric field in a vacuum. In the continuum limit, thisoperator is written in the form of the modal decomposition

E (+)(r, t) = i

√�

2ε0

1

(2π )3

∫ √ω(k)a(k)

× exp[i(k · r− ω(k)t)] d3k. (3.405)

Here a(k) and a†(k) are the photon annihilation and creation operators of a spa-tial mode with the wave vector k; the frequency ω(k) is given by the free-spacedispersion relation ω(k) = kc, with k = |k|. The operators a(k) and a†(k) obey thecanonical commutation relations

[a(k), a†(k′)] = (2π )3δ(k− k′)1, [a(k), a(k′)] = 0. (3.406)

The factor (2π )3 is not usual, but such particularities appear consistently in thereview article. Equation (3.405) determines the Heisenberg field operator E (+)(r, t)in all points r and t of the space–time as a solution of the initial-value problem,i.e. through the modal operators a(k) and a†(k) given at time t = 0 as Schrodingeroperators. For a complete quantum-mechanical description, we have to specify thedensity matrix of the field for the continuum set of modes k. In the Heisenbergrepresentation (3.405), this density matrix remains constant as time evolves. For awave travelling in the+z-direction, we would like to have a formula that determinesthe field operator at any point ρ in the transverse plane at coordinate z given the fieldoperator over the plane z = 0.

On comparison of relation (3.393) with relation (3.405), one sees that

a(z, ρ, t) = 1

(2π )3

∫ ∫ √ω(k)

k0a(k)

× exp{i[q · ρ + (kz − k0)z − (ω(k)− ω0)t]} d2q dkz . (3.407)

Page 165: Quantum Aspects of Light Propagation

156 3 Macroscopic Theories and Their Applications

The normalization of these operators is such that the free Hamiltonian of the elec-tromagnetic field can be written as

H0 = �ω0

c

Va†(z, ρ, t)a(z, ρ, t) d3r. (3.408)

The commutation relations (3.394) are not proved in this connection, but equal-timeones are derived,

[a(z, ρ, t), a†(z′, ρ ′, t)] = δ(r− r′)1, (3.409)

where

δ(r− r′) ≈(

1− i

k0

∂z− 1

2k20

∇2⊥

)δ(r− r′), (3.410)

with ∇2⊥ being the transverse Laplacian with respect to ρ. Here we have reproduced

only the expression derived in the quasimonochromatic and paraxial approxima-tions from the literature. In this approximation, the equation for the slowly varyingoperator a(z, ρ, t) reads

∂ta(z, ρ, t) =

(−c

∂z+ c

i

2k0∇2⊥

)a(z, ρ, t). (3.411)

An unpublished result of Sokolov, which is related to the equation for propaga-tion of a quantized field in a nonlinear parametric medium, is reproduced in Kolobov(1999). In part, it is based on the book of Klyshko (1988).

The positive-frequency operator of a quantized electric field in a transparentdielectric medium can be written in a form similar to that for a vacuum (Klyshko1988),

E (+)(r, t) = i

√�

2ε0

1

(2π )3

∫ξ (k)

√ω(k)a(k)

× exp[i(k · r− ω(k)t)] d3k. (3.412)

This differs from relation (3.405) in the factor ξ (k), which describes the strength ofthe field in the medium as compared to that in a vacuum. This constant is

ξ 2(k) = u(k)v(k)

c2 cos ρ(k). (3.413)

Here v(k) = cn(k) is the phase velocity of light in the medium, u(k) = ∂ω(k)

∂k is thegroup velocity, and ρ(k) is the so-called generalized anisotropy angle, that is, theangle between the electric field and the induction.

Page 166: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 157

The fact that the electromagnetic field is a vector field is not emphasized inKolobov (1999), but it is mentioned in respect of the book by Klyshko (1988).Relation (3.412) yet neglects the use of and summation over the appropriate parame-ter ν. The dispersion relation ω(k) is not single valued, but has at least two branches.These branches should be distinguished by a parameter μ. The annihilation opera-tors correspond to these branches. Relation (3.412) neglects the use of and summa-tion over the parameter μ too.

Using this notation one can introduce the slowly varying operator a(z, ρ, t) ofthe quantized field in the medium (cf. equation (3.393)),

E (+)(z, ρ, t) = iξ

√�ω0

2ε0cexp[i(k1z − ω0t)]a(z, ρ, t). (3.414)

Here we have denoted by k1 the wave number of the wave in the medium. The slowlyvarying operator a(z, ρ, t) is given by an equation identical to relation (3.407),

a(z, ρ, t) = 1

(2π )3

∫ ∫ √ω(k)

k0a(k)

× exp{i[q · ρ + (kz − k0)z − (ω(k)− ω0)t]} d2q dkz, (3.415)

but here ω(k) means a dispersion relation for the medium.One will describe the parametric interaction in the medium in terms of an effec-

tive Hamiltonian. It is assumed that a χ (2) nonlinear parametric medium fills a vol-ume V . The medium is illuminated by a monochromatic plane wave, the pump.The pump wave propagates in the +z-direction and has the frequency ωp and wavenumber kp,

E (+)p (z, ρ, t) = Ep exp[i(kpz − ωpt)]. (3.416)

We choose the frequency ωp of the pump wave in the form ωp = 2ω0 and considerthe amplitude Ep as a c-number, i.e. we neglect the quantum fluctuations of thepump wave.

Under usual assumptions the parametric interaction can be described by the fol-lowing effective Hamiltonian

Hint = i�n0g

c

Vexp[i(kp − 2k1)z][a†(z, ρ, t)]2 d3r+ H.c. (3.417)

Here n0 gives the density of active atoms in the parametric medium, and g is thestrength constant of the parametric interaction proportional to the amplitude Ep ofthe pump wave and the susceptibility constant χ (2) of the medium.

The evolution of the slowly varying amplitude operator a(z, ρ, t) in the paramet-ric medium is described by the following equation:

∂ta(z, ρ, t) = iω0a(z, ρ, t)+ i

�[H0 + Hint, a(z, ρ, t)]. (3.418)

Page 167: Quantum Aspects of Light Propagation

158 3 Macroscopic Theories and Their Applications

Here H0 is the free-field Hamiltonian in the medium. In terms of a(z, ρ, t) anda†(z, ρ, t) it is given by relation (3.408). One introduces the Fourier transform ofthe space–time photon annihilation operator a(z, ρ, t),

˜a(s, q,Ω) =∫ ∫ ∫

a(z, ρ, t) exp(iΩt) exp(−isz) exp(−iq · ρ) dt d2ρ dz

(3.419)

≡∫

e−isz ˜a(z, q,Ω) dz.

We express the Fourier transform ˜a(z, q,Ω) with the aid of a new operator ε(z, q,Ω),

˜a(z, q,Ω) = ε(z, q,Ω) exp{i[kz(q,Ω)− k1]z}, (3.420)

where

kz(q,Ω) =√

k2(ω0 +Ω)− q2, (3.421)

with q = |q| a z-component of the wave vector with frequency ω0 +Ω and spatialfrequency q. For ˜ε(s, q,Ω) similarly defined as ˜a(s, q,Ω) it holds that

˜ε(s, q,Ω) = ˜a(s + kz(q,Ω)− k1, q,Ω). (3.422)

One introduces the mismatch function Δ(q,Ω),

Δ(q,Ω) = kz(q,Ω)+ kz(−q,−Ω)− kp. (3.423)

One lets u = ∂ω(k1)∂k1

mean the group velocity of the wave in the crystal. On appropri-ate derivations, Kolobov (1999) presents the equation of propagation for the operatorε(z, q,Ω),

∂zε(z, q,Ω) = σ ε†(z,−q,−Ω) exp[iΔ(q,Ω)z], (3.424)

where σ = 2n0gu is the coupling constant of the parametric interaction.

When an active nonlinear medium is placed in a resonator, a description mayemploy discrete transverse modes of the cavity. Lugiato and Gatti (1993), Gatti andLugiato (1995), and Lugiato and Marzoli (1995) have adopted this approach. Letfl (ρ) mean these eigenmodes. The set of functions fl (ρ) satisfies both the conditionof orthonormality

∫f ∗l (ρ) fl ′(ρ) d2ρ = δll ′ (3.425)

and completeness

l

f ∗l (ρ) fl (ρ′) = δ(ρ − ρ ′). (3.426)

Page 168: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 159

One can expand the slowly varying field operator a(ρ, t) over the eigenmodes fl(ρ),

a(ρ, t) =∑

l

fl(ρ)al(t), (3.427)

where al (t) are operator-valued expansion coefficients that have the meaning ofphoton annihilation operators for the lth mode. From the commutation relations(3.394) together with relation (3.425) it is easy to see that al(t) and a†

l (t) obey thecommutation relation

[al(t), a†l ′ (t

′)] = δll ′δ(t − t ′)1. (3.428)

It is noted that the derivation is valid for the field operators outside the cavity. TheFourier transforms ˜al(Ω) are defined as

˜al(Ω) =∫ ∞

−∞exp(iΩt)al(t) dt. (3.429)

Noise spectrum of the photocurrent density for the lth mode as an analogue of

the noise spectrum (δi)2(q,Ω) is introduced in Kolobov (1999). It is not assumedthat the photocurrent density is uniform in space, but that it is stationary in time andthe photocurrent density fluctuations for different eigenmodes are uncorrelated.

We can express the space–time correlation function of the photocurrent density

operator (3.396) in terms of the noise spectra (δi)2l(Ω) of the individual eigenmodes

fl (ρ) of the cavity,

⟨1

2{δi(ρ, t), δi(ρ ′, t ′)}+

⟩=∑

l

fl(ρ) fl(ρ′)

× 1

∫ ∞

−∞(δi)2

l(Ω) exp[−iΩ(t − t ′)] dΩ. (3.430)

(i) Generation of multimode squeezed states of light

The generation of multimode squeezed states of light by a travelling-wave opti-cal parametric amplifier was described by Kolobov and Sokolov (1989a,b). As aresult of the parametric down-conversion, a pump photon ωp splits into signal andidler photons, with frequencies ω0 + Ω and ω0 − Ω, and wave vectors k(q,Ω)and k(−q,−Ω). Their transverse components are ±q and their z-componentsare kz(q,Ω) and kz(−q,−Ω), respectively, by relation (3.421). The evolution ofthe slowly varying operator ε(z, q,Ω) inside the crystal is described by equation(3.424). Solving this equation and respecting relation (3.420) between the operators˜a(z, q,Ω) and ε(z, q,Ω), one arrives at the following transformation

˜a(l, q,Ω) = U (q,Ω) ˜a(0, q,Ω)+ V (q,Ω) ˜a†(0,−q,−Ω), (3.431)

Page 169: Quantum Aspects of Light Propagation

160 3 Macroscopic Theories and Their Applications

with coefficients U (q,Ω) and V (q,Ω) equal to

U (q,Ω) = exp

{i

[kz(q,Ω)− k1 − Δ(q,Ω)

2

]l

}

×[

cosh(Γl)+ iΔ(q,Ω)

2Γsinh(Γl)

],

V (q,Ω) = exp

{i

[kz(q,Ω)− k1 − Δ(q,Ω)

2

]l

}

× σ

Γsinh(Γl), (3.432)

where

Γ =√|σ |2 − [Δ(q,Ω)]2

4. (3.433)

The functions U (q,Ω) and V (q,Ω) have the property

|U (q,Ω)|2 − |V (q,Ω)|2 = 1. (3.434)

At the input to the crystal, operators ˜a(0, q,Ω) and ˜a†(0, q,Ω) obey the free-field

commutation relation

[ ˜a(0, q,Ω), ˜a†(0, q′,Ω′)] = (2π )3δ(q− q′)δ(Ω−Ω′)1. (3.435)

The broad-band squeezing in a three-wave interaction was discussed by Caves andCrouch (1987) and in a four-wave interaction by Levenson et al. (1985) in the caseof co-propagation and by Yurke (1985) for counter-propagation. Equation (3.431)involves the spatial frequency q. It is assumed that, along with the pump wave, amonochromatic plane wave of frequency ω0 is incident normal to the input surfaceof the crystal. Upon leaving the crystal, this wave will serve as a local oscillatorwave with the complex amplitude β, which first enters as a wave with complexamplitude α and q = 0. Relation (3.431) is used,

β = |β| exp(iϕβ) = αU (0, 0)+ α∗V (0, 0). (3.436)

The type of noise modulation of the resultant field in space–time is determinedby the angle

θ (q,Ω) = ψ(l, q,Ω)− ϕβ. (3.437)

Phase modulation predominates for θ (q,Ω) = ±π2 and amplitude modulation for

θ (q,Ω) = 0, π . The mean of the photocurrent density operator and its noisespectrum is found from relations (3.395), (3.396), and (3.400). The mean of thephotocurrent density operator has the forms

〈i〉 = η|β|2 + η

(2π )3

∫ ∫|V (q,Ω)|2 d2q dΩ ≡ 〈i〉l + 〈i〉s, (3.438)

Page 170: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 161

where the subscript l refers to the local oscillator field and the subscript s indicatesthe spontaneous parametric down-conversion. One introduces the function

δ(q,Ω) = |V (q,Ω)|2. (3.439)

The mean of the photocurrent density operator 〈i〉s can be written as

〈i〉s = ηδs

TcSc, (3.440)

where

δs = 1

q2c Ωc

∫ ∫δ(q,Ω) d2q dΩ (3.441)

is the degeneracy parameter for spontaneous parametric down-conversion (Mandeland Wolf 1995), Tc = 2π

Ωcis its coherence time, Sc = ( 2π

qc)2 the coherence area, and

Ωc and qc the widths of the frequency and spatial frequency spectra of spontaneousparametric down-conversion.

The noise spectrum of the photocurrent density has the form

(δi)2(q,Ω) = 〈i〉 + 2η2|β|2 [δ(q,Ω)+ Re{exp(−2iϕβ)g(q,Ω)}]

+ η2

(2π )3

∫ ∫[δ(q′,Ω′)δ(q− q′,Ω−Ω′)

+ g∗(q′,Ω′)g(q− q′,Ω−Ω′)] d2q′ dΩ′, (3.442)

where

g(q,Ω) = U (q,Ω)V (−q,−Ω). (3.443)

Under homodyne detection, the down-conversion waves (q,Ω) and (−q,−Ω)modulate the local oscillator wave in space and time. With any angle ψ(z, q,Ω) fora while, slow quadrature components ˜aμλ(z, q,Ω), λ = c, s, are introduced with theproperty

˜a1c(z, q,Ω)+ i ˜a1s(z, q,Ω)

= exp[−iψ(z, q,Ω)] ˜a(z, q,Ω)+ exp[iψ(z, q,Ω)] ˜a†(z,−q,Ω), (3.444)

˜a2c(z, q,Ω)+ i ˜a2s(z, q,Ω)

= −i{exp[−iψ(z, q,Ω)] ˜a(z, q,Ω)− exp[iψ(z, q,Ω)] ˜a†(z,−q,Ω)

}. (3.445)

In other words, “complex quadrature components” are first defined. When the com-plex components are used, transformation (3.431) simplifies to the form

˜aμc(l, q,Ω)+ i ˜aμs(l, q,Ω)

= exp[iκ(q,Ω)] exp[±r (q,Ω)][ ˜aμc(0, q,Ω)+ i ˜aμs(0, q,Ω)], (3.446)

Page 171: Quantum Aspects of Light Propagation

162 3 Macroscopic Theories and Their Applications

where + corresponds to the component with μ = 1 and − to the component withμ = 2. The components ˜aμλ(0, q,Ω) at the input surface to the crystal are definedin the coordinate system with ψ(0, q,Ω) and the components ˜aμλ(l, q,Ω) at theoutput surface of the crystal are defined with ψ(l, q,Ω). These angles are

ψ(0, q,Ω) = 1

2arg[V (q,Ω)U−1(q,Ω)],

ψ(l, q,Ω) = 1

2arg[U (q,Ω)V (−q,−Ω)]. (3.447)

In relation (3.446), κ(q,Ω) and r (q,Ω) are two other squeezing parameters,

κ(q,Ω) = 1

2arg[U (q,Ω)U−1(q,Ω)],

exp[±r (q,Ω)] = |U (q,Ω)| ± |V (−q,−Ω)|. (3.448)

Leaving out the integral term in the noise spectrum of the photocurrent density(3.442), one can rewrite it in the form

(δi)2(q,Ω) = 〈i〉{1− η + η[

cos2[θ (q,Ω)] exp[2r (q,Ω)]

+ sin2[θ (q,Ω)] exp[−2r (q,Ω)]]}

. (3.449)

Maximum squeezing occurs at frequencies qm, Ωm, which fulfil the conditionΔ(qm,Ωm) = 0. They are said to belong to such a phase-matching surface. Toreduce shot noise to the highest extent, one chooses a complex amplitude of thelocal oscillator wave with θ (qm,Ωm) = ±π

2 . If the phase-matching condition is notperfectly met, the phase θ (q,Ω) is affected to the first order in Δ(q,Ω)lamp and thesqueezing parameter is not yet influenced to the first order.

In Kolobov (1999), the case of the frequency and angle-degenerate phase match-ing, Δ(0, 0) = 0, is considered. In the case of degenerate phase matching andk ′′Ω < 0 one infers that in the region of frequencies Ω < Ωm and spatial frequenciesq < qm the noise of the photocurrent density operator is reduced below the shot-noise level. In space–time language this can be said as follows. The frequencies Ωm

and qm determine the minimum time Tm and the minimum area of photodetectorSm, which are necessary for reducing fluctuations in the number of photoelectronsbelow the Poissonian limit.

In the case of nondegenerate phase matching in a crystal, when Δ(0, 0) > 0, onemust pay attention to nonzero carrier frequencies q and Ω. In fact, Δ(q,Ω) ≈ 0,when (a) q �= 0,Ω �= 0, (b) q �= 0,Ω = 0, and (c) q = 0,Ω �= 0. Three kinds ofmeasurement can be distinguished based on the three types of phase matching.

Yuen and Shapiro (1979) were the first to propose a degenerate mixing processas a possible source for squeezed light. A four-wave mixer can be set up either ina backward geometry, as proposed by Yuen and Shapiro (1979), or in a forwardgeometry, according to Kumar and Shapiro (1984). A backward four-wave mixercan produce multimode squeezed light with a much larger spatial bandwidth than aforward four-wave mixer and another scheme.

Page 172: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 163

The exposition is restricted to the backward four-wave mixing. The processoccurs in a transparent χ (3) nonlinear medium. It is assumed that the medium hasthe form of a plane slab the thickness of which (distance between two surfacesparallel to the ρ plane) is equal to l. Two counterpropagating plane monochro-matic pump waves E1 and E2 of angular frequency ω0 and wave vectors k1 andk2, respectively, illuminate the slab at a small angle to the z-axis. A quasiplane andquasimonochromatic probe wave of carrier frequency ω0 enters the medium fromthe left and propagates in the +z-direction. In the nonlinear interaction between thetwo pump waves and the probe wave, a phase conjugate wave is generated in themedium that propagates in the opposite direction to the probe wave (Fisher 1983).One describes the probe and conjugate waves by two corresponding slowly varyingoperators εp(z, ρ, t) and εc(z, ρ, t). Let kμ(q,Ω), μ = p, c, mean the wave vectorsof the probe and conjugate waves, respectively. One introduces the Fourier trans-forms of these space–time operators, ˜εμ(z, q,Ω), μ = p, c, as follows:

˜εμ(z, q,Ω) =∫ ∫

εμ(z, ρ, t) exp[i(Ωt − q · ρ)] dt d2ρ. (3.450)

These operators evolve in the nonlinear medium according to the equations

∂z˜εp(z, q,Ω) = −iκ ˜ε

†c(z,−q,−Ω) exp[−iΔ(q,Ω)z], (3.451)

∂z˜εc(z, q,Ω) = iκ ˜ε

†p(z,−q,−Ω) exp[−iΔ(q,Ω)z]. (3.452)

Here κ is a coupling constant proportional to the product of the two pump waveamplitudes and to the nonlinear susceptibility χ (3) of the medium, Δ(q,Ω) is aphase-mismatch function given by

Δ(q,Ω) = kpz(q,Ω)+ kcz(−q,−Ω)− k1z − k2z, (3.453)

with kp,cz(q,Ω) being the projections of the probe and conjugate wave vectors ontothe positive z-direction and k1,2z the corresponding projections for the pump waves.

The solution of equations (3.451) and (3.452) with the boundary conditions˜εp(z = 0, q,Ω) = ˜εp(0, q,Ω) and ˜εc(z = l, q,Ω) = ˜εc(l, q,Ω) may follow theclassical one (Fisher 1983). The input–output transformation, which yet differs fromthe solution by the inclusion of the incoupling and outcoupling beam splitters, ispresented in Kolobov (1999),

aout(q,Ω) ∝ U (q,Ω)ain(q,Ω)+ V (q,Ω)a†in(−q,−Ω), (3.454)

bout(q,Ω) ∝ U (q,Ω)bin(q,Ω)+ V (q,Ω)b†in(−q,−Ω). (3.455)

Notably one obtains two independent processes of multimode squeezing. Any ofthem can be compared with the optical parametric amplifier.

Page 173: Quantum Aspects of Light Propagation

164 3 Macroscopic Theories and Their Applications

The investigation of the difference is concentrated on comparison between thephase-mismatch functions Δ(q,Ω). (Here Δ ≡ ΔOPA, ΔFFWM, ΔBFWM.) This func-tion depends on the spatial frequency in the optical parametric amplification andthe forward four-wave mixing and does not depend on it in the backward four-wave mixing. From these properties one determines “spatial” squeezing bandwidth

q =√

k1l and q = ∞ in the paraxial approximation.

The frequency dependence can be used to filter the probe signal. One is led to theidea of a “realistic” medium and its “quantum nature”. Obviously, a nonlinearity ofthe equations is meant, but also a better quantization desired, which just as we sawat the beginning has not yet been employed.

Another source is based on a cavity, and it still can generate multimode squeezedstates. It is a subthreshold optical parametric oscillator, concretely this scheme in acavity with spherical mirrors (Lugiato and Marzoli 1995). Here we are provided bythe literature with another example, in which situation the paraxial approximationhas been used. Even though here a cavity is investigated, not a travelling wave, thecavity modes still seem to have been determined in the paraxial approximation.

For a cavity-based geometry a more natural language for the description of mul-timode squeezing is that of discrete eigenmodes of the resonator. In the case ofthe cavity with spherical mirrors, such a discrete eigenset is given by the Gauss–Laguerre modes

f pli (r, φ) = f pl (r )×{

cos(lφ) for i = 1,

sin(lφ) for i = 2,(3.456)

f pl (r ) = 2√2δl,0πw2

√p!

(p + l)!

√2r2

w2Ll

p

(2r2

w2

)exp

(− r2

w2

), (3.457)

where w is the waist of the beam and p, l = 0, 1, 2, , . . . are the radial and angularindices, respectively, r =

√x2 + y2 is the radial and φ is the angular variable.

The functions Llp are the Laguerre polynomials. The functions f pli (r, φ) satisfy the

conditions of orthonormality,

∫ 2π

0

∫ ∞

0f pli (r, φ) f p′l ′i ′(r, φ)r dr dφ = δpp′δll ′δi i ′ . (3.458)

The eigenfrequencies of these modes are given by

ωpl = ω00 + (2p + l)ζ, (3.459)

where ω00 is the lowest eigenfrequency of the resonator and the parameter ζ dependson the curvature of mirrors and the distance between them (Yariv 1989).

The source is described by the master equation,

∂tρ = 1

i�[Hint, ρ]+

2∑

i=1

p,l

ˆΛpli ρ, (3.460)

Page 174: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 165

where the term ˆΛpli ρ,

ˆΛpli ρ = γ(

2apli ρa†pli − a†

pli apli ρ − ρa†pli apli

)(3.461)

describes the damping of the mode pli due to cavity decay through the outcouplingmirror with the rate γ . The interaction Hamiltonian Hint is given by Lugiato andMarzoli (1995),

Hint = i�γ

2Ap

∫ 2π

0

∫ ∞

0

[A† 2(r, φ)− A2(r, φ)

]r dr dφ, (3.462)

where Ap is the coupling constant proportional to the nonlinear susceptibility χ (2)

of the medium and the amplitude of the pump wave.Instead of solving the master equation (3.460), we can write a set of indepen-

dent Langevin equations (Walls and Milburn 1994) for the annihilation and creationoperators apli (t) and a†

pli (t) inside the cavity,

apli (t) = −γ [(1+ iΔpl )apli (t)− Apa†pli (t)]+

√2γ cpli (t), (3.463)

where

Δpl = ωpl − ωs

γ, (3.464)

with ωpl and ωs the eigenfrequencies of the eigenmodes of the resonator and thefrequencies of signal photons, respectively. Every mode is damped and the rateconstant for each of the modes is the same. Such a simplification should still beexplained. The method of description seems to be known in quantum optics. Theoperators cpli (t) and c†pli (t) correspond to the operator-valued Langevin forces anddescribe the vacuum fluctuations entering the cavity through the outcoupling mir-rors. These operators obey the commutation relations

[cpli (t), c†pli (t′)] = δpp′δll ′δi i ′δ(t − t ′)1. (3.465)

In relation (3.462), a coupling constant is expressed as the product γ Ap, from whichwe conclude that 0 < Ap < 1. This property says that a subthreshold oscillator isbeing investigated.

Kolobov (1999) refers to Collet and Gardiner (1984) for the input–output rela-tions for the field operators bpli (t) in the wave outgoing from the cavity, cpli (t) ofthe vacuum fluctuations entering it, and apli (t) inside the cavity,

bpli (t) =√

2γ apli (t)− cpli (t). (3.466)

Using the Fourier transform

˜g pli (Ω) =∫ ∞

−∞exp(iΩt)gpli (t) dt, g = a, b, c, (3.467)

Page 175: Quantum Aspects of Light Propagation

166 3 Macroscopic Theories and Their Applications

to equations (3.463), we arrive at the following squeezing transformation betweenthe Fourier transforms of the incoming and outgoing operators:

˜bpli (Ω) = Upl (Ω)c pli (Ω)+ Vpl (Ω)c†pli (−Ω), (3.468)

with the coefficients

Upl (Ω) = [1− iΔpl(−Ω)][1− iΔpl (Ω)]+ A2p

[1+ iΔpl(Ω)][1− iΔpl (−Ω)]− A2p

,

Vpl (Ω) = 2Ap

[1+ iΔpl(Ω)][1− iΔpl (−Ω)]− A2p

, (3.469)

with Δpl (±Ω) = Δpl ∓ Ωγ

, which is a discrete equivalent of the multimode squeez-ing transformation (3.431).

The calculation of the photocurrent noise spectrum is not included, but the fol-lowing relation is presented (Lugiato and Marzoli 1995):

⟨1

2{δi(ρ, t), δi(ρ′, t ′)}+

⟩=∑

p,l

f pl (r ) f pl(r′) cos[l(φ − φ′)]

× 1

∫ ∞

−∞(δi)2

pl (Ω) exp[−iΩ(t − t ′)] dΩ, (3.470)

which is an analogue of relation (3.430). A relation holds

(δi)2pl (Ω) = 〈i(ρ, t)〉

{1+ 4Ap

(1+Δ2pl − A2

p − Ω2)2 + 4Ω2

× [2Ap + Re{exp(−2iϕβ)(1−Δ2pl + A2

p + Ω2 − 2iΔpl )}]},

(3.471)

where ϕβ is the phase of the local oscillator and Ω = Ωγ

is the dimensionless fre-quency and η = 1 (Collett and Walls 1985, Savage and Walls 1987).

(ii) Free propagation and diffraction of multimode squeezed light

With respect to the free propagation and diffraction of multimode squeezed lightit is shown that propagation in free space in general deteriorates the resolving powerof low-noise measurements with squeezed light. A lens allows one to compensatefor this deterioration and even further to improve the resolving power.

We will assume that the plane of photodetection lies at a distance L from theexit plane of the nonlinear crystal and is parallel to it. The slowly varying operators˜a(l, q,Ω) at the exit plane of the crystal and ˜a(l + L , q,Ω) at the photodetectionplane are for the free propagation related as (cf. equation (3.420)) follows:

Page 176: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 167

˜a(l + L , q,Ω) = exp{i[k(0)z (q,Ω)− k0]L} ˜a(l, q,Ω), (3.472)

where k(0)z (q,Ω) is the z-component of the wave vector in free space. Along with

the free propagation one is interested in the dependence of the field operator at theplane of photodetection on that at the input to the nonlinear crystal,

˜a(l + L , q,Ω) = U (q,Ω) ˜a(0, q,Ω)+ V (q,Ω) ˜a†(0,−q,−Ω), (3.473)

with the coefficients

U (q,Ω) = exp{i[k(0)z (q,Ω)− k0]L}U (q,Ω), (3.474)

and the like for V (q,Ω). It is a simple generalization of the above description. Newquantities are provided with a tilde.

Relation

θ (q,Ω) = θ (q,Ω)+ L

2[k(0)

z (q,Ω)+ k(0)z (−q,−Ω)− 2k0]

≈ θ (q,Ω)− q2 L

2k0, (3.475)

where a paraxial and quasimonochromatic approximation is assumed, says that theorientation angle (i.e. phase) of the squeezing changes more rapidly in dependenceon the spatial frequency than on the output from the crystal. The minimum areaSm of low-noise detection is proportional to lamp

k1+ 2Lk0

, where lamp is the amplification

length. The increase is related to the diffraction. The resolving power of the low-noise observation has decreased. The deterioration is reversible. Even the phaseshifts produced during wave propagation inside the nonlinear crystal can be com-pensated for.

For a lens of focal length f , provided that the object plane has the position −2 frelative to the lens and the image plane has the position 2 f relative to the lens, theoptical imaging is represented by the field operators

a(z + 4 f, ρ, t) = exp

[−i

k0ρ2

2 f

]a

(z,−ρ, t − 1

c

[4 f + ρ2

2 f

]c

). (3.476)

From this relation it follows that the noise spectrum (δi)2(q,Ω) has been conserved.From the results of the analyses performed it is natural to choose z = l, i.e. theoutput plane of the crystal.

Concerns with correct quantization call attention to the proposal of imaging someplane inside the crystal onto the detection plane. This imaging is understood as ageneral choice z = l + L , where L is negative. It suffices to choose

L = −lampk0

2kl, (3.477)

Page 177: Quantum Aspects of Light Propagation

168 3 Macroscopic Theories and Their Applications

for the phase θ (q,Ω) in the vicinity of the matching surface to become indepen-dent of spatial frequency. Thus geometrical imaging of the plane inside the crys-tal at the distance L given by (3.477) onto the photodetection plane broadens therange of spatial frequencies at which one has a noise reduction below the shot-noiselevel.

Kolobov (1999) remarks on what follows. An improvement of the frequencybehaviour of the noise spectrum can be achieved by inserting into the light beama slab of a dispersive medium with wave number k(1)(Ω). The length of the slabwill be

L (1) = −lampk ′′Ω

2k(1)Ω

′′ , (3.478)

if k(1)Ω

′′has the opposite sign to k ′′Ω.

In order to assess physical possibilities for low-noise measurements, the photo-electron number collected by a pixel with the area Sd during the time interval Td

is considered as an example. If it holds that Sd ≥ Sc and Td ≥ Tc, the result isindependent of Sd and Td. For high quantum efficiency, η ≈ 1, the statistics ofphotoelectrons is sub-Poissonian, when squeezing is significant. Here we concedethe efficiency of models simplified to several modes: The average number of photonsnecessary for a single low-noise measurement is given by a quantity, which we couldobtain on choosing as “average” model simplified to two modes.

Whereas the coherence time Tc limits the number of images, which can be trans-mitted in a time interval T , the coherence area Sc limits the number of modes onan illuminated spot of area S on the input to the nonlinear crystal, even though astatistical definition of the mode is peculiar. The scheme of homodyne detection isclosely related to holographic measurements.

(iii) Noiseless control of multimode squeezed light

With respect to the noiseless control of multimode squeezed light, the detectionof faint phase objects as proposed by Kolobov and Kumar (1993) is described. Thesub-shot-noise microscopy utilizes a Mach–Zehnder interferometer. The outgoinglight from the two ports of the second beam splitter is detected by two photodetectorarrays. As a natural generalization of the analysis in Caves (1981), the minimumdetectable spatially varying phase change is defined.

The amplitude modulation in space is not advantageous for creation of opticalimages with a regular (sub-Poissonian) photon statistics. One example of nonde-structive modulation in space is an opaque screen with apertures larger than thecoherence area of squeezed light Sc. A number of references for interference mixing,which provides such nondestructive modulation in time are presented. For general-ization and analogy in the case of spatial modulation, see Sokolov (1991a,b).

(iv) Spatially noiseless optical amplification of images

Page 178: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 169

Noiseless amplification has been defined in phase-sensitive amplifiers (see, forexample, Caves (1982)) and it should be extended to the spatial domain. Manyareas of physics would benefit from the possibility of noiseless amplification of faintoptical images. Astronomy and microscopy come to mind. Kolobov (1999) refers toKolobov and Lugiato (1995) for such a proposal. One considers a ring-cavity degen-erate optical parametric amplifier and monochromatic images. In general, spectralbandwidth of images should be within the bandwidth of the cavity employed. Theoptical parametric amplifier is combined with input and output lenses, which per-form the spatial Fourier transformation and broaden a narrow region of transversevectors q, which is an analogue of the band of (temporal) frequencies. The exposi-tion is self-contained.

Let a(ρ, t) and a†(ρ, t) mean the photon annihilation and creation operators inthe object plane P1 and let e(ρ, t) and e†(ρ, t) stand for the photon annihilationand creation operators in the image plane P4. Let bin(ξ , t) and bout(ξ , t) mean thefield operators in the input and the output planes of the optical parametric oscillator,respectively. The operator bin(ξ , t) is expressed through the operator a(ρ, t) in theobject plane by the following transformation performed by the lens L1:

bin(ξ , t) = 1

λ f

∫a(ρ, t) exp

[−i

λ fξ · ρ

]d2ρ, (3.479)

where f is the focal length of the lens and λ is the wavelength of the light. In aparaxial approximation, the slowly varying field operator b(ξ , t) of the cavity modeclosest to resonance with input signal is described by the equation

∂tb(ξ , t)− i

c

2k∇2⊥b(ξ , t) = −(κ + iΔ)b(ξ , t)

+ σ b†(ξ , t)+√

2κ bin(ξ , t). (3.480)

Here κ is the cavity decay constant equal to

κ = cT

2L, (3.481)

where T is the intensity transmission coefficient of the cavity outcoupling mirror, Lis the perimeter of the cavity, and c is the light velocity in a vacuum; the detuningparameter is defined as

Δ = ωc − ωs, (3.482)

where ωc is the longitudinal cavity frequency closest to the frequency ωs of thesignal field. In (3.480) σ is the constant of parametric interaction proportional to thepump amplitude and k is the wave number of the travelling wave inside the cavity,k = 2π

λ.

Page 179: Quantum Aspects of Light Propagation

170 3 Macroscopic Theories and Their Applications

The output field operator bout(ξ , t) is the sum of two waves, one of which isreflected from and another transmitted through the outcoupling mirror of the cavity,

bout(ξ , t) =√

2κ b(ξ , t)− bin(ξ , t). (3.483)

To express the output field operators in terms of the input operators one takes thespatio-temporal Fourier transform of b(ξ , t),

˜b(q,Ω) =∫ ∫

b(ξ , t) exp[i(Ωt − q · ξ )] d2ξ dt. (3.484)

The spatio-temporal Fourier transforms of bin(ξ , t) and bout(ξ , t) are similar.The transformation of the field amplitude from the object plane P1 to the input

plane P2 given by relation (3.479) is equivalent to the following relation betweenthe spatio-temporal Fourier transform bin(q,Ω) and the temporal Fourier transform˜a(ρ,Ω)

˜bin(q,Ω) = λ f ˜a

(−λ f

2πq,Ω

), (3.485)

where we have used

˜a(ρ,Ω) =∫

a(ρ, t) exp(iΩt) dt. (3.486)

Since the lens L2 has the same focal length as L1, we have an identical relationship

between the Fourier transforms ˜bout(q,Ω) and ˜e(ρ,Ω) in the output plane P3 and inthe image plane P4,

˜e(ρ,Ω) = λ f˜bout

(−λ f

2πρ,Ω

), (3.487)

˜e(ρ,Ω) =∫

e(ρ, t) exp(iΩt) dt. (3.488)

It can be derived that

˜e(ρ,Ω) = u(ρ,Ω) ˜a(ρ,Ω)+ v(ρ,Ω) ˜a†(−ρ,−Ω), (3.489)

with u(ρ,Ω) and v(ρ,Ω) given by

u(ρ,Ω) = [1− iδ(ρ,Ω)][1− iδ(ρ,−Ω)]+ |g|2[1+ iδ(ρ,Ω)][1− iδ(ρ,−Ω)]− |g|2 ,

v(ρ,Ω) = 2g

[1+ iδ(ρ,Ω)][1− iδ(ρ,−Ω)]− |g|2 . (3.490)

Page 180: Quantum Aspects of Light Propagation

3.3 Modes of Universe and Paraxial Quantum Propagation 171

Here one has introduced the dimensionless coupling strength g of the parametricinteraction,

g = σ

κ, (3.491)

and the dimensionless mismatch function δ(ρ,Ω),

δ(ρ,Ω) = Δ

κ− Ω

κ+(

ρ

ρ0

)2

, (3.492)

with ρ0 defined as

ρ0 = f

√λT

2π L. (3.493)

To ensure a linear amplification regime, the coupling strength g must be |g| < 1.In Kolobov (1999) it has been shown that multimode squeezed states of light

come about as a natural generalization of single-mode squeezed states. They can beproduced in experiments when just one spatial mode of the field is cut out by meansof a high-Q optical cavity. Travelling-wave configurations are most convenient forthe generation of multimode squeezed states. To observe them, one must employ adense array of photodetectors.

Many new physical phenomena are connected with multimode squeezing. Mul-timode squeezed states offer a few applications including optical imaging withsub-shot-noise sensitivity, sub-shot-noise microscopy, and noiseless amplificationof optical images. There are some other phenomena related to multimode squeez-ing, such as the similarity of homodyne detection of multimode squeezed states tothe scheme of optical holography, an application of these states to optical imagerecognition with photon-limited images (Morris 1989), a possibility to improve aquantum limit in optical resolution with the use of nonclassical light (den Dekkerand van den Bos 1997). Multimode squeezed states can be applied in the field ofoptical pattern formation, which studies the spatial and spatio-temporal phenomenathat arise in the structure of the electromagnetic field in the plane orthogonal to thedirection of propagation. For instance, the filamentation of a laser beam initiated byquantum fluctuations of light in its transverse area (Nagasako et al. 1997, Lugiatoet al. 1999).

Bjork et al. (2004) have shown that the use of entangled photon pairs in an imag-ing system can be simulated with a classically correlated source sometimes. Theyhave considered two schemes with “bucket detection” of one of the photons. In con-trast, entangled two-photon imaging may exhibit effects that cannot be mimickedby any classical source when bucket detection is not used (Strekalov et al. 1995).

Caetano and Souto Ribeiro (2004) have investigated theoretically and experimen-tally the transfer of the angular spectrum of the pump beam to the down-convertedbeams. They have demonstrated that the image of a given object placed in the pump

Page 181: Quantum Aspects of Light Propagation

172 3 Macroscopic Theories and Their Applications

can be formed in the twin beams by manipulating the entangled angular spectrumand performing coincidence detection.

Gatti et al. (2004) have analytically shown that it is possible to perform coher-ent imaging by using the classical correlation of two beams obtained by splittingthermal light. They have presented a formal analogy between two such classicallycorrelated beams and two entangled beams produced by parametric down conver-sion. The classical beams can qualitatively reproduce all the imaging properties ofthe entangled beams. These classical beams are spatially correlated both in the nearfield and in the far field even though to an imperfect degree.

Bache et al. (2004) have presented a theoretical study of ghost imaging whichuses balanced homodyne detection to measure signal and idler fields arising fromparametric down conversion. They have used a general model describing the three-wave quantum interaction with respect to finite size and duration of the pump pulse.They have shown that the signal–idler correlations contain the full amplitude andphase information about an object located in the signal arm, both in the near-field(object image) and the far-field (object diffraction pattern) cases. One may passfrom the far-field result to the near-field result by simply performing inverse Fouriertransformation. The analytical results are confirmed by numerical simulations.

Bennink et al. (2004) have reported two distinct experimental demonstrationsof coincidence imaging. They have shown that uncertainties of distance and meandirection of two classical fields must obey an inequality. With the use of entangledphotons they formed two images whose resolution had a product that was threetimes better than is possible according to classical diffraction theory. For the sake ofcomparison, a similar experiment was performed with light in a classical mixture ofstates (cf. Gatti et al. 2003). While the resolution of the image was good in the farfield, the uncertainty product obeyed the classical inequality in the near field.

Valencia et al. (2005) presented the first experimental demonstration of two-photon ghost imaging with a pseudothermal source. They have introduced the con-cepts of two-photon coherent and two-photon incoherent imaging. Similar to thecase of entangled states, a two-photon Gaussian thin lens equation connects theobject plane and the image plane. Specifically, the thermal source acts as a phaseconjugated mirror.

Altman et al. (2005) have probed the quantum image produced by parametricdown-conversion with a pump beam carrying orbital angular momentum. With onedetector fixed and the other scanning, the usual single-spot coincidence pattern ispredicted (Monken et al. 1998) to split into two spots, which has been demonstrated.

Mosset et al. (2005) have presented the first experimental demonstration of noise-less amplification of images that yielded spatially integrated intensity (of the pho-todetection process) for different lateral detector sizes. Achieving two-beam andsingle-beam conditions, they have compared phase-insensitive and phase-sensitiveschemes with theory.

Quantum imaging is a branch of quantum optics that investigates the ultimateperformance limits of optical imaging imposed by quantum mechanics. The use ofquantum-optical methods enables one to solve the problems of image formation,processing and detection with sensitivity and resolution which exceeds the limits

Page 182: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 173

of classical imaging. The most important theoretical and experimental results inquantum imaging can be found in Kolobov (2007).

3.4 Optical Nonlinearity and Renormalization

Abram and Cohen (1994) mainly applied a travelling-wave formulation of the theoryof quantum optics to the description of the self-phase modulation of a short coherentpulse of light. They seem to have been first to use a renormalization (Kubo 1962,Zinn-Justin 1989). The renormalized theory successfully describes the nonlinearchirp that the pulse undergoes in the course of its propagation and permits the cal-culation of the squeezing characteristics of self-phase modulation. The descriptionof the propagation of a short coherent pulse of light inside a medium that exhibitsan intensity-dependent refractive index (Kerr effect) has become relevant to opti-cal fibre communications, all-optical switching, and optical logic gates (Agrawal1989). Neglect of dispersion and the Raman and Brillouin scattering leads to thedescription of self-phase modulation. In classical theory it is derived that, in thecourse of its propagation, the pulse becomes chirped (i.e. different parts of the pulseacquire different central frequencies), which influences also its spectrum. Abramand Cohen (1994) have pointed out many difficulties in the investigation of thequantum noise properties of a light pulse undergoing self-phase modulation. Thetraditional cavity-based formalism truncates the mutual interaction among the spa-tial modes to a self-coupling of a single mode (or only a few modes) and cannotgive a reasonable approximation to the frequency spectrum produced by self-phasemodulation. In spite of the difficulties, papers based on a single-mode description ofa field indicated that the slowly varying approximation can produce squeezed light(Kitagawa and Yamamoto 1986, Shirasaki et al. 1989, Shirasaki and Haus 1990,Wright 1990, Blow et al. 1991) and others treated squeezing in solitons (Drummondand Carter 1987, Shelby et al. 1990, Lai and Haus 1989), an effect that was verifiedexperimentally (Rosenbluh and Shelby 1991). Blow et al. (1991) have shown thedivergence of nonlinear phase shift, which Abram and Cohen (1994) treat throughthe process of renormalization.

Let us review the basic features of the quantization of the electromagnetic fieldin a Kerr medium and discuss the relevance of the renormalization procedure to thetreatment of divergences of effective medium theories. We consider a transparent,homogeneous isotropic, and dispersionless dielectric medium that exhibits a nonlin-ear refractive index. We examine the situation similar to Abram and Cohen (1991).The Hamiltonian for the electromagnetic field in a Kerr medium is

H (t) =∫ {

1

2

[B2(z, t)+ ε E2(z, t)

]+ 3

4χ E4(z, t)

}dz, (3.494)

where the integration along the direction of propagation z is denoted explicitly, butintegration over the transversed directions x and y will be implicit. We use theHeaviside–Lorentz units for the electromagnetic field without � = c = 1, χ is the

Page 183: Quantum Aspects of Light Propagation

174 3 Macroscopic Theories and Their Applications

nonlinear (third-order) optical susceptibility. From the perspective of the substitu-tion of (3.25), the Hamiltonian (3.494) can be written as (Hillery and Mlodinow1984)

H (t) =∫ {

1

2

[B2(z, t)+ 1

εD2(z, t)

]− 1

4

χ

ε4D4(z, t)

}dz, (3.495)

where the displacement field

D(z, t) = ε E(z, t)+ χ E3(z, t). (3.496)

The canonical equal-time commutators are (cf. (3.28))

[ A(z1, t), D(z2, t)] = −i�cδ(z1 − z2)1, (3.497)

[B(z1, t), D(z2, t)] = −i�cδ′(z1 − z2)1. (3.498)

It is convenient to adopt a slowly varying operator picture in which the zeroth-order dynamics of the field governed by the linear medium Hamiltonian are alreadytaken into account exactly, while the optical nonlinearity can be treated within theframework of perturbation theory. In such a picture, a field operator Q(z, t) evolvinginside a nonlinear medium is related to the corresponding linear medium operatorQ0(z, t) by

Q(z, t) = U−1(t)Q0(z, t)U (t). (3.499)

The unitary transformation U (t) is given by

U (t) = T exp

[− i

∫ t

−∞H1(τ ) dτ

], (3.500)

where T ≡←−T denotes the time ordering and

H1(t) = −1

4

χ

ε4

∫D4

0(z, t) dz (3.501)

is the interaction Hamiltonian. The full nonlinear Hamiltonian (3.495) can be writ-ten as follows:

H (D0, B0) =H0(t)︷ ︸︸ ︷

H0(D0, B0)+H1(t), (3.502)

where H0(t) is the linear medium Hamiltonian. Following the traditional modalapproach to relation (3.499), Kitagawa and Yamamoto (1986) developed a single-mode treatment of the self-phase modulation. Clearly, such an investigation is valid

Page 184: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 175

only inside an optical cavity with a sparse mode structure. In this situation, the timeevolution cannot be interpreted as space progression. In developing a travelling-wave theory for self-phase modulation, Blow et al. (1991) obtained a solution simi-lar to the single-mode solution. They encountered a nonintegrable singularity uponnormal ordering, a thing what is termed in quantum field theory as “ultraviolet”divergence. To avoid an infinite nonlinear phase shift due to the Kerr interaction sosimply described, Blow et al. (1991, 1992) introduced a finite response time for thenonlinear medium as regularization in the Heisenberg picture. Alternatively, Hausand Kartner (1992) considered the group velocity dispersion for the propagation ofpulses in the medium as a regularization.

At any rate, the regularization is a subsequent sophistication of the simple modelas known in classical theory. The response time and the group velocity dispersionare necessary ingredients of a complete description of the propagation of the elec-tromagnetic excitation in fibres. But they have no influence on the effects associatedwith the vacuum fluctuations under study. A systematic way of dealing with thevacuum fluctuations in quantum field theory is the procedure of renormalization(Itzykson and Zuber 1980, Zinn-Justin 1989). The renormalization is known alsoin classical field theory. In order to obtain finite results, the procedure of renormal-ization redefines all the quantities that enter the Hamiltonian. The renormalizationpoint of view is that the new Hamiltonian is the only one we have access to. Itcontains the observable consequences of the theory and the parameters are the oneswe obtain from experiments. The bare quantities are only auxiliary parameters thatshould be eliminated exactly from the description (Stenholm 2000). The re-defined(renormalized) quantities are able to incorporate the (infinite) effects of the vacuumfluctuations. We will provide the definitions of broad-band electromagnetic fieldoperators and treat the propagation of light in a linear medium. The normal orderingis considered as the simplest renormalization, e.g. in the case of the effective linearHamiltonian

H0(t) = 1

2

∫ [B2

0 (z, t)+ 1

εD2

0(z, t)

]dz. (3.503)

To this end, the Hamiltonian H0(t) is to be written in terms of the creation andannihilation operators. The normal ordering allows us to subtract the vacuum-fieldenergy up to the first order from the effective Kerr Hamiltonian (3.495). However,when this Hamiltonian is used to describe propagation disregarding the richness ofthe quantum field theory, the normal ordering gives rise to additional divergencesthat can be attributed to the participation of the vacuum fields. Upon renormaliza-tion, involving also the refractive index, the divergences are removed.

As the Kerr nonlinearity involves the fourth power of the derivative ∂∂t A(z, t), it

cannot be in the general case renormalized to all orders with a finite number of cor-rections. Inspired by nonlinear optics, the slowly varying amplitude approximationdecouples counterpropagating waves and the renormalization to all orders becomespossible. All types of optical nonlinearity χ (n) give rise to divergences which requirethe renormalization. In the treatment of parametric down conversion (Abram and

Page 185: Quantum Aspects of Light Propagation

176 3 Macroscopic Theories and Their Applications

Cohen 1991), the problem of divergences and the need of normalization were notformulated.

In Abram and Cohen (1994), the broad-band electromagnetic field operators aredefined and the propagation of light in a nonlinear medium is treated. In the absenceof the optical nonlinearity, χ = 0 and the linear-medium displacement field has theusual proportionality relationship to the electric field,

D0(z, t) = ε E0(z, t). (3.504)

The magnetic field and the displacement field in the linear medium obey the equal-time commutation relation

[B0(z1, t), D0(z2, t)] = −i�cδ′(z1 − z2)1. (3.505)

The operators V±0 (cf. (3.80), (3.81) of Abram and Cohen (1991)) reappear as the

operators

ψ+(z, t) = 1√2ε

V+0 (z, t), (3.506)

ψ−(z, t) = 1√2ε

V−0 (z, t). (3.507)

For ψ±(z, t), the equations of motion in the Heisenberg picture may be calculatedby the use of commutator (3.505) as follows:

∂tψ±(z, t) = i

[H0, ψ±(z, t)

] = ∓v∂

∂zψ±(z, t), (3.508)

where v = c√ε

is the speed of light inside the dielectric exhibiting the refractiveindex ε. Their solutions are

ψ+(z, t) = ψ+(z − vt, 0), (3.509)

ψ−(z, t) = ψ−(z + vt, 0). (3.510)

The equal-time commutators of the copropagating field operators can be obtainedfrom the definition and commutator (3.505) as follows:

[ψ+(z1, t), ψ+(z2, t)] = −i�vδ′(z1 − z2)1, (3.511)

[ψ−(z1, t), ψ−(z2, t)] = i�vδ′(z1 − z2)1. (3.512)

For the counter propogating fields, the corresponding operators commute with eachother,

[ψ+(z1, t), ψ−(z2, t)] = 0. (3.513)

Page 186: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 177

Operators (3.506) and (3.507) permit us to express the linear medium Hamiltonian(3.503) as

H0(t) = 1

2

∫ [ψ2+(z, t)+ ψ2

−(z, t)]

dz, (3.514)

thus separating it into a sum of two mutually commuting partial operators, one foreach direction of propagation.

In the homogeneous medium it is possible to separate the electromagnetic fieldoperators ψ±(z, t) into positive- and negative-frequency parts

ψ±(z, t) = φ±(z, t)+ φ†±(z, t), (3.515)

defined as

φ±(z, t) = 1

2

[ψ±(z, t)± i

πV.p.

∫ ∞

−∞

ψ±(z′, t)

z − z′dz′]

, (3.516)

where V.p. denotes the Cauchy principal value. The operators φ†±(z, t) and φ±(z, t)

can be considered as creation and annihilation operators, respectively, for a right (orleft)-moving electromagnetic excitation which at time t is at point z. The equal-timecommutators of φ±(z, t) are somewhat complicated,

[φ±(z1, t), φ†±(z2, t)] = �v

2

∂z1

[P 1

z1 − z2∓ iπδ(z1 − z2)

]1,

[φ+(z1, t), φ†−(z2, t)] = 0, (3.517)

where P refers to the familiar generalized function P 1z . Nevertheless, an impor-

tant simplification results when only unidirectional propagation is considered. Onintroducing the operators

D(+)0 (z, t) =

√ε

2[φ+(z, t)+ φ−(z, t)], D(−)

0 (z, t) = [D(+)0 (z, t)]†, (3.518)

B(+)0 (z, t) = 1√

2[φ+(z, t)− φ−(z, t)], B(−)

0 (z, t) = [B(+)0 (z, t)]† (3.519)

and considering the relations

D0(z, t) =√

ε

2[ψ+(z, t)+ ψ−(z, t)],

B0(z, t) = 1√2

[ψ+(z, t)− ψ−(z, t)] (3.520)

and relation (3.515), we verify that

D0(z, t) = D(+)0 (z, t)+ D(−)

0 (z, t),

B0(z, t) = B(+)0 (z, t)+ B(−)

0 (z, t). (3.521)

Page 187: Quantum Aspects of Light Propagation

178 3 Macroscopic Theories and Their Applications

Using the new operators, the equal-time commutation relation (3.505) can suit-ably be modified as

[B(+)0 (z1, t), D(−)

0 (z2, t)] = − i

2�cδ′(z1 − z2)1. (3.522)

For a right-moving electromagnetic excitation, we observe that

φ+(+)(z1, t) =√

2B(+)0(+)(z1, t),

φ+(+)(z2, t) =√

2

εD(+)

0(+)(z2, t), (3.523)

where the subscript (+) refers to k > 0. Using (3.523), we obtain that

[φ+(z1, t), φ†+(z2, t)](+) = −i�vδ′(z1 − z2)1(+). (3.524)

Similarly, for a left-moving electromagnetic excitation, we note that

φ−(−)(z1, t) = −√

2B(+)0(−)(z1, t),

φ−(−)(z2, t) =√

2

εD(+)

0(−)(z2, t), (3.525)

where the subscript (−) refers to k < 0. From this

[φ−(z1, t), φ†−(z2, t)](−) = i�vδ′(z1 − z2)1(−). (3.526)

The electromagnetic creation and annihilation operators allow us to speak of thenormal order, for instance, when we write Hamiltonian (3.514) in the form

H0(t) =∫ [

φ†+(z, t)φ+(z, t)+ φ

†−(z, t)φ−(z, t)

]dz. (3.527)

We can define annihilation and creation wave-packet photon operators

F+(z, t) =∫

F(z − z)φ+(z, t) dz (3.528)

and

F†+(z, t) =

∫F∗(z − z)φ†

+(z, t) dz, (3.529)

respectively. Here F(z) is a complex function:

F(z) = 1√vK

exp(iK z)F(z), (3.530)

where vK is the central (carrier) frequency and F(z) is the wave-packet envelopefunction peaked at z = 0 and k = 0.

Page 188: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 179

On the usual assumption of narrow bandwidth and

i�v

∫F ′(z)F∗(z) dz = 1, (3.531)

where F ′ denotes the spatial derivative, we obtain that the operators F+ and F†+

follow the boson commutation relation[

F+(z, t), F†+(z, t)

]= 1. (3.532)

Let us remark that the commutation relation (3.532) is relation (A5) in Milburnet al. (1984), where the formalism of the counterdirectional coupling was derived orrather this pitfall underestimated.

Now a coherent pulse can be considered whose shape is described by ρF(z) witha scaling factor ρ. A coherent state appropriate to ρF is defined as

|ρF〉 = exp[ρ(

F †+ − F+

)]|0〉. (3.533)

It satisfies the “single-mode” eigenvalue equation

F+(z, t)|ρF〉 = ρ|ρF〉 (3.534)

and, at the same time, it obeys the approximate quantum field eigenvalue equation

φ+(z, t)|ρF〉 = �

√vKρ F∗(z − z) exp[−iK (z − z)]|ρF〉. (3.535)

The approximation made in the derivation of (3.535) has kindled the interest in theGlauber factorization conditions and the theory of coherence (see also Ledinegg(1966)).

When we examine right-moving pulses, we can introduce moving-frame coordi-nate

η = z − vt (3.536)

and simplify relation (3.509),

φ+(z, t) = φ(η, 0) ≡ φ(η), (3.537)

dropping the subscript+, whenever we use the coordinate η explicitly. Similarly, thecommutation relation (3.524) can be modified. The right-moving narrow-bandwidthwavepacket operators (3.528), (3.529) can now be written as

F(η) =∫

F(η − η)φ(η) dη, (3.538)

Page 189: Quantum Aspects of Light Propagation

180 3 Macroscopic Theories and Their Applications

where η = z − vt and

F†(η) =∫

F∗(η − η)φ†(η) dη. (3.539)

In the moving-frame representation F(η, t) = F(η, 0).It is feasible to find a connection with the approaches leading to narrow-band

field operators (a) contained in Shirasaki and Haus (1990), Drummond (1990), andBlow et al. (1990) and used in papers by Blow et al. (1991) and Shirasaki andHaus (1990). An important feature of these operators is that their commutator isa δ function

[ak0 (z1), a†k0

(z2)] = �vk0δ(z1 − z2)1. (3.540)

Under the same narrow-bandwidth condition, the commutator of the Abram–Cohenoperators, which is a delta function derivative, can be approximated by

δ′(z1 − z2) ≈ −ik0δ(z1 − z2). (3.541)

Abram and Cohen (1994) have analysed the approximations that enter the quan-tum treatment of propagation in a Kerr medium and outline the corresponding renor-malization procedure. The slowly varying amplitude approximation according toAbram and Cohen (1991) is used in Abram and Cohen (1994). For a Kerr medium,the interaction Hamiltonian H1(t) is expressed as

H1(t) = − χ

4ε4

∫D4

0(z, t) dz. (3.542)

According to (3.520), the interaction Hamiltonian can be written as

H1(t) = χ

16ε2

∫ [ψ+(z − vt)+ ψ−(z + vt)

]4dz. (3.543)

The exact Hamiltonian (3.495) may be written up to the first order in χ as

H (t) = H0(t)+ H1S+(t)+ H1S−(t)+ O(χ2), (3.544)

where

H1S±(t) = − χ

16ε2

∫ψ4±(z ∓ vt) dz (3.545)

are the parts of the Hamiltonian (3.543) that commute with H0(t).In terms of application of (3.537), no coordinate transformation has been

explained. In this case, the transformation leaves the time coordinate unchanged.

Page 190: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 181

In view of approximation (3.544), the equation of motion of a right-moving fieldoperator can be written in the interaction picture as follows:

∂tφ(η, t) = i

[H1S+(t), φ(η, t)

]. (3.546)

This first-order approximation to the equation of motion can be solved formallyusing the corresponding time-evolution operator (cf. (3.499))

US+(t) = T exp

[− i

∫ t

−∞H1S+(τ ) dτ

]. (3.547)

The classical slowly varying approximation has its quantum counterpart on a doubleassumption: (1) the initial state of the field is a narrow-bandwidth state and (2) thenonlinearity is weak enough so that the full nonlinear Hamiltonian (3.543) may beapproximated by its first-order stationary component H1S,

H1S = H1S+ + H1S−. (3.548)

Therefore, H1S+ will be referred to as the slowly varying amplitude Hamiltonian.Now we turn to the renormalization. In the framework of the rotating-wave

approximation, we obtain that

H1S+ = − χ

16ε2

∫6(φ†φ†φφ

)S dz, (3.549)

where

6(φ†φ†φφ

)S = φ†φ†φφ + φ†φφ†φ + φ†φφφ† + φφ†φ†φ + φφ†φφ† + φφφ†φ†.

(3.550)Upon the normal ordering, the perturbative Hamiltonian (3.544) for the electromag-netic field in a Kerr medium can be written as

H (t) =∫

φ†(z, t)φ(z, t) dz − κ

2

∫φ†(z, t)φ†(z, t)φ(z, t)φ(z, t) dz

− �κ Z∫

φ†(z, t)φ(z, t) dz, (3.551)

where κ = 3χ

4ε2 . Z is a function we give only asymptotically (cf. Abram and Cohen1994)

Z �Λ→∞

v

2πΛ2, (3.552)

Page 191: Quantum Aspects of Light Propagation

182 3 Macroscopic Theories and Their Applications

where Λ is a high-frequency cutoff. Whereas the first two terms in equation (3.551)are familiar, the third term, which is divergent, arises in the normal ordering proce-dure. For Λ fixed, this last term vanishes if �→ 0.

In the renormalization procedure, a formal series (in �) of “counterterms” isadded to the Hamiltonian in order to remove the divergences that arise upon nor-mally ordering the results of calculation (Itzykson and Zuber 1980). The Hamilto-nian itself exemplifies that it is not sufficient for removing divergences, but at thesame time renormalized parameters and renormalized field operators are introduced.In particular, a renormalized Kerr Hamiltonian HR(t) may be defined by introducinga counterterm of order � as follows:

HR(t) = H (t)+ 2�κ Z∫

φ†(z, t)φ(z, t) dz. (3.553)

The third term in equation (3.551) exchanges the sign and for Λ→∞ it is an infinitechange in the inverse of the refractive index. The renormalized field operators

φR(z, t) = √1+ �κ Z φ(z, t), (3.554)

φ†R(z, t) = √

1+ �κ Z φ†(z, t) (3.555)

are further quantities which, or at least whose Hermitian parts, etc. would relate toan experiment. Such a relationship is no more required from the bare quantities. Atthe same time, a renormalized refractive index is defined

nR = n

1+ �κ Z. (3.556)

The renormalized Kerr Hamiltonian can be written in terms of the renormalizedfield operators as

HR(t) = H0R(t)+ H1S+,R(t), (3.557)

with

H0R(t) =∫

φ†R(z, t)φR(z, t) dz, (3.558)

H1S+,R(t) =∞∑

j=0

�jκ j+1 H ( j)

1S+,R(t), (3.559)

where

H ( j)1S+,R(t) = (−1) j+1

2( j + 1)Z j

∫φ†R(z, t)φ†

R(z, t)φR(z, t)φR(z, t) dz; (3.560)

Page 192: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 183

here κ H (0)1S+,R(t) is the “usual” Kerr term and �

jκ j+1 H ( j)1S+,R(t) is the j th quantum

correction.Abram and Cohen (1994) have calculated the quantum noise properties of a

coherent pulse undergoing self-phase modulation in the course of its propagationby eliminating the vacuum divergences through the renormalization procedure. Theone-point averages were first determined. The detection of a light pulse by a bal-anced homodyne detector can be expressed in a moving frame through the measuredquantum operator

Mθ (η) = exp(iθ )√

KLO FLO(η − η)φ(η)+ H.c., (3.561)

where FLO(η) is the coherent amplitude of the local oscillator (LO) pulse peaked atη = η and θ is the phase difference between the local oscillator and signal pulses. Inhomodyne detection, the central frequency of the local oscillator is the same as thatof the incident pulse, KLO = K . This set-up measures simultaneously the expecta-tion values of the operators M0(η) and M π

2(η). For simplicity, we have neglected

the Kerr medium here, but it is included when we write φ(η; z) = φ(η, t = zv)

instead of φ(η). The instantaneous intensity of the signal pulse peaked at η = 0 isintroduced as

I (η) = vKρ2 F∗(η)F(η). (3.562)

It can be obtained that the operator φ(η; z) has the expectation value in the coherentstate

F∗(η; z) = 〈ρF |φ(η; z)|ρF〉= vKρF∗(η) exp[−iΘ(η; z)], (3.563)

where

Θ(η; z) = κvK zI (η) (3.564)

is the nonlinear phase shift produced by self-phase modulation. The nonlinear phaseΘ has exactly the same value as in classical nonlinear optics. The fact that equa-tion (3.563) obtained by invoking renormalization corresponds directly to what isobserved experimentally in a propagative configuration underlines the validity ofthis approach.

Two-point correlation functions were examined, in order to obtain the quantumnoise spectrum

S0(k) =∫ ∫

exp[ik(η1 − η2)]〈ρF |ΔMθ (η1)ΔMθ (η2)|ρF〉 dη2 dη1, (3.565)

where

ΔMθ (η) = Mθ (η)− 〈ρF |Mθ (η)|ρF〉 (3.566)

Page 193: Quantum Aspects of Light Propagation

184 3 Macroscopic Theories and Their Applications

and k is the spatial frequency at which the quantum noise is measured. In exper-iment, such a noise is detected at frequencies several orders of magnitude belowthe carrier optical frequency and its spectrum is considered to be flat throughoutthe typical bandwidth. The low-frequency noise spectrum Sθ (0) can be decomposedinto four terms

Sθ (k) ≈ Sθ (0) = S1 + S2 + exp(−2iθ)S3 + exp(2iθ )S4, (3.567)

with

S1 =∫

ILO(η − η)Θ2(η; z) dη, (3.568)

S2 =∫

ILO(η − η)[1+Θ2(η; z)] dη, (3.569)

S3 =∫

ILO(η − η)[iΘ(η; z)][1+ iΘ(η; z)] exp[2iΘ(η; z)] dη, (3.570)

S4 =∫

ILO(η − η)[−iΘ(η; z)][1− iΘ(η; z)] exp[−2iΘ(η; z)] dη, (3.571)

where ILO(η−η) is the local oscillator instantaneous intensity of the local oscillatorpulse. This result is similar to that obtained by linearizing the self-phase modula-tion exponential operator exp(iγ a†

k0ak0) around the mean field (Shirasaki and Haus1990).

It is appropriate to give a physical interpretation of the above results and todiscuss the case of squeezing that can be observed in the propagation of a coher-ent pulse. For narrow-bandwidth signals, the phase properties of quantum noise inequation (3.567) can be visualized by examining the field fluctuations. First, thequantum characteristic function is defined as

C(u, v) = 〈α| exp

{i∫

[uM π2(η)+ vM0(η)]

}|α〉, (3.572)

where |α〉 is the continuous-wave coherent state. Using the linked cluster theorem,the characteristic function (3.572) can be written in terms of connected averages.Two lowest order connected averages are feasible and describe the moments of theWigner distribution

W (q, p) = 1

4π2

∫ ∫C(u, v) exp[−i(pu + qv)] du dv. (3.573)

According to equation (3.563), the expectation value of the Wigner distribution is

q0 + ip0 =√

I eiθ , (3.574)

Page 194: Quantum Aspects of Light Propagation

3.4 Optical Nonlinearity and Renormalization 185

while the principal squeeze variances of the quantum noise are

2(BS ∓ |C |) = 1√1+Θ2 ±Θ

(3.575)

=√

1+Θ2 ∓Θ, (3.576)

where Θ ≡ Θ(η; z) and we have used the characteristics of quantum noise (Perinovaet al. 1991)

BS = 1

2

√1+Θ2, |C | = 1

2Θ. (3.577)

Moreover,

arg C = 2Θ+ arctan(Θ)− 3π

2. (3.578)

In the Abram–Cohen theory, for the case of a coherent beam of central frequency2×1015 Hz, bandwidth 100 GHz, and intensity 1 W propagating in a silica fibre, theGaussian noise is a good approximation up to nonlinear phase shifts of the order 103

rad.When the phase of the local oscillator is constant along the pulse profile, particu-

larly when the principal quadrature is measured at the peak of the pulse, the variancewill not be the same everywhere, the quadrature will not always be the principalquadrature due to the chirp. To circumvent this problem, the use of a matched localoscillator has been proposed such that its phase Θ(η) varies in a way that matchesthe signal chirp. Besides the Kerr effect, the Sagnac interferometer can be used forthis purpose (Shirasaki and Haus 1990, Blow et al. 1992, Bergman and Haus 1991).Alternatively, a local oscillator pulse that is much shorter than the signal pulse cansample only the central portion of the signal in order to measure the appropriatesqueezing.

Bespalov et al. (2002) have investigated the propagation of light in the (1 + 1)-dimensional approximation. They have paid attention to the two series expansionsof the index of refraction of an isotropic optical medium in Born and Wolf (1968).On these expansions they have based two wave equations, both with and withoutthe second space derivative term. They have presented a method to derive the non-linear wave equations suitable for describing dynamics of extremely short pulses.Although this analysis is completely classical, it does not exclude that a quantizationof the field and of the equations will be necessary in the near future.

Restriction to the transverse components and, finally, to the scalar wave equationis common. Lu et al. (2003) have studied the propagation of ultrashort pulsed beambeyond the paraxial approximation in free space. The nonparaxial corrections to anarbitrary paraxial solution are given in a series form. A comparison with rigorousnonparaxial results obtained by numerical method is carried out. Spatial and tempo-ral distributions are considered.

Page 195: Quantum Aspects of Light Propagation

186 3 Macroscopic Theories and Their Applications

In this book the “fully” relativistic quantum electrodynamics is not treated. Nev-ertheless, its importance for quantum optics is going to be appreciated in the nearestfuture. Shukla et al. (2004) have considered the nonlinear propagation of randomlydistributed intense short photon pulses in a photon gas. Fragmentation of incoherentphoton pulses in astrophysical contexts and in forthcoming experiments using veryintense short laser pulses has been predicted.

The renormalization and the Bogoliubov renormalization group are different con-cepts (Shirkov and Kovalev 2001). Kovalev et al. (2000) and Tatarinova and Garcia(2007) have expounded the renormalization-group approach to the problem of light-beam self-focusing.

Tatarinova and Garcia (2007) set the problem in the framework of the classicalnonlinear optics and so the renormalization is not needed. The propagation of alaser beam of intensity I in a nonlinear medium with a refractive index n0 + n(I ),where n0 is the linear refractive index, n(I ) is such that n(0) = 0, arbitrary in otherrespects, is studied. The case of nonlinear self-focusing accompanied by multipho-ton ionization has been explicitly analysed. The procedure of analytical solutionbegins with an approximate transformation of the nonlinear Schrodinger equationonto eikonal equations. Irrespective of these and some other approximations, theappropriate, easy, calculation provides results which are in good agreement withnumerical simulations.

3.5 Quasimode Theory

Glauber and Lewenstein (1991) have developed quantum optics of inhomogeneousmedia with linear susceptibilities. The topics treated have included the normal-modeexpansion and the plane-wave expansion. The authors have shown that plane-wavephotons can be related to the normal-mode ones within the framework of scatter-ing theory. They have used the quantization schemes discussed to determine thefluctuation properties of various field components. They have considered excitedatoms and changes in the spontaneous emission rates for both electric and magneticdipole transitions of the atoms within or near dielectric media. They have provideda quantum description of the transition radiation emitted by a charged particle inpassing from one dielectric medium to another.

Dalton et al. (1996) have carried out canonical quantization of the electromag-netic field and radiative atoms in passive, lossless, dispersionless, and linear dielec-tric media. The quantum Hamiltonian has been derived in a generalized multipoleform. Dalton et al. (1999b) have presented a macroscopic canonical quantizationof the electromagnetic field and radiating atom system in dielectric media basedon expanding the vector potential in terms of quasimode functions. The quasimodefunctions approximate the true mode functions of a classical optics device whenthey are obtained on the assumption of an ideal electric permittivity function andthe permittivity function describing the device does not deviate much from the idealone. Plane waves in Glauber and Lewenstein (1991) are such quasimodes. In the

Page 196: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 187

coupled-mode theory (see, e.g. Section 6.2) the “ideal” waveguide modes are alsosuch quasimodes.

Here we present part of the theory (Dalton et al. 1999b), which will be completedbelow. It is assumed that a classical linear optics device is described with the spa-tially dependent electric permittivity ε(R) (and the magnetic permeability μ(R)).The generalized Coulomb gauge condition for the vector potential A(R, t)

∇ · [ε(R)A(R, t)] = 0 (3.579)

is used. In a generalization of Helmholtz’s theorem (Dalton and Babiker 1997), avector field F(R, t) can be decomposed uniquely in the generalized transverse andlongitudinal components F(ε)

⊥ (R, t), F(ε)‖ (R, t) in the form

F(R, t) = F(ε)⊥ (R, t)+ F(ε)

‖ (R, t), (3.580)

with

∇ · [ε(R)F(ε)⊥ (R, t)] = 0, ∇ × F(ε)

‖ (R, t) = 0. (3.581)

The macroscopic Lagrangian is given by the relation

L ′(t) =∫

L′c(R, t) d3R, (3.582)

where the Lagrangian density L′c(R, t) is given by the relation

L′c(R, t) = 1

2ε(R)

[∂A(R, t)

∂t

]2

− 1

2μ(R)[∇ × A(R, t)]2. (3.583)

The conjugate momentum field Π(R, t) is obtained from the Lagrangian densityL′c(R, t) as

Π(R, t) = ε(R)∂A(R, t)

∂t. (3.584)

The Hamiltonian is

H ′(t) =∫ {

[Π(R, t)]2

2ε(R)+ [∇ × A(R, t)]2

2μ(R)

}d3R. (3.585)

We have the electric displacement field D(R, t),

D(R, t) = −Π(R, t). (3.586)

The true mode functions satisfy the generalized Helmholtz equation

∇ × 1

μ(R)[∇ × Ak(R)] = ω2

kε(R)Ak(R), (3.587)

Page 197: Quantum Aspects of Light Propagation

188 3 Macroscopic Theories and Their Applications

where ωk are real and positive angular frequencies and satisfy the generalizedCoulomb gauge condition

∇ · [ε(R)Ak(R)] = 0. (3.588)

The true mode functions satisfy the orthogonality and normalization conditionsrespecting ε(R) as a weight function,

∫ε(R)A∗

k (R) · Al (R) d3R = δkl . (3.589)

Generalized coordinates qk(t) and generalized momenta pk(t) can be introduced bythe relations

qk(t) =∫

ε(R)A∗k (R) · A(R, t) d3R, (3.590)

pk(t) =∫

A∗k (R) ·Π(R, t) d3R. (3.591)

These variables are complex. Expansions of the vector potential and conjugatemomentum field in terms of the true modes Ak(R) are

A(R, t) =∑

k

qk(t)Ak(R), (3.592)

Π(R, t) =∑

k

pk(t)ε(R)Ak(R). (3.593)

It is assumed that the exact electric permittivity and magnetic permeabilityfunctions do not deviate much from artificially chosen functions ε(R), μ(R). It isassumed that these functions produce quasimode functions Uα(R), idealized ver-sions of the true mode functions Ak(R). Let λα denote the angular frequency of thequasimode. One has equations

∇ × 1

μ(R)[∇ × Uα(R)] = λ2

αε(R)Uα(R), (3.594)

∇ · [ε(R)Uα(R)] = 0, (3.595)∫

ε(R)U∗α(R) · Uβ(R) d3R = δαβ, (3.596)

which are the generalized Helmholtz equation, the gauge conditions, and orthonor-mality conditions, respectively.

As the vector potential A(R, t) satisfies the generalized Coulomb gauge condition(3.579), the field ε(R)

ε(R) A(R, t) fulfils the generalized Coulomb gauge condition

∇ · [ε(R)F(R, t)] = 0, (3.597)

Page 198: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 189

where F(R, t), any field, can be chosen in the form ε(R)ε(R) A(R, t). Another choice is

based on the gauge transformation

A(R, t) = A(R, t)− ∇ψ(R, t), (3.598)

where ψ(R, t), an arbitrary function of R and t , must be specified (Glauber andLewenstein 1991), and the scalar potential

Φ(R, t) = ψ(R, t) (3.599)

need not be considered. Therefore, the expansion

ε(R)

ε(R)A(R, t) =

α,β

Qα(t)KαβUβ(R) (3.600)

exists, where the complicated form of the coefficients∑

α Qα(t)Kαβ may and maynot involve

Qα(t) =∫

ε(R)U∗α(R) · A(R, t) d3R (3.601)

and matrix elements Kαβ of a suitable matrix K. The vector potential is given as

A(R, t) =∑

α,β

Qα(t)Kαβ

ε(R)

ε(R)Uβ(R). (3.602)

Using expansion (3.602), one can write the Lagrangian (3.582), (3.583) as

L ′(t) = 1

2

α,β

Q∗α(t)(W−1)αβ Qβ(t)− 1

2

α,β

Q∗α(t)Vαβ Qβ(t), (3.603)

where

W = (KT )−1M−1(K∗)−1, (3.604)

V = K∗HKT , (3.605)

and

Mαβ =∫

ε(R)ε(R)

ε(R)U∗

α(R) · ε(R)

ε(R)Uβ(R) d3R, (3.606)

Hαβ =∫

1

μ(R)

[∇ × ε(R)

ε(R)U∗

α(R)

]·[∇ × ε(R)

ε(R)Uβ(R)

]d3R. (3.607)

Making a specific choice for K (Dalton et al. 1999b)

K = (M∗)−1, (3.608)

Page 199: Quantum Aspects of Light Propagation

190 3 Macroscopic Theories and Their Applications

one has

W = M, (3.609)

V = M−1HM−1. (3.610)

The generalized momentum coordinates Pα(t) for the electromagnetic field are

Pα(t) =∑

β

(M−1)αβ Qβ(t). (3.611)

The Hamiltonian is given by the relation

H ′(t) = 1

2

α,β

P∗α (t)Wαβ Pβ(t)+ 1

2

α,β

Q∗α(t)Vαβ Qβ(t). (3.612)

As the same Lagrangian is used and definition (3.584) does not depend on thegauge condition (3.588), the conjugate momentum field Π(R, t) is still expressedby equation (3.584). As from (3.602) it follows that

A(R, t) =∑

α,β

Qα(t)(M−1)βα

ε(R)

ε(R)Uβ(R), (3.613)

respecting (3.584) and exchanging α ↔ β, we obtain that

Π(R, t) =∑

α

Pα(t)ε(R)Uα(R). (3.614)

The classical generalized coordinates Qα(t) and generalized momenta Pα(t) arereplaced by quantum operators according to the prescriptions

Qα(t) → Qα(t), Q∗α(t) → Q†

α(t), (3.615)

Pα(t) → Pα(t), P∗α (t) → P†

α (t). (3.616)

The nonzero equal-time commutators are

[Qα(t), P†β (t)] = i�δαβ 1 = [Q†

α(t), Pβ(t)]. (3.617)

The vector potential A(R, t) and conjugate momentum field Π(R, t) now becomefield operators A(R, t), Π(R, t). Let us imagine the replacements (3.615) ((3.616))in the expression (3.602) ((3.614)) when the quantum operator A(R, t) (Π(R, t)) isintroduced. Similarly, the Hamiltonian given through equation (3.612) now becomesa quantum Hamiltonian H ′(t).

Page 200: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 191

It has the form H ′(t) = HQ(t)+ VQ−Q(t), where

HQ(t) = 1

2

α

[P†

α (t)Wαα Pα(t)+ Q†α(t)Vαα Qα(t)

], (3.618)

VQ−Q(t) = 1

2

α,β

α �=β

[P†

α (t)Wαβ Pβ(t)+ Q†α(t)Vαβ Qβ(t)

]. (3.619)

Dalton et al. (1999b) have defined an effective quasimode angular frequency μα

as follows:

μα =√

WααVαα. (3.620)

It may differ from λα in (3.594). As usual with quantum harmonic oscillators, anni-hilation and creation operators for each of the quasimodes are introduced as usuallinear combinations

Aα(t) =√

ηα

2�Qα(t)+ i

√1

2�ηα

Pα(t), (3.621)

A†α(t) =

√ηα

2�Q†

α(t)− i

√1

2�ηα

P†α (t), (3.622)

where

ηα =√

Vαα

Wαα

. (3.623)

The nonzero equal-time commutators are standard, [ Aα(t), A†β(t)] = δαβ 1. Defini-

tions (3.621) and (3.622) can be completed with the equations for A−α(t), A†−α(t),

where −α denotes that the operators are associated with the quasimode functionU∗

α(R).The relationship between the annihilation, creation operators ak , a†

k for the truemodes (see Dalton et al. (1996)) and the quantities just introduced can be obtainedfrom the expansions of two sets of functions ε(R)Ak and ε(R)Uα(R) in terms of theother. It is shown to involve a Bogoliubov transformation (Dalton et al. (1999c).

From Equations (3.621) and (3.622) one expresses the generalized coordinatesand momenta Qα(t), Pα(t) as follows:

Qα(t) =√

2ηα

[Aα(t)+ A†

−α(t)], (3.624)

Page 201: Quantum Aspects of Light Propagation

192 3 Macroscopic Theories and Their Applications

Pα(t) = 1

i

√�ηα

2

[Aα(t)− A†

−α(t)]. (3.625)

On substituting (3.624) and (3.625) into the Hamiltonians (3.618) and (3.619), theHamiltonians become

HQ(t) = �

α

[A†

α(t) Aα(t)+ 1

21

]μα, (3.626)

VQ−Q(t) = V RWAQ−Q(t)+ V non−RWA

Q−Q (t), (3.627)

where V RWAQ−Q(t) means the rotating-wave contribution,

V RWAQ−Q(t) = �

2

α,β

α �=β

(√

ηαηβ Mαβ + Vαβ√ηαηβ

)A†

α(t) Aβ(t), (3.628)

and V non−RWAQ−Q (t) stands for the nonrotating wave correction term,

V non−RWAQ−Q (t) = �

4

α,β

α �=β

(−√ηαηβ Mα,−β + Vα,−β√

ηαηβ

)A†

α(t) A†β(t)+ H.c. (3.629)

Similarly, the field operators A(R, t) and Π(R, t) should be expressed in terms ofannihilation and creation operators. One finds that

A(R, t) =∑

α,β

√�

2ηα

ε(R)

ε(R)

[Kαβ Aα(t)Uβ(R)+ K ∗

αβ A†α(t)U∗

β(R)], (3.630)

Π(R, t) =∑

α

1

i

√�ηα

2ε(R)

[Aα(t)Uα(R)− A†

α(t)U∗α(R)

]. (3.631)

The quantum Hamiltonian in the rotating wave approximation is H ′RWA(t) =

HQ(t) + V RWAQ−Q(t). Dalton et al. (1999b) have considered also an approximate form

of this Hamiltonian. A simplification is related to the approximation

μα ≈ λα (3.632)

or

μα ≈ λα + vαα, (3.633)

where we have added a term. Further,

V RWAQ−Q(t) ≈ �

α,β

α �=β

vαβ A†α(t) Aβ(t), (3.634)

Page 202: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 193

where

vαβ = 1

2

(√

λαλβ(M1)αβ + (H1)αβ√λαλβ

), (3.635)

(M1)αβ =∫

ε(R)

[ε(R)

ε(R)− 1

]2

U∗α(R) · Uβ(R) d3R, (3.636)

(H1)αβ =∫

1

μ0

{∇ ×

[ε(R)

ε(R)− 1

]U∗

α(R)

}

·{∇ ×

[ε(R)

ε(R)− 1

]Uβ(R)

}d3R. (3.637)

3.5.1 Relation to Quantum Scattering Theory

The phenomenon of scattering occurs in various situations in optics. In the classicalparticle mechanics, the scattering theory is a natural continuation and generaliza-tion of the analysis of collisions. It must be and has been reconstructed for wavefunctions in quantum mechanics. So the formulation in the Schrodinger picture isappropriate. The methods developed apply also to optical and acoustic scattering inclassical physics (Reed and Simon 1979).

In quantum field theory, the scattering theory resembles its simplified form forquantum mechanics, but with wave functions replaced by field operators. Accord-ingly, it is formulated in the Heisenberg picture. In spite of simplifications this grad-uation is present in quantum optics. First we will review basics of the Schrodingerpicture approach to the scattering theory and then outline the Heisenberg pictureapproach to this theory (Dalton et al. 1999a).

The single-channel scattering theory may be adequate for many applications inquantum optics. In this theory the Hamiltonian H (t, t ′) is written as a sum of anunperturbed Hamiltonian H0(t, t ′) and an interaction term V (t, t ′), t ′ = 0, t . Weshould denote more exactly H (t, t) in the Heisenberg picture and H (t, 0) in theSchrodinger picture. The first variable denotes the explicit time dependence of theHamiltonian. For simplicity it is assumed that H0(t, t ′), V (t, t ′), and H (t, t ′) aret-independent and the notation H0(t ′), V (t ′), and H (t ′), t ′ = 0, t , is used.

The state vector |ψ(t)〉 evolves as

|ψ(t)〉 = U (t)|ψ(0)〉, (3.638)

where the evolution operator U (t) given by

U (t) = exp

[− i

�H (0)t

](3.639)

is unitary.

Page 203: Quantum Aspects of Light Propagation

194 3 Macroscopic Theories and Their Applications

When a scattering experiment is described, the state vector |ψ(t)〉 should approachfreely evolving state vectors as t → ±∞, which are based on the so-called inputstates |ψin〉 and output states |ψout〉. So with the unitary free evolution operator U0(t)given by

U0(t) = exp

[− i

�H0(0)t

], (3.640)

we have the so-called asymptotic conditions (Taylor 1972, Newton 1966)

U0(t)|ψin〉U0(t)|ψout〉

}− |ψ(t)〉 → 0 as t →

{−∞,

+∞.(3.641)

The conditions (Taylor 1972, Newton 1966) that are sufficient for the asymptoticconditions to hold are that (‖ · · · ‖ are the norms of the state vectors)

∫ 0

−∞‖V U0(τ )|ψin〉‖ dτ < ∞ for a dense set of |ψin〉, (3.642)

∫ ∞

0‖V U0(τ )|ψout〉‖ dτ < ∞ for a dense set of |ψout〉. (3.643)

If the asymptotic conditions hold, then the Møller wave operators Ω± exist whichmap |ψin〉 and |ψout〉 onto the state vector at t = 0:

|ψ(0)〉 = Ω+|ψin〉= Ω−|ψout〉 (3.644)

and are defined through the relation

Ω+ = limt→−∞[U †(t)U0(t)]

= limt→−∞

{exp

[i

�H (0)t

]exp

[− i

�H0(0)t

]}(3.645)

and

Ω− = limt→+∞[U †(t)U0(t)]

= limt→+∞

{exp

[i

�H (0)t

]exp

[− i

�H0(0)t

]}. (3.646)

The Møller wave operators are isometric. They satisfy (Ω= Ω+ or Ω−) the relation

Ω†Ω = 1, (3.647)

but may not satisfy ΩΩ† = 1 (see below). So they may not be unitary.

Page 204: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 195

The scattering operator S maps the input vector |ψin〉 onto the output vector|ψout〉,

|ψout〉 = S|ψin〉, (3.648)

and from equations (3.644) and (3.647) it is obvious that it involves two Mølleroperators

S = Ω†−Ω+. (3.649)

It could be easily verified that the scattering operator S is unitary,

S† S = S S† = 1. (3.650)

Mapping (3.648) can be inverted,

|ψin〉 = S†|ψout〉. (3.651)

The Møller wave operators satisfy the important intertwining relation

H (0)Ω± = Ω± H0(0). (3.652)

From this relation and its Hermitian conjugate it follows that S and H0(0) commute,

S H0(0) = H0(0)S. (3.653)

In other words, the unperturbed Hamiltonian is invariant under the unitary transfor-mation S or the unperturbed energy is conserved in a scattering process.

The Møller wave operators are not unitary if there exist bound energy eigenstatesfor the Hamiltonian H . It can be derived that

ΩΩ† = 1− Λ, (3.654)

Λ is called unitary deficiency and is a sum of all the projectors onto the bound statesof H (0).

Some of the previous results simplify in the interaction picture. In this picturestate vector is defined through the equation

|ψI(t)〉 = U †0 (t)|ψ(t)〉. (3.655)

In physical systems the limits |ψI(∓∞)〉 may exist. On this assumption the simpli-fication occurs. Namely equations (3.641) and (3.648) become

|ψI(∓∞)〉 ={ |ψin〉,|ψout〉 (3.656)

Page 205: Quantum Aspects of Light Propagation

196 3 Macroscopic Theories and Their Applications

and

|ψI(+∞)〉 = S|ψI(−∞)〉. (3.657)

Schrodinger picture operators are transformed to interaction-picture ones via theequation

AI(t) = U †0 (t) AU0(t), (3.658)

where A is any Schrodinger operator. A possible time dependence of the operator Ahas not been designated. If this operator is time independent it is found that

i�d AI(t)

dt= [ AI(t), H0I(t)], (3.659)

where H0I(t) = U †0 (t)H0(0)U0(t). Especially, the Møller wave operators are associ-

ated with the interaction-picture operators ΩI±(t),

ΩI±(t) = U †

0 (t)Ω±U0(t) = exp

[i

�H0(0)t

]exp

[− i

�H (0)t

]Ω±, (3.660)

where we have used the intertwining relation. Taking the limits as t → ±∞ onesees that (Taylor 1972, Newton 1966)

ΩI+(+∞) = S, ΩI

+(−∞) = 1,

ΩI−(+∞) = 1, ΩI

−(−∞) = S†. (3.661)

In the case of ΩI+(t) equation (3.659) becomes

i�dΩI

+(t)

dt= VI(t)Ω

I+(t), (3.662)

where we have used both the intertwining relation and the equation VI(t) = HI(t)−H0I(t). The formal solution of the problem consisting in this equation with theboundary conditions (3.661) provides one with a Dyson expression (Taylor 1972,Newton 1966) for the scattering operator,

S = T exp

[− i

∫ ∞

−∞VI(t1) dt1

], (3.663)

where T means the time ordering.In the Schrodinger picture, scattering processes are often spoken of in terms of

transitions between initial and final states that are eigenstates of H0(0). So an initialstate |i〉 and a final state | f 〉 have the properties

H0(0)|i〉 = �ωi |i〉, H0(0)| f 〉 = �ω f | f 〉, (3.664)

Page 206: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 197

where ωi and ω f are frequencies. As conservation of the unperturbed energy holds,the matrix element 〈 f |S|i〉 is zero unless ω f = ωi . This fact is expressed using thetransition operator T (z) (Taylor 1972, Newton 1966), where z is a complex energyvariable. So

〈 f |S|i〉 = 〈 f |i〉 − 2π i

�δ(ω f − ωi )〈 f |T (�ωi + i0)|i〉. (3.665)

The T (z) operator is defined through the relation

T (z) = V (0)+ V (0)G(z)V (0), (3.666)

where

G(z) = [z1− H (0)]−1 (3.667)

is the resolvent operator. It obeys the Lippmann–Schwinger integral equation

T (z) = V (0)+ V (0)G0(z)T (z), (3.668)

where

G0(z) = [z1− H0(0)]−1 (3.669)

is the resolvent operator associated with H0(0).The Møller wave operators can also be related to the T (z) operator. So

Ω+ = 1+∫ ∞

−∞G0(�ω + i0)T (�ω + i0)δ

(ω1− H0(0)

)dω,

Ω− = 1+∫ ∞

−∞G0(�ω − i0)T (�ω − i0)δ

(ω1− H0(0)

)dω. (3.670)

Glauber’s theory of photodetection assumes multitime quantum correlation func-tions of the form (Glauber 1965)

G(t1, . . . , tn) = 〈ψ(0)|b1(t1)b2(t2) . . . bn(tn)|ψ(0)〉. (3.671)

We will define the input and output operators through the relations (Dalton et al.1999a)

bink (t) = Ω

†+bk(t)Ω+,

boutk (t) = Ω

†−bk(t)Ω−. (3.672)

Therefore, we read relations (8) and (10) in Dalton et al. (1999a) as follows:

G(t1, . . . , tn) = 〈ψin|bin1 (t1)bin

2 (t2) . . . binn (tn)|ψin〉,

G(t1, . . . , tn) = 〈ψout|bout1 (t1)bout

2 (t2) . . . boutn (tn)|ψout〉. (3.673)

Page 207: Quantum Aspects of Light Propagation

198 3 Macroscopic Theories and Their Applications

Further from the relations

(1− Λ)bk(t)(1− Λ) = Ω+bink (t)Ω†

+,

boutk (t) = Ω

†−bk(t)Ω− = Ω

†−(1− Λ)bk(t)(1− Λ)Ω−, (3.674)

it holds on substitution that

boutk (t) = Ω

†−Ω+bin

k (t)Ω†+Ω−. (3.675)

Using definition (3.649), we may write relation (3.675) in the form

boutk (t) = Sbin

k (t)S†. (3.676)

We may summarize that

bink (t)− U †

0 (t)bk(0)U0(t) → 0 as t →−∞,

boutk (t)− U †

0 (t)bk(0)U0(t) → 0 as t →+∞. (3.677)

Let us recall that operators in the interaction picture are introduced as

bIk(t) = U †

0 (t)bk(0)U0(t). (3.678)

The input and output operators are related to the interaction-picture operators forlong times,

bink (t)− bI

k(t) → 0 as t →−∞, bink (t)− S†bI

k(t)S → 0 as t →+∞,

boutk (t)− SbI

k(t)S† → 0 as t →−∞, boutk (t)− bI

k(t) → 0 as t →+∞. (3.679)

Dalton et al. (1999b) have outlined quasimode theory of the lossless beam splitterand Dalton et al. (1999d) have continued and extended it in relation to scatteringtheory. They have found references to the true mode and quasimode theories of thebeam splitter (see there).

They assume that the device consists of two trihedral pieces of glass of refractiveindex n (n >

√2) separated by a thin air gap of width d. The coordinate axes are

chosen such that refractive index equals n for |z| > d2 and it equals unity for |z| ≤ d

2 .To obtain quasimodes they choose ε(R) = n2ε0, μ(R) = μ0 everywhere. In this

case the quasimode functions are plane waves. For the sake of quantization, theyassume that the field is contained in a box of volume V = L3 and the quasimodefunctions

Uα(R) = 1√n2ε0V

eα exp(ikα · R), (3.680)

Page 208: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 199

where eα are polarization vectors and kα are wave vectors (eα ·kα = 0). The angularfrequencies are

λα = c

n‖kα‖. (3.681)

For the treatment to be simple, the authors may consider only two directions ofpropagation: one along 1√

2(ey − ez) and the other along 1√

2(ey + ez). In general, the

beam splitter is described by the Hamiltonian

H ′(t) = HQ(t)+ VQ−Q(t), (3.682)

where the unperturbed Hamiltonian HQ(t) and the coupling Hamiltonian VQ−Q(t)in the rotating-wave approximation are given by the relations

HQ(t) = �

α

μα A†α(t) Aα(t), (3.683)

VQ−Q(t) = �

α,β

α �=β

vαβ A†α(t) Aβ(t). (3.684)

The wave vectors kα for the quasimodes are

kα j = να j2π

L, j = x, y, z, (3.685)

where ναx , ναy , ναz are integers. The interesting directions are represented by ναx =0, ναy > 0, ναz = ∓ναy . On calculating the matrix elements (M1)αβ , (H1)αβ oneobtains that vαβ is zero unless

ναx = νβx ≡ νx , ναy = νβy ≡ νy (3.686)

and the polarization type (‖, ⊥) is the same. On the simplifying assumption itfollows that ναz = ∓νy , νβz = ∓νy , and λα = c

n νy2πL

√2 = λβ . So only the

quasimodes of the same angular frequency are coupled.The complete expression for the coupling constant (see Dalton et al. (1999d))

simplifies

vαβ = 1

2

(n2 − 1)

2Lsin

(2π

Lνyd

)

× c

n

√2

{(n2 − 1) (poln(α) = poln(β) =‖)

1 (poln(α) = poln(β) =⊥)(3.687)

for the appropriate values of α and β.

Page 209: Quantum Aspects of Light Propagation

200 3 Macroscopic Theories and Their Applications

Sums over quasimodes α with the same νx , νy reduce to sums over νy and the signof νz . The sums over νy can be converted to integral over ky using the prescription

νy

→ L

∫ ∞

0dky . (3.688)

Let us note that

L

2πvαβ → 1

2

(n2 − 1)

4πsin(kyd

) c

n

√2

{(n2 − 1) (poln =‖)

1 (poln =⊥). (3.689)

The application of quantum scattering theory to the beam splitter is justifiedin the usual situation where integrated one-photon and two-photon detection ratesare finite for incident light field states of interest (Dalton et al. 1999d). Withmodification made above we have not yet rederived the results of these authors.Also we are afraid that the appropriate operators do not converge in the rotating-wave approximation when only two directions of the incident light are consid-ered.

3.5.2 Mode Functions for Fabry–Perot Cavity

Dalton and Knight (1999a,b) have given a justification of the standard model ofcavity quantum electrodynamics in terms of a quasimode theory of macroscopiccanonical quantization. The quasimodes are treated for the representative case ofthe three-dimensional Fabry–Perot cavity. The form of the travelling and trappedmode functions for this cavity is derived in Dalton and Knight (1999a) and themode–mode coupling constants are calculated in Dalton and Knight (1999b). Theweak dependence of the coupling constants on the mode frequency differencedemonstrates that the conditions for Markovian damping of the cavity quasimodeare satisfied. We will speak of the atom–field interaction in the following subsec-tion.

A standard model used in cavity quantum electrodynamics and laser physics maybe pictured as follows. An optical cavity is produced by a perfect mirror and asemi-transparent mirror. Radiative atoms are located in the optical cavity. The atomsare coupled directly to a cavity quasimode, whose mode function is nonzero insidethe cavity and zero outside, with an atom–cavity coupling constant g. The cavityquasimode decays via Markovian damping with a rate constant Γc to certain externalquasimodes, the mode functions of which are nonzero outside the cavity and zeroinside, and which have the same axial wave vector as the cavity quasimode. Alsothe atom can decay directly via the Markovian damping with a rate constant Γ0 tocertain external quasimodes, with nonaxial wave vectors.

The standard model may be specified as a typical cavity model, the three-dimensional planar Fabry–Perot cavity. The cavity region I lies between a perfect

Page 210: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 201

mirror in the z = +l plane and a thin layer of dielectric material with dielectricconstant κ = n2 of thickness d , located between the z = 0 and z = −d planes(region II). The external region III lies between the dielectric layer plane at z = −dand a second perfect mirror in the z = −(L+d) plane. The external region length Lis much greater than the cavity length l, and both are larger than the dielectric layerthickness d.

It is assumed that the three regions constitute a rectangular cuboid with bound-aries also at x = ± L ′

2 , y = ± L ′2 . The mode functions and the necessary partial

derivatives of these functions must have the period L ′ in x and y, i.e. be invariantto transition from the plane x = − L ′

2 to the plane x = + L ′2 and from the plane

y = − L ′2 to the plane y = + L ′

2 .The electric permittivity function ε(R) for the true cavity is given as

ε(R) =⎧⎨

ε0 for −(L + d) ≤ z < −d,

κε0 for −d ≤ z < 0,

ε0 for 0 ≤ z ≤ l.(3.690)

An artificial cavity is described by a modified thickness d and a modified refractiveindex n. For the quasi-cavity the electric permittivity function ε(R) is given as

ε(R) =⎧⎨

ε0 for −(L + d) ≤ z < −d,

κε0 for −d ≤ z < 0,

ε0 for 0 ≤ z ≤ l,(3.691)

where the dielectric constant κ is related to the refractive index n through κ = n2.One makes the thin, strong dielectric approximation (Dalton and Knight 1999a). Inboth cases μ = μ0 everywhere, as there are no magnetic media involved. Obviously,the general form of the true mode functions and the quasimode functions is the same.We may interpret the notation of the true mode functions as general as far as it isuseful.

Each wave vector k is written in terms of its axial component kzez and its trans-verse component kτ ,

kτ = kx ex + kyey . (3.692)

It can be derived that the mode functions have the form

Ak(R) = exp[ikτ · (xex + yey)]Zk(z). (3.693)

Here it has been assumed that in (3.692)

kx = νx2π

L ′, νx = 0,±1,±2, . . . ,

ky = νy2π

L ′, νy = 0,±1,±2, . . . . (3.694)

Page 211: Quantum Aspects of Light Propagation

202 3 Macroscopic Theories and Their Applications

In (3.693)

Zk(z) =⎧⎨

αiei exp(ikiz z)+ αrer exp(ikrz z) region III,βtet exp(iktz z)+ βses exp(iksz z) region II,γueu exp(ikuz z)+ γvev exp(ikvz z) region I,

(3.695)

where αi, αr, βt, βs, γu, and γv are coefficients.The quantities, which are contained in relations (3.693) and (3.695), are defined

with respect to the ‖ and ⊥ polarizations. Dalton and Knight (1999a,b) investigatednot only the travelling modes but also the trapped modes. We refer to the originalwork for the latter.

In the ‖ polarization case we have the following parameters. In region III (n = ez)the wave vector of the forward-propagating wave is

ki = kω[τ sin(θ1)+ n cos(θ1)], (3.696)

where

kω = ωk

c, (3.697)

τ = ex cos φ + ey sin φ (3.698)

and that of the backward-propagating wave is

kr = kω[τ sin(θ1)− n cos(θ1)]. (3.699)

The polarization vectors are

ei = τ cos(θ1)− n sin(θ1), (3.700)

er = −τ cos(θ1)− n sin(θ1). (3.701)

The coefficients are(

αi

αr

)=(

ai

ar

)α0, (3.702)

where(

ai

ar

)= 1

2i

(exp[ikω(L + d) cos(θ1)]

exp[−ikω(L + d) cos(θ1)]

)(3.703)

and α0 will be characterized below.In region II the wave vectors are

kt = nkω[τ sin(θ2)+ n cos(θ2)], (3.704)

ks = nkω[τ sin(θ2)− n cos(θ2)], (3.705)

Page 212: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 203

where Snell’s law holds,

n sin(θ2) = sin(θ1). (3.706)

The polarization vectors are

et = τ cos(θ2)− n sin(θ2), (3.707)

es = −τ cos(θ2)− n sin(θ2). (3.708)

The coefficients are(

βt

βs

)=(

bt

bs

)β0 exp(iξ0), (3.709)

where(

bt

bs

)= 1

2

(exp(iφ0)

− exp(−iφ0)

)(3.710)

and β0 exp(iξ0) (β0 ≥ 0) will be characterized below.In region I the wave vectors are

ku = ki, kv = kr. (3.711)

The polarization vectors are

eu = ei, ev = er. (3.712)

The coefficients are(

γu

γv

)=(

gu

gv

)γ0, (3.713)

where(

gu

gv

)= 1

2i

(exp[−ikωl cos(θ1)]exp[ikωl cos(θ1)]

)(3.714)

and γ0 will be characterized in what follows.We have concentrated on the quasimodes. We have calculated the dependence of

α0 and β0 on γ0 and that of γ0 and β0 on α0. We have assumed that the γ0 dependenceis useful for γ0 � α0 and the α0 dependence is useful for α0 � γ0.

We obtain that

α0

γ0≈ cos[kω(L + d + l) cos(θ1)]

−Λ cos(θ1) cos[kω(L + d) cos(θ1)] sin[kωl cos(θ1)], (3.715)

Page 213: Quantum Aspects of Light Propagation

204 3 Macroscopic Theories and Their Applications

where Λ = n2kωd � 1,

β0 exp(iξ0)

γ0≈ −cos(θ1)

cos(θ2)sin[kωl cos(θ1)]. (3.716)

We should also assume that sin[kωl cos(θ1)] ≈ 0 or a resonant field.We obtain that

γ0

α0≈ cos[kω(l + L + d) cos(θ1)]

−Λ cos(θ1) cos[kωl cos(θ1)] sin[kω(L + d) cos(θ1)], (3.717)

β0 exp(iξ0)

α0≈ cos(θ1)

cos(θ2)sin[kω(L + d) cos(θ1)]. (3.718)

We should also assume that cos[kωl cos(θ1)] ≈ 0 or an external field far from reso-nance.

In the ⊥ polarization case we have the following parameters. In region III thewave vectors are given in relations (3.696) and (3.699). The polarization vectors are

ei = er = σ , (3.719)

where

σ = τ × n (3.720)

= ex sin φ − ey cos φ. (3.721)

The coefficients are given in relation (3.702), where(

ai

ar

)= 1

2i

(exp[ikω(L + d) cos(θ1)]

− exp[−ikω(L + d) cos(θ1)]

)(3.722)

and α0 will be characterized below.In region II the wave vectors are given in relations (3.704) and (3.705). The

polarization vectors are

et = es = σ . (3.723)

The coefficients are given in relation (3.709), where(

bt

bs

)= 1

2

(exp(iφ0)

exp(−iφ0)

)(3.724)

and β0 exp(iξ0) (β0 ≥ 0) will be characterized below.In region I the wave vectors are given in relation (3.711). The polarization vec-

tors are

eu = ev = ei. (3.725)

Page 214: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 205

The coefficients are introduced in relation (3.713), where(

gu

gv

)= 1

2i

(exp[−ikωl cos(θ1)]− exp[ikωl cos(θ1)]

)(3.726)

and γ0 will be characterized in what follows.We obtain that

α0

γ0≈ cos[kω(L + d + l) cos(θ1)]

−Λ[cos(θ2)]2

cos(θ1)cos[kω(L + d) cos(θ1)] sin[kωl cos(θ1)], (3.727)

β0 exp(iξ0)

γ0≈ − sin[kωl cos(θ1)]. (3.728)

We should also assume that sin[kωl cos(θ1)] ≈ 0 or a resonant field.We obtain that

γ0

α0≈ cos[kω(l + L + d) cos(θ1)]

−Λ[cos(θ2)]2

cos(θ1)cos[kωl cos(θ1)] sin[kω(L + d) cos(θ1)], (3.729)

β0 exp(iξ0)

α0≈ sin[kω(L + d) cos(θ1)]. (3.730)

We should also assume that cos[kωl cos(θ1)] ≈ 0 or an external field far from reso-nance.

These values have been obtained using wave optics, in which, for instance, thecoefficients in region III are connected to those in region I by the relation

(αi

αr

)= T

(γu

γv

), (3.731)

with

T ≈(

1− 12 iΛ cos(θ1) 1

2 iΛ cos(θ1)− 1

2 iΛ cos(θ1) 1+ 12 iΛ cos(θ1)

)(3.732)

for ‖ polarization and

T ≈(

1− 12 iΛ [cos(θ2)]2

cos(θ1) − 12 iΛ [cos(θ2)]2

cos(θ1)12 iΛ [cos(θ2)]2

cos(θ1) 1+ 12 iΛ [cos(θ2)]2

cos(θ1)

)(3.733)

for ⊥ polarization. The field must fulfil the boundary conditions

arαi − aiαr = 0,

gvγu − guγv = 0

}for both polarizations. (3.734)

Page 215: Quantum Aspects of Light Propagation

206 3 Macroscopic Theories and Their Applications

The eigenfrequencies ωk are given as solutions of the transcendental equation

(−ar ai)T

(gu

gv

)= 0, (3.735)

where relation (3.697) and the relations

cos(θ2) =√

1− |kτ |2n2k2

ω

, cos(θ1) =√

1− |kτ |2k2ω

(3.736)

may be mentioned, and |kτ |2 is a parameter.To achieve an approximate normalization of the near-resonant mode, we put

|γ0| =√

2

(L ′)2lε0(3.737)

independent of the polarization. When the external field is off a resonance, we put

|α0| =√

2

(L ′)2Lε0(3.738)

independent of the polarization.The coupling constants between different quasimodes are calculated according to

relations (3.636) and (3.637). The notation Uα(R) should be used for the quasimodefunctions which Aα(R) still denotes.

Dalton and Knight (1999b) have found that

Mαβ = Hαβ = 0 if ναx �= νβx or ναy �= νβy . (3.739)

They have also found that Mαβ and Hαβ are zero for quasimodes of differentpolarizations. They have shown that the coupling problem for modes in a three-dimensional Fabry–Perot cavity is equivalent to a similar problem in aone-dimensional Fabry–Perot cavity. The selection rules allow coupling betweenaxial cavity quasimodes and axial external quasimodes.

Coupling constants between a single axial cavity quasimode and axial externalquasimodes depend on the external quasimode frequency slowly. One may concludethat the conditions for irreversible Markovian damping of the cavity quasimode aresatisfied.

Analyses of cavities have been mentioned in the beginning of Section 3.3.1. Bar-nett and Radmore (1988) have shown that even the mode-strength function whichcharacterizes the true modes may be approximated using quasimodes and a phe-nomenological coupling. They have concluded that the approximation is good if thecavity is of sufficiently high quality and if the precise spatial dependence of the fielddoes not weigh.

Page 216: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 207

In the work of Garraway (1997a,b), atom-true field mode couplings are used as abasis for the pseudomode model. In certain situations quasimodes can be associatedwith pseudomodes (Dalton et al. 2003).

3.5.3 Atom–Field Interaction Within Cavity

First we complete what is needed for the description of a system of radiative atomsinteracting with the electromagnetic field in the presence of a neutral dielectricmedium on the assumptions made before. We add that the radiative atoms are sta-tionary and electrically neutral.

The radiative atoms possess charge density ρL(R, t), current density jL(R, t),polarization density PL(R, t), and magnetization density ML(R, t) given in termsof the positions rξα(t) and velocities rξα(t) of the charged particles forming theradiative atoms,

ρL(R, t) =∑

ξ,α

qξαδ(

R− rξα(t)),

jL(R, t) =∑

ξ,α

qξαδ(

R− rξα(t))

rξα(t). (3.740)

Here ξ = 1, 2, . . . lists different radiative atoms and α = 1, 2, . . . lists differ-ent particles within atom ξ . qξα, Mξα are the charge and mass for the ξα particle,respectively. One defines

PL(R, t) =∑

ξ,α

qξα

∫ 1

0[rξα(t)− Rξ (t)]

× δ(

R− Rξ (t)− u[rξα(t)− Rξ (t)])

du, (3.741)

ML(R, t) =∑

ξ,α

qξα

∫ 1

0u[rξα(t)− Rξ (t)]× [rξ (t)− Rξ (t)]

× δ(

R− Rξ (t)− u[rξα(t)− Rξ (t)])

du

+∑

ξ,α

qξα

∫ 1

0[rξα(t)− Rξ (t)]× Rξ (t)

× δ(

R− Rξ (t)− u[rξα(t)− Rξ (t)])

du, (3.742)

where Rξ (t) is the position of the centre of mass of the atom ξ , whose mass is Mξ .In the generalized Coulomb gauge condition, the scalar potential φ(R, t) satisfies ageneralized Poisson equation

∇ · [ε(R)∇φ(R, t)] = −ρL(R, t). (3.743)

Page 217: Quantum Aspects of Light Propagation

208 3 Macroscopic Theories and Their Applications

The macroscopic Lagrangian is given by the relation

L ′(t) = 1

2

ξ,α

Mξα r2ξα(t)− Vcoul(t)+

∫L′c(R, t) d3R, (3.744)

where the Lagrangian density L′c(R, t) is given by the relation

L′c(R, t) = 1

2ε(R)

[∂A(R, t)

∂t

]2

− 1

2μ(R)[∇ × A(R, t)]2

− P′L(R, t) · ∂A(R, t)

∂t+ML(R, t) · ∇ × A(R, t). (3.745)

In the Lagrangian Vcoul(t) is the Coulomb energy given by the relation

Vcoul(t) =∫

[ε(R)∇φ(R, t)]2

2ε(R)d3R (3.746)

and P′L(R, t) is the reduced polarization density:

P′L(R, t) = PL(R, t)− ε(R)∇φ(R, t). (3.747)

The conjugate momentum field Π(R, t) is obtained from the Lagrangian densityL′c(R, t) as

Π(R, t) = ε(R)∂A(R, t)

∂t− P′L(R, t) (3.748)

and the particle momenta are obtained from L ′(t) as

pξα(t) = Mξα rξα(t)

+ qξα

∫ 1

0uB(

Rξ (t)+ u[rξα(t)− Rξ (t)], t)

du × [rξα(t)− Rξ (t)].

(3.749)

The multipolar Hamiltonian is

H ′(t) =∑

ξ,α

p2ξα(t)

2Mξα

+ Vcoul(t)+∫

P′L2(R, t)

2ε(R)d3R

+∫ {

Π2(R, t)

2ε(R)+ [∇ × A(R, t)]2

2μ(R)

}d3R

+∫

Π(R,t) · P′L(R, t)

ε(R)d3R−

∫[∇ × A(R, t)] ·M′

L(R, t) d3R

+∑

ξ,α

q2ξα

2Mξα

{ ∫ 1

0u∇ × A

(Rξ (t)+ u

[rξα(t)− Rξ (t)

], t)

du

× [rξα(t)− Rξ (t)] }2

, (3.750)

Page 218: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 209

where the reduced magnetization density M′L(R, t) is given as

M′L(R, t) =

ξ,α

qξα

∫ 1

0u[rξα(t)− Rξ (t)]× pξα(t)

Mξα

× δ(

R− Rξ (t)− u[rξα(t)− Rξ (t)])

du. (3.751)

We have property (3.586).The reduced polarization density is given in terms of true mode functions as

P′L(R, t) =∑

k

ε(R)Ak(R)∫

PL(R′, t) · A∗k (R′) d3R′. (3.752)

Using expansion (3.602), one can write Lagrangian (3.744), with the Lagrangiandensity (3.745) as follows:

L ′(t) = 1

2

α,ξ

Mαξ r2αξ (t)− Vcoul(t)

+∑

α,ξ

qαξ rαξ (t) ·∫ 1

0uB(

Rξ (t)+ u[rξα(t)− Rξ (t)

], t

)du × [rξα(t)− Rξ (t)

]

+ 1

2

α,β

Q∗α(t)(W−1)αβ Qβ(t)− 1

2

α,β

Q∗α(t)Vαβ Qβ(t)

−∑

α

Q∗α(t)Nα(t), (3.753)

where

N(t) = K∗L(t),

Lα(t) =∫

ε(R)

ε(R)U∗

α(R) · PL(R, t) d3R. (3.754)

Making the choice (3.608) for K (Dalton et al. 1999b), one has

N(t) = M−1L(t). (3.755)

The generalized momentum coordinates Pα(t) for the electromagnetic field are

Pα(t) =∑

β

(M−1)αβ Qβ(t)− Nα(t). (3.756)

Page 219: Quantum Aspects of Light Propagation

210 3 Macroscopic Theories and Their Applications

The Hamiltonian is given by the relation

H ′(t) =∑

α,ξ

p2αξ (t)

2Mαξ

+ Vcoul(t)+ 1

2

α,ξ

N ∗α (t)Mαξ Nξ (t)

+ 1

2

α,β

P∗α (t)Mαβ Pβ(t)+ 1

2

α,β

Q∗α(t)Vαβ Qβ(t)

+∑

α,β

P∗α (t)Mαβ Nβ(t)+ 1

2

α,β

Q∗α(t)(M−1)αβ Rβ(t)

+ 1

2

α,β

Q∗α(t)Xαβ(t)Qβ(t), (3.757)

where

X(t) = M−1D(t)M−1, (3.758)

with

Rβ(t) = −∫

M′L(R, t) ·

[∇ × ε(R)

ε(R)U∗

β(R)

]d3R, (3.759)

Dαβ(t) =∑

μ,ξ

q2μξ

Mμξ

∫ ∫ ∫ 1

0u∫ 1

0u′δ(

R− Rξ (t)− u[rμξ (t)− Rξ (t)

] )

× δ(

R′ − R′ξ (t)− u′

[rμξ (t)− Rξ (t)

] )d u′du

×{[

rμξ (t)− Rξ (t)]2[∇ × ε(R)

ε(R)U∗

α(R)

]·[∇ × ε(R)

ε(R)Uβ(R)

]′

− [rμξ (t)− Rξ (t)] ·[∇ × ε(R)

ε(R)U∗

α(R)

] [rμξ (t)− Rξ (t)

]

·[∇ × ε(R)

ε(R)Uβ(R)

]′}d3R d3R′, (3.760)

where[∇ × ε(R)

ε(R) Uβ(R)]′

means that the vector R and the derivatives with respect

to its components are replaced by the vector R′ and the derivatives with respect tothe components of R′, respectively. The terms in the Hamiltonian are the particlekinetic energy, Coulomb energy, polarization energy, radiative energy (two terms),electric interaction energy, magnetic interaction energy, and diamagnetic energy.

The reduced polarization density can be expanded in terms of the quasimodefunctions ε(R)Uα(R) as

P′L(R, t) =∑

α

(M−1L(t))αε(R)Uα(R). (3.761)

Page 220: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 211

The quantization for the radiative atom charged particles is the familiar prescriptionsinvolving Hermitian operators

rαξ (t) → rαξ (t), pαξ (t) → pαξ (t), (3.762)

with the usual commutation rules applying.The full quantum multipolar Hamiltonian is

H ′(t) =∑

ξα

p2ξα(t)

2Mξα

+ Vcoul(t)+∫

P′L2(R, t)

2ε(R)d3R

+ �

α

μα

[A†

α(t) Aα(t)+ 1

21

]

+ �

2

α,β

α �=β

(√

ηαηβ Mαβ + Vαβ√ηαηβ

)A†

α(t) Aβ(t)

+ �

4

α,β

α �=β

(−√ηαηβ Mα,−β + Vα,−β√

ηαηβ

)A†

α(t) A†β(t)+ H.c.

+∫

Π(R, t) · P′L2(R, t)

ε(R)d3R−

∫ [∇ × A(R, t)] · M′

L(R, t) d3R

+∑

ξ,α

q2ξα

2Mξα

{ ∫ 1

0u∇ × A

(Rξ (t)1+ u

[rξα(t)− Rξ (t)1

], t)

du

× [rξα(t)− Rξ (t)1] }2

, (3.763)

where P′L(R, t), M′L(R, t) are given by equations (3.761) and (3.751) in the operator

form.The polarization energy term and the Coulomb energy term can be combined to

be equal to the sum of intra-atomic Coulomb and polarization energy terms plusintra-atomic contact energy terms. One has (Dalton and Babiker 1997)

Vcoul(t)+∫

P′L2(R, t)

2ε(R)d3R = V IA

coul(t)+ V IApol(t)+ Vcont(t). (3.764)

One defines ρLξ (R, t) by relation (3.740) with the sum over ξ omitted and PLξ (R, t)by (3.741) with the same modification. One modifies (3.743) and (3.747) as

∇ · [ε(R)∇φξ (R, t)] = −ρLξ (R, t) (3.765)

and

P′Lξ (R, t) = PLξ (R, t)− [ε(R)∇φξ (R, t)], (3.766)

Page 221: Quantum Aspects of Light Propagation

212 3 Macroscopic Theories and Their Applications

respectively. In (3.764)

V IAcoul(t) =

ξ

∫ [ε(R)∇φξ (R, t)

] · [ε(R)∇φξ (R, t)]

2ε(R)d3R, (3.767)

V IApol(t) =

ξ

∫ P′Lξ (R, t) · P′Lξ (R, t)

2ε(R)d3R, (3.768)

Vcont(t) =∑

ξ,η

ξ �=η

∫PLξ (R, t) · PLη(R, t)

2ε(R)d3R. (3.769)

To obtain the electric dipole approximation one neglects the magnetic and dia-magnetic interaction energy terms. The polarization density is given in its dipolarapproximation

PL(R, t) =∑

ξ

μξ (t)δ(

R− Rξ (t)), (3.770)

where μξ (t) is the dipolar moment for the ξ atom. The reduced polarization densitybecomes

P′L(R, t) =∑

ξ,α,β

(M−1)αβ

ε(

Rξ (t))

ε(

Rξ (t)) μξ (t) · U∗

β

(Rξ (t)

)ε(R)Uα(R). (3.771)

The atom–electromagnetic field interaction energy then assumes the forms

V E1A−F(t) =

ξ,α

1

i

√�ηα

2

ε(

Rξ (t))

ε(

Rξ (t))

×[

Aα(t)μξ (t) · Uα

(Rξ (t)

)− A†

α(t)μξ (t) · U∗α

(Rξ (t)

)], (3.772)

=∑

ξ

μξ (t) · Π(

Rξ (t), t)

ε(

Rξ (t)) . (3.773)

The quantum Hamiltonian in the electric dipole approximation and rotating waveapproximation is

H ′E1,RWA(t) = HA(t)+ HQ(t)+ V E1

A−F(t)+ V RWAQ−Q (t), (3.774)

with

HA(t) =∑

ξ,α

p2ξα(t)

2Mξα

+ V IAcoul(t)+ V IA

pol(t). (3.775)

Page 222: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 213

The coupling constants describing energy exchange processes between a radia-tive atom placed in the cavity and nonaxial external quasimodes vary slowly withthe external quasimode frequency. It follows that Markovian spontaneous emissiondamping occurs for the radiative atoms. On the contrary, their coupling with the(axial) cavity quasimodes consists in reversible photon exchanges as characterizedthrough one-photon Rabi frequencies.

In the analysis, the standard model in cavity quantum electrodynamics has beenconsidered. In the model the basic processes are described by a cavity damping rate,a radiative atom spontaneous decay rate, and an atom–cavity mode coupling con-stant. This model has been justified in terms of the quasimode theory of macroscopiccanonical quantization (Dalton and Knight 1999a,b).

3.5.4 Several Sets of Quasimodes

The quasimode theory of macroscopic quantization (Dalton et al. 1999b,c) has beengeneralized (Brown and Dalton 2001a,b). The generalization allows for the casewhere two or more quasipermittivities are introduced, along with their associatedsets of quasimode functions. This suggests problems such as reflection and refrac-tion at a dielectric boundary, the linear coupler, and the coupling of two opticalcavities. The theory comprises the above relations (3.579), (3.744), (3.745), (3.746),(3.747), (3.748), (3.586), and (3.750).

In some situations, such as a single laser cavity or a beam splitter, it suffices toconsider just a single quasipermittivity function in order to obtain suitable quasi-mode functions (Dalton et al. 1999b,c). A full quasimode treatment of the beamsplitter has been given in Dalton et al. (1999d). In other situations, the linear cou-pler (Lai et al. 1991) being an example, it is appropriate to construct quasimodefunctions via the introduction of two distinct quasipermittivities, each with its ownset of associated mode functions.

We assume N sets of quasimode functions U(l)α (R) (l = 1, . . . , N ), which are

defined as the solutions of N separate Helmholtz equations involving the quasiper-mittivities and quasipermeabilities ε(l)(R), μ(l)(R), respectively. With λ(l)

α the angu-lar frequency of the (l, α) quasimode, relations (3.594), (3.595), and (3.596) aregeneralized,

∇ × 1

μ(l)(R)[∇ × U(l)

α (R)] = (λ(l)α )2ε(l)(R)U(l)

α (R), (3.776)

∇ · ε(l)(R)U(l)α (R) = 0, (3.777)

∫ε(l)(R)U(l)∗

α (R) · U(l)β (R) d3R = δαβ. (3.778)

Expansion of the vector potential A(R) directly in terms of the quasimode func-tions U(l)

α (R) is not possible. Instead we can write

ε(R)

ε(R)A(R, t) =

l,m

α,β

Q(l)α (t)K (l,m)

αβ U(m)β (R), (3.779)

Page 223: Quantum Aspects of Light Propagation

214 3 Macroscopic Theories and Their Applications

which involves a double sum over all quasimodes. The square matrix K has becomea block matrix K, composed of K(l,m).

The quasimode functions U(l)α (R) for N > 1 are not all linearly independent. The

set of quasimodes arising from N different quasipermittivities is overcomplete. Itis solved by following an analogy with the theory of linear combinations of atomicorbitals (Coulson 1952). In that theory only the lower energy atomic orbitals areincluded. Here only quasimodes with the frequencies are retained that are importantfor the quantum optics system.

When applying the theory to situations where the true or quasi permittivitiesand permeabilities contain discontinuities, space integrals are cured with excludinginfinitesimal volumes containing these discontinuities from their domain.

Relation (3.753) is generalized,

L ′(t) = 1

2

α,ξ

Mαξ r2αξ (t)− Vcoul(t)

+∑

α,ξ

qαξ rαξ (t) ·∫ 1

0uB(

Rξ (t)+ u[rξα(t)− Rξ (t)], t)

du

× [rξα(t)− Rξ (t)]

+ 1

2

l,m

α,β

Q(l)∗α (t)(W−1)(l,m)

αβ Q(m)β (t)− 1

2

α,β

Q(l)∗α (t)V (l,m)

αβ Q(m)β (t)

−∑

α

Q(l)∗α (t)N (l)

α (t). (3.780)

Some of the matrices become block matrices of the form (given for an arbitrarycase Y)

Y =

⎜⎜⎜⎝

Y(1,1) Y(1,2) . . . Y(1,N )

Y(2,1) Y(2,2) . . . Y(2,N )

......

. . ....

Y(N ,1) Y(N ,2) . . . Y(N ,N )

⎟⎟⎟⎠ (3.781)

and column block matrices of the form (given for an arbitrary case C)

C =

⎜⎜⎜⎝

C(1,N )

C(2,N )

...C(N ,N )

⎟⎟⎟⎠ . (3.782)

A matrix Y becomes the block matrix Y and a matrix C becomes the column blockmatrix C, but the notation is not changed. Relations (3.606), (3.607), and (3.754)are generalized

M (l,m)αβ =

∫ε(R)

ε(l)(R)

ε(R)U(l)∗

α (R) · ε(m)(R)

ε(R)U(m)

β (R) d3R, (3.783)

Page 224: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 215

H (l,m)αβ =

∫1

μ(R)

[∇ × ε(l)(R)

ε(R)U(l)∗

α (R)

]

·[∇ × ε(m)(R)

ε(R)U(m)

β (R)

]d3R, (3.784)

L (l)α (t) =

∫ε(l)(R)

ε(R)U(l)∗

α (R) · PL(R, t) d3R. (3.785)

Relation (3.756) is generalized,

P (l)α (t) =

j

β

(M−1)(l,m)αβ Q(m)

β (t)− N (l)α (t). (3.786)

Relation (3.757) is generalized:

H ′(t) =∑

α,ξ

p2αξ (t)

2Mαξ

+ Vcoul(t)+ 1

2

l,m

α,ξ

N (l)∗α (t)M (l,m)

αξ N (m)β (t)

+ 1

2

l,m

α,β

P (l)∗α (t)M (l,m)

αβ P (m)β (t)+ 1

2

l,m

α,β

Ql∗α (t)V (l,m)

αβ Q(m)β (t)

+∑

l,m

α,β

P (l)∗α (t)M (l,m)

αβ N (m)β (t)

+ 1

2

l,m

α,β

Q(l)∗α (t)(M−1)(l,m)∗

αβ R(m)β (t)

+ 1

2

l,m

α,β

Q(l)∗α (t)X (l,m)

αβ (t)Q(m)β (t). (3.787)

The block matrix X(t) is given by (3.758), where the notation has the actual mean-ing. Relations (3.760) and (3.759) are generalized,

D(l,m)αβ (t) =

μ,ξ

q2μξ

Mμ,ξ

∫ ∫ ∫ 1

0u∫ 1

0u′δ(

R− Rξ (t)− u[rμξ (t)− Rξ (t)])

× δ(

R′ − R′ξ (t)− u′[rμξ (t)− Rξ (t)]

)du′ du

×{

[rμξ (t)− Rξ (t)]2

[∇ × ε(l)(R)

ε(R)U(l)∗

α (R)

]·[∇ × ε(m)(R)

ε(R)U(m)

β (R)

]′

− [rμξ (t)− Rξ (t)] ·[∇ × ε(l)(R)

ε(R)U(l)∗

α (R)

][rμξ (t)− Rξ (t)]

·[∇ × ε(m)(R)

ε(R)U(m)

β (R)

]′ }d3R d3R′, (3.788)

R(l)α (t) = −

∫M′

L(R, t) ·[∇ × ε(l)(R)

ε(R)U(l)∗

α (R)

]d3R. (3.789)

Page 225: Quantum Aspects of Light Propagation

216 3 Macroscopic Theories and Their Applications

Relation (3.761) is generalized:

P′L(R, t) =∑

l

α

(M−1L(t))(l)α ε(l)(R)U(l)

α (R). (3.790)

Replacements (3.615), (3.616) are adapted,

Q(l)α (t) → Q(l)

α (t), Q(l)∗α (t) → Q(l)†

α (t), (3.791)

P (l)α (t) → P (l)

α (t), P (l)∗α (t) → P (l)†

α (t). (3.792)

The nonzero equal-time commutators are (cf. (3.617))

[Q(l)α (t), P (m)†

β (t)] = i�δlmδαβ 1 = [Q(l)†α (t), P (m)

β (t)]. (3.793)

The electromagnetic field is real, which somewhat complicates the quantization(Brown and Dalton 2001a).

Relations (3.618) and (3.619) are generalized,

HQ(t) = 1

2

l

α

[P (l)†

α (t)W (l,l)αα P (l)

α (t)+ Q(l)†α (t)V (l,l)

αα Q(l)α (t)

], (3.794)

VQ−Q(t) = 1

2

×∑

l,m

α,β

(l,α)�=(m,β)

[P (l)†

α (t)W (l,m)αβ P (m)

β (t)+ Q(l)†α (t)V (l,m)

αβ Q(m)β (t)

]. (3.795)

Relations (3.621), (3.622), and (3.623) are generalized,

A(l)α (t) =

√η

(l)α

2�Q(l)

α (t)+ i

√1

2�η(l)α

P (l)α (t), (3.796)

A(l)†α (t) =

√η

(l)α

2�Q(l)†

α (t)− i

√1

2�η(l)α

P (l)†α (t), (3.797)

where

η(l)α =

√V (l,l)

αα

W (l,l)αα

. (3.798)

It follows simply from (3.793) that these annihilation and creation operators obeythe following equal-time nonzero commutation relations:

[ A(l)α (t), A(m)†

β (t)] = δlmδαβ 1. (3.799)

Page 226: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 217

Relations (3.624) and (3.625) are generalized,

Q(l)α (t) =

√�

2η(l)α

[A(l)

α (t)+ A(l)†−α (t)

], (3.800)

P (l)α (t) = −i

√�η

(l)α

2

[A(l)

α (t)− A(l)†−α (t)

]. (3.801)

Relation (3.626) is generalized,

HQ(t) = �

l

α

[A(l)†

α (t) A(l)α (t)+ 1

21

]μ(l)

α , (3.802)

Relation (3.620) is generalized,

μ(l)α =

√W (l,l)

αα V (l,l)αα . (3.803)

Let us recall that

VQ−Q(t) = V RWAQ−Q(t)+ V non−RWA

Q−Q (t), (3.804)

where V RWAQ−Q(t) is generalized,

V RWAQ−Q(t) = �

2

×∑

l,m

α,β

(l,α)�=(m,β)

⎝√

η(l)α η

(m)β M (l,m)

αβ + V (l,m)αβ√

η(l)α η

(m)β

⎠ A(l)†α (t) A(m)

β (t), (3.805)

and V non−RWAQ−Q (t) is also generalized,

V non−RWAQ−Q (t) = �

4

l,m

α,β

(l,α)�=(m,β)

×⎡

⎝−√

η(l)α η

(m)β M (l,m)

α,−β +V (l,m)

α,−β√η

(l)α η

(m)β

⎠ A(l)†α (t) A(m)†

β (t)+ H.c.

⎦ .

(3.806)

Relations (3.630) and (3.631) are generalized,

A(R, t) =∑

l,m

α,β

√�

2η(l)α

ε(m)(R)

ε(R)

×[

K (l,m)αβ A(l)

α (t)U(m)β (R)+ K (l,m)∗

αβ A(l)†α (t)U(m)∗

β (R)], (3.807)

Page 227: Quantum Aspects of Light Propagation

218 3 Macroscopic Theories and Their Applications

Π(R, t) = −i∑

α

√�η

(l)α

2ε(l)(R)

× [ A(l)α (t)U(l)

α (R)− A(l)†α (t)U(l)∗

α (R)]. (3.808)

The theory comprises the above relations (3.764), (3.767), (3.768), and (3.769).Relation (3.772), which comprises the annihilation and creation operators, is gener-alized,

V E1A−F(t) = −i

ξ

l,α

√�η

(l)α

2

[ε(l)(R)

ε(R)

]

Rξ (t)

×[

A(l)α (t)μξ (t) · U(l)

α

(Rξ (t)

)− A(l)†

α (t)μξ (t) · U(l)∗α

(Rξ (t)

)]. (3.809)

Let us recall that in the rotating wave and electric dipole approximations we canwrite the quantum Hamiltonian as

H ′E1,RWA(t) = HA(t)+ HQ(t)+ V E1

A−F(t)+ V RWAQ−Q (t). (3.810)

The coupling constants M (l,m)αβ , V (l,m)

αβ can be calculated from the matrices M andH using relations (3.609) and (3.610) (Brown and Dalton 2001a). For the usualcase where μ(l)(R)=μ(R) and the overlap between the set l and the set m of modefunctions is small, to good accuracy we have

M (l,m)αβ ≈

{1, i = j , α = β,

(M1)(l,m)αβ + (M2)(l,m)

αβ , otherwise,(3.811)

V (l,m)αβ ≈

{λ(l)2

α , i = j , α = β,

(H1)(l,m)αβ + (H2)(l,m)

αβ , otherwise.(3.812)

Relations (3.636), (3.637) are generalized and also modified,

(M1)(l,m)αβ =

∫ε(R)

[ε(l)(R)

ε(R)− 1

] [ε(m)(R)

ε(R)− 1

]U(l)∗

α (R)

· U(m)β (R) d3R, (3.813)

(M2)(l,m)αβ =

∫ε(R)

[ε(l)(R)

ε(R)+ ε(m)(R)

ε(R)− 1

]U(l)∗

α (R)

· U(m)β (R) d3R, (3.814)

Page 228: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 219

(H1)(l,m)αβ =

∫1

μ(R)

{∇ ×

[ε(l)(R)

ε(R)− 1

]U(l)∗

α (R)

}

·{∇ ×

[ε(m)(R)

ε(R)− 1

]U(m)

β (R)

}d3R, (3.815)

(H2)(l,m)αβ = −1

2

∫ε(R)

[λ(m)2

α

ε(l)(R)

ε(R)+ λ

(m)2β

ε(m)(R)

ε(R)

]

× U(l)∗α (R) · U(m)

β (R) d3R. (3.816)

Relation (3.632) is generalized,

μ(l)α ≈ λ(l)

α (3.817)

and relation (3.633) is also generalized,

μ(l)α ≈ λ(l)

α + v(l,l)αα . (3.818)

Relation (3.634) is generalized,

V RWAQ−Q(t) ≈ �

l,m

α,β

(l,α)�=(m,β)

v(l,m)αβ A(l)†

α (t) A(m)β (t). (3.819)

Relation (3.635) is generalized,

v(l,m)αβ = 1

2

{√λ

(l)α λ

(m)β

[(M1)(l,m)

αβ + (M2)(l,m)αβ

]

+ (H1)(l,m)αβ − (H2)(l,m)

αβ√λ

(l)α λ

(m)β

}. (3.820)

The foregoing theory has been applied to reflection and refraction at a dielectricinterface (Brown and Dalton 2001b). The true mode approach has continued theprevious literature, e.g. Allen and Stenholm (1992) or Carniglia and Mandel (1971).The analysis has been very thorough including the quantum scattering theory in theHeisenberg picture. The behaviour of the intensity for a localized one-photon wavepacket has been examined, which has exhibited agreement with the classical lawsof reflection and refraction. Such an accord is described also by the quantum theorybased on a microscopic model of the dielectric media (Hynne and Bullough 1990).

Here we will expound the quasimode approach in part. We shall assume thatspace has been divided into two regions. Region 1, which is formed by the pointswith z ≥ 0, is assumed to be filled with linear, homogeneous dielectric material ofrefractive index n1. Region 2, which consists of the points with z < 0, is assumedto contain material obeying the same restrictions, but with refractive index n2. Thepermittivity function for the system, ε(z), is then

Page 229: Quantum Aspects of Light Propagation

220 3 Macroscopic Theories and Their Applications

ε(z) ={

n21ε0, z ≥ 0,

n22ε0, z < 0.

(3.821)

We could try to use two quasipermittivity functions, one being n21ε0 in all space

and the other being n22ε0 in all space. With the two values, two sets of plane waves

are associated. The union of these sets does not enjoy the mutual orthogonalityof all functions. Instead we choose two sets of quasimodes, each set effectivelybeing restricted to just one of the regions. At a closer look, the functions are notconfined in one region, but are evanescent in the other region. An effective mutualorthogonality is present. A completeness of the union of these sets is also available.The spatially confined nature of these types of mode functions is also used whenapplying a quantum scattering theory approach to energy transfer from one regionto another.

For the reflection and refraction problems we choose the two quasipermittivityfunctions (Brown and Dalton 2001b),

ε(1)(z) ={

n21ε0, z ≥ 0,

(n2)2ε0, z < 0,(3.822)

ε(2)(z) ={

(n1)2ε0, z ≥ 0,

n22ε0, z < 0,

(3.823)

where the quasirefractive indices n1 and n2 are positive constants which fulfiln1, n2 � 1. The vanishingly small refractive index in one region means that allincident waves in the other region except some with angles of incidence smallerthan the critical angle produce only an evanescent wave in the region with negligiblerefractive index.

From the generalized Helmholtz equation (3.776), we can determine the formof the quasimode functions U(1)

α (R) and U(2)α (R), which are associated with the

quasipermittivities ε(1)(z) and ε(2)(z), respectively. We will treat the case of out-of-plane polarization. The quasimode functions are to an excellent approximationgiven by the formulae

U(1)α (R) ≈ N (1)

α

×{[√

r∗1 exp(ik(1)αi · R)+√r1 exp(ik(1)

αr · R)]Θ(z)

+ t1√

r∗1 exp(ik(1)αt · R)[1−Θ(z)]

}σ , (3.824)

U(2)α (R) ≈ N (2)

α

×{[√

r∗2 exp(ik(2)αi · R)+√r2 exp(ik(2)

αr · R)]

[1−Θ(z)]

+ t2√

r∗2 exp(ik(2)αt · R)Θ(z)

}σ , (3.825)

Page 230: Quantum Aspects of Light Propagation

3.5 Quasimode Theory 221

where

√rl = 1+ i tan(θ (l)

αi ) tanh(θ (l)αt′′)∣∣∣1− i tan(θ (l)

αi ) tanh(θ (l)αt′′)∣∣∣,

tl = 2

1− i tan(θ (l)αi ) tanh(θ (l)

αt′′)

, (3.826)

with

tan(θ (l)αi ) = |(k(l)

αi )τ |(k(l)

αi )z

,

tanh(θ (l)αt′′) = ±

√|(k(l)

αi )τ |2 − |k(l)αt |2

|(k(l)αi )τ |

, + for l = 1, − for l = 2,

|k(1)αt | =

n2

n1|k(1)

αi |, |k(2)αt | =

n1

n2|k(2)

αi |. (3.827)

Here Θ(z) is the step function and N (l)α are normalization constants appropriate to

the case where the evanescent wave has been neglected. We note that rl are complexunits and tl

√r∗l = |tl |, which may justify an alternative phase factor.

The usual formulae are obtained by multiplying the right-hand sides by√

rl ,which also simplifies them. Approximate expressions are obtained also by consid-ering rl = −1 and tl = 0. On the modification the two sets of the functions arecontinuous contrary to the notation. It is obvious that the subscript α should bereplaced by a variable k(l)

i . The corresponding z-component should be negative forl = 1 and it should be positive for l = 2.

In the case of the continuous sets, the scalar product of the functions simplifiesto a Dirac delta function when we choose

N (l)α =

(1

) 32 1

nl√

ε0. (3.828)

As this value does not depend on k(l)i , the subscript α has been preserved. We have

concluded that the discrete sets are not always obtained so easily as desired. Thecontinuous sets of the mode functions in the case where the evanescent waves havebeen neglected are discretized easily. The use of quantization box of volume L3 isimmediate. We assume that the propagation vectors k(l)

αi and k(l)αr obey

k(l)αi,X = k(l)

αr,X = N (l)α,X

L, (3.829)

where X = x, y and N (l)α,X are integers. We should complete the case X = z.

Page 231: Quantum Aspects of Light Propagation

222 3 Macroscopic Theories and Their Applications

For X = z the quantization box should not suggest the periodic boundary con-dition. We assume that the mode function vanish for z = 0 and z = L . Then ofcourse

k(l)αi,z = −k(l)

αr,z = N (l)α,z

π

L. (3.830)

The integer N (l)α,z should be negative for l = 1 and it should be positive for l = 2.

The modes with in-plane polarization are not considered for simplicity (Brown andDalton 2001b).

Page 232: Quantum Aspects of Light Propagation

Chapter 4Microscopic Theories

A divergence from the macroscopic theories emerges, when the polarization of themedium is described by separate equations. In the framework of this approach theelectric permittivity of the medium can be derived. The description of the fields canbe quantized. It seems that the separation of the equations for the medium polariza-tion is not a sufficient ground for the theory to be considered microscopic, but weadopt this nomenclature. It is important that the motion of the medium polarizationmay be damped and losses may be included. A quantum noise is considered for thefield commutators not to depend on the time.

Many papers have been devoted to the Green-function approach to the quantiza-tion of the electromagnetic field in a medium. As this theory rather begins with aquantum noise, it differs formally from the method of continua of harmonic oscil-lators. The equivalence between the Green-function approach and the method ofcontinua for media suitable for both approaches has been demonstrated, however.The Green-function approach has been elaborated on for various media, only theinclusion of a nonlinearity of the medium was under development in the course ofwriting this book.

The magnetic properties are usually neglected, but they must be included in thephenomenological quantum description of negative-index materials. Even thoughthe Casimir effect is not regularly connected with the propagation, an expression ofthe noise which is quantal in essence fits in the framework of the electromagnetic-field quantization.

4.1 Method of Continua of Harmonic Oscillators

Many scientists would call the following exposition a macroscopic theory for alossy medium, whereas we refer to a microscopic theory. A standard microscopicapproach is expected from the quantum theory of solids. Still, in the quantum theoryof solids, continua of harmonic oscillators have been considered, on which we shallconcentrate ourselves in what follows. In the framework of this model, one cansee the presence and correlation of fluctuations of the electric-field strength in thevacuum state of the field and the ground state of the matter.

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 4, C© Springer Science+Business Media, LLC 2009

223

Page 233: Quantum Aspects of Light Propagation

224 4 Microscopic Theories

4.1.1 Dispersive Lossy Homogeneous Linear Dielectric

Huttner and Barnett (1992a) have started from the observation that the macroscopicapproach to the theory of the electromagnetic field in a medium is a quantizationscheme that does accept dispersion, but not losses. So it does not deal with a fun-damental property of the susceptibility, the Kramers–Kronig relations. Losses inquantum mechanics are treated by coupling to a reservoir, and thus a quantizationscheme to describe the losses must introduce the medium explicitly.

A rigorous treatment is contained in the book of Klyshko (1988). Huttner andBarnett (1992a) use the model of Hopfield (1958) and Fano (1956), having firsttreated the quantization of light in a purely dispersive dielectric (Huttner et al. 1991)using a simple version of this model (Kittel 1987). Their analysis is restricted toa one-dimensional model and to transverse electromagnetic fields. After introduc-ing the Lagrangian densities, the effect of choice of the type of coupling betweenlight and matter on the definition of the conjugate variables for the componentsof the vector potential is discussed. The matter is not quite identical with thereservoir, but couplings rather form a chain, the radiation is coupled to the mat-ter (it is a field again) and the matter is coupled to the reservoir (it is a fieldof the dimension increased by unity). Diagonalization by the Fano technique isperformed (Fano 1961, Barnett and Radmore 1988), cf., (Rosenau da Costa et al.2000).

Huttner and Barnett (1992a) work, as usual, with fields in the reciprocal space.Only at the very beginning the radiation and the matter in the direct space, andthe reservoir in the Cartesian product of the direct and a reciprocal space are con-sidered. The total transition to a direct space for the reservoir is not usual, but isconceivable.

The description of the matter and reservoir is first diagonalized. This diagonal-ization gives rise to the (dressed) matter-field B(k, ω), whose operator exhibits thesame dependence on the wave vector and the frequency as the operator of the reser-voir field. It is proven that also the coupling constant dependent on at least thefrequency of the reservoir “elementary” mode fulfils the assumptions for furtherdiagonalization. This diagonalization gives origin to the field C(k, ω) for polari-tons. The operator of this field shares the dependence on the wave vector and thefrequency with the operator of the reservoir field. In contrast with the vacuum theory(the theory of the electromagnetic field in a vacuum) a macroscopic field emergesthis way whose operator depends also on the frequency.

The vector potential depends on the spatial coordinate and the time as usual andit has the form of the integral of the vector potential for a unit density of polaritonswith the wave vector k and the frequency ω multiplied by the polariton operatorC(k, ω). The appropriate relation contains the complex relative permittivity of themedium ε(ω) as a linear transform of the coupling constant g(ω) between the lightand the dressed matter-field B(k, ω). The complex relative permittivity ε(ω) fulfilsthe Kramers–Kronig relations.

Taking into account the frequency decompositions of the fields E(x, t) andB(x, t) (see, Huttner and Barnett (1992b)), one can introduce, in an “almost”

Page 234: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 225

conventional manner, the positive and negative propagating components. These dif-fer almost negligibly due to the imaginary part of the refractive index (see (3.506),(3.507), or (3.79)). That is to say that the fields c+(x, ω) and c−(x, ω) are introduced.Respecting the frequency decomposition of the field D(x, t), the fields c±(x, ω)are used and the spatial Langevin force f (x, ω) is introduced. Using these defini-tions, two Maxwell equations are transformed onto two spatial Langevin equations.The equal-space commutation relations between the operators at the frequencies ω

and ω′ can also be derived. From this, simple equal-space commutation relationsbetween the operators in the “application” times s and s ′ follow,

c±dir(x, s) = 1√2π

∫ ∞

0c±(x, ω) exp(−iωs) dω, (4.1)

and that may be why Huttner and Barnett (1992a) name the papers devoted tothe phenomenological approach to quantization (Levenson et al. 1985, Potasek andYurke 1987, Caves and Crouch 1987, Lai and Haus 1989, Huttner et al. 1990).

Let us note that Huttner and Barnett (1992b) in the introduction mention alsothe popular approach (Huttner et al. 1990), in which spatial progression equationsare derived and quantization of the field is performed imposing the equal-spacecommutation relations. In contrast with the macroscopic theories this technique isnot derived from a Lagrangian and has not been justified in terms of a canonicalscheme.

In Huttner and Barnett (1992b) the derivation of such equal-space commutationrelations is provided in the case of a linear dielectric. The canonical scheme andlosses cannot be easily unified, but this has been solved in Huttner and Barnett(1992b). The one-dimensional model has been expanded to three dimensions.

The Hamiltonian is first derived, then diagonalized, and the expansions of thefield operators are transformed. The propagation of light in the dielectric is anal-ysed, the field is expressed in terms of space-dependent amplitudes, and their spatialequations of evolution are obtained.

Huttner and Barnett (1992b) have started the canonical quantization from aLagrangian density

L = Lem + Lmat + Lres + Lint, (4.2)

where

Lem = ε0

2

−E︷ ︸︸ ︷(A+ ∇U ) 2 − 1

2μ0

B︷ ︸︸ ︷(∇ × A) 2 (4.3)

is the electromagnetic part which is expressed in terms of the vector potential A andthe scalar potential U ,

Lmat = ρ

2(X2 − ω2

0X2) (4.4)

Page 235: Quantum Aspects of Light Propagation

226 4 Microscopic Theories

is the polarization part, modelled by a harmonic-oscillator field X of frequency ω0

(the polarization field),

Lres = ρ

2

∫ ∞

0(Y2

ω − ω2Y2ω) dω (4.5)

is the reservoir part, comprising a field Yω of the continua of harmonic oscillatorsof frequencies ω, used to model the losses (reservoirs), and

Lint = −α(A · X+U∇ · X)− X ·∫ ∞

0v(ω)Yω dω (4.6)

is the interaction part with coupling constants α and v(ω). The interaction betweenthe light and the polarization field has the coupling constant α and the interactionbetween the polarization field and other oscillator fields used to model the losseshas the coupling constant v(ω). In general, α could be a tensor.

The displacement field is defined by

D(r, t) = ε0E(r, t)− αX(r, t). (4.7)

As U does not appear in the Lagrangian, U is not a proper dynamical variable, butit can be written in terms of the proper dynamical variable X. The former has anintegral expression and that is why we go to the reciprocal space. For example theelectric field is written as

E(r, t) = 1

(2π )32

∫E(k, t)eik·r d3k. (4.8)

We shall underline the newly introduced quantities in order to differentiatebetween the quantities in real and reciprocal spaces. Let us recall that E∗(k, t) =E(−k, t). It comprises both the annihilation and the creation operators, see below.

The total Lagrangian can be written in the form

L =∫ ′

(Lem + Lmat + Lres + Lint) d3k, (4.9)

where the prime means that the integration is restricted to half the reciprocal spaceand the Lagrangian densities become

Lem = ε0(|E|2 − c2|B|2),

Lmat = ρ(|X|2 − ω20|X|2),

Lres = ρ

∫ ∞

0(|Yω|2 − ω2|Yω|2) dω,

Lint = −α[A∗ · X+ A · X∗ + ik · (U ∗X−UX∗)]

−∫ ∞

0v(ω)(X∗ · Yω + X · Y∗

ω) dω. (4.10)

Page 236: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 227

As usual in quantum optics, we choose the Coulomb gauge, k · A(k, t) = 0, sothat the vector potential A is a purely transverse field. The scalar potential in thereciprocal space

U (k, t) = iα

ε0

(κ · X(k, t)

k

), (4.11)

where κ is a unit vector in the direction of k. The polarization field X and other oscil-lator fields Yω (the matter fields) are decomposed into transverse and longitudinalparts. For example X can be written as

X(k, t) = X⊥(k, t)+ X‖(k, t)κ (4.12)

and Yω can be expressed similarly. The total Lagrangian can then be written as thesum of two independent parts. The transverse part contains only transverse fieldsand is

L⊥ =∫ ′

(L⊥em + L⊥mat + L⊥res + L⊥int) d3k, (4.13)

where

L⊥em = ε0(|A|2 − c2k2|A|2),

L⊥mat = ρ(|X⊥|2 − ω20|X⊥|2),

L⊥res = ρ

∫ ∞

0(|Y⊥

ω |2 − ω2|Y⊥ω |2) dω,

L⊥int = −[αA · X⊥∗ +

∫ ∞

0v(ω)X⊥∗ · Y⊥

ω dω

]+ c. c. (4.14)

The longitudinal part, containing only longitudinal fields, is also given in Huttnerand Barnett (1992b). It can be derived that D is a purely transverse field. For con-venience, one can restrict oneself to transverse components of other fields and omitthe superscript ⊥.

Unit polarization vectors eλ(k), λ = 1, 2, are introduced, which are orthogonalto k and to one another, and the transverse fields are decomposed along them to get

A(k, t) =∑

λ=1,2

Aλ(k, t)eλ(k) (4.15)

and similar expressions for the other fields. L can now be used to obtain the com-ponents of the conjugate variables for the fields

−ε0 Eλ ≡ ∂L∂ A

λ∗ = ε0 Aλ, (4.16)

Page 237: Quantum Aspects of Light Propagation

228 4 Microscopic Theories

Pλ ≡ ∂L∂ X

λ∗ = ρ Xλ − αAλ, (4.17)

ω≡ δL

δYλ∗ω

= ρYλ

ω − v(ω)Xλ. (4.18)

The famous ambiguity is worth mentioning: The conjugate of A can be −ε0E (withthe coupling α

ρA · P), as well as −D (with the coupling E · X). Thus any gauge

determines a type of coupling.The Hamiltonian for the transverse fields is

H =∫ ′

(Hem +Hmat +Hint) d3k, (4.19)

where

Hem = ε0(|E|2 + c2k2|A|2) (4.20)

is the electromagnetic energy density, k being defined by k = √k2 + k2

c with kc ≡ωcc =

√α2

ρc2ε0,

Hmat =|P|2ρ

+ ρω20|X|2

+∫ ∞

0

[ |Qω|2

ρ+ ρω2|Yω|2 +

v(ω)

ρ(X∗ ·Q

ω+ c. c.)

]dω (4.21)

is the energy density of the matter fields, including the interaction between the polar-ization and the reservoirs and ω2

0 ≡ ω20+∫∞

0[v(ω)]2

ρ2 dω is the renormalized frequencyof the polarization field,

Hint =α

ρ(A∗ · P+ c. c.) (4.22)

is the interaction energy between the electromagnetic field and the polarization. Partof the interaction energy with the matter, namely α2

ρ|A|2, has already been classified

into (4.20).Fields are quantized in a standard fashion (Cohen-Tannoudji et al. 1989) by pos-

tulating equal-time commutation relations between the variables and their conju-gates

[ Aλ(k, t), E

λ′∗(k′, t)] = − i�

ε0δλλ′δ(k− k′)1, (4.23)

[Xλ(k, t), P

λ′∗(k′, t)] = i�δλλ′δ(k− k′)1, (4.24)

[Yλ

ω(k, t), Qλ′∗ω′

(k′, t)] = i�δλλ′δ(k− k′)δ(ω − ω′)1, (4.25)

where all quantized operators are denoted by a caret. As usual, the annihilationoperators are introduced.

Page 238: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 229

a(λ, k, t) =√

ε0

2�kc

[kc A

λ(k, t)− iE

λ(k, t)

], (4.26)

b(λ, k, t) =√

ρ

2�ω0

[ω0 X

λ(k, t)+ i

ρP

λ(k, t)

], (4.27)

bω(λ, k, t) =√

ρ

2�ω

[−iωY

λ

ω(k, t)+ 1

ρQ

λ

ω(k, t)

](4.28)

From the equal-time commutation relations for the fields (4.23), (4.24), (4.25),the equal-time commutation relations for the creation and annihilation operators

[a(λ, k, t), a†(λ′, k′, t)] = δλλ′δ(k− k′)1,

[b(λ, k, t), b†(λ′, k′, t)] = δλλ′δ(k− k′)1,

[bω(λ, k, t), b†ω′ (λ

′, k′, t)] = δλλ′δ(ω − ω′)δ(k− k′)1 (4.29)

are obtained. The normally ordered Hamiltonian for the transverse fields is

H = Hem + Hmat + Hint, (4.30)

where

Hem =∫ ∑

λ=1,2

�kca†(λ, k, t)a(λ, k, t) d3k, (4.31)

Hmat =∫ ∑

λ=1,2

{�ω0b†(λ, k, t)b(λ, k, t)+

∫ ∞

0�ωb†

ω(λ, k, t)bω(λ, k, t) dω

+ �

2

∫ ∞

0V (ω)

[b(λ, k, t)b†

ω(λ, k, t)

+ b†(λ,−k, t)b†ω(λ, k, t)+ H. c.

]dω

}d3k, (4.32)

Hint = i�

2

∫ ∑

λ=1,2

Λ(k)[a(λ, k, t)b†(λ, k, t)

+ a†(λ,−k, t)b†(λ, k, t)+ H. c.]

d3k, (4.33)

where V (ω) = v(ω)ρ

√ωω0

, Λ(k) ≡√

ω0ck2c

k, and the k integration has been restored to

the full reciprocal space. It is worth mentioning that the Maxwell–Lorentz equationscan be derived from the Hamiltonian.

It is important that the matter can be formally decoupled from the reservoir bythe Fano technique and a dressed matter field obtained. Following Fano (1961), thepolarization and reservoir parts of the Hamiltonian can be diagonalized. The dressed

Page 239: Quantum Aspects of Light Propagation

230 4 Microscopic Theories

matter field creation and annihilation operators B†(λ, k, ω, t) and B(λ, k, ω, t) areintroduced, respectively, which satisfy the usual equal-time commutation relations,

[B(λ, k, ω, t), B†(λ′, k′, ω′, t)] = δλλ′δ(k− k′)δ(ω − ω′)1, (4.34)

B(λ, k, ω, t) = α0(ω)b(λ, k, t)+ β0(ω)b†(λ,−k, t)

+∫ ∞

0

[α1(ω,ω′)bω′(λ, k, t)

+ β1(ω,ω′)b†ω′(λ,−k, t)

]dω′. (4.35)

The coefficients α0(ω), β0(ω), α1(ω,ω′), β1(ω,ω′) are defined as follows. It isinteresting that the diagonalization is performed once for the polarization and reser-voir parts of the Hamiltonian and once for the total Hamiltonian. The Hamiltonianexpressed in the modal annihilation operators is considered. From relation (4.35) itcan be seen that the diagonalization is performed independently for every pair ofthe counterpropagating modes of the polarization field (“the modes” here are onlyformally similar to those of the electromagnetic field) and that it is performed usinga Bogoliubov transformation. A useful definition of an “eigenoperator” is presentedBarnett and Radmore (1988),

[B(λ, k, ω, t), Hmat] = �ω B(λ, k, ω, t). (4.36)

The coefficients of the Bogoliubov transformation are calculated, that is to saythe formulae

α0(ω) =(

ω + ω0

2

)V (ω)

ω2 − ω20z(ω)

, (4.37)

where z(ω) is defined by

z(ω) = 1− 1

2ω0

[lim

ε→+0

∫ ∞

−∞

V(ω′)ω′ − ω + iε

dω′]

, (4.38)

β0(ω) =(

ω − ω0

2

)V (ω)

ω2 − ω20z(ω)

, (4.39)

α1(ω,ω′) = δ(ω − ω′)+(

ω0

2

)(V ∗(ω′)

ω − ω′ − i0

)V (ω)

ω2 − ω20z(ω)

, (4.40)

and

β1(ω,ω′) =(

ω0

2

)(V (ω′)ω + ω′

)V (ω)

ω2 − ω20z(ω)

(4.41)

are derived. In the study no constant V(ω) ≡ |V (ω)|2 occurs.As usual with the substitutions, we are also interested in the inverse transforma-

tion. It is given by the relations

b(λ, k, t) =∫ ∞

0

[α∗0 (ω)B(λ, k, ω, t)− β0(ω)B†(λ,−k, t, ω)

]dω (4.42)

Page 240: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 231

and

bω(λ, k, t) =∫ ∞

0

[α∗1 (ω′, ω)B(λ, k, ω′, t)− β1(ω′, ω)B†(λ,−k, ω′, t)

]dω′.

(4.43)The conditions

I ≡∫ ∞

0

[|α0(ω)|2 − |β0(ω)|2] dω = 1, (4.44)

I (ω,ω′) ≡∫ ∞

0

[α∗1 (ν, ω)α1(ν, ω′)− β1(ν, ω)β∗1 (ν, ω′)

]dν = δ(ω − ω′) (4.45)

for the coefficients of the Bogoliubov transformation seem to be familiar. It has beenshown that the diagonalization cannot be performed on the common assumption ofwhite noise (the Markov-type coupling). It is commented on free charges and aconducting medium being beyond the scope of Huttner and Barnett (1992a). Weneed not grieve for the assumption of the white noise. Without it, we are fartherfrom the original Lorentzian formulation, nothing more.

The diagonalization of the total Hamiltonian is formally very similar to the diago-nalization of its matter part. A dimensionless coupling constant ζ (ω)= i

√ω0[α0(ω)+

β0(ω)] is defined and the annihilation operators are introduced (by a Fano type oftechnique),

C(λ, k, ω, t) = α0(k, ω)a(λ, k, t)+ β0(k, ω)a†(λ,−k, t)

+∫ ∞

0

[α1(k, ω, ω′)B(λ, k, ω′, t)

+ β1(k, ω, ω′)B†(λ,−k, ω′, t)]

dω′, (4.46)

where the coefficients α0(k, ω), β0(k, ω), α1(k, ω, ω′), and β1(k, ω, ω′) are rathercomplicated and are derived in the form

α0(k, ω) =√

ω2c

kc

(ω + kc

2

)ζ (ω)

ω2 − k2c2 z(k, ω), (4.47)

where

z(k, ω) = 1− ω2c

2(kc)2

[lim

ε→+0

∫ ∞

−∞

|ζ (ω′)|2ω′ − ω + iε

]dω′, (4.48)

or alternatively

α0(k, ω) =√

ω2c

kc

(ω + kc

2

)ζ (ω)

ε∗(ω)ω2 − k2c2, (4.49)

where the complex relative permittivity ε(ω) is introduced

ε(ω) = 1+ 1

ω2

[k2c2 − (k2c2 + ω2

c )z∗(k, ω)], independent of k, (4.50)

Page 241: Quantum Aspects of Light Propagation

232 4 Microscopic Theories

β0(k, ω) =√

ω2c

kc

(ω − kc

2

)ζ (ω)

ε∗(ω)ω2 − k2c2, (4.51)

α1(k, ω, ω′) = δ(ω − ω′)+ ω2c

2

(ζ ∗(ω′)

ω − ω′ − i0

)ζ (ω)

ε∗(ω)ω2 − k2c2, (4.52)

and

β1(k, ω, ω′) = ω2c

2

(ζ ∗(ω′)

ω − ω′ − i0

)ζ (ω)

ε∗(ω)ω2 − k2c2. (4.53)

The operators C(λ, k, ω, t) and C†(λ, k, ω, t) also satisfy the usual commutationrelations,

[C(λ, k, ω, t), C†(λ′, k′, ω′, t)] = δλλ′δ(k− k′)δ(ω − ω′)1 (4.54)

and being operators for eigenmodes,

[C(λ, k, ω, t), H ] = �ωC(λ, k, ω, t), (4.55)

they have a harmonic time dependence

C(λ, k, ω, t) = C(λ, k, ω, 0)e−iωt . (4.56)

The vector potential is now given by

A(r, t) = 1

(2π )32

∫ √�ω2

c

2ε0

λ=1,2

eλ(k)

×∫ ∞

0

[ζ ∗(ω)

ω2ε(ω)− k2c2C(λ, k, ω, 0)e−i(ωt−k·r) + H. c.

]dω d3k. (4.57)

Relation (4.8) being modified for the operators expresses the operators A(r, t),E(r, t),... in terms of the operators A(k, t), X(k, t), Yω(k, t), E(k, t), P(k, t), Q

ω(k, t).

Let us note that on the substitution into (4.8) for A(k, t), E(k, t),... by the rela-tions

A(k, t) =√

2kcε0[a(k, t)+ a†(−k, t)],

E(k, t) = i

√�kc

2ε0[a(k, t)+ a†(−k, t)], (4.58)

. . . . . . . . . ,

Page 242: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 233

the annihilation and creation operators a(k, t), b(k, t), bω(k, t) and a†(k, t), b†(k, t),b†

ω(k, t), respectively, are introduced. On the substitution into the intermediate resultfor the operators b(k, t) and bω(k, t) by relations (4.42) and (4.43), the operatorsB(k, ω, t) are introduced. On the substitution into the intermediate result for theoperators a(k, t) by the relation (the slightly modified relation (4.2) from Huttnerand Barnett (1992b))

a(k, t) =∫ ∞

0

[α∗0 (k, ω)C(k, ω, t)− β0(k, ω)C†(−k, ω, t)

]dω (4.59)

and for the operators B(k, ω, t) by the relation

B(k, ω, t) =∫ ∞

0

[α∗1 (k, ω, ω′)C(k, ω, t)− β1(k, ω′, ω)C†(−k, ω, t)

]dω′,

(4.60)the operators C(k, ω, t) are introduced, which have the time dependence (4.56).

On the substitution into the intermediate result for C(k, ω, 0) by relation (4.46),the operators a(k, 0) and B(k, ω, 0) are introduced and on the substitution into theintermediate result for the operators B(k, ω, 0) by relation (4.35), the operatorsb(k, 0) and bω(k, 0) are introduced. On the substitution into the intermediate resultfor these operators by the formulae (4.26), (4.27), and (4.28), the operators A(k, 0),X(k, 0), Yω(k, 0), E(k, 0), P(k, 0), Q

ω(k, 0) are introduced. On the substitution into

the intermediate result for these operators by the relations

A(k, t) = 1

(2π )32

∫A(r, t)e−ik·r d3r,

E(k, t) = 1

(2π )32

∫E(r, t)e−ik·r d3r, (4.61)

. . . . . . . . . ,

the operators A(r, 0), E(r, 0),... are introduced. These substitutions solve the Cauchyor initial problem.

Huttner and Barnett (1992b) restrict themselves first to a one-dimensional casewhen describing the propagation in the dielectric. The vector potential is consideredin the simpler form

A(x, t) = 1√4π

∫ ∞

0A(ω)

[c+(x, ω, 0)e−iωt + c−(x, ω, 0)e−iωt + H. c.

]dω,

(4.62)where

A(ω) =√

�η(ω)

ε0Scω|n(ω)|2 , (4.63)

Page 243: Quantum Aspects of Light Propagation

234 4 Microscopic Theories

S is a cross-sectional area, n(ω) is the complex refractive index defined as the squareroot of the relative permittivity ε(ω) with a positive real part η(ω), and the operatorsc±(x, ω, t) are introduced

c±(x, ω, t) =√

Im{K (ω)}π

eiφ(ω)∫ ∞

−∞

C(k, ω, t)eikx

K (ω)∓ kdk, (4.64)

where the complex wave number K (ω) and the phase factor eiφ(ω) are expressed as

K (ω) = n(ω)ω

c, (4.65)

eiφ(ω) = ζ ∗(ω)

|ζ (ω)||n(ω)|n(ω)

, (4.66)

and Im{K (ω)} > 0.Since the magnetic field can be expressed similarly as the vector potential, the

spatial quantum Langevin equations of progression can be obtained as

∂ c±(x, ω, t)

∂x= ±iK (ω)c±(x, ω, t)±

√2Im{K (ω)} f (x, ω, t), (4.67)

where the Langevin-noise operator is

f (x, ω, t) = − i√2π

eiφ(ω)∫ ∞

−∞C(k, ω, t)eikx dk (4.68)

and it also enters a rather similar expression for the electric displacement operator.Equation (4.67) have been obtained from the Maxwell equations for the monochro-matic fields.

Huttner and Barnett (1992b) remind of the simple commutation relations

[ f (x, ω, t), f †(x ′, ω′, t)] = δ(x − x ′)δ(ω − ω′)1, (4.69)

further of the equal-space commutation relations

[c±(x, ω, t), c†±(x, ω′, t)] = δ(ω − ω′)1,

[c±(x, ω, t), c†∓(x, ω′, t)] = 0 (4.70)

and, finally, that c+(x, ω, t) commutes with all the Langevin operators f (x ′, ω′, t)and f †(x ′, ω′, t) for all x ′ > x , while c−(x, ω, t) commutes with all the Langevinoperators f (x ′, ω′, t) and f †(x ′, ω′, t) for all x ′ < x .

Jeffers and Barnett (1994) modelled the propagation of squeezed light through anabsorbing dispersive dielectric medium. Hradil (1996) considered “lossless” disper-sive dielectrics, i.e. dielectrics with a thin absorption line. He formulated a canoni-cal quantization of the electromagnetic field in a closed Fabry–Perot resonator witha dispersive slab.

Page 244: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 235

Wubs and Suttorp (2001) have solved the initial-value problem for the damped-polariton model formulated by Huttner and Barnett (1992a,b) and have found thatfor long times all field operators can be expressed in terms of the initial reser-voir operators. They have investigated the transient dynamics of the spontaneous-emission rate of a guest atom in an absorbing medium.

Hillery and Drummond (2001) have studied the scattering of the quantized elec-tromagnetic field from a linear dispersive dielectric in the limit of “thin” absorptionlines. The field is represented by means of the dual vector potential. Input–outputrelations are unitary and no additional quantum-noise terms are required. Equationsspecialized to the case of a dielectric layer with a uniform density of oscillators areusual expressions.

Janowicz et al. (2003) analyse radiative heat transfer between two dielectric bod-ies. Quantization of the electromagnetic field in inhomogeneous, dispersive, andlossy dielectrics is performed with the help of a procedure which is still attributedto Huttner and Barnett (1992b). Expectation value of the Poynting vector operatoris computed. To this end, two techniques suitable for nonequilibrium processes areutilized: the Heisenberg equation of motion and the diagrammatic Keldysh proce-dure. It is remarked that in nonlinear models the Keldysh formalism provides aframework for the perturbation expansion. The calculations fit into the developmentof the theory of thermal scanning microscopy.

4.1.2 Correlation of Ground-State Fluctuations

The quantization of the radiation imbedded in a dielectric with a space-dependentrefractive index has been expounded in the book by Vogel and Welsch (1994).A canonical quantization scheme for radiation fields in linear dielectrics with aspace-dependent refractive index has been developed by Knoll et al. (1987) andlater by Glauber and Lewenstein (1991). For application, see, for example, Knollet al. (1986, 1990, 1991), Knoll and Welsch (1992) and a related work (Knoll andLeonhardt 1992).

Gruner and Welsch (1995) have contributed to the stream of papers aiming ata description of quantum properties of the dispersive and lossy dielectrics includ-ing the vacuum fluctuations, i.e. fluctuations of radiation field in the ground stateof the coupled light–matter system. They study it in terms of a symmetrized cor-relation function. They try to expound and supplement the paper by Huttner andBarnett (1992b) from the point of view of the quantization of the phenomenologicalMaxwell theory.

First, the quantization of radiation in a dispersive and lossy dielectric is per-formed. This begins from the classical Maxwell equations (3.176) with (3.177) anda constitutive relation comprising an integral term,

D(r, t) = ε0

[E(r, t)+

∫ ∞

0χ (τ )E(r, t − τ ) dτ

], (4.71)

Page 245: Quantum Aspects of Light Propagation

236 4 Microscopic Theories

is transformed into the Fourier space to yield

D(r, ω) = ε0ε(ω)E(r, ω) (4.72)

and the Helmholtz equation is presented.The Huttner–Barnett quantization scheme is introduced with a diagonalized

Hamiltonian

H =∑

λ=1,2

∫ ∫ ∞

0�ωC†(λ, k, ω)C(λ, k, ω) dω d3k, (4.73)

which comprises a sum over λ which is absent from relation (3.14) of Huttner andBarnett (1992b). The effect of the medium is entirely determined by the complexpermittivity ε(ω). It still has no tensorial character. Let us remember relations (4.3),(4.5), (4.6), and (4.7) of Huttner and Barnett (1992b). In these relations, C(k, ω)

should be replaced by C(λ, k, ω) and the identity√

ω2c

2 ζ ∗(ω) = ω√

Im{ε(ω)} uti-lized.

The frequency-dependent field operators are introduced in the three-dimensionalcase. Not only the equal-time commutation relations, but even the most general onesare presented. The vector-field operators a(r, ω) and f(r, ω) have been introduced,the vector a(r, ω) being a generalization of the component c(x, ω) from Huttner andBarnett (1992b). The definition of the transverse δ function is presented as

δ⊥i j (r) = 1

(2π )3

∫ ∞

−∞

(δi j − ki k j

k2

)eik·r d3k, (4.74)

which can be interpreted also as a transverse projection of the columns of 3× 3 iden-tity matrix multiplied by the δ function. The transverse projection of the columns ofan identity matrix multiplied by other functions has proved to be useful, for example,the commutators [ai (r, ω), a j (r′, ω′)] can be expressed in terms of such projections.Here, the operator Δi j ,

Δi j F(r) =∫ ∞

−∞δ⊥i j (r− r′)F(r′) d3r′, (4.75)

is applied (the putting of an identity matrix to be multiplied by a function and sub-sequent transverse projection of the columns), but very complicated expressions areobtained.

Continuing the use of the operators a(r, ω) and f(r, ω), an analogue of relation(5.21) of Huttner and Barnett (1992b) (cf. (4.62) here) has been written. Then, ana-logues of their frequency decompositions of the vector-potential operators, electric-field strength operators, etc. have been presented. The operator constitutive equation(in the Fourier space)

D(r, ω) = ε0ε(ω)E(r, ω)− ε0F(ω)f(r, ω), (4.76)

Page 246: Quantum Aspects of Light Propagation

4.1 Method of Continua of Harmonic Oscillators 237

where

ε0F(ω) =√

πε0Im{ε(ω)} (4.77)

differs from the classical equation (4.72) by an additional term. On substituting intothe phenomenological Maxwell equations, the partial differential equation for theoperator a(r, ω) is obtained which is a Helmholtz equation with a right-hand side. Inthe three-dimensional case, there exists no decomposition into first-order equations.The canonical commutation relations are

[ Ai (r, t), E j (r′, t)] = − i�

ε0δ⊥i j (Δr)1, (4.78)

with the abbreviation

Δr = r− r′. (4.79)

A test of the consistency of the theory in the limit ε(ω)→ 1 has been accomplished.Let us recall the usual annihilation operators a(λ, k), which satisfy the commu-

tation relations

[a(λ, k), a†(λ′, k′)] = δλλ′δ(k− k′)1 (4.80)

and enter the expansion for A(r, 0).The operators a(r, ω) and f(r, ω) derived from C(λ, k, ω) are not independent

operators. Cf., Huttner and Barnett (1992b) who in the one-dimensional case intro-duce forward and backward-propagating fields and show that such a definitionensures the causal (one-sided) independence of the respective operators of the oper-ator f(r, ω). In the three-dimensional case, there exists no generalization of relation(4.64) and no equation for such quantities.

The theory is applied to the determination of the correlation of the ground-statefluctuations of the electric-field strength. The symmetric correlation function of theelectric-field strength

Kmn(Δr, τ ) = 1

2〈0|[Em(r, t + τ )En(r+Δr, t)

+ En(r+Δr, t)Em(r, t + τ )]|0〉 (4.81)

is considered. We remark that

Kmn(Δr, τ ) = �

4π2c2ε0nR(ω)

×∫ ∞

0cos(ωτ )Δi j

exp[−ω

c nI(ω)]

nR(ω)Δrsin[ω

cnR(ω)Δr

]dω, (4.82)

Page 247: Quantum Aspects of Light Propagation

238 4 Microscopic Theories

where Δr = |Δr| and

nI(ω) = Im{√

ε(ω)} = Im{n(ω)}, (4.83)

nR(ω) = Re{√

ε(ω)} = Re{n(ω)}. (4.84)

Restricting attention to optical frequencies within an interval of the width 2Δω,

ω0 −Δω < ω < ω0 +Δω, (4.85)

Δω

ω0� 1, (4.86)

where ω0 is an appropriately chosen centre frequency and assuming that dispersionand absorption are small on lengths of order of β−1,

β = ω

cnR(ω), (4.87)

and times of order of ω−1, Gruner and Welsch (1995) let

(βΔr )−1 � 1, (4.88)

(ωτ )−1 � 1. (4.89)

Further, they assume a transparent medium, such as a fibre, for which it may bejustified to put approximately

nR(ω) ≈ nR0 + nR1ω

ω0, (4.90)

nI(ω) ≈ nI(ω0) ≡ nI0. (4.91)

The influence of absorption, phase, and group velocities and group velocity dis-persion on the dynamics of the field fluctuations within a frequency interval (4.85)has been studied. The absorption causes a spatial decay of the correlation of thefield fluctuations. The light cone of strong correlation, which in empty space isdetermined by the speed of light in vacuum, is now given by the group velocityin the medium provided that the spatial distance is not too large. With increasingdistance, also the dispersion of the group velocity needs a consideration.

4.2 Green-Function Approach

On allowing for a frequency-dependent complex permittivity that is consistent withthe Kramers–Kronig relations and introducing a random operator noise source asso-ciated with the absorption of radiation, the classical Maxwell equations can be con-sidered as quantum operator equations. Their solution based on a Green-function

Page 248: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 239

expansion of the vector-potential operator seems to be a natural generalization ofthe mode expansion applicable to source-free radiation in nearly lossless dielectrics.

4.2.1 Dispersive Lossy Linear Inhomogeneous Dielectric

Gruner and Welsch (1996a) have expounded a quantization scheme which startswith phenomenological Maxwell equations instead of Lagrangian densities and isconsistent with the Kramers–Kronig relations and the familiar (equal-time) canoni-cal commutation relations for the vector potential and electric field. This is realizedfor homogeneous and inhomogeneous, especially, multilayered dielectrics.

In the phenomenological classical Maxwell theory, the equations comprise ε(ω),the frequency-dependent complex relative permittivity introduced phenomenologi-cally. This function has the analytical continuation in the upper complex half-plane,ε(Ω), which satisfies the relation

ε(−Ω∗) = ε∗(Ω). (4.92)

The real and imaginary parts of the relative permittivity satisfy the well-knownKramers–Kronig relations

Re{ε(ω)} − 1 = 1

πV.p.

∫ ∞

−∞

Im{ε(ω′)}ω′ − ω

dω′, (4.93)

Im{ε(ω)} = − 1

πV.p.

∫ ∞

−∞

Re{ε(ω′)} − 1

ω′ − ωdω′, (4.94)

where V.p. is the principal value of the integral.The quantization scheme is based on the Helmholtz equation with the source

term

Δ ˆA(r, ω)+K2(ω) ˆA(r, ω) = ˆjn(r, ω), (4.95)

where ˆA(r, ω) is the “Fourier transform” of the (known) operator vector-potential

A(r, t) and ˆjn(r, ω) is the “Fourier transform” of the operator-noise current. In fact,from the exposition it can be seen that the vector-potential operator is introduced bythe relation

A(r, 0) =∫ ∞

0

ˆA(r, ω, 0) dω + H. c., (4.96)

where quantum mechanically also the frequency-dependent operators can be timedependent, t = 0.

When Im{ε(ω)} > 0, a hypothetical addition of a nontrivial solution of thehomogeneous Helmholtz equation would violate the boundary condition at infinity.

Page 249: Quantum Aspects of Light Propagation

240 4 Microscopic Theories

Hence, the operator ˆA(r, ω)≡ ˆA(r, ω, t) is uniquely determined by a linear transfor-

mation of the source operator ˆjn(r, ω)≡ ˆjn(r, ω, t). This operator can be chosen inthe form (cf., Gruner and Welsch (1995))

ˆjn(r, ω) = F(ω)ω

c2f(r, ω), (4.97)

with F(ω) given in (4.77). The Hamiltonian H is diagonal in the operators f(r, ω),

H =∫ ∫ ∞

0�ωf†(r, ω) · f(r, ω) dω d3r, (4.98)

and these operators have the usual properties

[ fi (r, ω), f †j (r′, ω′)] = δ⊥i j (r− r′)δ(ω − ω′)1, (4.99)

[ fi (r, ω), f j (r′, ω′)] = [ f †i (r, ω), f †j (r′, ω′)] = 0. (4.100)

From the foregoing considerations it follows that (when all appropriate condi-tions are fulfilled) the operator of the vector potential can be defined by the relation

A(r, 0) =∫ ∞

0

∫G(r, r′, ω)ˆjn(r′, ω, 0) d3r′ dω + H. c., (4.101)

where the Green function G(r, r′, ω) satisfies the equation

ΔG(r, r′, ω)+ K 2(ω)G(r, r′, ω) = δ(r− r′) (4.102)

and the boundary condition that it vanishes at infinity. Another required property is

E(r, 0) = − ˙A(r, 0) (4.103)

and the canonical field commutation relations

[ Ai (r, 0), E j (r′, 0)] = − i�

ε0δ⊥i j (r− r′)1. (4.104)

Relation (4.104) must be verified by straightforward calculation. For the sake ofclarity, Gruner and Welsch (1996a) illustrate this procedure in linearly polarizedradiation propagating in the x direction. Relation (4.104) are replaced by the relation

[ A(x, 0), E(x ′, 0)] = − i�

Aε0δ(x − x ′)1, (4.105)

Page 250: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 241

where A is the normalization area perpendicular to the x direction. It is shown thatwhen losses in the dielectric may be disregarded, Im{ε(ω)} → 0, the concept ofquantization through the mode expansion can be recognized. The operators f (x, ω)are replaced by the operators a±(x, ω) (it would be possible to introduce the opera-tors a(x, k = ±Re{n(ω)}ω

c )), which satisfy the commutation relations

[a±(x, ω), a†±(x ′, ω′)] = exp

[−Im{n(ω)}ω

c|x − x ′|

]δ(ω − ω′)1, (4.106)

[a±(x, ω), a†∓(x ′, ω′)] = 2Im{n(ω)}ω

cexp

[∓iβ(ω)

ω

c(x + x ′)

]

× exp[−Im{n(ω)}ω

c|x − x ′|

] sin[Re{n(ω)}ω

c |x − x ′|]

Re{n(ω)}ωc

× θ [±(x − x ′)]δ(ω − ω′)1, (4.107)

where θ (x) is the Heaviside function. These operators become independent of x inthe limit Im{n(ω)}ω

c |x − x ′| → 0. As the commutation relation (4.105) is in anobvious contradiction with a macroscopic approach, it is important that Gruner andWelsch (1996a) have derived the relation

[ AΔω(x, 0), EΔω(x ′, 0)] = − i�

AεR(ωc)ε0δ(x − x ′)1, (4.108)

where ωc is the centre frequency for suitably defined operators, AΔω(x, 0), EΔω(x, 0).The theory further reveals that the weak absorption gives rise to space-dependentmode operators that spatially progress according to quantum Langevin equations inthe direct space. As could be expected, the operators a±(x, ω), as the forward- andbackward-propagating fields, are governed by quantum Langevin equations, but itholds that the operator-valued Langevin noise is space dependent,

F±(x, ω) = ±1

i

√2Im{n(ω)}ω

cexp

[∓iRe{n(ω)}ω

cx]

f (x, ω). (4.109)

In other words, the operators a±(x, ω) progress in space.As an example of inhomogeneous structure, two bulk dielectrics with a common

interface are considered. The problem of determining a classical Green functionreappears. The verification of the commutation relation (4.105) is performed bystraightforward calculation, which is more complicated. A general proof of this rela-tion is not present, causality reasons are only pointed out. There exists a straightfor-ward generalization of the quantization method based on a mode expansion (Khos-ravi and Loudon 1991, 1992, Agarwal 1975). The behaviour of short light pulsespropagating in a dispersive absorbing linear dielectric with a special attention tosqueezed pulses has been studied (Schmidt et al. 1996).

Page 251: Quantum Aspects of Light Propagation

242 4 Microscopic Theories

4.2.2 Dispersive Lossy Nonlinear InhomogeneousDielectric

Emphasizing the important differences from the linear model, the Lagrangian andHamiltonian for the nonlinear dielectric are introduced by Schmidt et al. (1998).The Lagrangian density (4.2) has been denoted by Ll(r) and this relation with Lreplaced by Ll(r) has been utilized in the Lagrangian density in the relation

L(r) = Ll(r)+ Lnl(r), (4.110)

where moreover

Lnl(r) = f [X(r)]. (4.111)

While in the linear case it is sufficient to quantize only the transverse fields, in thenonlinear case such a procedure would result in a loss of generality. The result ofthe substitution from relations (4.31), (4.32), (4.33) into relation (4.30) which wehave denoted as H , we denote here as H⊥

l . The total Hamiltonian can be written as

H = Hl + Hnl, (4.112)

where the nonlinear interaction term Hnl is given by

Hnl = −∫

f [X(r)] d3r (4.113)

and the Hamiltonian Hl that governs the linear dynamics can be written as

Hl = H ‖l + H⊥

l , (4.114)

where

H ‖l =

∫ (�ω0b†(‖, k, t)b(‖, k, t)+

∫ ∞

0�ωb†

ω(‖, k, t)bω(‖, k, t) dω

)d3k

+ �

2

∫ ∫ ∞

0V (ω)[b†(‖,−k, t)+ b(‖, k, t)][b†

ω(‖, k, t)bω(‖,−k, t)] dω d3k

(4.115)

and the components b(‖, k, t), bω(‖,−k, t) must be appropriately defined (see,Schmidt et al. (1998)) for b‖(k), b‖(k, ω)). In general, Hnl couples the transverseand longitudinal fields, cf., relation (4.12).

Schmidt et al. (1998) have derived evolution equations for the field operators andshown that additional noise sources appear in the nonlinear terms. Linear relation-ships between quantum (operator-valued) fields are introduced following Huttner

Page 252: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 243

and Barnett (1992b) as well as Gruner and Welsch (1995). The relations hold for alltimes and both in linear and in nonlinear cases.

Schmidt et al. (1998) do not attempt at diagonalization of the nonquadraticalHamiltonian H , relation (4.112), the notation of which is still the same as of theHamiltonian in (4.30). They avoid the difficulty with the generalization of the def-initions (4.46) and (4.35) to the nonlinear functions of the operators a(k), B(k, ω),b(k), bω(k).

We now approach the following representations of the matter fields. The longitu-dinal matter field X‖(r) can be expressed in terms of the field f‖(r, ω) as

X‖(r) =√

2ρω0

∫ ∞

0[α∗0 (ω)− β∗0 (ω)]f‖(r, ω) dω + H. c., (4.116)

the transverse matter field X⊥(r) can be expressed in terms of the field f(r, ω) as

X⊥(r) =∫ ∞

0

ˆX⊥(r, ω) dω + H. c., (4.117)

where

ˆX⊥(r, ω) = ε0

α

[−iω[ε(ω)− 1] ˆA(r, ω)+

√�

πε0Im{ε(ω)} f(r, ω)

], (4.118)

with ˆA(r, ω) being connected with the field f(r, ω) as the solution of equation (4.95)and the explicit relation (4.97), and the vector-potential field

A(r) =∫ ∞

0

ˆA(r, ω) dω + H. c. (4.119)

If the validity of the expressions (4.119), (4.116), and (4.117) is related to the timeevolution of the kind of (4.56), we may be afraid that this correctness will not endurethe change to the nonlinear case. This change is reflected in the equations of motionfor the basic fields and the vector-potential field in the Heisenberg picture,

i�∂

∂tf‖(r, ω) = [f‖(r, ω), H ] = �ωf‖(r, ω)+ [f‖(r, ω), Hnl], (4.120)

i�∂

∂tf(r, ω) = [f(r, ω), H ] = �ωf(r, ω)+ [f(r, ω), Hnl], (4.121)

i�∂

∂tˆA(r, ω) = [ ˆA(r, ω), H] = �ω ˆA(r, ω)+ [ ˆA(r, ω), Hnl]. (4.122)

To the number of the relations, which nevertheless hold in linear and nonlinear cases,the nonhomogeneous Helmholtz equation belongs

Δ ˆA(r, ω)+ K2(ω) ˆA(r, ω) = ω

c2

√�

πε0Im{ε(ω)} f(r, ω). (4.123)

Page 253: Quantum Aspects of Light Propagation

244 4 Microscopic Theories

Here it is noticed that f(r, ω)≡ f(r, ω, t), ˆA(r, ω)≡ ˆA(r, ω, t). Respecting the nota-tion K 2(ω) = c−2ω2ε(ω) (cf., (4.65)), we can see that

K 2(ω) ˆA(r, ω, t) = K2(ω ˆ1) ˆA(r, ω, t), (4.124)

where ˆ1 is the identity superoperator and relation (4.122) implies that

ωˆ1 = ˆ1

∂t+ 1

�H×

nl , (4.125)

where we, for the sake of clarity, write ∂∂t to the right from the notation ˆ1 and the

action of H×nl on an operator O is defined by

H×nl O ≡ [Hnl, O]. (4.126)

Relation (4.125) can be written in the form

Δ ˆA(r, ω, t)+ K2(

ˆ1∂

∂t+ 1

�H×

nl

)ˆA(r, ω, t) = ω

c2

√�

πε0Im{ε(ω)} f(r, ω, t),

(4.127)where the elimination of the field X(r) using relations (4.116) and (4.118) indicates

new noise sources. All of the fields f‖(r, ω), f(r, ω), ˆA(r, ω) obey the nonlineardynamics. By integration of (4.127) over ω, an equation adequate to the linear andnonlinear cases is obtained. The wealth of operator-valued fields serves the expres-sion of the dispersion and absorption in the nonlinear medium. The basic equationsare applied to the one-dimensional case and propagation equations for the slowlyvarying field amplitudes of pulse-like radiation are derived. The scheme is relatedto the familiar model of classical susceptibilities and applied to the problem of prop-agation of quantized radiation in a dispersive and lossy Kerr medium. In the lineartheory it is possible to separate the two transverse polarization directions from eachother and from the longitudinal direction. As has already been stated, this is notpossible for nonlinear media. In practice, in a single-mode optical fibre, only onetransverse polarization direction will be excited. Then the total Hamiltonian (4.112)can be reduced to a one-dimensional single-polarization form.

Let us consider the propagation in the x direction of plane waves polarized in they direction. The one dimensionality of the problem permits one to decompose thefield ˆA(x, ω) into components ˆA+(x, ω) and ˆA−(x, ω), respectively, propagating inthe positive and negative x-directions,

ˆA(x, ω) = ˆA+(x, ω)+ ˆA−(x, ω), (4.128)

where ˆA±(x, ω) are the solutions of spatial equations of progression

∂xˆA±(x, ω) = ±iK (ω) ˆA±(x, ω)∓ iN

√Im{ε(ω)}

ε(ω)f (x, ω), (4.129)

Page 254: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 245

with the normalization factor N =√

4πε0Ac2 . We remark that ˆA±(x, ω)≡ ˆA±(x, ω, t).

Similarly as from relations (4.123) and (4.127), one can arrive from relation (4.129)at relation

∂xˆA±(x, ω, t) = ±iK

(ˆ1

∂t+ 1

�H×

nl

)ˆA±(x, ω, t)

∓ iN√

Im{ε(ω)}ε(ω)

f (x, ω, t). (4.130)

In analogy to (4.119), the operators A(+)± (x) can be introduced,

A(+)± (x) =

∫ ∞

0

ˆA±(x, ω) dω. (4.131)

Integrating (4.130) over ω, an equation appropriate to the linear and nonlinear casesis obtained.

Adequately to the derived equations which we consider to be mere approxima-tions in the nonlinear case, Schmidt et al. (1998) study the narrow-bandwidth fieldcomponents and narrow-bandwidth pulses. The theory has been applied to narrow-bandwidth pulses propagating in a dielectric with a Kerr-like nonlinearity.

4.2.3 Elaboration of Linear Theory

Dung et al. (1998) have developed three-dimensional quantization presented in partin Gruner and Welsch (1996a) concerning dispersive and absorbing inhomogeneousdielectric medium. The approach directly starts with the Maxwell equations in thefrequency domain for the macroscopic electromagnetic field. It is shown that theclassical Maxwell equations together with the constitutive relations except relation(4.71) can be transferred to quantum theory. On considering the charge and currentdensities, one concentrates oneself on the noise-charge and noise-current densities.The operator-valued noise-charge density ˆρ and the operator-valued noise-current

density ˆj are introduced, which are related to the operator-valued noise polariza-tion ˆP,

ˆρ(r, ω) = −∇ · ˆP(r, ω), (4.132)ˆj(r, ω) = −iω ˆP(r, ω). (4.133)

It follows from relations (4.132) and (4.133) that ˆρ and ˆj fulfil the equation of con-tinuity:

∇ · ˆj(r, ω) = iω ˆρ(r, ω). (4.134)

Page 255: Quantum Aspects of Light Propagation

246 4 Microscopic Theories

The source term ˆj is related to a bosonic vector field f by the relation like (4.97). Thecommutation relation (4.100) remains valid and relation (4.99) must be modified tothe form

[ fi (r, ω), f †j (r′, ω′)] = δi jδ(r− r′)δ(ω − ω′)1. (4.135)

It is pointed out that the current density ˆj is not transverse, because the wholeelectromagnetic field is considered. Hence, the vector field f assumed here is nottransverse as well and the spatial δ function in relation (4.135) is an ordinary δ

function instead of a transverse δ function.Relation (4.96) is an integral representation of the vector-potential operator.

Dung et al. (1998) start from the partial differential equation

∇ × ∇ × ˆE(r, ω)− ω2

c2ε(r, ω) ˆE(r, ω) = iωμ0

ˆj(r, ω), (4.136)

whose solution can be represented as (here and in part of what follows we use adifferent notation)

ˆE(r, ω) = iωμ0

∫G(r, s, ω) · ˆj(s, ω) d3s, (4.137)

where G(r, s, ω) is the tensor-valued Green function of the classical problem. Itsatisfies the equation

[∇r∇r − 1

(Δr + ω2

c2ε(r, ω)

)]·G(r, s, ω) = δ(r− s)1 (4.138)

together with appropriate boundary conditions. Dung et al. (1998) have derivedcommutation relations

[Ei (r), Bk(r′)] = �

πε0εkmj

∂x ′m

∫ ∞

−∞

ω

c2Gi j (r, r′, ω) dω, (4.139)

where εkmj is the Levi-Civita tensor and

Gi j (r, r′, ω) = ei ·G(r, r′, ω) · e j , (4.140)

and

[Ei (r), Ek(r′)] = 0 = [Bi (r), Bk(r′)]. (4.141)

In the sense of the Helmholtz theorem there exists a unique decomposition of the

electric field ˆE into a transverse part ˆE⊥ and a longitudinal part, ˆE‖, i.e. the Coulomb

gauge can be introduced, where ˆE⊥ = iω ˆA and ˆE‖ = −∇ ˆϕ. In the Coulomb gauge,

Page 256: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 247

the vector and scalar potentials ˆA and ˆϕ, respectively, are related to the electricfield as

ˆAi (r, ω) = 1

∫δ⊥i j (r− s) ˆE j (s, ω) d3s, (4.142)

∂xi

ˆϕ(r, ω) = −∫

δ‖i j (r− s) ˆE j (s, ω) d3s, (4.143)

where δ⊥i j and δ‖i j are the components of the transverse and longitudinal tensor-

valued δ functions

δ⊥(r) = δ(r)1+∇∇(4π |r|)−1, (4.144)

δ‖(r) = −∇∇(4π |r|)−1. (4.145)

It is recalled that A(r) and ε0˙A(r) are canonically conjugated field variables. On

the contrary, the complexity of the commutation relation (4.139) suggests that the“canonical” commutators are not so simple as we would expect by the definition.The commutation relation between the vector potential and the scalar potential is ascomplicated, when one and only one of these quantities is differentiated with respectto the time or comprises such a derivative. The simple commutation relations are

[ Ai (r), A j (r′)] = 0 = [ ˙Ai (r), ˙A j (r′)], (4.146)

[ϕ(r), ϕ(r′)] = 0 = [ϕ(r), ˙Ai (r′)]. (4.147)

Then, the theory is applied to the bulk dielectric such that the dielectric functioncan be assumed to be independent of space, ε(r, ω)= ε(ω) for all r. In this case, thesolution of equation (4.138) that satisfies the boundary condition at infinity is (cf.,Tomas 1995)

G(r, r′, ω) = [∇r∇r + K 2(ω)1]

K−2(ω)g(|r− r′|, ω), (4.148)

where

g(r, ω) = exp[iK (ω)r ]

4πr. (4.149)

Relation (4.139) can be simplified as

[Ei (r), Bk(r′)] = − i�

ε0εikm

∂xmδ(r− r′)1, (4.150)

and the “canonical” commutator corresponds to the definition

[ Ai (r), ˙A j (r′)] = i�

ε0δ⊥i j (r− r′)1. (4.151)

Page 257: Quantum Aspects of Light Propagation

248 4 Microscopic Theories

Moreover,

[ϕ(r), A j (r′)] = 0. (4.152)

The commutation relations presented are equal-time Heisenberg picture ones andtherefore it is emphasized that they are conserved. To make contact with the earlierwork, Dung et al. (1998) define the vectors

f⊥(r, ω) =∫

δ⊥(r− s) · f(s, ω) d3s, (4.153)

f‖(r, ω) =∫

δ‖(r− s) · f(s, ω) d3s. (4.154)

The commutation relations (4.135) and (4.100) imply that

[ f ⊥(‖)i (r, ω), ( f ⊥(‖)

j (r′, ω′))†] = δ⊥(‖)i j (r− r′)δ(ω − ω′)1, (4.155)

[ f ⊥(‖)i (r, ω), f ⊥(‖)

j (r′, ω′)] = [ f ⊥i (r, ω), ( f ‖j (r′, ω′))†] = 0. (4.156)

The representation of transverse vector potential simplifies to

ˆA(r, ω, 0) = μ0

∫g(|r− r′|, ω)ˆj⊥(r′, ω, 0) d3r′. (4.157)

It can be derived that the scalar potential operator

ˆϕ(r, ω, 0) = 1

4πε0ε(ω)

∫ ˆρ(s, ω, 0)

|r− s| d3s, (4.158)

where ˆρ(r, ω, 0) = (iω)−1∇ · ˆj‖(r, ω, 0).Another application is the quantization of the electromagnetic field in an inhomo-

geneous medium that consists of two bulk dielectrics with a common interface. Thedetermination of the tensor-valued Green function for three-dimensional configura-tion of dielectric bodies is a very involved problem, in general. Dung et al. (1998)return to the simple configuration which was mentioned in Gruner and Welsch(1996a). It is referred to Tomas (1995) for the classical treatment of multilayer struc-tures. It is shown that for the configuration under study, the commutation relations(4.150), (4.151), and (4.152) hold.

The necessity of a new calculation of the quantum electrodynamical commu-tation relations for a new three-dimensional configuration (cf., Dung et al. 1998)is not absolute. Scheel et al. (1998) have proven that the fundamental equal-timecommutation relations of quantum electrodynamics are preserved for an arbitrarilyspace-dependent Kramers–Kronig dielectric function.

Let us recall that the complex-valued dielectric function ε(r, ω) depends on fre-quency and space,

ε(r, ω) → 1, if ω →∞. (4.159)

Page 258: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 249

It is assumed that the real part (responsible for dispersion) and the imaginary part(responsible for absorption) are related to each other according to the Kramers–Kronig relations, because of causality. This also implies that ε(r, ω) is a holomorficfunction in the upper complex half-plane of frequency

∂ω∗ε(r, ω) = 0, Im ω > 0. (4.160)

Scheel et al. (1998) study relation (4.139). By comparison of the right-hand sides ofthis relation and relation (4.150), they arrive at the identity to be proven

−∫ ∞

−∞

ω

c2G(r, r′, ω) dω × ←

∇r′ = −iπ1δ(r− r′)× ←∇r′ . (4.161)

Here the left arrow means that the operators ∂∂x ′m

will first be written as←∂

∂x ′min the

expansion of the Hamilton operator, ∇, with this upper limit.Based on the partial differential equation (4.138) for the tensor-valued Green

function, an integral equation will be presented in what follows. The partial differ-ential equation and the boundary condition at infinity determine the Green functionuniquely. By comparison of relation (4.137) with a constitutive relation, we couldderive that iμ0ωGi j (r, s, ω) are holomorphic functions of ω in the upper complexhalf-plane, i.e.

∂ω∗[ωGk j (r, s, ω)

] = 0, Im ω > 0, (4.162)

with

ωGk j (r, s, ω) → 0 if |ω| → ∞. (4.163)

Second derivation of the Cauchy–Riemann equation (4.162) consists in the appli-cation of ∂

∂ω∗ to relation (4.138). The left-hand side of relation (4.162) is then theunique solution of the homogeneous problem.

Knoll and Leonhardt (1992) calculate the time dependent, let us say a direct-space Green function. This could be useful in the combination with a time-dependent(direct-space) noise fdir(r, s). Scheel et al. (1998) have derived the relation

iμ0ωGi j (r, s, ω) = Di j (r, s, ω) =∫ ∞

0eiωτ Di j (r, s, τ ) dτ, (4.164)

where Di j (r, s, τ ) are components of the tensor-valued response function thatcausally relates the electric field E(r, t) to an external current jext(s, t − τ ), so that

Di j (r, s, τ ) = 1

∫ ∞

−∞e−iωτ Di j (r, s, ω) dω

= −μ0∂

∂τGi jdir(r, s, τ ), (4.165)

Page 259: Quantum Aspects of Light Propagation

250 4 Microscopic Theories

where

Gi jdir(r, s, τ ) ≡ 1

∫ ∞

−∞e−iωτ Gi j (r, s, ω) dω (4.166)

is the direct-space Green function.From the theory of partial differential equations it is known (see, e.g. Garabe-

dian (1964)) that there exists an equivalent formulation of the problem in terms

of an integral equation. On introducing ε0(ω) ≡∫

ε(r,ω) d3r∫d3r , an appropriately space-

averaged reference relative permittivity, the integral equation for the tensor-valuedGreen function can be written as

G(r, s, ω) = G(0)(r, s, ω)+∫

K(r, v, ω) ·G(v, s, ω) d3v, (4.167)

where

G(0)(r, s, ω) = [1−∇r∇s K−2(s, ω)]g(|r− s|, ω), (4.168)

K(r, v, ω) = [∇rg(|r− v|, ω)][∇v ln(K 2(v, ω)

)]

+ [K 2(v, ω)− K 20 (ω)]g(|r− v|, ω)]1. (4.169)

Here g(r, 0) ≡ g0(r, 0) is given by (4.149), where K (ω) ≡ K0(ω),

K 2(r, ω) = ω2

c2ε(r, ω), (4.170)

K 20 (ω) = ω2

c2ε0(ω). (4.171)

It can be seen that the components of the kernel Kik(r, v, ω) are holomorphic func-tions of ω in the upper complex half-plane, with

Kik(r, v, ω) → 0 if |ω| → ∞. (4.172)

To prove the fundamental commutation relation (4.150), we first decompose thetensor-valued Green function into two parts,

G(r, s, ω) = G1(r, s, ω)+G2(r, s, ω), (4.173)

where G1(r, s, ω) satisfies the integral equation

G1 = G(0)1 +

∫K ·G1 d3v, (4.174)

with

G(0)1 (r, s, ω) = g(|r− s|, ω)1, (4.175)

G2(r, s, ω) = Γ(r, s, ω)←∇s. (4.176)

Page 260: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 251

In relation (4.176) Γ is the solution of the integral equation

Γ = Γ(0) +∫

K · Γ d3v, (4.177)

with

Γ(0)(r, s, ω) = −∇r[K−2(s, ω)g(|r− s|, ω)]. (4.178)

Scheel et al. (1998) derive that iμ0ωG1 and μ0ω2Γ are the Fourier transforms of

the response functions to the noise-current density and the noise-charge density,respectively.

Combining relations (4.173) and (4.176) and recalling that←∇r′ ×

←∇r′ = 0, we see

that the left-hand side of relation (4.161) can be rewritten as

−∫ ∞

−∞

ω

c2G(r, r′, ω) dω × ←

∇r′ = −∫ ∞

−∞

ω

c2G1(r, r′, ω) dω × ←

∇r′ . (4.179)

Thus, only the noise-current response function iμ0ωG1 contributes to commutator(4.139). Multiplying the integral equation (4.174) by the function ω

c2 and integrat-ing over ω, we obtain as the derivation of relation (4.105) from the holomorphicproperties of the tensors K and ωG1 that

∫ ∞

−∞

ω

c2G1(r, r′, ω) dω = iπ1δ(r− r′). (4.180)

The outer product of this equation and the operator (−←∇r′ ) can be taken and together

with relation (4.179) implies relation (4.161).In addition, it is shown that the scheme also applies to media with both absorption

and amplification (in a bounded region of space). An extension of the quantizationscheme to linear media with bounded regions of amplification is given and theproblem of anisotropic media is briefly addressed, for which the permittivity is asymmetric complex tensor-valued function of ω,

εi j (r, ω) = ε j i (r, ω). (4.181)

Extensions of previous work on the propagation in absorbing dielectrics tooklinear amplification into account (Jeffers et al. 1996, Matloob et al. 1997, Artoniand Loudon 1998).

Knoll et al. (1999) investigated quantum-state transformation by dispersive andabsorbing four-port devices. Under the usual assumptions on the dielectric per-mittivity, quantization of the Hamiltonian formalism of the electromagnetic fieldusing a method close to the microscopic approach was performed by Tip (1998).A proper definition of band gaps in the periodic case and a new continuity equa-tion for energy flow were obtained, and an S-matrix formalism for scattering from

Page 261: Quantum Aspects of Light Propagation

252 4 Microscopic Theories

absorbing objects was worked out. In this way the generation of Cerenkov andtransition radiation have been investigated. A path-integral formulation of quan-tum electrodynamics in a dispersive and absorbing dielectric medium has beenpresented by Bechler (1999) and has been applied on the microscopic level to thequantum theory of electromagnetic fields in dielectric media. Results concerningquantum electrodynamics in dispersing and absorbing dielectric media have beenreviewed by Knoll et al. (2001). Tip et al. (2001) have proven the equivalence oftwo methods for quantization of the electromagnetic field in general dispersing andabsorbing linear dielectrics: the Langevin-noise-current method and the auxiliaryfield method.

Petersson and Smith (2003) have illustrated the role of evanescent waves inpower calculations for counterpropagating beams. In classical optics the field ofa beam can be represented in terms of its plane-wave spectrum (Smith 1997, Clem-mow 1996). Counterpropagating and “counter-evanescent” plane waves are definedrelative to a selected plane. When a line current is placed over a dielectric slab, it isappropriate to insert a plane between the line current and the slab. The time-averagepower passing through a plane is a sum of powers contributed by the propagatingplane waves and by “counter-evanescent” pairs of plane waves with the same trans-verse components of their wave vectors.

Suttorb and Wubs (2004) have provided a microscopic justification of the phe-nomenological quantization scheme for the electromagnetic field in inhomogeneousdielectric due to Gruner and Welsch (1995) (cf., references in subsection 4.2.3).Matloob (2004a) has paid attention to a damped harmonic oscillator. He has useda macroscopic Langevin equation for it. A canonical quantization scheme for theLangevin equation has been provided. A macroscopic electromagnetic field hasbeen quantized in a homogeneous linear isotropic dielectric by the association ofa damped quantum-mechanical harmonic oscillator with each mode of the radia-tion field. Matloob (2004b) has introduced a particular damped harmonic oscilla-tor. He has used an appropriate form of the macroscopic Langevin equation. Thecanonical quantization scheme has been followed. A macroscopic electromagneticfield has been quantized in a linear isotropic permeable dielectric medium by asso-ciating a damped quantum-mechanical oscillator with each mode of the radiationfield.

In Matloob (2005) a homogeneous medium is assumed that is isotropic in itsrest frame. One works with positive frequency parts of the fields. The Minkowskirelations are presented which are generalized constitutive relations for uniformlymoving media. The electric induction vector depends also on the magnetic strengthvector and the magnetic-induction vector depends also on the electric strength vec-tor. Using the Minkowski relations the Maxwell–Minkowski equations are derived.The field vectors E and B are expressed in terms of the vector potential in the Weylgauge. A time-independent wave equation with the noise polarization and noisemagnetization for the vector potential is derived. It is shown that the constitutiverelations may be convenient, with the electric induction vector independent of themagnetic strength vector and the magnetic-induction vector independent of the elec-tric strength vector, on considering the anisotropy of the material in the laboratory

Page 262: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 253

frame. The Green tensor is studied in reciprocal and spatial coordinate space. Thefields are quantized by expressing the noise-current density in terms of two infinitesets of appropriately chosen bosonic field operators. The vacuum field fluctuation isexpressed.

4.2.4 Optical Field at Dielectric Devices

Matloob et al. (1995) provided expressions for the electromagnetic-field operatorsfor three geometries: an infinite homogeneous dielectric, a semi-infinite dielectric,and a dielectric slab. A microscopic derivation has shown that a canonical quantumtheory of light at the dielectric–vacuum interface is possible Barnett, Matloob, andLoudon (1995).

A simple quantum theory of the beam splitter, which can be applied to a Fabry–Perot resonator, was introduced by Barnett et al. (1996) and developed by Barnettet al. (1998).

Artoni and Loudon (1997) applied the Huttner–Barnett scheme for quantizationof the electromagnetic field in dispersive and absorbing dielectrics for the calcu-lations of the effects of perpendicular propagation in a dielectric slab and to theproperties of the incident light pulse. Their approach has provided a deeper under-standing of antibunching (Artoni and Loudon 1999). Brun and Barnett (1998) con-sidered an experimental set-up using a two-photon interferometer, where insertionof a dielectric into one or both arms of the interferometer is essential.

Suggestive is a comparative study of fermion and boson beam splitters (Loudon1998). Fermions can be studied in analogy with bosons (Cahill and Glauber 1999).

Di Stefano et al. (1999) extended the field quantization to these material systemswhose interaction with light is described, near a medium boundary, by a nonlocalsusceptibility. Di Stefano et al. (2000) have developed a quantization scheme forthe electromagnetic field in dispersive and lossy dielectrics with planar interface,including propagation in all the spatial directions, and considering both the trans-verse electric and the transverse magnetic polarized fields. Di Stefano et al. (2001a)have presented a one-dimensional scheme for the electromagnetic field in arbitraryplanar dispersing and absorbing dielectrics, taking into account their finite extent.They have derived that the complete form of the electric-field operator includesa part that corresponds to the free fields incident from the vacuum towards themedium and a particular solution which can be expressed by using the classicalGreen-function integral representation of the electromagnetic field. By expressingthe classical Green function in terms of the classical light modes, they have obtaineda generalization of the method of modal expansion (e.g. Knoll et al. (1987)) toabsorbing media. Di Stefano et al. (2001b) have based an electromagnetic-fieldquantization scheme on a microscopic linear two-band model. They have derivedfor the first time a noise-current operator for general anisotropic and/or spatiallynonlocal media, which can be described only in terms of an appropriate frequency-dependent susceptibility.

Page 263: Quantum Aspects of Light Propagation

254 4 Microscopic Theories

The Green-tensor formalism is well suited to studying the behaviour of thequantized electromagnetic field in the presence of dispersing and absorbing bodies.Especially, it has been applied successfully to the study of input–output relations(Gruner and Welsch 1996b). As a continuation of this work, Khanbekyan et al.(2003) studied the quantized field in the presence of a dispersing and absorbingmultilayered planar structure (shortly, multilayer plate). Three-dimensional input–output relations have been derived for frequency components of the electric-fieldoperator in the transverse reciprocal space. Input–output relations for frequencycomponents of this operator in the direct space have been given as well. The condi-tions have been stated, under which the input–output relations can be expressed interms of bosonic operators. These relations have been discussed for the case of theplate being surrounded by vacuum. The theory applies to effectively free fields andthose created by active atomic sources inside and/or outside the plate.

Khanbekyan et al. (2003) consider n − 1 layers with thicknesses d j , j =1, . . . , n − 1, the region on the left of the plate ( j = 0), and the region on theright of the plate ( j = n). The permittivity is

ε(z, ρ, ω) =n∑

j=0

λ j (z)ε j (ω), independent of ρ, (4.182)

where ρ = (x, y) and

λ j (z) ={

1, if z ∈ j th region,0, otherwise.

(4.183)

For simplicity, we shift the fields G(r, r′, ω), E(r, ω), and f(r, ω) and all the regionsalong the z-axis such that the j th region has the left boundary plane going throughthe origin ( j > 0) or the right boundary plane going through the origin ( j = 0). Theresult of the shift of respective fields will be denoted by G( j j ′)(r, r′, ω), E( j)(r, ω),and f( j)(r, ω). If Θ(z) is the unit-step function, then

[Θ(z − z′)]( j j ′) ={

Θ( j − j ′) for j �= j ′,Θ(z − z′) for j = j ′, (4.184)

[Θ(z′ − z)]( j j ′) ={

Θ( j ′ − j) for j �= j ′,Θ(z′ − z) for j = j ′. (4.185)

Since the Green tensor depends only on the difference ρ−ρ ′, it can be representedas a two-dimensional Fourier integral

G( j j ′)(r, r′, ω) = 1

(2π )2

∫eik·(ρ−ρ ′)G( j j ′)(z, z′, k, ω) d2k, (4.186)

where k = (kx , ky) is the wave vector parallel to the interfaces and

G( j j ′)(z, z′, k, ω) =∫

e−ik·σ G( j j ′)(z, z′, σ ≡ ρ − ρ ′, ω) d2σ . (4.187)

Page 264: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 255

The electric field operator E( j)(r, ω) may be written as a twofold Fourier transform:

E( j)(r, ω) = 1

(2π )2

∫eik·ρ E( j)(z, k, ω) d2k, (4.188)

where

E( j)(z, k, ω) =∫

e−ik·ρ E( j)(z, ρ, ω) d2ρ, (4.189)

and the bosonic field operator may be written in the integral form as

f( j)(r, ω) = 1

(2π )2

∫eik·ρ f( j)(z, k, ω) d2k, (4.190)

where

f( j)(z, k, ω) =∫

e−ik·ρ f( j)(z, ρ, ω) d2ρ. (4.191)

The Green tensor G( j j ′)(z, z′, k, ω) by the paper (Tomas 1995) may be written as

G( j j ′)(z, z′, k, ω) = −ezδ j j ′

k2j

ezδ(z − z′)+ g( j j ′)(z, z′, k, ω), (4.192)

where

g( j j ′)(z, z′, k, ω) = i

2

q=p,s

σq{E j>

q (z, k, ω)Ξ j j ′q E j ′<

q (z′,−k, ω)[Θ(z − z′)]( j j ′)

+ E j<q (z, k, ω)Ξ j ′ j

q E j ′>q (z′,−k, ω)[Θ(z′ − z)]( j j ′)}, (4.193)

(σp = 1, σs = −1). In equation (4.193)

E j>q (z, k, ω) = e( j)

q+(k)eiβ j (z−d j ) + rqj/ne( j)

q−(k)e−iβ j (z−d j ), (4.194)

E j<q (z, k, ω) = e( j)

q−(k)e−iβ j z + rqj/0e( j)

q+(k)eiβ j z (4.195)

and

Ξ j j ′q = 1

βntq0/n

tq0/j e

iβ j d j

Dq j

tq0/j ′e

iβ j ′d j ′

Dq j ′, (4.196)

with

Dq j = 1− rqj/0rq

j/ne2iβ j d j , (4.197)

Page 265: Quantum Aspects of Light Propagation

256 4 Microscopic Theories

(d0 = dn = 0) and

β j =√

k2j − k2 = β ′j + iβ ′′j (β ′j , β ′′j ≥ 0) (4.198)

(k = |k|), where

k j =√

ε j (ω)ω

c= k ′j + ik ′′j (k ′j , k ′′j ≥ 0), (4.199)

and tqj/j ′ = β j

β j ′tq

j ′/j and rqj/j ′ , respectively, the transmission and reflection coefficients

between the regions j ′ and j . The unit vectors e( j)q±(k) in equations (4.194) and

(4.195) are the polarization unit vectors for transverse electric (q = s) and transversemagnetic (q = p) waves

e( j)s±(k) = k

k× ez, (4.200)

e( j)p±(k) = 1

k j

(∓β j

kk+ kez

). (4.201)

If both the incoming fields incident on the two boundary planes of the plate areknown, as well as the fields generated inside the plate, one can calculate the fieldsoutgoing from the two boundary planes by means of input–output relations. Theserelations are valid also for evanescent-field components.

To obtain generally valid input–output relations, we restrict our attention to theelectric-field operator. This operator in front of the structure (superscript 0) andbehind the structure (superscript n) is decomposed in terms of input and outputamplitude operators

E(0)(z, k, ω) =∑

q=p,s

[e(0)

q+(k)E (0)qin(z, k, ω)

+ e(0)q−(k)E (0)

qout(z, k, ω)], (4.202)

E(n)(z, k, ω) =∑

q=p,s

[e(n)

q−(k)E (n)qin(z, k, ω)

+ e(n)q+(k)E (n)

qout(z, k, ω)]. (4.203)

Here, the operators

E (0)qin(z, k, ω) = −μ0ω

2β0eiβ0z

∫ z

−∞e−iβ0z′ j(0)(z′, k, ω) · e(0)

q+(k) dz′ (4.204)

and

E (n)qin(z, k, ω) = −μ0ω

2βne−iβn z

∫ ∞

zeiβn z′ j(n)(z′, k, ω) · e(n)

q−(k) dz′ (4.205)

Page 266: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 257

are input amplitude operators, where the Green function is of a very simple form.The operators

E (0)qout(z, k, ω) = e−iβ0z E (0)

qout(k, ω)

+ μ0ω

2β0e−iβ0z

∫ 0

zeiβ0z′ j(0)(z′, k, ω) · e(0)

q−(k) dz′ (4.206)

and

E (n)qout(z, k, ω) = eiβn z E (n)

qout(k, ω)

+ μ0ω

2βneiβn z

∫ z

0e−iβn z′ j(n)(z′, k, ω) · e(n)

q+(k) dz′ (4.207)

are output amplitude operators, where the Green function has the complicated formand is “hidden” in the input–output relations

(E (0)

qout(k, ω)E (n)

qout(k, ω)

)=(

rq0/n(k, ω) tq

n/0(k, ω)tq0/n(k, ω) rq

n/0(k, ω)

)(E (0)

qin(k, ω)

E (n)qin(k, ω)

)

+n−1∑

j=1

( j)q0+(k, ω) φ

( j)q0−(k, ω)

φ( j)qn+(k, ω) φ

( j)qn−(k, ω)

)(E ( j)

q+(k, ω)E ( j)

q−(k, ω)

). (4.208)

They relate the output amplitude operators at the boundary planes of the plate to theinput amplitude operators at these planes,

E (0)qin,out(k, ω) = E (0)

qin,out(z, k, ω)∣∣∣z=0−

, (4.209)

E (n)qin,out(k, ω) = E (n)

qin,out(z, k, ω)∣∣∣z=0+

, (4.210)

and to the amplitude operators associated with the layers

E ( j)q±(k, ω) = −μ0ω

2β j

∫ d j

0e∓iβ j z′ j( j)(z′, k, ω) · e( j)

q±(k) dz′, (4.211)

j = 1, 2, . . . , n − 1, which may have been better denoted say by F ( j)q±(k, ω), since

E ( j)q+(k, ω) �= E ( j)

q+(z, k, ω)∣∣∣z=d−j

. (4.212)

E ( j)q−(k, ω) �= E ( j)

q−(z, k, ω)∣∣∣z=0+

. (4.213)

Page 267: Quantum Aspects of Light Propagation

258 4 Microscopic Theories

In equation (4.208), the coefficients at the operators (4.211) read

φ( j)q0+ =

tqj/0e2iβ j d j

Dq jrq

j/n, φ( j)q0− =

tqj/0

Dq j, (4.214)

φ( j)qn+ =

tqj/neiβ j d j

Dq j, φ

( j)qn− =

tqj/0eiβ j d j

Dq jrq

j/n. (4.215)

It is worth noting that any two planes z = z(0) ≤ 0− and z = z(n) ≥ 0+ forj = 0 and j = n, respectively, also can be used in principle for a formulation of theinput–output relations. The theory applies to effectively free fields and those createdby active atomic sources inside and/or outside the plate.

4.2.5 Modification of Spontaneous Emission by Dielectric Media

Scheel et al. (1999a) have found quantum local-field corrections appropriate tothe spontaneous emission by an excited atom. Dung et al. (2000) have developeda formalism for studying spontaneous decay of an excited two-level atom in thepresence of arbitrary dispersing and absorbing dielectric bodies. They have shownhow the minimal-coupling Hamiltonian simplifies to a Hamiltonian in the dipoleapproximation. The formalism is based on a source-quantity representation of theelectromagnetic field in terms of the tensor-valued Green function of the classicalproblem and appropriately chosen bosonic quantum fields. All relevant informationabout the bodies such as form and dispersion and absorption properties is containedin the tensor-valued Green function. This function has been available for variousconfigurations such as planarly, spherically, and cylindrically multilayered media(Chew 1995).

The theory has been applied to the spontaneous decay of a two-level atom placedat the centre of a three-layer spherical microcavity, the wall being modelled by aLorentz dielectric. The tensor-valued Green function of the configuration has beenknown (Li et al. 1994). The calculations have been performed on the assumptionof a dielectric with a single resonance. For simplicity, it has been assumed that theatom is positioned at the centre of the cavity. Weak and strong couplings are studiedand in the study of the strong couplings both the normal-dispersion range and theanomalous-dispersion range associated with the band gap are considered. Whereasin the range of normal dispersion, the cavity input–output coupling dominates thestrength of the atom–field interaction, the significant effect within the band gap isthe photon absorption by the wall material.

Dung et al. (2001) have studied nonclassical decay of an excited atom near adispersing and absorbing microsphere of given complex permittivity that satisfiesthe Kramers–Kronig relations laying emphasis on a Drude–Lorentz permittivity.Among others, they have found a condition on which the decay becomes purely non-radiative. Dung et al. (2002a,b) have given a rigorous quantum-mechanical deriva-tion of the rate of intermolecular energy transfer in the presence of dispersing and

Page 268: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 259

absorbing media with spatially varying permittivity. They have applied the theory tobulk material, multislab planar structures, where they also have made comparisonwith experiments, and to microspheres. They have shown that the minimal-couplingscheme and the multipolar-coupling scheme yield exactly the same form of the rateformula.

Tip (2004) has used his auxiliary field method to obtain various equivalentHamiltonians for charged particles interacting with absorptive dielectrics. In twosteps, the representations cease to manifest the generalized Coulomb gauge used,but it remains in one term, concentrated in a wave operator. It has also been shownfor excited atoms in a photon crystal with transition frequency in a band gap thattheir states do not decay radiatively.

For a transparent dielectric, theoretical studies can take a traditional approach.Inoue and Hori (2001) have developed a formalism of quantization of electro-magnetic fields including evanescent waves based on the detector-mode functionaldefined in terms of those for the widely used triplet modes. They have evaluatedatomic and molecular radiation near a dielectric boundary surface. Matloob andPooseh (2000) have discussed a fully quantum-mechanical theory of the scatteringof coherent light by a dissipative dispersive slab. Matloob and Falinejad (2001) havecalculated the Casimir force between two dielectric slabs by using the notion of theradiation pressure associated with the quantum electromagnetic vacuum. Specifi-cally, they have used the fact that only the field correlation functions are neededfor the evaluation of vacuum radiation pressure on an interface. Matloob (2001) haspostulated an electromagnetic-field Lagrangian density at each point of space–timeto be of an unfamiliar form comprising the noise-current density. He has expressedthe displacement D(r, t) merely in terms of the electric-field E(r, t ′), t ′ ≤ t , withoutadding a noise polarization term.

In the framework of a semiclasical approach, Paspalakis and Kis (2002) havestudied the propagation dynamics of N laser pulses interacting with an (N+1)-levelquantum system (one upper state and N lower states). Assuming the system to bein a superposition state of all of the lower levels initially they have determined theconditions of complete opacity or transparency of the medium. The coupling ofpulses is most interesting in the limit of parametric generation.

A simplified approach to the quantization is sufficient for the theory of theradiation pressure on dielectric surfaces (Loudon 2002). Loudon (2003) continues(Loudon 2002) with two changes or extensions. Instead of a planar pulse he consid-ers Laguerre–Gaussian light beams. He considers the transfer of angular momentumto a dielectric. He may issue from the book (Allen et al. 2003) and arrive at the paper(Padgett et al. 2003).

New forces are produced by a pulse of Laguerre–Gaussian light in comparisonwith a plane-wave pulse, which produces only a longitudinal force. These are radialand azimuthal forces. A simplification is achieved by assuming that the modal func-tion has zero radial index. The pulse is assumed to contain a single photon. In casean interface is considered the propagation direction of the pulse is assumed to beperpendicular to the surface. If the pulse is propagated into a dielectric, then it isassumed that the dielectric is – weakly – attenuating to ensure that the model need

Page 269: Quantum Aspects of Light Propagation

260 4 Microscopic Theories

not complicate by including any exit or reflection, or finiteness of the dielectric inthe direction of propagation.

Loudon (2003) gives Laguerre–Gaussian light in terms of the Lorentz-gauge vec-tor potential. He does not speak of the associated Laguerre polynomials, but it isobvious that the degree of a polynomial is zero, when he restricts himself to theradial index p = 0.

The theory is quantized. The single-photon pulse is represented by the statevector |1〉. The spectrum of the photon wave packet is a narrow-band Gaussianfunction. The author introduces the normal-order Poynting operator :S(r, t):, withz-component denoted by :Sz(r, t):. If the dispersion is ignored, the author can writethe expectation value

〈1|:Sz(r, t):|1〉 = �ω0c

L

√2

πexp

[−2c2

L2

(t − ηz

c

)2]|u|2, (4.216)

where L is a conventional length of the pulse, ω0 is a central frequency of the wavepacket, η ≡ η(ω0) is a refractive index, and u ≡ uk0,l(r) is the modal function,k0 = η(ω0)ω0

c is an angular wave number, and l is the orbital angular-momentumquantum number. The time integral of equation (4.216) is

∫ ∞

−∞〈1|:Sz(r, t):|1〉 dt = �ω0|u|2. (4.217)

Further, the author constructs the normally ordered angular-momentum densityoperator. He introduces the Lorentz force-density operator :f(r, t): and lets : fz(r, t):denote the longitudinal component of this operator. He determines that

〈1|: fz(r, t):|1〉 = −2�ω0c(η2 − 1)

L3

√2

π

(t − ηz

c

)

× exp

[−2c2

L2

(t − ηz

c

)2]|u|2. (4.218)

Similarly, he lets : fρ(r, t): denote the radial component of the operator :f(r, t):.As usual, ρ =

√x2 + y2. He determines that

〈1|: fρ(r, t):|1〉 = 2�ω0(η2 − 1)√2πηL

( |l|ρ− 2ρ

w20

)exp

[−2c2

L2

(t − ηz

c

)2]|u|2.

(4.219)The radial force compresses the dielectric towards the cylinder of radius ρ0 =w0

√|l|2 , where w0 is the beam waist.

Page 270: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 261

Similarly, he lets : fφ(r, t): denote the azimuthal component of the operator:f(r, t):. As usual, φ = arg(x + iy). He determines that

〈1|: fφ(r, t):|1〉 = −4�c2(η2 − 1)

ηL3

√2

π

(t − ηz

c

)

× exp

[−2c2

L2

(t − ηz

c

)2](

l

ρ− σ |l|

ρ+ 2σρ

w20

)|u|2, (4.220)

where σ is the spin angular-momentum quantum number of the beam.The author extends the theory to the case, where space is divided into two regions

with a dielectric of real refractive index η0(ω) at z < 0 and a dielectric of a complexrefractive index

n(ω) = η(ω)+ iκ(ω) (4.221)

at z > 0. The modification of the result (4.217) at z > 0 is∫ ∞

−∞〈1|:Sz(r, t):|1〉 dt = �ω0 exp

(−2ω0κz

c

)4η0η

(η0 + η)2 + κ2|u|2, (4.222)

where η0, η, and κ are evaluated at frequency ω0.The author gives particular attention to the transfer of longitudinal and angular

momentum to the dielectric from light incident from free space. He lets the totalforce on dielectric in {z > 0} and at time t be represented by the force operator

F(t) =∫

z>0f(r, t) dr (4.223)

and lets Fz(t) denote the longitudinal component of this operator. The time-integratedforce, or the total linear momentum transfer to the dielectric, is

∫ ∞

−∞〈1|:Fz(t):|1〉 dt = 2�ω0

c

⎧⎪⎪⎨

⎪⎪⎩

η2 + κ2 − 1

(η + 1)2 + κ2︸ ︷︷ ︸

surface

+ 2

(η + 1)2 + κ2︸ ︷︷ ︸

bulk

⎫⎪⎪⎬

⎪⎪⎭

= 2�ω0

c

η2 + 1+ κ2

(η + 1)2 + κ2︸ ︷︷ ︸

total

. (4.224)

In pursuit of the torque the author first introduces the operator that represents thedensity of the z-component of the torque on the dielectric :gz(r, t):. Next he lets thetotal torque on the dielectric in {z > 0} and at time t be represented by the torqueoperator

Gz(t) =∫

z>0gz(r, t) dr. (4.225)

Page 271: Quantum Aspects of Light Propagation

262 4 Microscopic Theories

The time-integrated torque, or total angular-momentum transfer on dielectric, is

∫ ∞

−∞〈1|:Gz(t):|1〉 dt = 4�(l + σ )η

(η + 1)2 + κ2

⎧⎪⎪⎨

⎪⎪⎩

η2 + κ2 − 1

η2 + κ2︸ ︷︷ ︸

surface

+ 1

η2 + κ2︸ ︷︷ ︸

bulk

⎫⎪⎪⎬

⎪⎪⎭

= 4�(l + σ )η

η2 + 1+ κ2︸ ︷︷ ︸

total

. (4.226)

It is shown that it is meaningful to divide the total transfer of linear momentuminto surface reflected, surface transmitted, and bulk transmitted contributions pro-vided a photon has passed through. From this the shift of a slab of mass M maybe calculated, due to a normally incident single-photon wave packet. Similarly (butfor κ = 0 only) the angular rotation of the dielectric slab, whose moment of inertiaaround the z-axis is denoted I may be calculated, due to such a wave packet.

A scheme for transferring quantum states from the propagating light fields toa macroscopic, collective vibrational degree of freedom of a massive mirror hasbeen proposed (Zhang et al. 2003). The proposal may realize an Einstein–Podolsky–Rosen state in position and momentum for a pair of massive mirrors at distinctlocations by exploiting a nondegenerate optical parametric amplifier. Loudon et al.(2005) have paid attention to the photon drag effect. The momentum transfer fromlight to a dielectric material has been calculated by evaluation of the relevant Lorentzforce. The photon drag effect is named after the generation of currents or electricfields in semiconductors.

Leonhardt and Piwnicki (2001) have analysed the propagation of slow light inmoving media in the case where the light is monochromatic in the laboratory frame.The extremely low group velocity is caused by the electromagnetically inducedtransparency of an atomic transition. Lombardi (2002) has re-examined the physicalsignificance of different velocities which can be introduced for a wave train. Ough-stun and Cartwright (2005) have compared the group velocity with the instantaneouscentroid velocity of the pulse Poynting vector for an ultrashort Gaussian pulse. Verylong pulses that are well tuned to a region of anomalous dispersion do not havesuperluminal peak velocity of a real physical significance. Yanik and Fan (2005)have formulated basic principles that underlie stopping and storing light coherentlyin many different physical systems. Following a brief discussion of one of knownatomic stopping-light schemes, an all-optical scheme has been analysed in detail.

Tip (2007) has studied the properties of atoms close to an absorptive dielectricusing his quantized form of the phenomenological Maxwell equations. The authorhas treated the coupling of atoms with longitudinal modes in detail. The atomicinteraction potential changes from the Coulomb one to a static potential, i.e. onethat obeys a Poisson equation with zero-frequency limit of the permittivity. Thelongitudinal interactions of atoms with absorptive dielectrics are responsible fornonradiative decay of the atoms. It has been found that the Hamiltonian used by

Page 272: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 263

Dung et al. (2002b) is unitarily equivalent to a special case of the Hamiltonian usedby Tip (2007).

4.2.6 Left-Handed Materials

The interest in “left-handed” materials is reflected both in the theory and in theexperiment. Veselago (1967, 1968) was the first to address the question of propaga-tion of the electromagnetic waves in the medium with both the permittivity ε0ε andthe permeability μ0μ negative. Veselago has shown that, although such materials arenot available in the nature, their existence is not excluded by the Maxwell equationsMarkos 2005).

First we summarize the basic electromagnetic properties of left-handed materials.The propagation of the electromagnetic wave, E = E0ei(k·r−ωt), is described by thewave equation

[k2 − ω2

c2εμ

]E(r, t) = 0, (4.227)

where k is the wave vector and ω is the frequency,

k2 = ω2

c2εμ. (4.228)

Propagation is possible if k2 > 0. This is always true in dielectrics, where μ = 1and ε is real, while no propagation is possible in metals, the permittivity of whichis negative. Materials with both ε and μ negative allow the propagation of electro-magnetic waves.

The Maxwell equations simplify to the form

k× E = ωμ0μH, k×H = −ωε0εE. (4.229)

Using them in the identities

k · (E×H) = H · (k× E) = E · (H× k), (4.230)

we can derive that

k · (E×H) = ωμ0μH2 = ωε0εE2, (4.231)

or

k · S < 0, (4.232)

where

S = E×H (4.233)

Page 273: Quantum Aspects of Light Propagation

264 4 Microscopic Theories

is the Poynting vector. It means that the wave vector k and the Poynting vector Shave opposite directions.

The three vectors E, H, and k are a left-handed set of vectors, contrary to conven-tional materials, where they are a right-handed set. This inspired Veselago to namethe materials with both ε and μ negative left-handed materials.

Veselago also has pointed out that the left-handed material must be dispersive,ε or μ must depend on the frequency of a monochromatic field, since the time-averaged energy density of the electromagnetic field,

〈U 〉 = 1

2

{ε0

∂ [ω0ε(ω0)]

∂ω0〈E2〉 + μ0

∂ [ω0μ(ω0)]

∂ω0〈H2〉

}, (4.234)

would be negative. Here the electromagnetic field is assumed to be quasimonochro-matic, with a centre or carrier frequency ω0 > 0. The time averaging is done overthe period of the carrier, 2π

ω0. The original Gaussian units of measurement can be

respected by the replacements ε0 → 14π

, μ0 → 14π

. From the Kramers–Kronigrelations it follows that ε and μ must be complex. Strictly speaking, we shouldspeak of the medium with both Re ε and Re μ negative. Many results have beenderived on the assumption of a transparent medium, i.e. that Im ε and Im μ may beneglected.

The energy density (4.234) is positive, since (Landau et al. 1984)

∂(ωε)

∂ω> 0,

∂(ωμ)

∂ω> 0. (4.235)

The permittivity and permeability determine the index of refraction, n ≡ n(ω),and the relative impedance, Z ≡ Z (ω), by the relations

n = ±√εμ, Z =√

μ

ε. (4.236)

It is assumed that the complex square root has a positive real part, or nonnegativeimaginary part if the real part vanishes. The plus sign in (4.236) is used when thesquare root has a positive imaginary part and the minus sign in (4.236) is used whenthe square root has a negative imaginary part. Even though the assumption thatIm ε = Im μ = 0 is frequently used and it invalidates the rule of the sign, we mayassume Im ε > 0 or Im μ > 0 small and apply the rule. Due to continuity, the minussign is used in (4.236) for ε < 0 and μ < 0.

We introduce a phase velocity vector

vp ≡ ω

s · ks, (4.237)

where s is the direction of the Poynting vector S. As the wave vector

k = ω

cns, (4.238)

Page 274: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 265

we have let vp denote cn s. On writing the wave vector k in the forms

k = ±ω

c

√εμ s, (4.239)

we obtain that

∂k∂ω

= 1

2c

[Z

∂(ωε)

∂ω+ Z−1 ∂(ωμ)

∂ω

]s, s · ∂k

∂ω> 0, (4.240)

where Z still means the relative impedance, not the coordinate. Now we introducea group-velocity vector

vg = 1

s · ∂k∂ω

s, s · vg > 0. (4.241)

It has the same direction as the Poynting vector not only in the right-handed media,but also for the left-handed materials.

For n < 0 or, more generally, Re n < 0, the negative refraction occurs. Let usillustrate that the derivation of the Snell law is valid also for a negative refractiveindex. We consider a planar interface in the plane z = 0 between the half-spacesz < 0 and z > 0. The half-space z < 0 is free and the half-space z > 0 is filledwith a left-handed material. We introduce n1 = 1 and n2 = n. We assume anincident monochromatic wave with E+1 = E+10ei(k+1 ·r−ωt), a reflected wave with E−1 =E−10ei(k−1 ·r−ωt), and a transmitted one with E+2 = E+20ei(k+2 ·r−ωt). Here k+1 , k−1 , and k+2are the respective wave vectors. The respective Poynting vectors may be denotedby S+1 , S−1 , and S+2 . Let s+1 , s−1 , and s+2 mean their directions. Quite reasonably, wesuppose that s+1z > 0, s−1z < 0, and s+2z > 0 both for the right-handed media and theleft-handed materials.

We consider k+1y = 0. Then k−1y = 0 and k+2y = 0 by the isotropy. Still it holds thatk−1x = k+1x , k+2x = k+1x . Let us note that k+1x = ω

c n1s+1x = ωc s+1x , k+2x = ω

c n2s+2x = ωc ns+2x .

From this, ns+2x = s+1x . Therefore, the negative refraction occurs when the refractiveindex n is negative.

A planar slab of a material with ε = −1 and μ = −1 can be compared with alens. Its imaging is described by the equation a+b = l, where a > 0 is the distancefrom the object to the front plane, b > 0 is the distance from the rear plane to theimage, and l is the thickness of the slab. The image is real and direct, though notamplified.

It is worth noting that the left-handed material can enhance incident evanescentwaves (Pendry 2000). As we will show below, the slab of a material with ε =μ = −1 does not reflect light. These remarkable properties have led to the term aperfect lens for the planar slab.

Artificial structures were first proposed, which have negative permittivity andpermeability in the microwave region of frequencies. A periodic array of very thinmetallic wires has the negative permittivity. We let a mean the spatial period of

Page 275: Quantum Aspects of Light Propagation

266 4 Microscopic Theories

the lattice and r mean the radius of wires. Pendry et al. (1996) have expressed theresponse of this medium to the external electric field parallel to the wires (of atwo-dimensional lattice) using the effective permittivity (a three-dimensional latticehas been considered first)

εeff ≡ εeff(ω) = 1− ω2p

ω(ω + iγe), (4.242)

where

ωp =√

2πc

a√

ln( ar )

(4.243)

is the plasma angular frequency and γe is the absorption parameter.Similarly, it was predicted in Pendry et al. (1999) that a periodic array of split-

ring resonators behaves as a medium with negative magnetic permeability. Theresponse of the regular lattice to the external magnetic field perpendicular to theplane of the ring is given by the effective permeability

μeff ≡ μeff(ω) = 1− Fω2

ω2 − ω20 + iΓω

. (4.244)

Here ω0 is the resonant frequency and Γ is the absorption parameter. The parameterF is the filling factor for the split ring. Combination of both structures gives riseto the material with both negative permittivity and permeability—the left-handedmaterial.

The reports on first experiments, e.g. (Shelby et al. 2001) were followed by somecriticism. For instance, it was argued that the negative refraction is ruled out by thecausality principle (Valanju et al. 2002). Absorption was suggested as an alternativeexplanation of the experiment with negative refraction (Sanz et al. 2003).

Numerical simulations of the transmission of the electromagnetic waves throughthe left-handed medium offered an independent possibility to verify the theoreticalpredictions (Ziolkowski and Heyman 2001, Markos and Soukoulis 2002a,b). For-mulae for the transmission amplitude, t, and the reflection amplitude, r, of theelectromagnetic wave through a homogeneous slab read

1

t= cos(nkl)− i

2(Z + Z−1) sin(nkl), (4.245)

r

t= − i

2(Z − Z−1) sin(nkl), (4.246)

where k = ωc conventionally. Especially, t = exp(inkl), r = 0 for ε = μ = −1.

On the assumption that the wavelength of the electromagnetic wave is much largerthan the spatial period of the left-handed structure and on using relations (4.245)and (4.246), the index of refraction and the impedance have been derived from thenumerical data (Smith et al. 2002). The origin of absorption has been traced up(Markos et al. 2002).

Page 276: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 267

For applications it would be very interesting to pass from the microwave to theoptical frequencies, or from the GHz to THz region. Pendry (2003) has provided anapproval of the contemporary reports on experimental proofs of some properties ofthe materials with negative refractive index. In 2007, a progress has been reportedon Dolling et al. (2007).

Photonic crystals have been analysed in theory and numerical calculations(Notomi 2000, Foteinopoulou et al. 2003, Foteinopoulou and Soukoulis 2003) andused in experiments (Cubukcu et al. 2003). Optical frequencies have been assumed,but it has not been asked whether the effective permittivity and the effective perme-ability may be defined.

Foteinopoulou et al. (2003) present in fact all the interesting results of Veselago(1968) in the lossless case. We cite only the time-averaged energy flux 〈S〉 = 〈U 〉vg

and the time-averaged momentum density 〈p〉 = 〈U 〉ω

k.They concentrated themselves on clarification of some of the controversial issues.

Their numerical calculations have described a two-dimensional photonic crystal.For the photonic crystal system a frequency range exists for which the effectiverefractive index is negative. So, a wave hitting the photonic crystal interface for thatfrequency will undergo negative refraction similar to a wave hitting the interface ofa homogeneous medium with negative index n. It is worth noting that the effectivemedium is two-dimensionally isotropic. Otherwise, an effect similar to the negativerefraction may occur in a right-handed medium.

In their simulations, a finite extent line source was placed outside a photonic crys-tal at an angle of 30◦. The source starts emitting at t = 0 an almost monochromaticTE wave. The source is adjusted to generate a Gaussian beam.

Foteinopoulou et al. (2003) have used finite-difference time-domain simulationsto study the time evolution of an electromagnetic wave as it reaches the interface.The wave is trapped temporarily at the interface, reorganizes, and, after a long time,the wave exibits the negative refraction.

Shen (2004) has defined a frequency-independent effective rest mass of a photon.Letting meff denote this mass, we may find it from the relation

m2effc

4

�2= − res

ω=0ωn2(ω). (4.247)

If the left-handed medium can be modelled as two-time derivative Lorentz material(Ziolkowski 2001), then m2

eff ≥ 0 for the electromagnetic parameters characteristicof (Ruppin 2000).

Ruppin (2002) has obtained a modification of formula (4.234) for the time-averaged energy density, which respects the absorption in the medium, but isrestricted to a relative permittivity ε(ω) and a relative permeability μ(ω) of thefollowing forms:

ε ≡ ε(ω) = 1+ ω2p

ω2r − ω2 − iΓeω

, Γe > 0,

Page 277: Quantum Aspects of Light Propagation

268 4 Microscopic Theories

μ ≡ μ(ω) = 1− Fω20

ω2 − ω20 + iΓhω

, Γh > 0. (4.248)

At the same time, it is a generalization of the result obtained by Loudon (1970) inthe case of no magnetic dispersion. The time-averaged energy density is

〈U 〉 = 1

2

[ε0

(ε′ + 2ωε′′

Γe

)〈|E|2〉 + μ0

(μ′ + 2ωμ′′

Γh

)〈|H|2〉

], (4.249)

where ε′ = Re ε, ε′′ = Im ε, μ′ = Re μ, μ′′ = Im μ, conventionally.Veselago (2002) has presented a miniature review of the progress in negative-

index materials. The subject of that paper which comprises such a review is theformulation of Fermat’s principle. The formulation stating that the optical length isstationary is correct.

For simplicity, we may consider differentials only for a Euclidean space. The dif-ferential of the optical length is the differential of the Euclidean length multipliedby the refractive index. If a variation of a path is restricted to the positive-indexmedium, the optical length of the path taken by a ray of light in travelling betweentwo points is a local minimum. If a variation is restricted to the negative-indexmedium, the optical length of the sought path is a local maximum.

Pendry (2003) has provided an approval of the contemporary reports on experi-mental proofs of some properties of the materials with negative refractive index.

Naqvi and Abbas (2003) have noted that principle of duality has a somewhat dif-ferent form in negative-index materials. Engheta (1998) has expressed a continuoustransition from the original solution (α = 0) to the dual solution (α = 1) in the caseof the positive-index medium

Efd = E cos(α

π

2

)+ Z0 ZH sin

π

2

),

Z0 ZHfd = −E sin(α

π

2

)+ Z0 ZH cos

π

2

). (4.250)

For the negative refractive index, a form of Maxwell’s equations suggests a differenttransformation

Efd = E cos(α

π

2

)− Z0 ZH sin

π

2

),

Z0 ZHfd = E sin(α

π

2

)+ Z0 ZH cos

π

2

). (4.251)

The two transformations are identified when, in the version for the negative-indexmaterial, the replacement α ↔ −α is made.

Marques et al. (2002) have measured transmission in a hollow metallic waveg-uide. They have found that it behaves similarly as the periodic array of thin metallicwires with its negative effective permittivity both when unloaded and when loaded.After a similarity to the left-handed medium had been achieved, the waveguidetransmitted waves at about 6 GHz. A standard analysis of a metallic waveguide

Page 278: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 269

which still utilizes the effective magnetic permeability does not admit a radiationmode and contradicts the experiment. Another calculation has associated a modewith the split-ring-resonator medium anisotropy (Kondrat’ev and Smirnov 2003).The transmission may have been radiationless. Marques et al. (2002) worked withthe wavelength of 5 cm, while they had the longest waveguide with l = 36 mm.

Quite remarkably, the hollow waveguide behaves as a one-dimensional plasma

with effective permittivity εeff≡ εeff(ω) = ε0

(1− ω2

cω2

). The assertion that the trans-

mission of electromagnetic waves occurs due to the split-ring-resonator mediumanisotropy has been dismissed by experiment (Marques et al. 2003).

Zharov et al. (2003) have noticed that so far properties of left-handed materialsin the nonlinear regime of wave propagation have not been studied. They assumethat a metallic structure is embedded into a nonlinear dielectric with a permittivitywhich depends on the strength of the electric field in a general way, εD ≡ εD(|E|2).As an application they consider a linear dependence which corresponds to the Kerrnonlinearity. The effective nonlinear dielectric permittivity εeff(|E|2) is found to bea sum of the earlier result (4.242) and a third-order nonlinear term, εD(|E|2).

The effective magnetic permeability of the composite structure (for F � 1)

μeff(H) = 1+ Fω2

ω20NL(H)− ω2 + iΓω

(4.252)

differs from the earlier result (4.244) by the dependence of the eigenfrequencies ofoscillations on the magnetic field. It holds that

ω0NL(H) = ω0 X, (4.253)

where X is one of the stable roots of the equation

|H |2 = αA2 E2c

(1− X2)[(X2 −Ω2)2 +Ω2γ 2]

X6, (4.254)

where H is an appropriate component of the field H, α = ±1 stands for a focusingor defocusing nonlinearity, respectively, Ec is a characteristic electric field, A2 =16 ε3

D0ω20h2

c2 , εD0 = εD(0), and γ = Γω0

. h is the width of the ring.Huang and Gao (2003) have been motivated by the paper (Chui and Hu 2002).

They have investigated the effective refractive index spectra of a granular composite,in which metallic magnetic inclusions are embedded into the host medium. They cal-culate the effective permittivity and the effective permeability as well based on theClausius–Mossotti relation (Grimes and Grimes 1991). Numerical results show that,by controlling the volume fraction of dispersive spherical particles in nondispersivehost medium, a composite medium which is left handed in a certain frequencyregion can be prepared. They investigate a three-phase composite. Especially, byembedding dielectric and magnetic granules into the host medium and controllingthe volume fractions of the two sorts of the granules, the left-handed compositemedium can be realized.

Page 279: Quantum Aspects of Light Propagation

270 4 Microscopic Theories

A wave packet description allows an easy grasp of negative refraction (Huangand Schaich 2004). The interesting properties of the left-handed materials have beenillustrated on the realistic assumption of losses and incidence of a Gaussian beam(Cui et al. 2004). It has been demonstrated that a negative-index material allows anultrashort pulse to propagate with minimal dispersion (D’Aguanno et al. 2005). Thenegative-index material has been described with a lossy Drude model

ε(ω) = 1− 1

ω(ω + iγe), μ(ω) = 1−

(ωpm

ωpe

)2 1

ω(ω + iγm), (4.255)

where ω = ωωpe

is the normalized frequency, ωpe and ωpm are the respective electric

and magnetic plasma frequencies, and γe = γe

ωpeand γm = γm

ωpeare the respective

electric and magnetic loss terms normalized with respect to the electric plasmafrequency. Attention has been paid to the group-velocity dispersion parameterβ2 = d2k

dω2 [Agrawal (1995)]. The frequencies for which β2 = 0 are plotted as zerogroup-velocity dispersion points in figures. It has been noted that these points arerelated to ω < ωpe and that no zero group-velocity dispersion point is present whenωpm

ωpe= 1.

For a macroscopic quantization, Milonni (1995) selects any frequency region,where absorption is negligible, i.e. away from absorption resonances. He assumesa uniform (or homogeneous) dielectric medium. The formulation simplifies and,although restricted in their range of validity, the results are applicable to a widerange of interesting and practical situations. He derives the electromagnetic energydensity in a form which resembles the time-averaged linear dispersive energy(3.186). But the magnetic term of the energy density is expressed using thefrequency-dependent magnetic permeability, i.e. at the same level of generality asthe electric term.

For some frequency ω a mode function Fω(r) can be considered which satisfiesthe transversality condition and the Helmholtz equation

∇ · Fω(r) = 0, (4.256)

∇2Fω(r)+ ω2

c2ε(ω)μ(ω)Fω(r) = 0 (4.257)

and appropriate boundary conditions. It is assumed that the mode function is nor-malized,

∫|Fω(r)|2 d3r = 1. (4.258)

The monochromatic components at frequency ω of the fields are

E(r, t) = Re{Cωα(t)Fω(r)}, (4.259)

B(r, t) = Re{−i

c

ωCωα(t)∇ × Fω(r)

}, (4.260)

Page 280: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 271

D(r, t) = Re[ε(ω)Cωα(t)Fω(r)], (4.261)

H(r, t) = Re

{− i

μ(ω)

c

ωCωα(t)∇ × Fω(r)

}, (4.262)

where α(t) = α(0) exp(−iωt). The fields are expressed with these relations as inthe classical optics, with an amplitude α(t), but also with a normalization constantCω for the electric strength field. The energy is determined as the integrated energydensity in the form

〈H〉 = n(ω)

8πμ(ω)|Cω|2|α(t)|2 d

dω[n(ω)ω]. (4.263)

The choice

Cω =√

4πμ(ω)

n(ω) d[n(ω)ω]dω

(4.264)

of the normalization constant gives the energy in the form

〈H〉 = 1

2|α(t)|2, (4.265)

which corresponds to a harmonic oscillator. For the quantization of this oscillatorwe do the replacement α(t) →√

2�ω a(t) where a(t) is an annihilation operator, in(4.259)–(4.262).

Milonni (1995) does mention (Drummond 1990), but he does not expound, orcomment on, the first and the second time derivative of the electric displacementwhich we obtain in (3.206) on substituting the first of equations (3.207) for Eν .

The approach of Drummond (1990) enables one to respect the inhomogeneity ofthe medium as noted in (Milonni 1995). Let us remark that this approach leads alsoto twice as many annihilation and creation operators as expected. Ignoring that alsothe number of such narrow-band fields may be greater than one, this ratio reminds usof the microscopic model assuming an oscillator medium with a single resonance.

As different systems of units have been utilized we will compare some formulaeof Milonni (1995) with those of Drummond (1990). Restricting himself to a singlecarrier frequency (ν = 1), Drummond (1990) has presented the following expan-sions:

D1(x, t) = i∑

k,λ

√√√√ �∂ω1

k∂k

2V kζ1(ω1k )

(k× e1

kλa1kλeik·x−iω1

k t − H. c.)

, (4.266)

B1(x, t) = −i∑

k,λ

μω1k

√√√√ �∂ω1

k∂k

2V kζ1(ω1k )

(e1

kλa1kλeik·x−iω1

k t − H. c.)

, (4.267)

Page 281: Quantum Aspects of Light Propagation

272 4 Microscopic Theories

where V is the quantization volume and (cf., (3.220))

ω1k = k

√ζ1(ω1

k )

μ. (4.268)

Regarding wide bandwidths or a simplification merely, Drummond (1990) has com-pleted the following expansion:

Λ(x, t) =∑

k,λ

√�

∂ωk∂k

2V kζ (ωk)

(akλekλeik·x−iωk t + H. c.

), (4.269)

where

∂ωk

∂k= ωk

k

[1− ωk ζ

′(ωk)

2ζ (ωk)

]−1

. (4.270)

Using relation (4.266), abandoning the Taylor expansion as in relation (4.269)and dividing by ε, we obtain a normalization constant in the form

1

2Cω

√2�ω =

√�n(ω)ω

2ε d[n(ω)ω]dω

. (4.271)

Since μr

n = nεr

and for different units in the relations under consideration, the

replacement 2π ↔ 12ε0

is performed, we have also

1

2Cω

√2�ω =

√μ(ω)2π�ω

n(ω) d[n(ω)ω]dω

(4.272)

in conformity with Milonni (1995).The normalization constant is obvious from the expression

E(r, t) = 1

2Cω

√2�ω

[aωFω(r)+ a†

ωF∗ω(r)]. (4.273)

For a multimode field, the electric-field operator (4.273) is replaced by

E(r, t) =∑

β

√√√√ 2π�ωβμ(ωβ)

n(ωβ) d[n(ωβ )ωβ ]dωβ

[aβ(t)Fβ(r)+ a†

β(t)F∗β(r)], (4.274)

where Fβ(r) is a mode function for mode β obeying the transversality condition andthe Helmholtz equation

∇2Fβ(r)+ ω2β

c2ε(ωβ)μ(ωβ)Fβ(r) = 0. (4.275)

Page 282: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 273

For an effectively unbounded medium plane-wave modes can be used,

Fβ(r) → i√V

ekλ exp(ik · r), (4.276)

with V a quantization volume and ekλ (λ = 1, 2) a unit polarization vector orthogo-nal to k.

On introducing the notation

γk = d[n(ωk)ωk]

dωk= c

vg(ωk)(4.277)

and assuming ekλ to be real for simplicity, we can write the operators for E, B, D,and H fields in the forms

E(r, t) = i∑

k,λ

√2π�ωkμ(ωk)

n(ωk)γk V

× [akλ(t) exp(ik · r)− a†kλ(t) exp(−ik · r)]ekλ, (4.278)

B(r, t) = i∑

k,λ

√2π�μ(ωk)c2

ωkn(ωk)γk V

× [akλ(t) exp(ik · r)− a†kλ(t) exp(−ik · r)]k× ekλ, (4.279)

D(r, t) = i∑

k,λ

√2π�ωkn(ωk)ε(ωk)

γk V

× [akλ(t) exp(ik · r)− a†kλ(t) exp(−ik · r)]ekλ, (4.280)

H(r, t) = i∑

k,λ

√2π�c2

ωkn(ωk)μ(ωk)γk V

× [akλ(t) exp(ik · r)− a†kλ(t) exp(−ik · r)]k× ekλ. (4.281)

A comparison with the paper (Huttner et al. 1991) can be made. It is not so easy asthe previous excercise, because sums are to be compared with integral expressions.In the sum (4.280) the factor

√2π�ωkn(ωk)ε(ωk)vg(ωk)

cV=√

2π�ωkε(ωk)vg(ωk)

vp(ωk)V(4.282)

is used. In the integral expression, n2± is involved instead of ε(ωk) (for μ(ωk) = 1).

Again the replacement 2π ↔ ε02 must be done. The macroscopic quantization

(Milonni 1995) leads only to optical polaritons, while the microscopic quantization,

Page 283: Quantum Aspects of Light Propagation

274 4 Microscopic Theories

although it is frequently named also a “macroscopic one”, introduces also acousticalpolaritons.

Milonni and Maclay (2003) have applied the theory of Milonni (1995) to theeffects of a negative-index medium on an excited guest atom, the Doppler effect,radiative recoil, and spontaneous and stimulated radiation rates, and also to thespectral density of thermal radiation.

A quantization scheme for the electromagnetic field interacting with atomicsystems in the presence of dispersing and absorbing magnetodielectric media iscontained in Dung et al. (2003). The magnetodielectric media include left-handedmaterial.

The spontaneous decay of an excited two-level atom is influenced by the envi-ronment. The atom embedded in a homogeneous, purely electric medium has thedecay rate

Γ = nΓ0, (4.283)

where n = √ε is the refractive index and Γ0 is the decay rate in free space. It

follows also from Fermi’s golden rule. Now the magnetodielectric medium has thedecay rate

Γ = μnΓ0, (4.284)

where the refractive index n = √εμ. This should be verified for the material with

both μ and n negative. Moreover, the expression (4.284) should be generalized toinclude the magnetodielectric absorption.

It is assumed that the magnetodielectric medium is causal and linear. It is char-acterized by a relative permittivity ε(r, ω) and a relative permeability μ(r, ω). Forinstance,

ε(r, ω) = 1+ ω2Pe(r)

ω2Te(r)− ω2 − iωγe(r)

, (4.285)

μ(r, ω) = 1+ ω2Pm(r)

ω2Tm(r)− ω2 − iωγm(r)

, (4.286)

where ωPe(r), ωPm(r) are the coupling strengths, ωTe(r), ωTm(r) are the transverseresonance frequencies, and γe(r), γm(r) are the absorption parameters. For nota-tional convenience, the spatial argument has been omitted.

The quantization has been performed by generalizing the theory expounded alsoin this book, Section 4.1.2. Let ˆE(r, ω), . . . be the operators of the electric strength,etc. in frequency space. Especially, let ˆP(r, ω) and ˆM(r, ω), respectively, be theoperators of the polarization and the magnetization in frequency space.

The operator-valued Maxwell equations are very similar to the classical ones. Wepresent the electric constitutive relation

ˆP(r, ω) = ε0[ε(r, ω)−1] ˆE(r, ω)+ ˆPN(r, ω), (4.287)

Page 284: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 275

with ˆPN(r, ω) being the noise polarization associated with the electric losses due tomaterial absorption, and the magnetic constitutive relation

ˆM(r, ω) = κ0[1− κ(r, ω)] ˆB(r, ω)+ ˆMN(r, ω), (4.288)

where κ0 = μ−10 , κ(r, ω) = μ−1(r, ω), and ˆMN(r, ω) is the noise magnetization

associated with magnetic losses. From the viewpoint of the theory the noise termsguarantee that the field commutators do not depend on Re{ε(r, ω)} and Re{κ(r, ω)}.

We present also the wave equation for ˆE(r, ω), the right-hand side of whichincludes the noise polarization ˆPN(r, ω) and the noise magnetization ˆMN(r, ω),

∇ × κ(r, ω)∇ × ˆE(r, ω)− ω2

c2ε(r, ω) ˆE(r, ω) = iωμ0

ˆjN(r, ω), (4.289)

where

ˆjN(r, ω) = −iω ˆPN(r, ω)+ ∇ × ˆMN(r, ω) (4.290)

is the noise current.Dung et al. (2003) consider the solution

ˆE(r, ω) = iωμ0

∫G(r, r′, ω)ˆjN(r′, ω) d3r′, (4.291)

where G(r, r′, ω) is the classical Green tensor obeying the equation

∇ × κ(r, ω)∇ ×G(r, r′, ω)− ω2

c2ε(r, ω)G(r, r′, ω) = δ(r− r′)1. (4.292)

Cf., (4.136), (4.137), and (4.138) above. In Section 4.2.7 will be referred to calcula-tions, in which the following property of the Green tensor is used, G(r, r′,−ω∗) =G∗(r, r′, ω).

All of the commutation relations follow from the choice

ˆPN(r, ω) = i

√�ε0

πIm{ε(r, ω)} fe(r, ω) (4.293)

and

ˆMN(r, ω) =√−�κ0

πIm{κ(r, ω)} fm(r, ω) (4.294)

and from the commutation relations for the fundamental bosonic vector fields(λ, λ′ = e, m)

[ fλi (r, ω), f †λ′ j (r′, ω′)] = δλλ′δi jδ(r− r′)δ(ω − ω′)1, (4.295)

[ fλi (r, ω), fλ′ j (r′, ω′)] = 0. (4.296)

Page 285: Quantum Aspects of Light Propagation

276 4 Microscopic Theories

So far the time dependence has not been considered, but the complete notationshould be as in (4.157) and (4.158) above. We may note that the vector potential

ˆA(r, ω, 0) = 1

iωˆE⊥(r, ω, 0), (4.297)

where ˆE⊥(r, ω, 0) means the transverse part of ˆE(r, ω, 0), cf. (4.142). We consideralso the scalar potential with the property as quantized in (4.143)

− ∇ ˆϕ(r, ω, 0) = ˆE‖(r, ω, 0), (4.298)

where ˆE‖(r, ω, 0) means the longitudinal part of ˆE(r, ω, 0), cf. (4.143).

Having defined ˆA(r, ω, 0), we may introduce also

ˆB(r, ω, 0) = ∇ × ˆA(r, ω, 0). (4.299)

On dividing by iωμ0, equation (4.289) yields

∇ × ˆH(r, ω, 0)+ iω ˆD(r, ω, 0) = 0, (4.300)

where we have used also the two constitutive relations. The definition of the operatorof the magnetic induction may be rewritten as an operator-valued Maxwell equation

∇ × ˆE(r, ω, 0) = iω ˆB(r, ω, 0). (4.301)

On the scalar multiplication of equation (4.289) with the ∇ operator from the left,we obtain that

− ω2

c2∇ ·

[ε(r, ω) ˆE(r, ω)

]= iωμ0(−iω)∇ · ˆPN(r, ω), (4.302)

or

∇ ·[ε0ε(r, ω) ˆE(r, ω)+ ˆPN(r, ω)

]= 0, (4.303)

or

∇ · ˆD(r, ω) = 0, (4.304)

where

ˆD(r, ω) = ε0ε(r, ω) ˆE(r, ω)+ ˆPN(r, ω). (4.305)

By definition,

∇ · ˆB(r, ω) = 0. (4.306)

It can be shown that the equal-time commutation relations

[Ei (r, t), E j (r′, t)] = 0 = [Bi (r, t), B j (r′, t)], (4.307)

[ε0 Ei (r, t), B j (r′, t)] = −i�εi jk∂kδ(r− r′)1 (4.308)

Page 286: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 277

are preserved, where ∂k means the partial derivative with respect to the kth compo-nent of the vector r.

The interaction of charged particles with the medium-assisted electromagneticfield is studied using Hilbert space on which the components of the bosonic fieldsfλi (r, ω) and operators rα and pα act. Here rα and pα are, respectively, the positionand the canonical momentum operator of the αth particle of mass mα and chargeqα . The charge density

ρA(r, t) =∑

α

qαδ(r1− rα(t)

)(4.309)

and the scalar potential of the particles

ϕA(r, t) =∫

ρA(r′, t)

4πε0|r− r′| d3r′ (4.310)

are introduced.In the minimal-coupling scheme and for nonrelativistic particles, the total Hamil-

tonian reads

H (t) = �

λ=e,m

∫ ∫ ∞

0ωf†λ(r, ω, t) · fλ(r, ω, t) dω d3r

+∑

α

1

2mα

[pα(t)− qαA(rα(t), t)

]2

+ 1

2

∫ρA(r, t)ϕA(r, t) d3r+

∫ρA(r, t)ϕ(r, t) d3r. (4.311)

The third and last terms can be written in the forms

1

2

∫ρA(r, t)ϕA(r, t) d3r = 1

2

α

α′α �=α′

qαqα′

4πε0|rα(t)− rα′ (t)| , (4.312)

∫ρA(r, t)ϕ(r, t) d3r =

α

qαϕ(rα(t), t

). (4.313)

In this new situation the old operators should be denoted as E0(r, t), B0(r, t),D0(r, t), H0(r, t), except the bosonic vector fields fλ(r, t). Then we introduce thefield operators in the presence of charge particles

E(r, t) = E0(r, t)−∇ϕA(r, t), B(r, t) = B0(r, t), (4.314)

D(r, t) = D0(r, t)− ε0∇ϕA(r, t), H(r, t) = H0(r, t), (4.315)

Page 287: Quantum Aspects of Light Propagation

278 4 Microscopic Theories

the new operators. The exposition in the literature may have been more explicit,

because the old operators have not been renamed and the new operators are �E (r, t),etc.

These operators obey the time-independent and time-dependent Maxwell equa-tions, where

jA(r, t) = 1

2

α

[(rα(t)

)·δ(r1− rα(t)

)+ δ(r1− rα(t)

) (rα(t)

)·](4.316)

is the operator of the current density of the particles.In the literature it has been shown that the Hamiltonian (4.311) generates the

time-dependent Maxwell equations and the Newton equations of motion for thecharged particles.

The authors treat the spontaneous decay of an excited two-level atom. Theysolve the problem of the time development of the state |ψ(t)〉, |ψ(0)〉 = |{0}〉|u〉,whose alternative are the quantum states |1λ(r, ω)〉|l〉. Here |l〉 is the lower statewhose energy is set equal to zero and |u〉 is the upper state of energy �ωA. ωA

is a transition frequency, |1λ(r, ω)〉 ≡ f†λ(r, ω)|{0}〉. In a formal expression of thestate |ψ(t)〉, Cu(t), Cel (r, ω, t) and Cml(r, ω, t) are the respective coefficients. Thesecoefficients satisfy linear differential equations and the initial conditions Cu(0) = 1and Cλl(r, ω, 0) = 0. In the differential equations, both the tensor-valued Greenfunction and the vector dA, the transition dipole moment, occur.

As expected, the solution reminds us of the Weisskopf–Wigner theory. In theintegro-differential equation the vector dA and the tensor-valued Green function stilloccur in a relatively simple fashion. The decay rate Γ can be expressed in terms ofthem and the shifted transition frequency ωA is utilized.

The case of nonabsorbing bulk material is treated first. On this assumption

Γ = Re[μ(ωA)n(ωA)]Γ0, (4.317)

where

Γ0 = ω3Ad2

A

3�πε0c3(4.318)

is the free-space decay rate, but taken at the shifted transition frequency.When ε(ωA) and μ(ωA) have opposite signs, then the refractive index is purely

imaginary and Γ = 0. In contrast, for nonabsorbing left-handed material, the rate Γ

is given by relation (4.317) without the notation Re.The case of atom in a spherical cavity is treated second. The cavity may be con-

ceived as a way of removing the singularity of the tensor-valued Green function.Forms of Γ

Γ0valid for an atom at an arbitrary position inside a spherical free-space

cavity surrounded by an arbitrary spherical multilayer material environment may beapplied (Dung et al. 2003, Li et al. 1994, Tai 1994).

Page 288: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 279

In fact, those formulae are specialized for the atom situated at the centre ofthe cavity and the otherwise homogeneous material environment. Results of Scheelet al. (1999a, 2000) have been amended. The ratio of the decay rates has been calcu-lated in the literature as a function of the (shifted) atomic transition frequency ωA.Large cavities are considered first. Then smaller ones are characterized. The numberof clear-cut cavity resonances decreases as the radius of the cavity decreases. A cav-ity is considered whose radius is much smaller than the transition wavelength. Everytime a comparison is made between (a) dielectric matter, (b) magnetic matter, and(c) magnetodielectric matter. The decay rate in the cases (a) and (b) inside a bandgap is low and, in the case (a), it also increases much due to each cavity resonance.In the case (c) this rate is low, if the band gap may belong either to the permittivityor to the permeability. Also the role of the resonances is similar to the case (a)or (b). But in the overlap of the electric and magnetic band gaps, the decay rateincreases.

Using their results, the authors may address also the local-field corrections. Thedescription of an atom should not be derived using the macroscopic field, even ifthe description, such as relation (4.317), does not involve the field. The theory ofthe authors directly applies to the real-cavity model. But simplifying assumptionsare too weak to work for the magnetodielectric matter. The complexity relies onthe dependence on z = R ωA

c . The expansion in powers of z must begin with aterm proportional to R−3, continue with R−1, and an already constant term and theO(R) term. If the material absorption may be disregarded, or when ε and μ maybe taken for real numbers, the terms proportional to R−3 and R−1 may be left out.Hence

Γ ≈[

3ε(ωA)

1+ 2ε(ωA)

]2

Re[μ(ωA)n(ωA)]Γ0. (4.319)

For strong dielectric absorption and R small we have a different approximation

Γ ≈ 9 Im{ε(ωA)}|1+ 2ε(ωA)|2

(c

ωA R

)3

Γ0. (4.320)

The decay may be regarded as being purely radiationless (Dung et al. 2003).Felbacq and Bouchitte (2005) have found a unified theoretical approach to the

left-handed materials. They have used a renormalization group analysis, which takesinto account the coupling between each resonator. They have checked the theoreticalresults numerically. They can explain the result by Pokrovsky and Efros (2002) that,by embedding wires in a medium with negative μ, one does not get a left-handedmedium.

Ozbay et al. (2007) have provided 17 references to reports on metamaterialsappropriate to a wide range of operating frequencies such as radio, microwave,millimetre wave, far-infrared, mid-infrared, near-infrared frequencies, and even vis-ible wavelengths. In that article results obtained from experiments at the microwavefrequencies have been reported.

Page 289: Quantum Aspects of Light Propagation

280 4 Microscopic Theories

Roppo et al. (2007) have studied the pulsed second-harmonic generation inpositive- and negative-index media. In the positive-index media and in the presenceof phase mismatch two forward-propagating components of the second-harmonicare generated (Bloembergen and Pershan 1962). In the pulsed generation, the sec-ond harmonic signal comprises a pulse which walks off (and is recognized in muchwork) and a second pulse which is “captured” and propagates under the pumppulse.

4.2.7 Application to Casimir Effect

In this subsection we will pay attention to papers which apply the quantization ofthe electromagnetic field in dispersive and absorbing media to the Casimir effect.Casimir’s work had its origin in a problem of colloidal chemistry, namely, the sta-bility of hydrophobic suspensions of particles in dilute aqueous electrolytes (Spar-naay 1989, Milonni 1994). Such suspensions are said to be stable if the particlesdo not coagulate. The particles are charged. Each particle is surrounded by ions ofopposite charge. We expect that a repulsive force between the particles separatedby a distance d increases more rapidly than an attractive force as d → LD + 0,where LD is a Debye length. We realize also that the repulsive force between theseparticles decreases more rapidly than the attractive force as d →∞. The attractiveforce should be obtained by integrating the pairwise forces between atoms, assum-ing an interatomic force given by the London–van der Waals interaction (London1930).

Now we mention the original idea of Casimir. In 1948 he gave expression for theattractive force per unit area

FC(d) = �cπ2

240d4, (4.321)

where d is a distance between two uncharged, perfectly conducting parallel plates.Milonni (1994) has reviewed a standard calculation of the Casimir force. It is thecalculation of the difference between the zero-point field energies for finite andinfinite plate separations. This difference is interpreted as the potential energy ofthe system. In calculating this energy, the Euler–Maclaurin summation formula isused. The initial difference between a divergent sum over modes of the confinedfield and a divergent integral is modified to a difference between a convergent sumand a convergent integral, featuring a formal dependence on a function f (k). Herek is the wavenumber of a mode. Finally it emerges that the result does not dependon the function f (k), if f (k) satisfies some conditions. The attractive force betweenthe plates is then obtained as the derivative of the potential energy with respect tothe distance d on changing the sign.

The same result has been obtained by considering the radiation pressure (Milonni1988). In calculating it, the Euler–Maclaurin summation formula has been applied.Let us note that only a d-independent factor is determined here. The same expression

Page 290: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 281

can be obtained also by a modification of the standard calculation, so the derivativewith respect to the distance and with the changed sign can be taken independent offinding the constant. As the function f (k) depends on km ≈ 1

a0, where a0 is the Bohr

radius, we see that a different function f (x) depends on kmd ≈ da0

, a number of theBohr radii spanning the distance between the plates.

These calculations are presented in Chapters 2 and 3 in Milonni (1994). Muchlater, in Chapter 7, he mentions forces between dielectric slabs. In the early 1950s,predictions of microscopic theories did not agree with experimental results. Milonni(1994) mentions the Lifshitz macroscopic theory (Lifshitz 1956). He does notexpound this theory in fact. He indicates that some results follow the Casimirapproach. The force between two semi-infinite dielectric slabs separated by a differ-ent dielectric medium or vacuum is derived. The case of vacuum between slabs hasbeen treated by Lifshitz. The general case of a dielectric medium between slabs hasbeen treated by Schwinger et al. (1978).

The physical basis of Lifshitz’s calculations is not so difficult to understand. Hewas first to use a random field, K(r, t), corresponding to some real or complex noisepolarization. At zero temperature, this field satisfies the fluctuation–dissipation rela-tion (the Gaussian system)

〈Ki (r, t)K j (r′, t)〉 = 2� Im{ε(ω)}δi jδ(r− r′) . (4.322)

Let us recall that the (Kronecker and Dirac) delta functions in the commutatorbetween noise polarizations are multiplied by �

πε0 Im{ε(ω)}. From relation (7.69),

p. 233 in Milonni (1994), it can be seen that K(r, t) = 4πP(r, t) in the Gaus-sian system of units. Therefore, it would be appropriate to convert K(r, t) into the

International System of units similarly as D(r, t), using the factor√

ε04π

. Then the

fluctuation–dissipation relation becomes

〈Ki (r, t)K j (r′, t)〉 = �ε0

2πIm{ε(ω)}δi jδ(r− r′) (SI). (4.323)

We recognize the right-hand side as a half of the appropriate commutator (K(r, t) =P(r, t) in the SI). Exactly, K(r, t) ≡ K(r, ω, t) and we miss the functional factorδ(ω−ω′) in the Lifshitz theory. Lifshitz (1956) then calculates the force in terms ofthe Maxwellian stress tensor, which we present here in the form

T(r, t) = ε0E(r, t)E(r, t)+ 1

μ0B(r, t)B(r, t)

− 1

2

[ε0E2(r, t)+ 1

μ0B2(r, t)

]1. (SI). (4.324)

In Kupiszewska and Mostowski (1990) and Kupiszewska (1992), the Casimireffect is studied on a restriction to the one-dimensional version. This means that

Page 291: Quantum Aspects of Light Propagation

282 4 Microscopic Theories

only wave vectors normal to the surface are taken into account in the calculations.It is also assumed that the temperature is zero, T = 0.

In Kupiszewska and Mostowski (1990), the Casimir effect for the case of two nonabsorbing dielectric slabs has been studied in detail. The electromagnetic field hasbeen quantized in the presence of a dielectric medium. The use of the Maxwellianstress tensor has been considered. It has been noted that the value of the appropriatecomponent of the stress tensor is equal to the energy density for the one-dimensionalcalculation. An infinite expression has been regularized by means of an exponentialcutoff function. The Casimir force has been calculated in the limit of semi-infiniteslabs. A result has been provided for any slab thickness, but for a small reflectioncoefficient.

In Kupiszewska (1992), the absorbing dielectric slabs have been considered. Themedium has been modelled as a continuous field of quantum harmonic oscillatorsinteracting with a heat bath. The atoms and the electromagnetic field have beendescribed with equations of motion and the long-time solution has been found. Asa generalization of the previous result, two contributions to the Casimir effect havebeen distinguished.

Weigert (1996) has considered several modes between the perfectly conductingmetallic plates to be in a squeezed vacuum state. At a given time instant the smallestpossible expectation value of the energy in a neighbourhood of one of the mirrorsis obtained through a calculation. It has been proposed to generate squeezed modesinside such a cavity and to measure an increase of the Casimir force. Weigert (1996)admits that the state of the system is not stationary, but he does not consider just theLorentz force, only the stress tensor. He calls it energy stress tensor, but he calculatesonly with a stress tensor.

The theoretical work on the Casimir force outweighs experimental work. Onregarding dielectric measurement, an accuracy of better than 10% has been reported(Lamoreaux 1997). An introductory guide to the literature on the Casimir force hasbeen published (Lamoreaux 1999).

Electromagnetic field quantization in an absorbing medium has been readdressed,and the Casimir effect both for two lossy dispersive dielectric slabs and betweentwo conducting plates was analysed by Matloob (1999a,b) and by Matloob et al.(1999). Matloob and Falinejad (2001) have investigated the Casimir effect betweentwo dispersive absorbing slabs in three dimensions. The dielectric function of theslabs has been assumed to be an arbitrary complex function of frequency satisfy-ing the Kramers–Kronig relations. The Maxwellian stress tensor has been used toevaluate the vacuum radiation pressure of the electromagnetic field on each slab interms of vacuum expectation values. These averages have been expressed using thefluctuation–dissipation theorem and Kubo’s formula (Landau and Lifshitz 1980).A simple relation to the imaginary part of a tensor-valued Green function has beenrecognized. So, the infinities of the stress tensor and the regular expression whichdiverges to them have been obvious. No explicit electromagnetic field quantizationhas been made. In a certain step of calculations the infinities cancel. Attention hasbeen paid to various limits of the general expression and to the Lorentz model of the

Page 292: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 283

dielectric function. The effect of finite temperature on the Casimir force betweentwo dielectric slabs has also been considered.

Kurokawa and Wakayama (2002) have introduced a Casimir energy for a com-pact Riemann surface of genus at least 2 and have related it to the Selberg zetafunction. The scope of the paper has been the application of the Selberg trace for-mula to such a Riemann surface similar to methods of mathematical physics andquantum chaos.

da Silva et al. (2002) have generalized the so-called thermofield dynamics viaan analytic continuation of the Bogoliubov transformations. It has been achievedthat a field in arbitrary confined regions of space and time is described. In the caseof an electromagnetic field, the energy-momentum tensor has been subjected to thegeneralized Bogoliubov transformation. The Casimir effect has been calculated forzero and nonzero temperature. The generalized Bogoliubov transformation has beenapplied also to the description of the field fulfilling the Dirichlet boundary condi-tions (at a conducting plate) and the Neumann boundary conditions at a permeableplate (the Casimir–Boyer model).

Tomas (2002) has considered the Casimir effect in a dispersive and absorbingmultilayered system using the Minkowski stress tensor method. He has calculatedthe Casimir force in a lossless dispersive layer of an otherwise absorbing multilayerby employing the quantized field operators as emerge from the scheme expoundedin this chapter. He has presented the expression obtained and has compared it withthe result of Zhou and Spruch (1995) who had applied the surface mode summationmethod to purely dispersive media. As an illustration he has calculated the Casimirforce on a dielectric slab in a planar cavity with realistic mirrors. The differencebetween Casimir energies in two distinct layers has been established and the differ-ence between Casimir forces in two such layers has been presented provided thattheir refractive indices are equal.

Boyer (2003) has presented a model, where physical ideas are transparent and thecalculations allow easy numerical evaluation. The model has no direct connectionwith experiment. One-dimensional analogues of three-dimensional concepts andtheir properties are studied. A simplified thermodynamics is evoked. He assumesa one-dimensional box of length L at zero temperature T = 0. But the one-dimensionality assumption reaches so far that all the virtual photons, if considered,should have the same, or just the opposite direction of the wave vector. He introducesthe Casimir energy ΔUzp(x, L) for the case, where a partition is present in the boxat a position x . He considers boundary conditions, let us say for intervals (0, x)and (x, L), namely, the Dirichlet or Neumann boundary conditions. The Dirichletcondition corresponds to a perfectly conducting boundary condition describing aperfectly conducting material in three spatial dimensions and the Neumann condi-tion is simplified from an infinitely permeable boundary condition describing aninfinitely permeable medium for electromagnetic waves. The boundary conditionsat x ′ = 0 and x ′ = L are enforced by the walls. In fact, x cannot be used for thecoordinate, which is denoted by x ′ instead. The boundary condition at x ′ = x isenforced by the partition.

Page 293: Quantum Aspects of Light Propagation

284 4 Microscopic Theories

Boyer (2003) lets α be 0 or 1, where α = 0 for like boundary conditions forpartition and walls, and α = 1 for unlike boundary conditions. He finds that

ΔUzp(x, L) = −π�c

(1

24− α

16

)(1

x+ 1

L − x− 4

L

). (4.325)

For α = 0, we may state that, off the centre of the box, forces act which repel thepartition from the nearest wall. He speaks of an attractive force between the partitionand the walls. For α = 1, we may say that, from the centre of the box, attractiveforces act. He mentions a repelling force between the partitions and the walls. Theforce between a conducting plate and a permeable plate was given, e.g. in Boyer(1974). The zero-point-energy limit is contrasted by the high-temperature energy-equipartition limit. This corresponds to the Rayleigh–Jeans spectrum of radiation.Then

ΔURJ(x, L , T ) = 0. (4.326)

He discusses also the Casimir forces at finite temperature.Emig (2003) has developed a novel approach for calculating the Casimir forces

between periodically deformed objects. Theories for realistic geometries have beendeveloped in response to high-precision measurements (Mohideen and Roy 1998,Chan et al. 2001a,b, Bressi et al. 2002). The theories do not comprise rigorous, non-perturbative methods for calculating the force. The simplest and commonly usedapproximation is the proximity force theorem. For corrugated metal plates, it failsat a small corrugation length. A different approximation is the pairwise summationof renormalized retarded van der Walls forces. However, Lifshitz’s theory for dielec-tric bodies demonstrates that, in general, the interaction cannot be obtained from apairwise summation. The results do not agree with those from the zeta-functionmethod (Barton 2001) in a situation, where this method can be used. Emig (2003)has considered the force between a rectangular corrugated plate and a flat one. Thisgeometry cannot be treated by perturbation theory due to the rectangular edges. Theforce has been found by the non-perturbative method.

It was respected that in the most precise experiments the Casimir force betweenrough metallic plates was measured (Genet et al. 2003). It has been only one ofthe real conditions which differ from the ideal situation and assumptions of the the-ory. Others are imperfect reflection, nonzero temperature, and a geometry differentfrom the parallel plates. The temperature effect has been neglected, because it issignificant at large distances, while roughness corrections are more necessary at thesmallest distances typical of the experiments. The proximity force approximationhas been tested on the case where the force is measured between a plane and asphere. The approximation leads to correct results when the radius of the sphere ismuch larger than the distance of closest approach. In the case of metallic plates,the proximity force approximation is only valid for the roughness spectrum con-taining small enough wave numbers. While mean number of waves spanning theinterplate distance (multiplied by 2π ) may be informative of the accuracy of the

Page 294: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 285

approximation, Genet et al. (2003) have proposed a specific roughness sensitivityand have considered its expectation value.

Many problems are formulated when the perfectly conducting plates of Casimirare replaced by other perfectly conducting surfaces. It can be utilized that theCasimir problem is not modified or generalized to dielectric media at the sametime. For example, the rectangular cavity has been considered by Lukosz (1971)and Maclay (2000). The generalization to a system of conducting shells has alsobeen realized, cf., (Plunien et al. 1986). The rectangular cavity of sides (a1, a2, a3)depends on the three parameters, Λ ≡ (a1, a2, a3). The system of conducting shellsdepends on another system of parameters Λ.

Mazzitelli et al. (2003) have computed the Casimir interaction energy betweentwo concentric cylinders. To this end they have used approximate semiclassicalmethods and the exact mode-by-mode summation method. They characterize amethod according to Schaden and Spruch (1998, 2000). In this method the zero-point radiation is described with trajectories of a particle, and so as a real radiation.They mention the well-known decomposition of the spectral density into a smoothterm and the oscillating contribution. Periodic orbit theory relates oscillations in thequantum level density of a given Hamiltonian to the periodic orbits in the corre-sponding classical system. They derive an energy approximation using the periodicorbit theory,

E sem = −��c

4πa2

√b

a

w≥0

v≥v(w)

fvw

1

v4N

(b

a, v, w

), (4.327)

where � is the “quantization” length, a, b, a < b, are radii of the cylinders, v(w) isthe least positive integer v such that cos

(π w

v

)> a

b , fvw is 1 for v = 2, w = 0 andis 2 otherwise,

N (α, v,w) ≡√[

α − cos(π w

v

)] [α cos

(π w

v

)− 1]

[1+ α2 − 2α cos

(π w

v

)]2 . (4.328)

They further write

E sem = E semw=0 + E sem

w≥1, (4.329)

where E semw=0 (E sem

w≥1) are obtained from relation (4.327), where condition w ≥ 0 isreplaced by the condition w = 0 (w ≥ 1). They show that, for b

a ∼ 1, E sem ∼ E semw=0,

where

E semw=0 = −�

�cπ3

360

√ab

(b − a)3. (4.330)

Mazzitelli et al. (2003) mention the proximity theorem (Derjaguin and Abrikosova1957, Derjaguin 1960). For its application they assume two parallel plates of area A,

Page 295: Quantum Aspects of Light Propagation

286 4 Microscopic Theories

not of different areas. This difference does not suggest decision whether the largeror the smaller area should be chosen. Still the relation

EP = −�cπ2

720

A

(b − a)3(4.331)

is applied to the plates “wound” into a cylinder. Here A?= 2π�a, 2π�b. Obviously,

the theory of periodic trajectories suggests the choice A = 2π�√

ab. But only thenumerical calculation shows that the approximation is relatively good for 1 < b

a <

4. Further they compute the exact Casimir energy for the coaxial cylinders using themode-by-mode summation method (Nesterenko and Pirozhenko 1997). The finalresult has the form

Eex = E12 − 0.01356

(1

a2+ 1

b2

)��c, (4.332)

where

E12 = − ��c

2π2a2

×∫ ∞

0

n

∫ ∞

kz

√k2

z − y2d

dyln [Fn12(iy, 1, α)] dy dkz, (4.333)

with

Fn12(iy, 1, α) =[

1− In(y)Kn(αy)

In(αy)Kn(y)

] [1− I ′n(y)K ′

n(αy)

I ′n(αy)K ′n(y)

], (4.334)

α = ba , In(z) and Kn(z) are the modified Bessel functions and the MacDonald func-

tions, respectively. Again on the condition α ∼ 1 the relations simplify

Eex ∼ E12 ∼ −��cπ3

360a2

1

(α − 1)3+ O

[1

(α − 1)2

]. (4.335)

The semiclassical approximation is valid. In contrast, the semiclassical energy foran isolated cylinder vanishes and the exact energy for a cylinder of radius a is

EC = −0.01356��c

a2. (4.336)

They present also numerical results.Ahmedov and Duru (2003) have calculated the Casimir energies with respect

to the previous work such as Mazzitelli et al. (2003) and Høye et al. (2001). Letus consider the region between two close coaxial cylinders. We assume that thecylinders have the radii r0 < r1. Then the Casimir energy per unit height is

Page 296: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 287

Ecyl = − π3 R

720Δ3

(1+ 15

2π2

Δ2

R2

), (4.337)

where Δ ≡ r1 − r0, R = √r0r1.

Let us imagine a flat space which is periodic in the z coordinate unlike theEuclidean space. Let us consider the region between two cylinders or tori in thisspace provided that the axis of these cylinders is the z axis. Then the Casimirenergy is

Etor = − π3 RL

720Δ3

(1+ 15

2π2

Δ2

R2

), (4.338)

where L is the length of the flat space measured parallel to the z axis.Let us analyse the ring with a rectangular cross-section. The Casimir energy is

Ebox ≈ − π3 RL

720Δ3+ Rζ (3)

16Δ2, (4.339)

where L is the height of the cylinders and ζ (z) is the Riemann ζ -function.Let us consider two close concentric spheres. We assume that the spheres have

the radii r0 < r1. Then the Casimir energy is

Esph = − π3 R2

360Δ3

(1+ 5Δ2

4π2 R2

). (4.340)

Let us evaluate two close coaxial cones. We assume that the cones have the apexangles θ0 < θ1 ≤ π

2 . Then the Casimir energy per unit volume is

E ≈ − Θπ3

720r4Δ3, (4.341)

where Δ ≡ θ1 − θ0, Θ ≡ √sin θ0 sin θ1, and r is the distance from the com-

mon vertex. Dividing the right-hand side by 2πΘΔ to “correct” the energy density(Ahmedov and Duru 2003), we obtain that

E = − π2

1440r4Δ4+ O(Δ−3). (4.342)

This is similar to the solution of the wedge problem (Deutsch and Candelas 1979),

E = − 1

1440r4Δ2

(π2

Δ2− Δ2

π2

), (4.343)

where Δ is the angle between the half-planes of the boundary.Geyer et al. (2003) have begun with the state of the research of the Casimir effect.

They have also mentioned some difficulty with calculations of the temperature effecton the Casimir force between real metals of finite conductivity. They distinguish fivedifferent approaches. According to the fifth approach, the description of the thermalCasimir force can be obtained by the Leontovich surface impedance boundary con-dition.

Page 297: Quantum Aspects of Light Propagation

288 4 Microscopic Theories

Three domains of frequencies are distinguished: The region of the normal skineffect for low frequencies, the region of the anomalous skin effect, or relaxationdomain for higher frequencies, and the region of the infrared optics for yet higherfrequencies. In the region of the anomalous skin effect and in the relaxation domainmetal cannot be described by any dielectric permittivity depending only on the fre-quency. The space dispersion is also essential.

Otherwise, the theoretical basis is as follows. Boundary conditions are introduced

Et = Z (ω)Bt × n, (4.344)

where Z (ω) is the surface impedance of the conductor, Et and Bt are the tangentialcomponents of the (Fourier transformed) electric and magnetic fields, and n is theunit normal vector to the surface (pointing inside the metal). For an ideal metal wehave Z ≡ 0 and for real nonmagnetic metals |Z | � 1 holds (Landau et al. 1984).

The surface impedance is determined over the whole frequency axis, even thoughit is different in each of the three domains. It is respected that the main contributionto the Casimir free energy and force is given by the frequency region centred aroundthe so-called characteristic frequency ωc = c

2a , where a is the space separationbetween two metal plates. Relation (43) in Geyer et al. (2003) is an analogue ofrelation (8.62) in the book (Milonni 1994). An approximate expression (45), whichis not reproduced here, has been derived for the case of a sphere above a plate madeof a real metal.

Geyer et al. (2003) remind of the fact that at the temperature T = 0 only theanomalous skin effect and infrared optics occur. For Au the transition frequencyΩ = 6.36 × 1013 rad/s is obtained. They determine the characteristic frequencyωc at each separation distance, and then they fix the proper impedance function.In a figure, which is not reproduced here, graphs of both impedance functions areplotted. A “transition” impedance function does not exist evidently. The correctionfactors E(a)

E (0)(a) to the Casimir energy agree quite well in the region of the infraredoptics and in the transition region. In the region of the anomalous skin effect, theresults due to the right and wrong choice differ significantly. They further presentnumerical results for T = 70 K and T = 300 K. At these temperatures, the normalskin effect occurs already, but only at larger separations. The relative thermal cor-rection (Geyer et al. 2003) is calculated for 0 < a ≤ 5 μm and only by the use ofthe impedance of infrared optics and of anomalous skin effect.

Raabe et al. (2003) have underscored that one-dimensional quantization schemesare not rigorous enough when the Casimir force between absorbing multilayerdielectrics is calculated. At the beginning they warn that the “mode summation”method, which was employed by H. Casimir himself, cannot be generalized to thecase of absorbing bodies, because in such bodies there are no modes. Then theycharacterize three procedures:

(1) The electromagnetic field and the material bodies are treated macroscopically.Explicit field quantization is not performed, but the field correlation functionsare written down in conformity with statistical thermodynamics.

Page 298: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 289

(2) The electromagnetic field and the material bodies are quantized at a micro-scopic level. The bodies are described by appropriate model systems. Simplify-ing assumptions are made.

(3) The electromagnetic field and the material bodies are described macroscopi-cally as in the first procedure. But the medium-assisted electromagnetic field isquantized by using an infinite set of appropriately chosen bosonic basic fields.

Raabe et al. (2003) have reserved the first method for Lifshitz (1955, 1956). Inthis context they have mentioned (Schwinger et al. 1978) and have characterized thepaper (Matloob and Falinejad 2001). The mentions about the second method com-prise the note that the calculations were carried out only for one-dimensional sys-tems. Let us refer only to Kupiszewska and Mostowski (1990), Kupiszewska (1992).The third method is used in Raabe et al. (2003), but the authors also refer to Tomas(2002). Then they add two further methods. One is the surface-mode approach in thenonretarded limit (van Kampen et al. 1968) and including retardation (see referencesin the cited paper). The other is the scattering approach (Jaekel and Reynaud 1991).

Raabe et al. (2003, 2004) have reproduced the essential traits of their quantizationscheme. Then they describe the multilayer structure. They consider n − 1 layers ofthicknesses dl > 0, l = 1, . . . , n − 1. These layers have the boundaries zl , l =1, . . . , n, which have the properties zl+1 = zl +dl , l = 1, . . . , n−1. They introducez0 = −∞, zn+1 = +∞, and so, inclusive of the substrate and the superstrate, thereare n + 1 layers. The permittivity is

ε(r, ω) = εl(ω) for zl < z < zl+1, l = 0, . . . , n. (4.345)

For the tensor-valued Green function, we are referred to the paper (Tomas 1995).From the expression of the Green tensor it follows that it conserves its form on theintervals (zl, zl+1) × (zl ′ , zl ′+1), l ′ = 0, . . . , n. If both spatial arguments are in thesame layer, l = l ′, we let Gl (r, r′, ω) denote the form G(r, r′, ω). We introduce thescattering part

Gscatl (r, r′, ω) = Gl(r, r′, ω)−Gbulk

l (r, r′, ω) for r′ �= r, (4.346)

where Gbulkl (r, r′, ω), the bulk part, is the solution for the case that the medium of

the lth layer fills up the whole space. The values of the scattering part of the Greentensor for r′ = r are obtained in the coincidence limit of the position vectors. It isassumed that a “cavity”, which separates walls, is the j th layer, 1 < j < n − 1.The walls are composed of j − 1 layers l = 1, . . . , j − 1 and of n − 1 − j layersl = j + 1, . . . , n − 1. If some of the walls are semi-infinite, the numbering maydiffer a little.

Before the Casimir force is calculated from the stress tensor, a more generaltensor

T(r, r′, t) = Te(r, r′, t)+ Tm(r, r′, t)

− 1

21 Tr{Te(r, r′, t)+ Tm(r, r′, t)} (4.347)

Page 299: Quantum Aspects of Light Propagation

290 4 Microscopic Theories

is defined, where 1 is the second-rank unit tensor and

Te(r, r′, t) = 〈D(r, t)E(r′, t)〉, (4.348)

Tm(r, r′, t) = 〈B(r, t)H(r′, t)〉(r �= r′). (4.349)

The expectation values are calculated in thermal equilibrium, a stationary state ofthe field.

(i) Basic equation.

For finite temperatures T , they employ the statistical operator

ρ = Z−1 exp

(− H

kBT

), (4.350)

where

Z = Tr

{exp

(− H

kBT

)}, (4.351)

and kB is the Boltzmann constant. In relations (4.348) and (4.349), ρ may be writtenexplicitly. On substituting relation (4.350) into the modified relation (4.348), we get

Te(r, r′) = �

π

∫ ∞

0coth

(�ω

2kBT

)ω2

c2Im{ε(r, ω)G(r, r′, ω)} dω (4.352)

and on substituting relation (4.350) into the modified relation (4.349), we obtain

Tm(r, r′) = − �

π

∫ ∞

0coth

(�ω

2kBT

)∇ × Im{G(r, r′, ω)} ×←−∇′ dω. (4.353)

On the left-hand side of relations (4.352) and (4.353), the time argument t has beendropped, since the right-hand sides do not depend on t .

Although the generalization to a “cavity” containing dielectric medium has beenknown, Raabe et al. (2003, 2004) have restricted themselves to the free spacebetween two stacks. The Casimir force (per unit area) is given by the zz-componentof the stress tensor (4.324).

We may modify relations (4.347), (4.352), and (4.353) by the way of writing thesuperscripts scat. Then we introduce the stress tensor

Tscat(r) = limr′→r

Tscat(r, r′). (4.354)

With respect to a layer, it is suitable to introduce the tensor

Tscatj (r) = lim

r′→rTscat

j (r, r′). (4.355)

Page 300: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 291

Now a number of concepts and pieces of notation are introduced. Propagation con-stants

βl = βl (q, ω) =√

ω2εl(ω)

c2− q2 (4.356)

and reflection coefficients for σ -polarized waves at the top (+) and bottom (−) ofthe j th layer are defined that are

rσn+ = 0, σ = s, p (4.357)

and are calculated from the recurrence relations

r sl+ =

(βl

βl+1− 1

)+(

βl

βl+1+ 1

)exp (2iβl+1dl+1) r s

(l+1)+(

βl

βl+1+ 1

)+(

βl

βl+1− 1

)exp (2iβl+1dl+1) r s

(l+1)+, (4.358)

and

rpl+ =

(βl

βl+1− εl

εl+1

)+(

βl

βl+1+ εl

εl+1

)exp (2iβl+1dl+1) rp

(l+1)+(

βl

βl+1+ εl

εl+1

)+(

βl

βl+1− εl

εl+1

)exp (2iβl+1dl+1) rp

(l+1)+. (4.359)

The coefficients rσl− are

rσ0− = 0 (4.360)

and the recurrencies for the others are analogous, which are formally obtained fromrelations (4.358) and (4.359) on replacements

l �→ l, l + 1 �→ l − 1 (4.361)

and on the change of the subscript + of the reflection coefficient to −. Also denom-inators of the fractions for multiple reflections

Dσ l = Dσ l(q, ω) = 1− rσl+rσ

l− exp (2iβldl ) (4.362)

are introduced. Finally,

T scatzz, j = − �

2π2

∫ ∞

0coth

(�ω

2kBT

)

× Re

{∫ ∞

0qβ j exp

(2iβ j d j

)∑

σ

D−1σ j rσ

j−rσj+ dq dω

}. (4.363)

As T scatzz, j does not depend on the space point in the j th layer, the argument r has been

dropped.

Page 301: Quantum Aspects of Light Propagation

292 4 Microscopic Theories

(ii) Imaginary frequencies.

We introduce

ξm = 2mπkBT

�, m integer. (4.364)

Since the permittivity is positive on the positive imaginary frequency axis, we intro-duce

κ j =√

ξ 2ε j (iξ )

c2+ q2. (4.365)

Exploiting the analytical properties of the ω integrand in relation (4.363), we arriveat an expression of the integral with respect to ω by a residue series. Finally

T scatzz, j =

kBT

π

∞∑

m=0

(1− 1

2δm0

)

×[∫ ∞

0qκ j exp

(−2κ j d j)∑

σ

D−1σ j rσ

j−rσj+ dq

]

ω=iξm

, (4.366)

which may be regarded as a generalization of the famous Lifshitz formula Lifshitz(1955), (1956). For m = 0, the term with ω = 0 is peculiar and it should be replacedby the limit ω → 0+. To obtain T scat

zz, j in the zero-temperature limit, we may sim-ply repeat the derivation from relation (4.363). Of course, replacement

∑∞m=0 →

2πkBT dξ can be realized in relation (4.366).

(iii) One-dimensional systems.

A comparison with the three-dimensional case is made only for T = 0. Contraryto the three-dimensional description, the sum with respect to σ is omitted, sincein the one-dimensional system normal incidence occurs and the description canbe restricted to a single polarization. Further one of the integrals is replaced by amultiplication with a constant,

1

4π2

∫d2q �→ 1

A∑

q

, (4.367)

where A is the normalization area.Also analytical expressions for specific distance laws in the zero-

temperature limit are derived. For example, it is shown that the Casimir forcebetween two single-slab walls behaves asymptotically as d−6 instead of d−4 in the

Page 302: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 293

large-distance asymptotic regime. Results for single-slab walls for periodic multi-layer wall structure are illustrated in figures.

Chen et al. (2003) study the difference of the thermal Casimir forces at differenttemperatures between real metals. If the temperatures are fixed, the difference of theCasimir forces increases with a decrease of the separation distance. The configura-tions of two parallel plates and a sphere above a plate are considered.

In the case of two parallel plates, they utilize a perturbation result from the paperBordag et al. (2000) to express the thermal Casimir force denoted by Fpp(a, T ),where a is the separation distance and T is a temperature. They concentrate on thedifference

ΔFpp ≡ ΔFpp(a, T1, T2) = Fpp(a, T2)− Fpp(a, T1), (4.368)

where T1 and T2 are temperatures. For example, for Au, T1 = 300 K and T2 = 350K, |ΔFpp| decreases with an increase of a. For an ideal metal |ΔFpp| does not dependon a.

In the case of a sphere above a plate, they use a perturbation result after Klimchit-skaya and Mostepanenko (2001). The thermal Casimir force is denoted by Fps(a, T ).They study the difference

ΔFps ≡ ΔFps(a, T1, T2) = Fps(a, T2)− Fps(a, T1). (4.369)

For example, for Au, T1 = 300 K and T2 = 350 K, |ΔFps| decrease with an increaseof a. For an ideal metal |ΔFps| is a linear function of a. Then Chen et al. (2003)compare the chosen approach with that, e.g. after the paper (Brevik et al. 2002)(further references see Chen et al. 2003). They fix a = 0.5 μm and T1 = 300 K,while T1 ≤ T2 ≤ 350 K. The chosen approach exhibits an increase |ΔFps| with thetemperature T2. The difference is negative both for a real and for an ideal metal. Thealternative approach provides a negative difference for an ideal metal, but a positivedifference (more than six times larger at T2 = 350) for a real metal.

Iannuzzi and Capasso (2003) have published a comment on the paper (Kennethet al. 2002). They believed that, at distances relevant to Casimir force measurementsand to nanomachinery, the Casimir force between two slabs in vacuum was alwaysattractive. They have referred also to (Bruno 2002), a paper devoted to an attractiveCasimir magnetic force.

Kenneth et al. (2003) have replied to the comment (Iannuzzi and Cappasso 2003).They have declared the consensus that exploring the possible existence or design ofmaterials with nontrivial magnetic properties for obtaining a repulsive Casimir forceis important.

Action of the Casimir force on magnetodielectric bodies embedded in media hasbeen analysed in (Raabe and Welsch 2005). The consistency of expressions derivedin the framework of the macroscopic theory with microscopic harmonic-oscillatormodels is shown. It could be startling that here Raabe and Welsch (2005) declarethe macroscopic quantum electrodynamics themselves. We have chosen in this bookthat their theory is named microscopic just as the theory due to Hopfield (1958) and

Page 303: Quantum Aspects of Light Propagation

294 4 Microscopic Theories

Huttner and Barnett (1992a,b). It is consensual, even though with a reservation,which can be found below relation (68), which is not reproduced here, namely, thatthe level of representation is rather a mesoscopic one (Raabe and Welsch 2005). Itmay be controversial that they do not use micro- or mesoscopic for the model withthe two auxiliary fields fe(r, ω), fm(r, ω).

The exposition begins with the classical Maxwell equations with charges andcurrents, but without the constitutive relations. It is worthwhile to mention that theLorentz force density

f(r) = ρ(r)E(r)+ j(r)× B(r) (4.370)

is written as

f(r) = ∇ · T(r)− ε0∂

∂t[E(r)× B(r)], (4.371)

where T(r) is the stress tensor,

T(r) = ε0E(r)E(r)+ 1

μ0B(r)B(r)− 1

2

[ε0E2(r)+ 1

μ0B2(r)

]1. (4.372)

The integral of the Lorentz force density f(r) over some space region (volume) Vgives the total electromagnetic force F acting on the matter inside V

F =∫

Vf(r) d3r. (4.373)

On integrating both sides of equation (4.371) over V we obtain that

F =∫

∂VT(r) · da(r)− ε0

d

dt

VE(r)× B(r) d3r, (4.374)

where da(r) is an infinitesimal surface element. If the volume integral on the right-hand side of this equation does not depend on time, then the total force reduces tothe surface integral

F =∫

∂VdF(r), (4.375)

where

dF(r) = da(r) · T(r) = T(r) · da(r). (4.376)

The tensor T(r) may be decreased by a constant term, i.e. a constant, spaceindependent tensor (it suffices to consider the position on the surface ∂V ). Raabeand Welsch (2005) have commented on the role of Minkowski’s stress tensor, which

Page 304: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 295

is considered in much work devoted to the related topic, be it under this name or onlyas a “stress tensor”.

Relation (4.371) can be generalized to characterize the density of a generalizedLorentz force

f(r, r′)+ f(r′, r) = ∇ r+r′2· {T(r, r′)+ T(r′, r)}

− ε0∂

∂t

[E(r)× B(r′)+ E(r′)× B(r)

], (4.377)

where f(r, r′) means the density of a generalized Lorentz force

f(r, r′) = ρ(r)E(r′)+ j(r)× B(r′), (4.378)

∇ r+r′2= ∇ + ∇′ ≡ ∇r + ∇r′ , (4.379)

and T(r, r′) is a generalized stress tensor

T(r, r′) = ε0

[E(r)E(r′)− 1

21E(r) · E(r′)

]

+ 1

μ0

[B(r)B(r′)− 1

21B(r) · B(r′)

]. (4.380)

Raabe and Welsch (2005) expound the quantum theory of the electromagneticfield as described also in this book in Section 4.2.6. They have presented also com-mutation relations between the charge density operator, the current density operator,and the electromagnetic-field operators. They have calculated correlation functionsof some operators in thermal states of the field. Calculation of the expectation valueof the Lorentz force, which is not reproduced here, follows.

In our opinion, the operator of the Lorentz force has not been presented in suchan explicit form as its expectation value. The latter is denoted by the same notationas the corresponding classical stress tensor. As the expectation value of the stresstensor operator is infinite before a quantum correction, the notation is generalizedto the form T(r, r′) (just the same notation as in the classical theory),

T(r, r′) = ε0〈E(r)E(r′)〉 + 1

μ0〈B(r)B(r′)〉

− 1

21[ε0〈E(r) · E(r′)〉 + 1

μ0〈B(r) · B(r′)〉

]. (4.381)

In other words, we can quantize relation (4.380) to introduce a generalized stresstensor T(r, r′) and to write relation (4.381) in the form

T(r, r′) = 〈T(r, r′)〉. (4.382)

Page 305: Quantum Aspects of Light Propagation

296 4 Microscopic Theories

To make contact with microscopic approaches, Raabe and Welsch (2005) con-sider a harmonic-oscillator medium and derive that the (steady-state) Lorentz forceacting on such a medium in some space region V is

F = limt→∞

V〈ρ(r, t)E(r, t)+ j(r, t)× B(r, t)〉 d3r, (4.383)

where again the charge density operator ρ(r, t) and the current density operatorj(r, t) are appropriately expressed. They apply the theory to a planar magnetodi-electric structure. Its definition is specific in that homogeneity of the dielectric in aninterspace (“cavity”) 0 < z < d j is required, where the subscript j has been used inconformity with Raabe et al. (2003, 2004).

Raabe and Welsch (2005) give the relevant stress tensor element Tzz(r) in theinterspace 0 < z < d j . Their relations (75) and (76) for this element, which arenot reproduced here, are rather complicated. They include also a criticism of basingthe calculations on Minkowski’s stress tensor (Tomas (2002), which leads to a rela-tively simple relation (81), which is not repeated here too. It can be believed that thewarning is helpful, since one prefers simpler formulae to more complicated ones,provided that the simpler ones are not wrong.

To calculate the Casimir force on a plate in a nonempty cavity, Raabe and Welsch(2005) choose five regions, finite (1, 2, 3) or semi-infinite (0, 4). Let us assume thatthe two walls and the plate are almost perfectly reflecting. The generalization ofCasimir’s well-known formula is

F = �cπ2

240

√μ

ε

(2

3+ 1

3εμ

)(1

d43

− 1

d41

), (4.384)

where dk are thicknesses of regions k = 1, 3. Let us also assume that μ = 1. Thenthe counterpart of the previous formula based on the Minkowski stress tensor is

F (M) = �cπ2

240

1√ε

(1

d43

− 1

d41

), (4.385)

which is just one of the formulae underlying the critique.Pitaevskii (2006) defends the validity of the paper (Dzyaloshinskii et al. 1960),

which has been disqualified or underestimated by the criticism in Raabe and Welsch(2005). It is important for the theory of the van der Waals–Casimir forces inside adielectric fluid. The tensor of the van der Waals–Casimir forces was obtained bysummation of an appropriate set of Feynman diagrams for the free energy and itsvariation with respect to the density (Dzyaloshinskii and Pitaevskii 1959). On thecondition of mechanical equilibrium this tensor differs from a Minkowski-like oneby a constant tensor. Dzyaloshinskii et al. (1960) obtained the same force betweensolid bodies, separated by a dielectric fluid, as Barash and Ginzburg (1975) andSchwinger et al. (1978). Pitaevskii (2006) discusses the reason why, in his opinion,the approach of Raabe and Welsch (2005) is incorrect. Raabe and Welsch (2006)

Page 306: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 297

maintain their position that the Casimir force should be calculated on the basis ofthe Lorentz force.

The paper (Leonhardt and Philbin 2007a) is interesting for its use of the notion ofa transformation medium. This notion seems to belong to classical optics essentiallyand to be formed after the general relativity theory. A concise quantum theory oflight in spatial transformation media has been developed in (Leonhardt and Philbin2007b). Leonhardt and Philbin (2007a) calculate the Casimir force for a dispersivemedium in their set-up inspired by Casimir’s original idea. They consider two per-fect conductors with a metamaterial sandwiched in between. The repulsive Casimirforce of a left-handed material may balance the weight of one of the mirrors, lettingit levitate on zero-point fluctuations. The simple formula for the Casimir force hasbeen compared with the result of the more sofisticated Lifshitz theory.

Chen et al. (2006) have measured the Casimir force between a gold-coated sphereand two Si samples of higher and lower resistivity. The lowering of resistivity cor-responded to enhancement of carrier density by several orders of magnitude. Eachmeasurement was compared with theoretical results using the Lifshitz theory withdifferent dielectric permittivities (Bordag et al. 2001, Chen et al. 2005, 2006, Lam-oreaux 2005) and found to be consistent with this theory.

Lenac and Tomas (2007) have considered the Casimir effect between metallicplates assuming them to be dispersive and lossless and separated by a medium withthe (Gaussian) unit permittivity. They have taken two very different permittivitiesfor media outside the plates, i.e. ε2 = 1 or ε2 = ∞ (perfect conductor). They haveanalysed the contributions of system eigenmodes with great attention to surfaceplasmon polariton modes. When the separation between the metallic plates is small,the surface plasmon polariton modes influence the Casimir effect dominantly exceptthe case of thin layers that are supported by a highly reflective medium.

Messina and Passante (2007a) have calculated the Casimir–Polder force densityon an uncharged, perfectly conducting plate placed in front of a neutral atom. To thisaim first-order perturbation theory and the quantum operator associated to the clas-sical electromagnetic stress tensor have been used. The result of Casimir and Polder(1948) has been rederived by integration of the force density. This integration isnot an argument against the well-known nonadditivity of the Casimir–Polder forces(Milonni 1994 and references therein), and it has been discussed appropriately.

Munday and Capasso (2007) have performed precision measurements of theCasimir–Lifshitz force between two metal surfaces (gold) separated by a fluid(ethanol). For this situation, the measured force is attractive and is approximately80% smaller than the force predicted for ideal metals in vacuum. The results werefound to be consistent with Lifshitz’s theory.

There exists a geometry well suited to the aim of an accurate theory–experimentcomparison, namely, that with parallel and periodic corrugations of the metallicsurfaces. The Casimir force is a superposition of the usual normal component and alateral one in this situation. In general, vacuum-induced torque is present (Rodrigueset al. 2006a).

Rodrigues et al. (2007a) have studied the lateral Casimir force arising betweentwo corrugated metallic plates. They assume that corrugations are imprinted on both

Page 307: Quantum Aspects of Light Propagation

298 4 Microscopic Theories

plates with the same period and along the same direction, but with a spatial mis-match. They have used the scattering theory in a perturbative expansion in powersof the corrugation amplitudes. The result is valid provided that these amplitudesare smaller than L (mean separation distance), λC (corrugation wavelength), andλP (plasma wavelength). Limiting cases such as the proximity-force approximationlimit and the perfect reflection limit are recovered when the length scales L , λC, andλP obey some specific orderings.

In the development of ever smaller atomic magnetic traps carbon nanotubes havebeen considered to become the elementary building blocks. It is well known that anatom held in a magnetic trap near an absorbing dielectric surface will undergo ther-mally induced spin–flip transitions. Some of these transitions lead to trapping losses.Fermani et al. (2007) have calculated atomic spin-flip lifetimes and have estimatedtunneling lifetime corresponding to the sum of the Casimir–Polder potential and themagnetic trapping potential. Their analysis indicates that the Casimir–Polder forceis the dominant loss agent.

Fulling (2007) have presented results on the Casimir force in one-dimensionalpiston models. These models are applications of quantum graphs (Roth 1985, Kuch-ment 2004). They have characterized the quantum star graphs mainly. A finite quan-tum graph consists of B one-dimensional undirected bonds or edges of length L j

( j = 1, . . . , B) and some vertices. Either end of each bond ends at one of thesevertices, and the valence of a vertex is defined as the number of bonds meetingthere. At the univalent vertices either a Dirichlet or a Neumann boundary conditionis imposed. For instance, the space may consist of B one-dimensional rays of largelength L attached to a central vertex. In each ray a piston is located a distance a fromthe vertex. At the central vertex the field has the Kirchhoff (generalized Neumann)behaviour. In fact, the pistons are treated as univalent vertices. If at each piston thefield obeys the Neumann boundary condition, then the force is (� = 1 = c)

F = (B − 3)π

48a2. (4.386)

When B = 1 or 2, the result is related to an ordinary Neumann interval of lengtha or 2a, respectively. When B > 3, the force is repulsive. The pistons will tendto move outward. Fulling et al. (2007) have discussed a periodic-orbit approach tocalculations of the Casimir forces. They have also numerically examined the rate ofconvergence of the periodic-orbit expansion.

Rodriguez et al. (2007a,b) have developed a numerical method to compute theCasimir forces in arbitrary geometries, for arbitrary dielectric and metallic materials.They have based their approach on the familiar result due to Lifshitz and Pitaevskii(1980), Dzyaloshinskii et al. (1961), and Pitaevskii (2006). The Casimir force isobtained in terms of the stress tensor integrated over space and imaginary frequency.The vacuum expectation value of the stress tensor is calculated in terms of the Greenfunction, which is automatically regularized on application of the finite-differencemethod to solve for the Green function. The geometries that have been consideredhave the property that the bodies have not a contact and they are in the free space.

Page 308: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 299

Then it is innocent to evoke the Minkowski stress tensor, since a contour or surfacearound the body of interest lies in the free space. But also the Maxwellian stresstensor is named.

Messina and Passante (2007b) have paid attention to fluctuations of the Casimir–Polder force between a neutral atom and a perfectly conducting wall. They havemade use of the method of time-averaged operators introduced by Barton and widelyused by him in his papers on fluctuations of the Casimir forces for macroscopic bod-ies (Barton 1991a,b). They have also calculated the Casimir–Polder force fluctua-tions for an atom between two conducting walls. This situation has been investigatedalready by Barton (1987). The force operator has been derived from an effectiveinteraction Hamiltonian (Passante et al. 1998). To this end the effective interactionenergy operator is differentiated with respect to the distance from the atom to thewall.

Intravaia et al. (2007) remind that the Casimir effect, at short distances, isdominated by the coupling between the surface plasmons that are present on twometallic mirrors (Van Kampen et al. 1968). The Casimir energy is calculated interms of quasielectrostatic (or nonretarded) field modes. When the mirror separa-tion increases, retardation must be taken into account. Intravaia et al. (2007) use themethod of Schram (1973). They choose the dielectric function

ε(ω) = 1− ω2p

ω2(4.387)

to describe the metal. They calculate dispersion relations for the relevant modesnumerically. They distinguish bulk modes, propagating cavity modes, and evanes-cent modes. For the TE polarization all modes are propagating, but for the TMpolarization two modes are evanescent in at least some range of wave vectors. Thesemodes are referred to as “plasmonic”. A plasmonic contribution to the Casimirenergy is denoted by Ep. The short-distance asymptotics is − A�ωp

L2 , where A is the

area of the mirrors. The large-distance asymptotics is + A�√

ωpcL5/2 . This is balanced

in the total Casimir energy by the contribution of photonic modes (cavity and bulkmodes), which yields the negative, binding, energy again.

Passante and Spagnolo (2007) have evaluated the Casimir–Polder potential bet-ween two atoms in the presence of an infinite perfectly conducting plate and atnonzero temperature. They assume the wall located at z = 0 and let rA and rB

denote the positions of atoms A and B, respectively. First they outline the methodused by reproducing the Casimir–Polder potential energy between two atoms in athermal field

WAB(R) = �c

π

∫ ∞

0k3αA(k)αB(k) coth

(�ck

2kBT

)

× V(k, R) : τ (k, R) dk, (4.388)

(cf. Wennerstrom et al. 1999), where R = |R| is the distance between the two atoms,R = rB − rA, k is a wavenumber, αA(k) (αB(k)) is the dynamical polarizability of

Page 309: Quantum Aspects of Light Propagation

300 4 Microscopic Theories

the atom A (B) (Power and Thirunamachandran 1993),

V(k, R) = 1

R3

{(1− 3

RR

RR

)[cos(k R)+ k R sin(k R)]

−(

1− RR

RR

)k2 R2 cos(k R)

}, (4.389)

τ (k, R) =(

1− RR

RR

)sin(k R)

k R

+(

1− 3RR

RR

)[cos(k R)

k2 R2− sin(k R)

k3 R3

]. (4.390)

Then they derive and discuss their results for the retarded atom–atom Casimir–Polder interaction when both a thermal field and a boundary condition are present.The interaction energy is

WAB(R, R) = WAB(R)+ WAB(R)

− �c

π

∫ ∞

0k3αA(k)αB(k) coth

(�ck

2kBT

)

× σ :[τ (k, R) · V(k, R)+ V(k, R) · τ (k, R)

]dk, (4.391)

where R = |R| is the distance between one atom and the image of the other atomreflected on the plate, R = rB−σ ·rA, and σ is the reflection tensor on the conductingplate, supposed orthogonal to the z axis.

The analysis of most Casimir force experiments using a sphere-plate geome-try has relied on the proximity-force approximation (PFA), which expresses theCasimir force between a sphere and a flat plate in terms of the Casimir energybetween two parallel plates. Krause et al. (2007) have conducted an experimentalassessment of the range of applicability of the proximity-force approximation. Theyhave measured the Casimir force and force gradient between a gold-coated plate andfive gold-coated spheres with different radii using a microelectromechanical torsionoscillator. Specifically, according to the proximity-force approximation, the Casimirforce between a sphere of radius R and a flat plate separated by a distance z � Rcan be written as

F(z) ≈ FPFA(z) ≡ 2π REpp(z), (4.392)

where Epp(z) is the Casimir energy per unit area between two parallel plates sepa-rated by the distance z.

If the bodies are smooth and perfectly conducting, the exact Casimir force maybe expanded in powers of z

R (Scardicchio and Jaffe 2006),

FCasimir(z, R) = 2π REpp(z)

[1+ β

z

R+ O

(z2

R2

)], (4.393)

Page 310: Quantum Aspects of Light Propagation

4.2 Green-Function Approach 301

where β is a dimensionless parameter and the Landau notation O( f (x)) means thatO( f (x))

f (x) is bounded for x → 0. An effective pressure Peff(z, R) may be expandedsimilarly, but a new dimensionless parameter is denoted by β ′. The roughness andconductivity effects can be included and the modified notation, β(z) and β ′(z),respects a general dependence on z. For separations z < 300 nm, Krause et al.(2007) have found that |β ′(z)| < 0.4 at the 95% confidence level.

Rodrigues et al. (2006b) have presented a novel theoretical approach to the lateralCasimir force beyond the regime of validity of the proximity-force approximation.They have related the results of the new approach to the measured values (Chenet al. 2002a,b). Unfortunately, the novel approach also has its region of validity andthe illustration chosen does not fit into it. Besides, the complete proximity-forceapproximation has led to a happy coincidence of the theoretical value of 0.33 pNwith the average of measured values of 0.32 pN, while the expectation value belongsto the interval 0.32 ± 0.077 pN (at 95% confidence). This situation is reflected inthe comment (Chen et al. 2007) and the reply (Rodrigues et al. 2007b).

Emig (2007) has explored the lateral Casimir force between two parallel peri-odically patterned metal surfaces. It is assumed that the surfaces are set into rel-ative oscillatory motion so that their normal distance is a periodic function oftime. This scenario resembles the so-called ratchet systems (Reimann (2002). Itis demonstrated that the system allows for directed lateral motion of the surfaces.These results show that Casimir interactions offer contactless translational actuationschemes for nanomechanical systems.

Emig et al. (2007) have developed a systematic method for computing theCasimir energy between arbitrary compact dielectric objects. Casimir interactionsare completely characterized by the scattering matrices of the individual bodies.As an example they compute the Casimir energy between two identical dielectricspheres at any separation.

The thermal part of the Casimir force was subject to discussions (Milton 2004).On the assumption of real metals the Lifshitz formula is used. For LT large com-pared with �c

kB, where L is the distance between the parallel plates and T is the

temperature of the system, the assumption of ideal metals leads to a result which istwice the Lifshitz one.

Svetovoy (2007) has analysed the repulsive thermal Casimir force between twometals and a metal and a high-permittivity dielectric. The repulsion discussed insuch work has the meaning of a negative thermal correction to the force at zerotemperature, but the total force is always attractive. The force is calculated usingthe Lifshitz formula written via real frequencies (Landau and Lifshitz 1963). Twocontributions of the fluctuating fields, propagating waves and evanescent waves, aredistinguished. For both material configurations the repulsive s-polarized evanescent-wave contribution dominates for LT small. Here L is the distance between parallelplates and T is the temperature of the system. In this case, the force between idealmetals is attractive and small (Mehra 1967, Brown and Maclay 1969). The idealmetal is rather the limit case of a superconductor than of a normal metal (Antezzaet al. 2006).

Page 311: Quantum Aspects of Light Propagation

302 4 Microscopic Theories

Rizzuto (2007) considers a neutral two-level atom uniformly accelerated in adirection parallel to an infinite mirror and calculates the atom-wall Casimir–Polderinteraction between the accelerated atom and the mirror. The mirror is modelled asDirichlet boundary conditions on a massless scalar field.

The author evaluates the vacuum fluctuation (vf) and radiation reaction (rr) con-tributions to the atom-wall Casimir–Polder interaction energy. First she expressesonly the contributions to the radiative shifts of the atomic levels.

Let us assume the atom to be at rest. The Casimir–Polder interaction energybetween the atom at rest and the wall is obtained by considering only the z0-dependent terms, E (vf)

CP and E (rr)CP , in the vacuum fluctuation and in the radiation

reaction contributions, respectively,

E (vf)CP = μ2

64π2c2z0

[2 f

(ω0,

2z0

c

)− π cos

(2ω0z0

c

)], (4.394)

E (rr)CP =

μ2

64πc2z0cos

(2ω0z0

c

), (4.395)

where ω0 corresponds to the energy difference of the levels of the atomic system,�ω0, z0 is the distance of the atom from the mirror, and

f (α, β) =∫ ∞

0

sin(αx)

x + βdx . (4.396)

In the case of a uniformly accelerated atom, with the acceleration in a direc-tion parallel to the reflecting plate, a generalization of relations (4.394) and (4.395)yields

E (vf)CP = μ2

64π2c2z0 N

{2 f

(ω0,

2c

asinh−1

( z0a

c2

))

− π cos

(2ω0c

asinh−1

( z0a

c2

))

− a

ω0c

[1− cos

(2ω0c

asinh−1

( z0a

c2

))]}, (4.397)

E (rr)CP =

μ2

64πc2z0 Ncos

(2ω0c

asinh−1

( z0a

c2

)), (4.398)

where N =√

1+ ( z0ac2

)2and a is the proper acceleration of the atom.

Page 312: Quantum Aspects of Light Propagation

Chapter 5Microscopic Models as Related to MacroscopicDescriptions

The models expounded in Chapter 4 are often labelled as macroscopic, since appar-ently they do not allow one to consider the Clausius–Mossotti or the Lorentz–Lorenzrelation. Even though we shall not, even here, expound this interesting subject,which, however, is notorious in the classical theory, we shall mention the quantummodels, whose microscopic character cannot be doubted. The role of macroscopicaverages, which is analyzed in the classical theory as well, is discussed from thequantal viewpoint.

5.1 Quantum Optics in Oscillator Media

A quantum-optical experimental setup may consist of active and passive devices,active devices to generate light of certain properties (e.g. nonclassical light) andpassive ones to modify and apply it. It is an interesting and nontrivial problem tostudy how quantum statistical properties of light are influenced by passive opticaldevices like mirrors, resonators, beam splitters, or filters.

Knoll and Leonhardt (1992) have continued also the paper (Knoll et al. 1987),where the medium is nondispersive and lossless, but they now intend to considerdispersion and losses.

On introducing the Hamiltonian for the complete system, the Heisenberg equa-tions of motion for field operators and medium (not source) quantities are derived.The complete system under consideration consists of the subsystems: optical field,medium atoms, and sources. The field is described by the electric-field-strengthoperator E(x, t) and the electromagnetic vector-potential operator A(x, t) in theCoulomb gauge. The medium is modelled by the damped harmonic oscillators{qμ(t), pμ(t)}, namely, the oscillators coupled to reservoirs composed of bath oscil-lators {qμB(t), pμB(t)}, the quanta of whose energy may be, for example, phonons.The medium oscillators are localized at xμ, they have the same mass m and the elas-ticity (force) constant k. The bath oscillators are characterized by masses m B andthe elasticity constants kB and the coupling is expressed by the coupling constantsσB . The atomic sources are described by a current operator j (x, t), but its dynamics

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 5, C© Springer Science+Business Media, LLC 2009

303

Page 313: Quantum Aspects of Light Propagation

304 5 Microscopic Models as Related to Macroscopic Descriptions

need not be specified, e is the electron charge. For simplicity, a one-dimensionalmodel is considered only.

The Hamiltonian of the complete system is

H (t) = HR(t)+ HM(t)+ HRS(t)+ HS(t), (5.1)

where HR(t) is the Hamiltonian of the optical field, HM(t) is that of the mediumatoms, and HRS(t) describes the interaction between the optical field and sources,

HR(t) =∫

ε0

2

⎣E2(x, t)+ c2

(∂ A(x, t)

∂x

)2⎤

⎦ dx, (5.2)

HM(t) =∑

μ

{[ pμ(t)− e A(xμ, t)]2

2m+ k

2q2

μ(t)

+∑

B

[p2

μB(t)

2m B+ kB

2q2

μB(t)+ σB

2

[qμ(t)− qμB(t)

]2]}

, (5.3)

HRS(t) = −∫

j (x, t) A(x, t) dx, (5.4)

and HS(t) is the Hamiltonian of the atomic sources which is left unspecified. Theusual commutation rules are

[ A(x, t),−ε0 E(x ′, t)] = i�δ(x − x ′)1,

[qμ(t), pμ′(t)] = i�δμμ′ 1,

[qμB(t), pμ′B ′ (t)] = i�δμμ′δB B ′ 1. (5.5)

The Heisenberg equations of motion for field operators, medium operators, and bathoperators have been obtained. As a result of a Wigner–Weisskopf approximationfor the interaction of medium oscillators with the bath operators, quantum Langevinequations have been obtained. By eliminating the medium quantities from equationsfor field operators and further by a usual procedure, a generalized wave equationfor the vector potential is obtained. Using a Green function, this wave equation issolved.

The decomposition of the time-ordered quantum correlation functions into time-ordered correlation functions of the source operators and the free-field operators hasbeen derived. The notation of positive time-ordering T ≡ T+ (see, (2.110)) and theordering symbol O ≡ O+ are used. The property (2.120) and the consequence forthe time-ordered quantum correlations have been recalled.

The time-dependent Green function for a dielectric layer as the simplest opticaldevice is calculated. The field behind the layer is discussed and represented by thenegative-frequency part of the field, its expectation value and the normally orderedquadrature variance determined for the sake of squeezing analysis.

Page 314: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 305

5.2 Problem of Macroscopic Averages

Assuming that the medium is constituted by atoms, which do not form a continuum,we can aproximate a continuum using the macroscopic averaging. This idea is illus-trated by a simple model. It is indicated that many material constants would requireappropriately complicated derivations. Inclusion of losses enhances requirementson the microscopic model as well, even though we will restrict our discourse ona one-dimensional standing wave in a cavity. The macroscopic averaging can begeneralized even to this case.

5.2.1 Conservative Oscillator Medium

Dutra and Furuya (1997) have investigated a simple microscopic model for theinteraction between an atom and radiation in a linear lossless medium. It is a guesttwo-level atom inside a single-mode cavity with a host medium composed of othertwo-level atoms that are approximated by harmonic oscillators. There is an inten-tion to show, in general, that the ordinary quantum electrodynamics suffices, atleast in principle, and there is no need to quantize the phenomenological classicalMaxwell equations. If a macroscopic description is possible, it should appear asan approximation to the fundamental microscopic theory under certain conditions.Such a “macroscopic” approximation is obtained and conditions of its validity arederived.

All of the medium harmonic oscillators are represented by a collective harmonicoscillator, then two modes of a polariton field are defined. The macroscopic averageis regarded as filtering out higher spatial frequencies. The field that influences theguest atom is modified and the characteristics of the effect of such a microscopicfield are calculated. A condition is pointed out under which the contribution of theatoms to the quantum noise appears only through a frequency-dependent dielectricconstant. An effective description is obtained by leaving out the polariton modewhich is approximately equal to the collective mode of the medium.

Dutra and Furuya (1997) have introduced a microscopic model of a materialmedium that they have adopted, N two-level atoms having the same resonance fre-quency ω0 in a single-mode cavity of the resonance frequency ω. They considera guest two-level atom of the resonance frequency ωa and strongly coupled to thefield so that it will not be approximated by a harmonic oscillator. The operator ofthe displacement field in the cavity is given by the relation

D(x, t) =√

�ωε0

Lsin(ω

cx)

[a(t)+ a†(t)], (5.6)

where L is the length of the cavity. It is assumed that ω and L for the single-modecavity are in the relation

ω

c= π

L. (5.7)

Page 315: Quantum Aspects of Light Propagation

306 5 Microscopic Models as Related to Macroscopic Descriptions

The operator of the polarization of the medium is given by

P(x, t) =N∑

j=1

d jδ(x − x j )[b j (t)+ b†j (t)], (5.8)

where

d j =√

2ω0m0q j (5.9)

are electric dipole moments in eigenstates |1,s〉 j of the operators [b js(t)+ b†js(t)],

large s, weakly converging to [b j (t) + b†j (t)] for s → ∞ (s, the greatest number

of quanta, has been introduced on the normalization ground). In (5.9), m0 is aneffective mass, q j are effective charges, products d j [b j (t) + b†

j (t)] are the electric-dipole-moment operators of the atoms of the medium that are located at x j . Theoperators a(t), a†(t), b j (t), and b†

j (t) satisfy the bosonic commutation relations, inparticular,

[b j (t), b†j ′ (t)] = δ j j ′ 1, (5.10)

[b j (t), b j ′ (t)] = 0, (5.11)

and a(t), a†(t) commute with b j (t), b†j (t).

The Hamiltonian is given by the relation

H (t) = �ωa†(t)a(t)+ �ω0

N∑

j=1

b†j (t)b j (t)− 1

ε0

∫D(x, t)P(x, t) dx

+ �ωa

2σz(t)+ �Ω[a(t)+ a†(t)][σ (t)+ σ †(t)], (5.12)

where σz(t) and σ (t) are the pseudo-spin operators and

Ω = −d

√ωma

ε0L�sin(ω

cxa

)(5.13)

is the Rabi frequency of the guest atom located at xa whose electric dipole moment(in the eigenstate |1〉 of the operator [σ (t)+ σ †(t)]) is d and the effective mass ma.We notice that

− 1

ε0

∫D(x, t)P(x, t) dx = �

N∑

j=1

g j (ω)[a(t)+ a†(t)][b j (t)+ b†j (t)], (5.14)

Page 316: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 307

where

g j (ω) = −d j

√ωm0

ε0L�sin(ω

cx j

). (5.15)

In simplifying the Hamiltonian, Dutra and Furuya (1997) denote

G(ω) =√√√√

N∑

j ′=1

[g j ′(ω)]2, (5.16)

the coupling constant between the field and the collective harmonic oscillatordescribed by the annihilation operator

B(t) = 1

G(ω)

N∑

j=1

g j (ω)b j (t). (5.17)

In the Hamiltonian (5.12), the self-energy terms (Cohen-Tannoudji et al. 1989) havebeen neglected. This results in the condition 4[G(ω)]2 ≤ ωω0. Further, the originalproblem is reduced to the case of a single atom coupled to two polariton modes. The

frequencies of these modes are denoted by Ω1 and Ω2 so that Ω1 � ω√

1− 4 [G ′(ω)]2

ω20

,

ω and Ω2 → ω0 when ω → 0, ∞, respectively, with

G ′(ω) =√

ω0

ωG(ω). (5.18)

In other words, Ω1 < Ω2 for ω < ω0, Ω1 > Ω2 for ω > ω0. The dressed operators aredenoted by ck(t) and c†k(t), k = 1, 2, and the operators a(t) and B(t) are expressedin their terms (Chizhov et al. 1991).

The problem of the extra quantum noise introduced by the atoms of the mediumis discussed. In the case when the atoms of the medium are only weakly coupled tothe field, i.e. G ′(ω), ω � ω0, it holds that (Glauber and Lewenstein 1991)

〈{Δ[a(t)+ a†(t)]}2〉 ≈ √εr, (5.19)

where Δ[a(t) + a†(t)] = a(t) + a†(t) − 〈a(t) + a†(t)〉1, and εr is the relative per-mittivity

εr ≈ 1+ 4[G ′(ω)]2

ω20

. (5.20)

From the relation (5.19), the variance of the operator of the electric displacementfield D(x, t) (cf., (Glauber and Lewenstein 1991)) can be calculated. The adoption

Page 317: Quantum Aspects of Light Propagation

308 5 Microscopic Models as Related to Macroscopic Descriptions

of the continuous distribution of atoms in the medium instead of the realistic discreteone implies a greater variance (Rosewarne 1991).

Let us address the problem of macroscopic averages. The macroscopic theoriesof quantum electrodynamics in nonlinear media, have often, by the “definition”avoided discussing the macroscopic averaging procedure. The quantum-mechanicalaveraging advocated by Schram (1960) removes the quantum fluctuations fromthe macroscopic theory. The problem of the macroscopic theory what should bethe macroscopic averaging procedure resisted, for many years, the solution. It wasLorentz who in the beginning of the twentieth century first tried such a derivation,cf., chapters (de Groot 1969, van Kranendonk and Sipe 1977). Robinson (1971,1973) has proposed a different kind of macroscopic average. He regards a macro-scopic description as a description where the spatial Fourier components of the fieldvariables above some limiting spatial frequency k0 are irrelevant. Dutra and Furuya(1997) take the Fourier components with the spatial frequencies above ω

c for irrel-evant in a macroscopic description. They arrive at the following expression for themacroscopic polarization

ˆP(x, t) = −2G(ω)

√�ε0

ωLm0sin(ω

cx)

[B(t)+ B†(t)]. (5.21)

The macroscopic “average” does not change the operator of the electric displace-ment field D(x, t) and the macroscopic electric-field-strength operator is given bythe relation

ˆE(x, t) = 1

ε0[D(x, t)− ˆP(x, t)]. (5.22)

The calculation of the variance of a quantity typical of the operator ˆE(x, t) is larger

than ε− 3

2r , which agrees with Rosewarne’s result (Rosewarne 1991). Thus, it has been

shown that the contribution from the atoms to the quantum noise of the field doesnot restrict itself to inclusion of the dielectric constant. We are going to report thesuitable macroscopic theory of electrodynamics in a material medium which doesnot suffer from the problems which are discussed here.

It is shown that under certain conditions a macroscopic description incorporatingthe frequency dependence of the relative permittivity provides a good approxima-tion. In this domain, Milonni’s result has been recovered (Milonni 1995). The guestatom is not affected by the polariton mode if the frequency of the atom is far fromΩ2. An analysis of the probability of this mode inducing transitions shows that sucha probability is negligible when

|Ω22 − ω2| Ω

Ω2

√Ω2ω0

(Ω22 − ω2)2 + 4ω0ωG2

� |Ω2 − ωa|. (5.23)

Page 318: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 309

In the regime described by the relation (5.23), the leaving out of the polaritonmode described by c2 and c†2 and the macroscopic averaging leads to the macro-scopic Hamiltonian

Hmac(t) = �Ω1c†1(t)c1(t)+ �ωa

2σz(t)− d

ε0Dmac(xa, t)[σ (t)+ σ †(t)], (5.24)

where

Dmac(x, t) =√

�Ω1ε0εr√

εr

Lγsin

(√εr

Ω1

cx

)[c1(t)+ c†1(t)], (5.25)

with

γ = d

dΩ1(Ω1

√εr), (5.26)

is the macroscopic displacement field. By the relation (5.26), γ is the ratio betweenthe speed of light in the vacuum and the group velocity in the medium. The macro-scopic polarization Pmac(x, t) is given by the relation (5.21) simplified by leavingout its c2(t), c†2(t) polariton component. Then, from the relation

Emac(x, t) = 1

ε0

[Dmac(x, t)− Pmac(x, t)

], (5.27)

the macroscopic electric field is obtained. It is stated that the results of Dutra andFuruya (1997) for the macroscopic fields coincide with those derived by Milonni(1995) for the case of one and more modes.

de Lange and Raab (2006) recall series decompositions of D and H comprisingmacroscopic densities of multipole moments

Di = ε0 Ei + Pi − 1

2∇ j Qi j + 1

6∇k∇ j Qi jk + . . . , (5.28)

Hi = μ−10 Bi − Mi + 1

2∇ j Mi j + . . . , (5.29)

where Pi (Qi j , Qi jk, . . .) is an electric dipole (quadrupole, octupole, . . . ) densityand Mi (Mi j , . . .) is a magnetic dipole (quadrupole, . . . ) density, and, for monochro-matic fields and on concentration on nonmagnetic media,

Pi = αi j E j + 1

2ai jk∇k E j + . . .+ 1

ωG ′

i j B j + . . . , (5.30)

Qi j = aki j Ek + . . . , Qi jk ≈ 0, (5.31)

Mi = − 1

ωG ′

j i E j + . . .+ χi j B j + . . . , Mi j ≈ 0, (5.32)

Page 319: Quantum Aspects of Light Propagation

310 5 Microscopic Models as Related to Macroscopic Descriptions

where αi j , ai jk, . . . , G ′i j , . . . are material constants. For harmonic fields, these decom-

positions can be written as

Di = ε0 Ei + αi j E j + 1

2ikkai jk E j − 1

2ik j aki j Ek + . . .

−iG ′i j B j + . . . , (5.33)

Hi = −iG ′j i E j + . . .+ μ−1

0 Bi − χi j B j + . . . . (5.34)

The authors show a lack of translational invariance. On using the Maxwell equa-tions, these series can be recast into the form

Di = ε0 Ei + αi j E j + . . .− i

(G ′

i j −1

2ωε jklakli

)B j + . . . , (5.35)

Hi = −i

(G ′

j i −1

2ωεiklakl j

)E j + . . .+ μ−1

0 Bi + . . . , (5.36)

which contains the origin-independent material constants.For monochromatic fields and to describe magnetic media, it is necessary that

relations (5.30), (5.31), and (5.32) are extended by other terms,

Pi = 1

ωα′i j E j + 1

2ωa′i jk∇k E j + . . .+ Gi j B j + . . . , (5.37)

Qi j = − 1

ωa′ki j Ek + . . . , (5.38)

Mi = G ji E j + . . .+ 1

ωχ ′i j B j + . . . . (5.39)

The original terms are included in the ellipses. Here α′i j , α′i jk , . . ., Gi j , χ ′i j , . . .

are other material constants. For harmonic fields, the above decompositions canbe extended as

Di = −iα′i j E j + 1

2kka′i jk E j + 1

2k j a

′ki j Ek + . . .+ Gi j B j + . . . , (5.40)

Hi = −G ji E j + . . .+ iχ ′i j B j + . . . . (5.41)

The translational invariance is assessed also here. Using the Maxwell equations,these series can be recast into the form

Di = −iα′i j E j + 1

3kk(a′i jk + a′jki + a′ki j )Ei + . . .

+(

Gi j − 1

3Gllδi j − 1

6ωε jkla

′kli

)B j + . . . , (5.42)

Page 320: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 311

Hi =(−G ji + 1

3Gllδi j + 1

6ωεikla

′kl j

)E j + . . .

+ (iχ ′i j + ?)B j + . . . , (5.43)

where ? means nonmagnetic polarizability densities of electric multipole order 8 andmagnetic multipole order 4. This form comprises the origin-independent materialconstants.

5.2.2 Kramers–Kronig Dielectric

Dutra and Furuya (1998a) have pointed out that the Huttner–Barnett model at thestage after the diagonalization of the Hamiltonian for the polarization field and reser-voirs can operate with a larger class of dielectric functions than that admitted by theoriginal microscopic model. At this stage, the relative permittivity is expressed independence on the dimensionless coupling function ζ (ω),

ε(ω) = 1+ ω2c

2ωlim

ε→+0

∫ ∞

−∞

|ζ (ω′)|2(ω′ − ω − iε)ω′

dω′, (5.44)

where the notation from Section 4.1.1 has been used. For example, the permittivityobtained in the Lorentz oscillator model (Klingshirn 1995) can be recovered byadopting

ζ (ω) = ±i2ω√

ω

ω2 − ω20 − i2κω

√κ√π

, (5.45)

where ω0 means the frequency of the damped oscillations and κ is the frequency-independent absorption rate. The relative permittivity for the original Huttner–Barnett microscopic model is of the form

ε(ω) = 1− ω2c

ω2 − ω20 + ω0

2 F(ω), (5.46)

where

F(ω) ≡ limε→+0

∫ ∞

−∞

|V (ω′)|2ω′ − ω − iε

dω′. (5.47)

It is indicated that, in the Lorentz oscillator model, equation (5.46) yields the solu-tion

F(ω) = i4ωκ

ω0, (5.48)

Page 321: Quantum Aspects of Light Propagation

312 5 Microscopic Models as Related to Macroscopic Descriptions

but the integral equation (5.47) with this left-hand side (≡ replaced by =) is notsolvable to yield a coupling function V (ω). This is the main difficulty, because fromthe relation

ζ (ω) = i√

ω0ωV (ω)

ω2 − ω20 + ω0

2 F∗(ω), (5.49)

we obtain that

|V (ω)|2 = 4ωκ

πω0(5.50)

or we could determine

v(ω) = ρ

√ω0

ωV (ω) = ±ρ

√κ√π

. (5.51)

This presents a restriction of the Huttner–Barnett microscopic model, which is nev-ertheless mentioned by Huttner and Barnett (1992a,b).

5.2.3 Dissipative Oscillator Medium

Let us recall that in the microscopic model, the electromagnetic-field operators aregiven by integrals both over k and ω. Huttner and Barnett (1992a) say themselvesthat they lose the relationship between the frequency and the wave vector k. Thisobservation is relative to the macroscopic theories, where the Dirac delta functionsuitable for the expression of such a relationship is never replaced by another (gen-eralized) function. The quantities such as (4.49), (4.51), (4.52), and (4.53) are for-mulated in dependence on the relative permittivity ε(ω). Dutra and Furuya (1998b)suggest a simplification of the expression

ε(ω) = 1+ ω2c

2ωlim

ε→+0

∫ ∞

−∞

ξ (ω′)ω′ − ω − iε

dω′, (5.52)

where (Dutra and Furuya 1998b)

ξ (ω) = ω0ω|V (ω)|2

|ω2 − ω20 + ω0

2 F(ω)|2 , (5.53)

with F(ω) defined by the relation (5.47). They try to calculate the relative permit-tivity for the Huttner–Barnett microscopic model by means of classical electrody-namics. The Huttner–Barnett approach is applied to the particular case, where thecoupling strength is a slowly varying function of frequency.

In continuation of Dutra and Furuya (1997), a modified version of a simplemodel takes account of absorption. The inclusion of losses necessarily introduces

Page 322: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 313

a continuum of modes in the model. Consequences are minimized by the adoptionof the standard elimination of the reservoir.

The interaction between the radiation field and the medium is described by adipole-coupling Hamiltonian, where the canonically conjugated field is the dis-placement field instead of a minimal-coupling Hamiltonian, where the canonicallyconjugated field is the electric field. For simplicity, a Lorentzian shape for |V (ω)|2is assumed, given by the relation

V (ω) = iΔ

ω − ω0 + iΔ

√κ

π, (5.54)

where ω0 � Δ� κ . The Hamiltonian incorporating absorption is assumed to be

H (t) = �ωca†(t)a(t)+ �ω0

N∑

j=1

b†j (t)b j (t)− 1

ε

∫D(x, t)P(x, t) dx

+ �

∫ ∞

0

N∑

j=1

ΩW †j (Ω, t)W j (Ω, t) dΩ

+ �

∫ ∞

0

N∑

j=1

[V (Ω)b†

j (t)W j (Ω, t)+ V ∗(Ω)W †j (Ω, t)b j (t)

]dΩ, (5.55)

where ωc means the same as ω in the relations (5.6), (5.7), etc., W †j (Ω, t), W j (Ω, t)

are reservoir creation and annihilation operators that commute with every otheroperator except for the commutation relation

[W j (Ω, t), W †j ′ (Ω

′, t)] = δ j j ′δ(Ω−Ω′)1. (5.56)

Substituting the relations (5.6) and (5.8) into the Hamiltonian (5.55) and introduc-ing appropriate collective operators, the total Hamiltonian becomes a sum of twouncoupled Hamiltonians

H (t) = H1(t)+ H2(t). (5.57)

The second Hamiltonian H2(t) describes (N − 1) damped collective excitationsof the medium to which the field does not couple. The field and the single dampedcollective excitations of the medium, to which the field couples, are described bythe Hamiltonian H1(t) alone. This Hamiltonian is given by the relation

H1(t) = Hem(t)+ Hmat(t)+ Hint(t), (5.58)

where

Hem(t) = �ωca†(t)a(t) (5.59)

Page 323: Quantum Aspects of Light Propagation

314 5 Microscopic Models as Related to Macroscopic Descriptions

is the Hamiltonian of the field,

Hmat(t) = �ω0 B†(t)B(t)+ �

∫ ∞

0ΩY †(Ω, t)Y (Ω, t) dΩ

+ �

∫ ∞

0

[V (Ω)B†(t)Y (Ω, t)+ V ∗(Ω)Y †(Ω, t)B(t)

]dΩ (5.60)

is the Hamiltonian of medium, and

Hint(t) = �G(ωc)[a†(t)+ a(t)

] [B(t)+ B†(t)

](5.61)

is their interaction Hamiltonian. The collective annihilation operators B(t) andY (Ω, t) are given by the relations

B(t) =N∑

j=1

φ j b j (t) (5.62)

and

Y (Ω, t) =N∑

k=1

φk Wk(Ω, t), (5.63)

where

φ j = g j (ωc)

G(ωc)(5.64)

and g j (ωc) and G(ωc) are given by equations (5.15) and (5.16), respectively, with ω

replaced by ωc. Dutra and Furuya (1998b) have a (conventional) strictly microscopicmodel, where the medium is not continuous, but discrete. They admit the practicalityof the macroscopic average of the physical quantities. Following Robinson (1971,1973), they understand the macroscopic averaging as filtering out of higher spatialfrequencies.

A classical Hamiltonian is considered which is identical to the relation (5.58)except that a(t), B(t), and Y (Ω, t) will be replaced by a(t)√

�, B(t)√

�, and Y (Ω,t)√

�. The

standard elimination of the reservoir variables (which corresponds to the standardtreatment in the quantum theory, but with the bonus that the reservoir not beinginitially excited leads to a great simplification) is performed and the real variables

D(t) = a(t)+ a∗(t), (5.65)

B(t) = −i[a(t)− a∗(t)], (5.66)

P(t) = −i[B(t)− B∗(t)], (5.67)

X (t) = B(t)+ B∗(t) (5.68)

Page 324: Quantum Aspects of Light Propagation

5.2 Problem of Macroscopic Averages 315

are introduced. The variables D(t), B(t), and X (t) are related to the electric displace-ment field D(x, t), the magnetic field M(x, t), and the polarization field P(x, t) by

D(x, t) =√

ε0ωc

Lsin(ωc

cx)D(t), (5.69)

M(x, t) = −c

√ε0ωc

Lcos

(ωc

cx)B(t), (5.70)

P(x, t) = −2G ′(ωc)√

ε0

ω0Lsin(ωc

cx)X (t). (5.71)

From the classical Hamiltonian, the classical equations of motion for X (t) and P(t)are obtained,

d

dtX (t) = −κX (t)+ ω0P(t), (5.72)

d

dtP(t) = −ω0X (t)− κP(t)+ 2GD(t). (5.73)

These equations along with equations of motion for D(t) and B(t), which are notgiven here, admit a solution oscillating at the frequency ω, with the property thatdX (t)

dt = −iωX (t), dP(t)dt = −iωP(t), dD(t)

dt = −iωD(t). All the solutions have theproperty

X (t) = − 2ω0G(ωc)

ω20 + κ2 − ω2 − i2κω

D(t). (5.74)

From the relation

D(x, t) = ε(ω)[D(x, t)− P(x, t)], (5.75)

we obtain that

ε(ω) = 1+ 4[G ′(ωc)]2

ω′20 − ω2 − i2κω

, (5.76)

where

ω′20 = ω2

0 + κ2 − 4[G ′(ωc)]2 (5.77)

is the modified resonance frequency of the medium.The further topic in (Dutra and Furuya 1998b) is essentially the relation (4.76)

due to Gruner and Welsch (1995). In particular, it is shown that also in the case ofthe simple microscopic model of the medium used by Dutra and Furuya (1998b),it is possible to diagonalize first Hmat(t) and then the total Hamiltonian (5.58). Thediagonal form of the Hamiltonian Hmat(t) is achieved in terms of the continuous

Page 325: Quantum Aspects of Light Propagation

316 5 Microscopic Models as Related to Macroscopic Descriptions

operators B(ν, t),

B(ν, t) = α(ν)B(t)+∫ ∞

0β(ν,Ω)Y (Ω, t) dΩ, (5.78)

where α(ν) and β(ν,Ω) are some coefficients. The diagonal form of the total Hamil-tonian (5.58) is achieved in terms of the continuous operators A(ω, t),

A(ω, t) = α1(ω)a(t)+ α2(ω)a†(t)

+∫ ∞

0

[β1(ω, ν)B(ν, t)+ β2(ω, ν)B†(ν, t)

]dν, (5.79)

where α1(ω), α2(ω), β1(ω, ν), β2(ω, ν) are some coefficients. Suitable operators,namely, those of the electric displacement field and of the macroscopic electricstrength field are defined, such that

ˆE(x, t) = 1

ε0ε(ω)D(x, t)+ 2

√�

ω0ε0Lsin(ωc

cx)

× G ′(ωc)

[∫ ∞

0α∗(ω) A(ω, t) dω + H. c.

]. (5.80)

The difference from the relation (4.76) arises, because Dutra and Furuya (1998b)have only a single mode of the field, use a dipole-coupling Hamiltonian instead ofa minimal-coupling Hamiltonian, and have defined their field operators in terms ofdifferent quadratures of the annihilation and creation operators.

Using the microscopic approach, Hillery and Mlodinow (1997) devoted them-selves to the standard optical interactions and derived an effective Hamiltoniandescribing counterpropagating modes in a nonlinear medium. On considering mul-tipolar coupled atoms interacting with an electromagnetic field, a quantum theoryof dispersion has been obtained whose dispersion relations are equivalent to thestandard Sellmeir equations for the description of a dispersive transparent medium(Drummond and Hillery 1999).

Independently, the theory of light propagation in a Bose–Einstein condensate anda zero-temperature noninteracting Fermi–Dirac gas has been developed (Javanainenet al. 1999). Ruostekoski (2000) has theoretically studied the optical propertiesof a Fermi–Dirac gas in the presence of a superfluid state. He also considereddiffraction of atoms by means of light-stimulated transitions of photons betweentwo intersecting laser beams. Optical properties could possibly signal the presenceof the superfluid state and determine the value of the Bardeen–Cooper–Schriefferparameter in dilute atomic Fermi–Dirac gases. Crenshaw and Bowden (2000a,b)have derived effects of the Lorentz local fields on spontaneous emission in dielec-tric media. Bloch–Langevin operator equations have been obtained for two-levelatoms embedded in a host dielectric medium using the macroscopic and microscopicquantizations and the macroscopic formulation has been criticized (Crenshaw andBowden 2002).

Page 326: Quantum Aspects of Light Propagation

5.3 Single-Photon Models 317

Crenshaw (2003) has presented a real-space derivation of the macroscopic quan-tum Hamiltonian from the microscopic quantum electrodynamic model of a dielec-tric. Crenshaw (2004) has transformed the macroscopic real-space Hamiltonian tomomentum space. The microscopic model has been reduced to the macroscopicHamiltonian (Ginzburg 1940) by way of the reciprocal-space model of the field in adielectric (Hopfield 1958).

Cerboneschi et al. (2002) have shown that the electromagnetically induced trans-parency is related to very small group velocities for the probe pulse also in an opensystem. The modifications of the atomic momentum produced by laser interactionshave been taken into account.

Tanaka et al. (2003) have observed a negative delay (positive advance) of the peakof an optical pulse in case the pulse is tuned to the anomalous dispersion region. Thenegative velocity of the peak is not the velocity of energy flow.

5.3 Single-Photon Models

Guo (2007) has discussed propagation of one incident photon through a medium asthe multiple-scattering process from the medium. The medium is assumed to be anensemble of identical two-level atoms. Interaction with a two-level test atom outsidethe medium is considered as well. It is assumed that all the atoms are in the groundstate when t = 0. Initially, there is one photon in a mode. The system is treated inthe Schrodinger picture. An integral form of the evolution is presented.

As in Guo (2005), the kernel of the integral transformation is expanded. It isassumed that counter-rotating terms are superfluous and will be ignored. Photonpropagation through the atomic ensemble results in states S1, S1→2, S1→2→3,...,which come from the first-order scattering of the incident photon, from the second-and third-order scatterings of the incident photon, respectively. The photon is scat-tered into the states (modes) |1α〉, |1β〉, |1γ 〉,... in the time order. The author refers toclassical works (Mandel and Wolf 1995) and (Loudon 2000) for an ambiguity of thephoton phase. But a possibility for the photon to become entangled with the atomhas been declared.

The author expresses the time-dependent probability amplitude A for the testatom to transit from its ground state to an excited state. On the assumption ofisotropy and uniformity of the medium, the conservation of polarization and wave-number of the photon is derived. On these conditions A may be written in a formreminding of the expression for the electric field E(r, t) in Guo (2002). The factor4παatomk2n0 reappears as 4πnk2

0 Patom, where k0 = ω0c , �ω0 is the energy differ-

ence between the energy states of a two-level medium atom, n is the density of theatoms, and Patom is the resonant component of the polarizability of the atoms in theensemble.

Berman (2007) has considered the problem of a source atom radiating into amedium of dielectric atoms using a microscopic model. The model system con-sists of a source atom embedded in a dielectric. The source atom is modelled as a

Page 327: Quantum Aspects of Light Propagation

318 5 Microscopic Models as Related to Macroscopic Descriptions

two-level atom with a ground, lower, state |1〉 and an excited, upper, state |2〉. Thesestates are separated in frequency by ω0. The dielectric is contained within a sphereof radius R0, ω0

R0c � 1. This medium consists of a uniform distribution of atoms.

A bath density may be introduced and N is let to denote this density. Bath atom jis modelled as a four-level atom in the inverted tripod configuration with a groundstate |g〉( j) and three excited states |m〉( j), m = −1, 0, 1. The frequency separationbetween the ground state and any of the three excited states is ω. It is assumed that|Δ|ω� 1, where Δ = ω0 − ω, Δ �= 0. The positive frequency component of the

electric-field operator at the space point R, |R| = R, is

E+(R) = i∑

k,λ

√�ωk

2ε0Ve(λ)

k akλeik·R, (5.81)

where akλ is an annihilation operator for a photon having the propagation vector kand the polarization e(λ)

k , ωk = kc, V is the quantization volume, and

e(1)k ≡ e(1)(k) = cos(θk) cos(φk)ex + cos(θk) sin(φk)ey − sin(θk)ez,

e(2)k ≡ e(2)(k) = − sin(φk)ex + cos(φk)ey (5.82)

are unit polarization vectors, with

θk ≡ θ (k) = cos−1

(ez · k

|k|)

,

φk ≡ φ(k) = arg[(ex + iey) · k

]. (5.83)

In relation (5.81) the time is not written, since the Schrodinger approach has beenadopted. The source atom is excited by a pulse of a classical driving field, and theaverage field amplitude and the intensity radiated by the source atom are determinedto the first order in the dielectric density N . The expectation value of the field isdenoted as 〈E+(R, t)〉, where the time is given along with the position again. Thisexpectation value is

〈E+(R, t)〉 = i∑

k,λ

√�ωk

2ε0Ve(λ)

k ei(k·R−ωk t)bkλ(t)b∗1,0(t), (5.84)

where b1,0(t) is the amplitude to find all atoms in their ground states and to findno photons in the field and bkλ(t) is the amplitude to find all atoms in their groundstates and a photon having the wave vector k and the polarization λ in the field. It isassumed that the exciting field is weak enough to approximate

b1,0(t) ≈ 1. (5.85)

Page 328: Quantum Aspects of Light Propagation

5.3 Single-Photon Models 319

Berman (2007) has found the average field

〈E+(R, t)〉 ≈ −μω20(sin θ )e(1)ei(k0 R−ω0t)

4πε0c2 R

×{

1+ 2i1

ω0

∂t+ iδnk0 R0 − δn

R0

c

∂t

}b(0)

2,0

(t − R

c

), (5.86)

where θ ≡ θ (R), e(1) ≡ e(1)(R), k0 = ω0c , δn = − 1

4πε0

(2π Nμ′2

�Δ

), μ (μ′) is a charac-

teristic of the source atom (a bath atom), and b(0)2,0(t) is the excited state amplitude in

the absence of the medium (Berman 2004).Macroscopic theories are expected to give

〈E+(R, t)〉d = μ(sin θ )e(1)eik0(R+δn R0)

4πε0c2 R

× ∂2

∂t2

[b(0)

2,0

(t − R

c− δn R0

c

)e−iω0t

], (5.87)

where the subscript d stands for dielectric, δn = n − 1, and n is the index ofrefraction in the medium. If |k0δn R0| � 1 and | ( δn R0

c

)∂∂t b(0)

2,0

(t − R

c

) | � 1, thisexpression can be expanded to the first order in δn as

〈E+(R, t)〉d ≈ −μω20(sin θ )e(1)ei(k0 R−ω0t)

4πε0c2 R

×{

1+ 2i1

ω0

∂t+ iδnk0 R0 − 3δn

R0

c

∂t

}b(0)

2,0

(t − R

c

), (5.88)

which does not differ seriously from relation (5.86). Using relation (5.85), Berman(2007) has derived that the average field intensity is equal to

〈E−(R, t) · E+(R, t)〉 = |〈E+(R, t)〉|2, (5.89)

where E−(R, t) = E†+(R, t) and 〈E+(R, t)〉 is given by relation (5.86). He has given

a microscopic derivation of the fact that the field intensity propagates with a reducedvelocity in the medium. He has tried to elucidate the nature of the retardation mecha-nism and has shown that the modifications of the emitted field are closely correlatedwith nearly forward scattering in the medium.

Page 329: Quantum Aspects of Light Propagation

Chapter 6Periodic and Disordered Media

The periodic and disordered media have first been studied in the condensed-matterphysics. In this physics field, electrons are paid attention to. The concepts of peri-odic and disordered media depend on whether electrons or photons are investigated.Nevertheless there is an analogy between the wave function of a single electron andthe classical electromagnetic field. Therefore it was possible to achieve many resultsfor periodic and disordered media in photonics using this similarity.

The periodic media are conceptually simpler. Here we shall mainly speak ofthe macroscopic approach to quantization of the electromagnetic field in a peri-odic medium, but we shall also mention the papers, which have utilized specificapproaches. Finite periodic media can be coupling media of free-space modes. Inapplications the corrugated waveguides are important. Photonic crystals are infiniteor finite periodic media. The literature on one-dimensional periodic media abounds,since also fabrication of such media is simpler.

We shall mention the quantization of the electromagnetic field in a disorderedmedium mainly in connection with various physical studies. The macroscopicapproach to the quantization of the electromagnetic field suffices usually, but adescription of the disordered or random medium is not easy.

We shall cite the papers, whose authors have restricted themselves to the quantuminput–output relations and the detection theory. This is also related to applicationof results of further fields of the quantum physics. Many papers have reported therandom lasers. Even though we intend to review rather the theory, we see that we canonly provide a partial review. In the essence of the matter, it is that much theoreticalwork is not published as the optical physics.

6.1 Quantization in Periodic Media

Media whose dielectric constant is periodic are fabricated as microstructured materi-als with promising photonic band-gap structures. A number of methods for studyingthese media originate from the solid-state theory, where they are used in investiga-tion of ordinary, electronic crystals. This is a possible explanation why periodicmedia are called photonic crystals.

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 6, C© Springer Science+Business Media, LLC 2009

321

Page 330: Quantum Aspects of Light Propagation

322 6 Periodic and Disordered Media

The idea of photonic crystals was tested by the early experiments of Yablonovitchand Gmitter (1989). Since then the periodic media have been treated not only in thebooks devoted to optics (Born and Wolf 1999, Yeh 1988, Pedrotti and Pedrotti 1993)but also in monographs on the photonic crystals, e.g. (Joannapoulos 1995).

For illustration, we will treat a one-dimensional photonic crystal as Mishra andSatpathy (2003), whose contribution consists in an interesting method of solution,which is not reproduced here. We combine the assumption of a periodic mediumwith that of a rectangular waveguide.

Waveguides are useful optical devices. An optical circuit can be made using themand various optical couplers and switches. Classical theory of optical waveguidesand couplers has been elaborated in 1970s (Yariv and Yeh 1984), nonclassical lighthas been proposed as a source for improving performance, and the quantum theoryhas been gaining importance. Recently, quantum entanglement has been pointedout as another resource. Quantum descriptions may be very simple, but essentially,they ought to be based on a perfect knowledge of quantization. By way of para-dox, quantization is based on classical normal modes. Therefore, it is appropriateto concentrate ourselves on normal modes of rectangular mirror waveguide. It willbe assumed that the waveguide is filled with homogeneous refractive medium. Asthis has the only effect of changing the speed of light, it will be assumed that anonhomogeneity in a finite segment of the waveguide is present to model a coupler.

6.1.1 Classical Description of Electromagnetic Field

Vast literature has been devoted to the solution of the Maxwell equations, and theirvalue for the wave and quantum optics cannot be denied. Depending on the systemof physical units used, the Maxwell equations have several forms. Let us mentiononly two of them, appropriate to the SI units and the Gaussian units. The time-dependent vector fields, which enter these equations, are E(x, y, z, t), the electricstrength vector field, and B(x, y, z, t), the magnetic induction vector field. In fact,other two fields are used, but they can also be eliminated through the so-calledconstitutive relations. Saying this we make some simplifying assumptions, but weare tacit about them. The so-called monochromaticity assumption

E(x, y, z, t) = E(x, y, z; ω) exp(iωt),

B(x, y, z, t) = B(x, y, z; ω) exp(iωt) (6.1)

allows one to treat the time-independent Maxwell equations.As announced, we restrict ourselves to a rectangular mirror waveguide. We

assume that it has an infinite length, a width 2ax , and the height 2ay . The coordinatesystem is chosen so that the z-axis is the axis of the waveguide and the x-, y-axesare parallel with sides of the waveguide.

We deal with nonvanishing solutions of the time-independent Maxwell equations

∇ × 1

μ(x, y, z; ω)B(x, y, z; ω)− iωε(x, y, z; ω)E(x, y, z; ω) = 0, (6.2)

Page 331: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 323

∇ × E(x, y, z; ω)+ iωB(x, y, z; ω) = 0, (6.3)

∇ · [ε(x, y, z; ω)E(x, y, z; ω)] = 0, (6.4)

∇ · B(x, y, z; ω) = 0, (6.5)

where E(x, y, z; ω) and B(x, y, z; ω) are vector-valued functions in a domain

G = {(x, y, z) : −ax < x < ax ,−ay < y < ay,−∞ < z < ∞} (6.6)

and ω > 0 is a parameter. The desired solutions are to obey the boundary conditions

n(x, y, z)× E(x, y, z; ω) = 0, (6.7)

n(x, y, z) · B(x, y, z; ω) = 0, (6.8)

where n(x, y, z) is any unit outer-pointing normal vector at the point (x, y, z) ∈ ∂G,with

∂G = {(x, y, z) :(−ax ≤ x ≤ ax ∧ |y| = ay ∧ −∞ < z < ∞)

∨ (|x | = ax ∧ −ay ≤ y ≤ ay ∧ −∞ < z < ∞)}. (6.9)

The boundary conditions (6.7) and (6.8) are a formal expression of the fact that thewalls of the waveguide are perfect mirrors.

Here μ(x, y, z; ω) = μ0, ε(x, y, z; ω) is a function defined up to a finite numberof z-values such that

ε(x, y, z; ω) = ε0εr0, for z < 0, z > L ,

= ε0εr(z), for 0 < z < L , (6.10)

with μ0 > 0, μ0 = 4π × 10−7 Hm−1 the free-space magnetic permeability, ε0 > 0the free-space electric permittivity, εr0 > 0, εr(z) are relative electric permittivitiesof the medium. The medium electric permittivity ε(z) = ε0εr(z) has a period Λ,Λ | L , or ε(z) = ε(z + Λ), and is a positive function,

∫ L0 ε(z) dz = Lε0εr0. It is a

formal expression of the idea that a plate with a periodically modulated permittivityis contained in the waveguide.

6.1.2 Modal Functions

(i) Homogeneous refractive medium

We assume that the waveguide is filled with a homogeneous nonmagnetic refractivemedium, which is also nondispersive and lossless for simplicity. This assumptionholds on infinite intervals (−∞, 0) and (L ,∞) in the z-coordinate. On the finite

Page 332: Quantum Aspects of Light Propagation

324 6 Periodic and Disordered Media

interval (0, L), the medium is not homogeneous, but it is periodic. On average, itselectric permittivity equals to that of the homogeneous medium assumed.

For illustration, we will consider examples where solutions have finite norms, inSection 6.1.4. Let us assume that the relation ε(x, y, z; ω)= ε0εr0 holds everywherein G. We will express the solution in the form

E(x, y, z; ω) = E(x, y) exp(−ikzz),

B(x, y, z; ω) = B(x, y) exp(−ikzz), (6.11)

with kz �= 0. In analogy with the electromagnetic-field theory, from equations (6.2),(6.3), (6.4), and (6.5) we derive the time-independent wave equation

(Δ+ ω2

v2

)C = 0, (6.12)

where

v = 1√ε0εr0μ0

, (6.13)

and C ≡ C(x, y, z; ω) stands for E and B substitutionally. Respecting (6.11), wemay rewrite (6.12) in the form

(∂2

∂x2+ ∂2

∂y2

)C+

(ω2

v2− k2

z

)C = 0. (6.14)

Introducing the notation Er (x, y), Br (x, y), r = x, y, z, for the components ofthe vectors E(x, y), B(x, y), respectively, we have (Greiner 1998, p. 366)

Ex (x, y) = α cos

(mπ

2ax(x + ax )

)sin

(nπ

2ay(y + ay)

),

Ey(x, y) = β sin

(mπ

2ax(x + ax )

)cos

(nπ

2ay(y + ay)

),

Ez(x, y) = iγ sin

(mπ

2ax(x + ax )

)sin

(nπ

2ay(y + ay)

), (6.15)

Bx (x, y) = iα′ sin

(mπ

2ax(x + ax )

)cos

(nπ

2ay(y + ay)

),

By(x, y) = iβ ′ cos

(mπ

2ax(x + ax )

)sin

(nπ

2ay(y + ay)

),

Bz(x, y) = γ ′ cos

(mπ

2ax(x + ax )

)cos

(nπ

2ay(y + ay)

), (6.16)

Page 333: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 325

where m, n = 0, 1, . . . ,∞. Equation (6.14) yields the relation among m, n, kz ,and ω,

(mπ

2ax

)2

+(

2ay

)2

+ k2z =

ω2

v2. (6.17)

No solution of this kind exists if ω < ωg , where

ωg = ωmn = v

√(mπ

2ax

)2

+(

2ay

)2

. (6.18)

We can specify two linearly independent solutions for m, n = 0, 1, . . . ,∞,m + n ≥ 1,

αTE = iωky

k2x + k2

y

γ ′TE, α′TE =kx kz

k2x + k2

y

γ ′TE,

βTE = −iωkx

k2x + k2

y

γ ′TE, β ′TE =kykz

k2x + k2

y

γ ′TE, (6.19)

γTE = 0, γ ′TE = γ ′TE,

αTM = kx kz

k2x + k2

y

γTM, α′TM = 1

v2

iωky

k2x + k2

y

γTM,

βTM = kykz

k2x + k2

y

γTM, β ′TM = 1

v2

−iωkx

k2x + k2

y

γTM, (6.20)

γTM = γTM, γ ′TM = 0.

Here TE means transverse electric and TM transverse magnetic. For m = 0, a TEsolution exists, but no TM solutions exist; for n = 0, the same occurs and otherwise,both solutions exist. In (6.19) and (6.20), kx , ky are abbreviations, kx ≡ mπ

2ax, ky ≡

nπ2ay

. Let us remark that γ ′TE and γTM are complex parameters.

(ii) Plate with a periodically modulated permittivity

We generalize the first of relations (6.11) to the form

E(x, y, z; ω) = E(x, y)u(z), (6.21)

where u(z) is an unknown function, but E(x, y) is given by relations (6.15). We dis-tinguish the case of a TE solution from the case of a TM solution. In the latter case,this form cannot be required, since in the case of a TM solution, E(x, y) dependson kz . In fact the components kx , ky are meaningful in the inhomogeneous mediumwith the z-dependence of the dielectric constant, not the component kz .

In the case of a TE solution, the relation

∇ · E(x; ω) = 0 (6.22)

Page 334: Quantum Aspects of Light Propagation

326 6 Periodic and Disordered Media

holds, where x ≡ (x, y, z). The Maxwell equations are equivalent to a Helmholtzequation. It is derived that the function u(z) obeys the ordinary differential equation

d2

dz2u(z)+ k2

z (z)u(z) = 0, −∞ < z < ∞, (6.23)

where k2z (z) = k2(z)− (k2

x + k2y), with k2(z) ≡ ω2

c2 εr(z). A solution will be “sewn” ofa solution of the equation

d2

dz2u(z)+ k2

z u(z) = 0, −∞ < z < ∞ (6.24)

and a solution of the equation

d2

dz2u(z)+ k2

z (z)u(z) = 0, −∞ < z < ∞, (6.25)

where k2z (z) = k2(z)− (k2

x + k2y), with k2(z) ≡ ω2

c2 εr(z).The general solution of equation (6.25) has the form

u(z) = cB fB(z)eikBz + cF fF(z)e−ikBz, (6.26)

where cB and cF are arbitrary complex numbers, eikBΛ is a Bloch factor, with either0 < kB < π

Λor Im kB > 0. The functions fB(z) and fF(z) fulfil the conditions

fB(0) = fF(0) = 1 (6.27)

and have the period Λ. On differentiating relation (6.26) with respect to z, it can beseen that also d

dz u(z) has the same form in principle. And so the Bloch functions

u(z) = fB(z)eikBz, fF(z)e−ikBz (6.28)

are to be determined as the solutions of Equation (6.25) having along with therespective derivatives some initial data u(0), du

dz

∣∣0

and obeying the conditions

u(Λ) = λu(0),du

dz

∣∣∣∣z=Λ

= λdu

dz

∣∣∣∣z=0

, (6.29)

where λ is a complex number. From the fact that equation (6.25) does not containddz u(z), it can be derived that the transition matrix from the initial data u(0), du(z)

dz

∣∣∣z=0

to the respective values at z = Λ is unimodular. And so λ = exp (±ikBΛ). Sincethe transition matrix is real, kB is either real or pure imaginary.

Page 335: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 327

For illustration we may consider a Kronig–Penney dielectric (Mishra and Satpathy2003). In this case, it holds

kz(z) ={

k1z, 0 < z < Λ2 ,

k2z,Λ2 < z < Λ.

(6.30)

The equation for kB is obtained in the form

cos(kBΛ) = cos

(k1z

Λ

2

)cos

(k2z

Λ

2

)

−1

2

(k1z

k2z+ k2z

k1z

)sin

(k1z

Λ

2

)sin

(k2z

Λ

2

). (6.31)

We assume that the general solution of equation (6.23) has the form

u(z) =

⎧⎪⎪⎪⎪⎨

⎪⎪⎪⎪⎩

c(−)B eikz z + c(−)

F e−ikz z, for z < 0,

c(0)B fB(z)eikBz + c(0)

F fF(z)e−ikBz, for 0 < z < L ,

c(+)B eikz z + c(+)

F e−ikz z, for z > L .

(6.32)

Here fB(z), fF(z) are periodic functions such that fB(0) = fF(0) = 1. The solu-tion along with its first derivative is continuous at z = 0, L , which can be expressedas relations between c(−)

B , c(−)F , c(0)

B , c(0)F , c(+)

B , c(+)F . These coefficients are arbitrary

otherwise.The continuity of the (original) function at z = 0, L can be written as

c(−)B + c(−)

F = c(0)B + c(0)

F , c(0)B eikB L + c(0)

F e−ikB L = c(+)B eikz L + c(+)

F e−ikz L . (6.33)

The continuity of the derivative of the function at z = 0, L can be expressed as therelations

kzc(−)B − kzc

(−)F = fBc(0)

B + fFc(0)F ,

fBc(0)B eikB L + fFc(0)

F e−ikB L = kzc(+)B eikz L − kzc

(+)F e−ikz L , (6.34)

where

fB = −id

dz

[fB(z)eikBz

]∣∣∣∣0

, fF = −id

dz

[fF(z)e−ikBz

]∣∣∣∣0

. (6.35)

One of parametrizations of the solution of equation (6.23) is a dependence on c(−)B ,

c(−)F , which is

Page 336: Quantum Aspects of Light Propagation

328 6 Periodic and Disordered Media

⎝c(0)

B

c(0)F

⎠ = 1

fF − fB

×⎛

⎝fF − kz fF + kz

− fB + kz − fB − kz

⎝c(−)

B

c(−)F

⎠ , (6.36)

⎝c(+)

B

c(+)F

⎠ =⎛

⎝sBB sBF

sFB sFF

⎝c(−)

B

c(−)F

⎠ , (6.37)

where

sBB = 2

De−ikz L cos(kBL)kz

(− fF + fB)

+ i2

De−ikz L sin(kBL)

(k2

z − fB fF), (6.38)

sFF = 2

Deikz L cos(kBL)kz

(− fF + fB)

− i2

Deikz L sin(kBL)

(k2

z − fB fF), (6.39)

sBF = −i2

De−ikz L sin(kBL)

(kz + fB

) (kz + fF

), (6.40)

sFB = i2

Deikz L sin(kBL)

(kz − fB

) (kz − fF

), (6.41)

with

D = 2kz(− fF + fB

). (6.42)

Another of parametrizations of the solution (6.32) is a dependence on c(+)B , c(−)

F .Of remaining four coefficients, we first determine only c(−)

B , c(+)F . Their form is inter-

esting as an input–output relation

⎝c(−)

B

c(+)F

⎠ =⎛

⎝t ′ r

r ′ t

⎝c(+)

B

c(−)F

⎠ , (6.43)

where

r = − sBF

sBB, r ′ = sFB

sBB, t ′ = 1

sBB, (6.44)

t = sBBsFF − sFBsBF

sBB= 1

sBB, (6.45)

Page 337: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 329

since the determinant of the “scattering matrix” is∣∣∣∣sBB sBF

sFB sFF

∣∣∣∣ = 1. (6.46)

Finally, we may also express c(0)B and c(0)

F ,

⎝c(0)

B

c(0)F

⎠ = t

fF − fB

(fF − kz e−i(kB+kz )L

(fF + kz

)

− fB + kz ei(kB−kz )L(− fB − kz

))⎛

⎝c(+)

B

c(−)F

⎠ . (6.47)

In the case of a TM solution, we proceed quite generally. We introduce unknownfunctions α(z), β(z), γ (z) such that

Ex (x, y, z; ω) = α(z) cos [kx (x + ax )] sin[ky(y + ay)

],

Ey(x, y, z; ω) = β(z) sin [kx (x + ax )] cos[ky(y + ay)

],

Ez(x, y, z; ω) = iγ (z) sin [kx (x + ax )] sin[ky(y + ay)

]. (6.48)

The functions α(z) and β(z) are linearly dependent,

− kyα(z)+ kxβ(z) = 0, −∞ < z < ∞. (6.49)

Equation (6.23) is replaced by

d

dz

[kxα(z)+ kyβ(z)

] = −iγ (z)k2z (z),

d

dz

[k2(z)γ (z)

] = −ik2(z)[kxα(z)+ kyβ(z)

], (6.50)

where k(z) = k, kz(z) = kz for z < 0, z > L and k(z) = k(z), kz(z) = kz(z) for0 < z < L , where k(z), kz(z) have the period Λ. The equations

d

dz

[kxα(z)+ kyβ(z)

] = −iγ (z)k2z ,

d

dz

[k2γ (z)

] = −ik2[kxα(z)+ kyβ(z)

](6.51)

have a general solution

kxα(z)+ kyβ(z) = cBkzeikz z + cFkze

−ikz z,

k2γ (z) = −cBk2eikz z + cFk2e−ikz z, (6.52)

where cB and cF are arbitrary complex numbers. Let us note properties of theequations

Page 338: Quantum Aspects of Light Propagation

330 6 Periodic and Disordered Media

d

dz

[kxα(z)+ kyβ(z)

] = −iγ (z)k2z (z),

d

dz

[k2(z)γ (z)

] = −ik2(z)[kxα(z)+ kyβ(z)

]. (6.53)

The general solution of equations (6.53) has the form

kxα(z)+ kyβ(z) = cB fB⊥(z)eikBz + cF fF⊥(z)e−ikBz,

k2(z)γ (z) = cB fB3(z)eikBz + cF fF3(z)e−ikBz, (6.54)

where cB and cF are arbitrary complex numbers. The functions fB⊥(z), fF⊥(z) fulfilthe conditions

fB⊥(0) = fF⊥(0) = kz . (6.55)

It can be expected, e.g., that the inclusion of magnetic properties, with the neglect offrequency dependence, should lead to a small complication and a similar treatmentof the TE solution.

For α(z), β(z), γ (z), the Bloch factor eikBΛ is of importance. For the Kronig–Penney dielectric, the Bloch factor can be determined from the equation

cos(kBΛ) = cos

(k1z

Λ

2

)cos

(k2z

Λ

2

)

− 1

2

(k2

1k2z

k22k1z

+ k22k1z

k21k2z

)sin

(k1z

Λ

2

)sin

(k2z

Λ

2

). (6.56)

It is well known (Mishra and Salpathy 2003) that the relations (6.31) and (6.56) canbe written as a single one when appropriate notation is adopted. In fact, we observethe change

k1z → k21

k1z, k2z → k2

2

k2z. (6.57)

We assume a general solution, which is sewn from a solution of equations (6.51) forz < 0, from a solution of equations (6.53) for 0 < z < L , and from another solutionof equations (6.51) for z > L . Then we obtain modified formulae (6.44) and (6.45)adopting the changes

kz → k2

kz, fB →− 1

kzfB3(0), fF →− 1

kzfF3(0). (6.58)

Quantization of the electromagnetic field propagating in an ideal uniform waveg-uide is (for a fixed transverse mode) assumed in the article (Marinescu 1992), and itis shown that the appropriate scalar field obeys a Klein–Gordon type equation. Evena Dirac-type equation for the waveguides is considered in that article.

Page 339: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 331

6.1.3 Method of Coupled Modes

To any ω > min{ω10, ω01}, there exists only a finite number of modal functions(6.11), with (6.15) and (6.16). We will let J (ω), J (ω) ⊂ J , denote the correspondingindex set. We can partition this set into disjoint sets,

J (ω) = J+(ω) ∪ J0(ω) ∪ J−(ω), (6.59)

where J+(ω) is the set of indices that have kz > 0, J0(ω) is the set of indices thathave kz = 0, and J−(ω) is the set of indices that have kz < 0.

To any modal function E jin (x) with jin ∈ J+(ω), one (on the basis of physi-cal knowledge of propagation) looks for a solution of the problem formulated inSection 6.1.2 as follows. If z < 0 then

E(x) =∑

j∈{ jin}∪J−(ω)

A j (0)E j (x), (6.60)

B(x) =∑

j∈{ jin}∪J−(ω)

B j (0)B j (x), (6.61)

where A j (0) = B j (0) = 1 for j = jin. If z > L then

E(x) =∑

j∈J+(ω)

A j (L)E j (x), (6.62)

B(x) =∑

j∈J+(ω)

B j (L)B j (x). (6.63)

The coupled-mode method determines a form of the solution even for z ∈ [0, L] andit finds complex numbers A j (0), j ∈ J−(ω), and A j (L), j ∈ J+(ω). This method isapproximate. If 0 ≤ z ≤ L then

E(x) =∑

j∈J+(ω)∪J−(ω)

A j (z)E j (x), (6.64)

B(x) =∑

j∈J+(ω)∪J−(ω)

B j (z)B j (x). (6.65)

When the dependence of A j (z), B j (z) on the z-coordinate is weak (their variation isslow), we may consider

EF,B(x) =∑

j∈J±(ω)

A j (z)E j (x), (6.66)

BF,B(x) =∑

j∈J±(ω)

B j (z)B j (x). (6.67)

Page 340: Quantum Aspects of Light Propagation

332 6 Periodic and Disordered Media

Here

A j (z) =[∫ ∣∣E j (x⊥, z)

∣∣2 d2x⊥

]−1 ∫E∗j (x⊥, z) · EF,B(x)d2x⊥. (6.68)

Similarly B j (z) is determined.We assume TE waves. A possible use of the method for TM waves means

another, rough, approximation. We concentrate on the electric field and, therefore,we solve the Helmholtz equation

∇2E(x)+ εr(z)ω2

c2E(x) = 0. (6.69)

We will expound a means of describing the transformation

E(x) �→ [εr(z)− εr0]E(x), 0 ≤ z ≤ L . (6.70)

In fact we assume that

[εr(z)− εr0]E(x) = E′(x) =∑

j∈J+(ω)∪J−(ω)

A′j (z)E j (x), (6.71)

{[εr(z)− εr0] E(x)}F,B = E′F,B(x) =∑

j∈J±(ω)

A′j (z)E j (x), (6.72)

where

A′j (z) =[∫ ∣∣E j (x⊥, z)

∣∣2 d2x⊥

]−1 ∫E∗j (x⊥, z) · E′F,B(x)d2x⊥. (6.73)

Particularly,

[εr(z)− εr0]E j (x) =∑

j∈J+(ω)∪J−(ω)

K j j ′ (z)E j ′(x), (6.74)

{E′j (x)

}F,B

=∑

j ′∈J±(ω)

K j j ′ (z)E j ′(x), (6.75)

where

K j j ′ (z) =[∫ ∣∣E j ′ (x⊥, z)

∣∣2 d2x⊥

]−1 ∫E∗j ′ (x⊥, z) · {E′j (x)

}F,B

d2x⊥, (6.76)

i.e.,

A′j (z) =∑

j ′∈J+(ω)∪J−(ω)

K j ′ j (z)A j ′(z). (6.77)

Page 341: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 333

Further, the coupled-mode theory comprises an approximate expression (Yarivand Yeh 1984, Section 6.4, p. 197)

∇2E(x) =∑

j∈J+(ω)∪J−(ω)

h j (x), (6.78)

where

h j (x) = ∇2 A j (z)E j (x)+ 2∇A j (z) ·∇E j (x)+ A j (z)∇2E j (x)

≈ −2ik jzd

dzA j (z)E j (x)− εr0

ω2

c2A j (z)E j (x), (6.79)

or

∇2E(x) ≈ −2i∑

j∈J+(ω)∪J−(ω)

k jzd

dzA j (z)E j (x)− εr0

ω2

c2E(x). (6.80)

On substituting it into (6.69) and using (6.71), (6.77), we obtain a system of differ-ential equations

− 2ik jzd

dzA j (z) = −ω2

c2

j ′∈J+(ω)∪J−(ω)

K j ′ j (z)A j ′(z), (6.81)

to be solved on the boundary conditions

A j (0) ={

1 for j = jin,0 for j ∈ J+(ω)\{ jin}, (6.82)

A j (L) = 0 for j ∈ J−(ω). (6.83)

We will reduce the equations to the form

k jz

|k jz|d

dzA j (z) = − i

2

ω2

c2

j ′∈J+(ω)∪J−(ω)

K j ′ j (z)

|k jz| A j ′ (z). (6.84)

On physical grounds, it is expected that

j ′∈J+(ω)∪J−(ω)

k jz

|k jz| |A j (z)|2 (6.85)

is independent of z. We have

Page 342: Quantum Aspects of Light Propagation

334 6 Periodic and Disordered Media

k jz

|k jz|d

dzA j (z)A∗j (z)+ c.c.

= − i

2

ω2

c2

j ′∈J+(ω)∪J−(ω)

K j ′ j (z)

|k jz| A j ′ (z)A∗j (z)+ c.c. ≡ 0, (6.86)

where c.c. means complex conjugation, or the expected property is

K j ′ j (z)

|k jz| = K ∗j j ′(z)

|k j ′z| . (6.87)

6.1.4 Normalized Modes of the Electromagnetic Field

The normalization of modal functions of the electromagnetic field, which is madefor the sake of quantization, can be based, in optics, on a simple connection of thevector potential with the electric-field strength vector. This connection follows fromthe use of the so-called Coulomb gauge. First, we assume that only the electromag-netic field is present in the cavity (in the so-called nonrelativistic approximation). Inquantum optics, the simple instance of a perfectly closed cavity or resonator is oftenconsidered, which can be described mathematically in terms of the subset G of theusual Euclidean space R3 with the boundary ∂G.

In optics, the quantization is a definition of the vector potential operator A(x, t)by the relation

A(x, t) =∑

j∈J

[A(phot)

j (x, t)a j + A(phot)∗j (x, t)a†

j

], (6.88)

where J is an index set, and the photon annihilation and creation operators a j anda†

j in the j th mode fulfil the commutation relations

[a j , a†j ′ ] = δ j j ′ 1, [a j , a j ′ ] = [a†

j , a†j ′ ] = 0. (6.89)

Further

A(phot)j (x, t) =

√�

2ε0ω ju j (x) exp(−iω j t), (6.90)

with � the reduced Planck constant, ε0 the vacuum (free-space) electric permittivity,ω j and u j (x) satisfying the Helmholtz equation

∇2u j (x)+ ω2j

c2u j (x) = 0, (6.91)

the transversality condition

∇ · u j (x) = 0, (6.92)

Page 343: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 335

and the boundary conditions

nx × u j (x)∣∣∂G = 0, (6.93)

nx ·[∇ × u j (x)

]∣∣∂G = (nx ×∇) · u j (x)

∣∣∂G = 0, (6.94)

where nx is the normal vector at the point x. It is required that the modal functionsu j (x) be orthogonal and normalized as expressed by the relation

Gu j (x) · u j ′ (x) d3x = δ j j ′ . (6.95)

From relation (6.90), it is seen that the harmonic time dependence has the formexp(−iωt), not exp(iωt) as in relation (6.1). We have adopted this change althoughit may be a source of obscurity.

(i) Empty rectangular cavity

For illustration, we will assume that

G = {x : −ax < x < ax , −ay < y < ay, −az < z < az}, (6.96)

where ax , ay, az are positive. It can be proved that the index set J is a collection ofj = (nx , ny, nz, s), where nr ∈ 0, 1, . . . ,∞, r = x, y, z, s = TE, TM, nx > 0 ors = TE, ny > 0 or s = TE, and nz > 0 or s = TM, nx + ny + nz ≥ 2.

The solutions ω j have the form

ω j = c√

k2x + k2

y + k2z , (6.97)

with

c = 1√ε0μ0

, kr = nrπ

2ar, (6.98)

and the solutions u j (x) are connected with the classical solutions

E jx (x) = α j cos

(nxπ

2ax(x + ax )

)sin

(nyπ

2ay(y + ay)

)

× sin

(nzπ

2az(z + az)

),

E jy(x) = β j sin

(nxπ

2ax(x + ax )

)cos

(nyπ

2ay(y + ay)

)

× sin

(nzπ

2az(z + az)

),

Page 344: Quantum Aspects of Light Propagation

336 6 Periodic and Disordered Media

E jz(x) = γ j sin

(nxπ

2ax(x + ax )

)sin

(nyπ

2ay(y + ay)

)

× cos

(nzπ

2az(z + az)

), (6.99)

to the equivalent boundary-value problem

∇ · E j = 0, ∇ × B j + iω j

c2E j = 0, (6.100)

∇ · B j = 0, ∇ × E j − iω j B j = 0, (6.101)

nx · B j (x, ω j )∣∣∂G = 0, nx × E j (x, ω j )

∣∣∂G = 0. (6.102)

Here

α j = −iω jky

k2x + k2

y

γ ′j , β j = iω jkx

k2x + k2

y

γ ′j ,

γ j = 0, for s = TE, (6.103)

α j = − kx kz

k2x + k2

y

γ j , β j = − kykz

k2x + k2

y

γ j , for s = TM, (6.104)

with γ ′j , γ j complex parameters.The connecting relation is

u j (x) = −i

√2ε0

�ω jE(phot)

j (x), (6.105)

where

E(phot)j (x) =

r=x,y,z

E (phot)jr (x)er , (6.106)

with E (phot)jr (x) given by the formulae (6.99), (6.103), (6.104), in which

γ ′j = 2

√�

ε0ω j

√k2

x + k2y

(1+ δkx 0)(1+ δky 0)(1− δkz0)Vζ ′j , (6.107)

γ j = 2c

√�

ε0ω j

√k2

x + k2y

(1− δkx 0)(1− δky 0)(1+ δkz0)Vζ j (6.108)

are substituted, V = 8ax ayaz , |ζ ′j | = |ζ j | = 1.It can be easily derived that the vector-valued functions u j (x), j ∈ J , satisfy a

completeness relation,

Page 345: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 337

j∈J

u j (x)u∗j (x′) = δ(x− x′)1−∇x∇x′G(x, x′), (6.109)

where G(x, x′) is a Green’s function for the Laplace operator (the Dirichlet problem).

(ii) Rectangular cavity filled with a refractive medium

In this case, the quantization can be performed according to the relations (6.88),(6.89), (6.90), with ω j and u j (x) satisfying the Helmholtz equation

∇2u j (x)+ εr0

ω2j

c2u j (x) = 0, (6.110)

where εr0 is the relative electric permittivity of the medium, the transversality con-dition (6.92), and the boundary conditions (6.93) and (6.94). It is required that themodal functions u j (x) be orthogonal and normalized as expressed by the relation

Gεr0u∗j (x) · u j ′ (x) d3x = δ j j ′ . (6.111)

For illustration, we will assume that G is defined by (6.96). The solutions ω j

have the form

ω j = v

√k2

x + k2y + k2

z , (6.112)

with

v = 1√ε0εr0μ0

, kr = nrπ

2ar, r = x, y, z, (6.113)

and the solutions u j (x) are connected with the solutions of the form (6.99) to theequivalent boundary-value problem

∇ · E j = 0, ∇ × B j + iεr0ω j

c2E j = 0, (6.114)

∇ · B j = 0, ∇ × E j − iω j B j = 0, (6.115)

nx · B j (x, ω j )∣∣∂G = 0, nx × E j (x, ω j )

∣∣∂G = 0. (6.116)

The connecting relation is (6.105) with (6.106), where E (phot)jr (x) are given by the

formulae (6.99), (6.103), (6.104), in which

Page 346: Quantum Aspects of Light Propagation

338 6 Periodic and Disordered Media

γ ′j = 2

√�

ε0εr0ω j

√k2

x + k2y

(1+ δkx 0)(1+ δky 0)(1− δkz0)Vζ ′j , (6.117)

γ j = 2c

√�

ε0εr0ω j

√k2

x + k2y

(1− δkx 0)(1− δky 0)(1+ δkz0)Vζ j (6.118)

are substituted.It can be easily derived that the vector-valued functions u j (x), j ∈ J , satisfy a

completeness relation∑

j∈J

εr0u j (x)u∗j (x′) = δ(x− x′)1−∇x∇x′G(x, x′), (6.119)

where G(x, x′) is a Green’s function for the Laplace operator (the Dirichlet problem).

(iii) Rectangular waveguide filled with a refractive medium and located in a flatspace

We will consider a subset G = G⊥ × S1(−az ≤ z < az), with the boundary ∂G =∂G⊥ × S1(−az ≤ z < az), of a flat non-Euclidean space R2 × S1(−az ≤ z < az),where S1(−az ≤ z < az) means a topological circle of the length 2az . In thiscase, the quantization can be performed according to the relations (6.88), (6.89),(6.90), with ω j and u j (x) satisfying the Helmholtz equation of the form (6.110), thetransversality condition (6.92), and the boundary conditions (6.93) and (6.94). It isrequired that the modal functions u j (x) be orthogonal and normalized as expressedby the relation (6.111).

For illustration, we will assume that

G = {x : −ax < x < ax ,−ay < y < ay,−az ≤ z < az}, (6.120)

where ax , ay, az are positive. It can be proved that the index set J is a collection ofj = (nx , ny, nz, s), where nr ∈ {0} ∪ N, r = x, y, nz ∈ Z, s = TE, TM, nx > 0 ors = TE, ny > 0 or s = TE, and nx + ny ≥ 1.

The solutions ω j have the form

ω j = v

√k2

x + k2y + k2

z , (6.121)

with

v = 1√ε0εr0μ0

, kr = nrπ

2ar, r = x, y, kz = nzπ

az, (6.122)

and the solutions u j (x) are connected with the classical solutions

E jx (x) = α j cos

(nxπ

2ax(x + ax )

)sin

(nyπ

2ay(y + ay)

)

× exp

(inzπ

az(z + az)

),

Page 347: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 339

E jy(x) = β j sin

(nxπ

2ax(x + ax )

)cos

(nyπ

2ay(y + ay)

)

× exp

(inzπ

az(z + az)

),

E jz(x) = iγ j sin

(nxπ

2ax(x + ax )

)sin

(nyπ

2ay(y + ay)

)

× exp

(inzπ

az(z + az)

), (6.123)

to the equivalent boundary-value problem of the forms (6.114), (6.115), (6.116).The connecting relation is (6.105) with (6.106), where E (phot)

jr (x) are given by theformulae (6.123), (6.103), (6.104), in which

γ ′j =√

2�

ε0εr0ω j

√k2

x + k2y

(1+ δkx 0)(1+ δky 0)Vζ ′j , (6.124)

γ j = c

√2�

ε0εr0ω j

√k2

x + k2y

(1− δkx 0)(1− δky 0)Vζ j (6.125)

are substituted. It can be easily derived that the vector-valued functions u j (x), j ∈ J ,satisfy a completeness relation (6.119). Usually the scalar product is considered

Gεr(z)u∗(x) · u(x) d3x. (6.126)

With these solutions of the problem defined, they have finite norms.

(iv) Rectangular waveguide filled with a homogeneous refractive medium

We will consider a subset G = G⊥ × R1, with the boundary ∂G = ∂G⊥ × R1, ofthe usual Euclidean space R3. In optics, the quantization may be a definition of thevector potential operator A(x, t) by the relation

A(x, t) =∑

j⊥∈J⊥

∫ ∞

−∞

[A(phot)

j⊥ (x, kz, t)a j⊥ (kz)

+ A(phot)∗j⊥ (x, kz, t)a†

j⊥ (kz)]

dkz, (6.127)

where J⊥ is an index set and the photon annihilation, and creation operators a j⊥ (kz)and a†

j⊥ (kz) in the mode ( j⊥, kz) fulfil the commutation relations

[a j⊥ (kz), a†j ′⊥

(k ′z)] = δ j⊥ j ′⊥δ(kz − k ′z)1,

[a j⊥ (kz), a j ′⊥ (k ′z)] = [a†j⊥(kz), a†

j ′⊥(k ′z)] = 0. (6.128)

Page 348: Quantum Aspects of Light Propagation

340 6 Periodic and Disordered Media

Further

A(phot)j⊥ (x, kz, t) =

√�

2ε0ω j⊥ (kz)u j⊥ (x, kz) exp[−iω j⊥(kz)t], (6.129)

with ε0 the vacuum (free-space) electric permittivity, ω j⊥(kz) andu j⊥ (x, kz) satisfying the Helmholtz equation

∇2u j⊥ (x, kz)+ εr0

ω2j⊥(kz)

c2u j⊥ (x, kz) = 0, (6.130)

the transversality condition

∇ · u j⊥ (x, kz) = 0, (6.131)

and the boundary conditions

nx × u j⊥ (x, kz)∣∣∂G = 0, (6.132)

nx ·[∇ × u j⊥ (x, kz)

]∣∣∂G = (nx ×∇) · u j⊥ (x, kz)

∣∣∂G = 0, (6.133)

where nx is the normal vector at the point x. It is required that the modal functionsu j⊥ (x, kz) be orthogonal and normalized as expressed by the relation

Gεr0u∗j⊥kz

(x, kz) · u j ′⊥k ′z (x, kz) d3x = δ j⊥ j ′⊥δ(kz − k ′z). (6.134)

For illustration, we will assume that

G = {x : −ax < x < ax ,−ay < y < ay,−∞ < z < ∞}, (6.135)

where ax , ay are positive. It can be proved that the index set J⊥ is a collection ofj⊥ = (nx , ny, s), where nr ∈ {0} ∪ N, r = x, y, s = TE, TM, nx > 0 or s = TE,ny > 0 or s = TE, and nx + ny ≥ 1.

The solutions ω j⊥(kz) have the form

ω j⊥ (kz) = v

√k2

x + k2y + k2

z , (6.136)

with

v = 1√ε0εr0μ0

, kr = nrπ

2ar, r = x, y, (6.137)

and the solutions u j⊥ (x, kz) are connected with the classical solutions

E j⊥x (x, kz) = α j⊥(kz) cos

(nxπ

2ax(x + ax )

)

× sin

(nyπ

2ay(y + ay)

)exp (ikzz) ,

Page 349: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 341

E j⊥ y(x, kz) = β j⊥ (kz) sin

(nxπ

2ax(x + ax )

)

× cos

(nyπ

2ay(y + ay)

)exp (ikzz) ,

E j⊥z(x, kz) = iγ j⊥(kz) sin

(nxπ

2ax(x + ax )

)

× sin

(nyπ

2ay(y + ay)

)exp (ikzz) , (6.138)

to the equivalent boundary-value problem

∇ · E j⊥ = 0, ∇ × B j⊥ + iεr0ω j⊥ (kz)

c2E j⊥ = 0, (6.139)

∇ · B j⊥ = 0, ∇ × E j⊥ − iω j⊥(kz)B j⊥ = 0, (6.140)

nx · B j⊥

∣∣∂G = 0, nx × E j⊥

∣∣∂G = 0. (6.141)

Here E j⊥ ≡ E j⊥ (x, kz, ω j⊥ (kz)), B j⊥ ≡ B j⊥ (x, kz, ω j⊥ (kz)),

α j⊥ (kz) = −iω j⊥ (kz)ky

k2x + k2

y

γ ′j⊥ , β j⊥ (kz) = iω j⊥(kz)kx

k2x + k2

y

γ ′j⊥ ,

γ j⊥ (kz) = 0, for s = TE, (6.142)

α j⊥(kz) = − kx kz

k2x + k2

y

γ j⊥ , β j⊥ (kz) = − kykz

k2x + k2

y

γ j⊥ ,

γ j⊥(kz) = γ j⊥ , for s = TM, (6.143)

with γ ′j⊥ , γ j⊥ complex parameters.The connecting relation is

u j⊥ (x, kz) = −i

√2ε0

�ω j⊥E(phot)

j⊥ (x, kz), (6.144)

where

E(phot)j⊥ (x, kz) =

r=x,y,z

E (phot)j⊥r (x, kz)er , (6.145)

with E (phot)j⊥r (x, kz) given by the formulae (6.138), (6.142), (6.143), in which

γ ′j⊥ (kz) =√

ε0εr0ω j⊥ (kz)

√k2

x + k2y

(1+ δkx 0)(1+ δky 0)πV⊥ζ ′j⊥ , (6.146)

γ j⊥(kz) = c

√�

ε0εr0ω j⊥ (kz)

√k2

x + k2y

(1− δkx 0)(1− δky 0)πV⊥ζ j⊥ (6.147)

Page 350: Quantum Aspects of Light Propagation

342 6 Periodic and Disordered Media

are substituted, V⊥ = ax ay , |ζ ′j⊥| = |ζ j⊥| = 1.The vector-valued functions u j⊥ (x, kz), j⊥ ∈ J⊥, satisfy a completeness relation,

j⊥∈J⊥

∫ ∞

−∞εr0u j⊥ (x, kz)u∗j⊥ (x′, kz) dkz

= δ(x− x′)1−∇x∇x′G(x, x′), (6.148)

where G(x, x′) is a Green’s function for the Laplace operator (the Dirichlet problem).

(v) Rectangular waveguide filled with a nonhomogeneous refractive medium

Quantum electrodynamics in periodic dielectric media has been treated severaltimes (Caticha and Caticha 1992, Kweon and Lawandy 1995, Tip 1997). For thequantization true modal functions may be useful which we have found in theSection 6.1.2(ii). They have not been presented just as needed, but complex con-jugated. We must consider another difficulty that they are not orthogonal.

6.1.5 Quantization in Linear NonhomogeneousNonconducting Medium

Tip (1997) reminds one of the fact that a quantization of the electromagnetic fieldis made for a complete quantum description of interaction of radiation with atomsor molecules. These however may be placed in a “photonic material”, and so aquantization of the electromagnetic field in the material is purposeful.

Tip (1997) points out the fact that photonic crystals are classical dielectric mediawith a periodic dielectric permittivity. The periodicity may give rise to a band struc-ture. If a band gap is present and an embedded atom has a transition frequencyin the gap, a single-photon emission is inhibited. However, even if no gap develops,decay rates may differ much from their free values. Free decay rates can be obtainedthrough an application of Fermi’s golden rule. Generalization for any electric per-mittivity and magnetic permeability uses the so-called local density of states relatedto the classical electromagnetic field again.

The author expounds a general approach to the quantization of linear evolutionequations. Such an equation for F(t) from a separable, real Hilbert space has theform

∂ F(t)

∂t= N F(t)− G(t), N † = −N . (6.149)

The matter is a quantization of the quantity F(t).Tip (1997) assumes a linear, nonhomogeneous, nonconducting material medium

including external currents or a Schrodinger quantum particle system. We havemostly adopted the notation and have changed it in part only. In particular,

Page 351: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 343

ε0 = μ0 = 1 upon the choice of units. The case ε = μ = 1 is called the vacuum caseand, with arbitrary ε and μ, one may still speak of a free electromagnetic field. Wewill still use the asterisk for complex conjugates, the centred dot in multiplicationof matrices, which indicates that they are treated as tensors.

In general, an n-component field over the space Rd × R is considered, whichsatisfies the equation

∂tf(x, t) = M (x,∇) · f(x, t), (6.150)

where x ∈ Rd , and M ≡ M (x,∇) is an n × n matrix, whose entries are real partialdifferential operators with, in general, variable coefficients. There exists a bounded,real, invertible matrix ρ(x) with bounded inverse such that

M† · ρ2(x)+ ρ2(x) ·M = 0, (6.151)

or the “energy”

E = 1

2

∫[ρ(x) · f(x, t)]2 d3x (6.152)

can be introduced. Considering

F(x, t) = ρ(x) · f(x, t), (6.153)

N = ρ ·M · ρ−1, (6.154)

where N ≡ N (x,∇), we obtain the equations

∂tF(x, t) = N (x,∇) · F(x, t), (6.155)

N† = −N. (6.156)

On introducing, in addition, a norm

‖F(x, t)‖2 = 2E, (6.157)

equation (6.155) describes a unitary time evolution on the real space Hr =L2(Rd , dx; Rn)

F(x, t) = exp(Nt) · F(x, 0). (6.158)

Tip (1997) intends to work with a real Hilbert space as long as possible. Hebegins with two examples of field equations with a conservation law. Proceed-ing with Maxwell’s equations for a nonconducting material medium, he writes theMaxwell equations, which are equations for real waves originally. He assumes that

Page 352: Quantum Aspects of Light Propagation

344 6 Periodic and Disordered Media

ε(x) and μ(x) are real, smooth, bounded from below and above by positive con-stants.

For J = 0, the energy is conserved,

E = 1

2

∫ {ε(x)[E(x, t)]2 + 1

μ(x)[B(x, t)]2

}d3x

= 1

2

∫|F(x, t)|2 d3x, (6.159)

where

F(x, t) =(√

ε E(x, t)1√μ

B(x, t)

)=(

F1(x, t)F2(x, t)

). (6.160)

Relation (6.154) becomes

N =(

0 N12

N21 0

)=(

0 − 1√ε(ε · ∇) 1√

μ1√ε(ε · ∇) 1√

μ0

), (6.161)

where ε is the Levi-Civita pseudotensor, or

N = W · N0 ·W, (6.162)

where

N0 =(

0 −ε · ∇ε · ∇ 0

), W =

(1√ε1 0

0 1√μ

1

). (6.163)

The orthogonal eigenprojector of the matrix N at the eigenvalue 0 is

P0 =(

P1 00 P2

)=(√

ε∇[∇ · ε∇]−1∇√ε 00

√μ∇[∇ · μ∇]−1∇√μ

). (6.164)

Here [∇ · ε∇]−1 and [∇ · μ∇]−1 can be expressed by an integral transform in thevacuum case.

Tip (1997) pays attention to the Helmholtz operators and to the scattering theory,which is not reproduced here. The Helmholtz operators are

H1 = −N12N21 = 1√ε

(ε · ∇) · 1

μ(ε · ∇)

1√ε,

H2 = −N21N12 = 1√μ

(ε · ∇) · 1

ε(ε · ∇)

1√μ

. (6.165)

Let uλα denote eigenvectors of H1,

H1 · uλα = λ2uλα, (6.166)

Page 353: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 345

where λ ≥ 0 and α distinguish eigenvectors at the same eigenvalue. As H1 is a realoperator, u∗λα differs from some uλβ by a constant factor only. We can always usereal eigenvectors. As H2 = N−1

12 ·H1 ·N12 = N21 ·H1 ·N−121 , the eigenvector H2 can

be obtained for instance as N21 · uλα .The exposition of the Lagrange formalism is introduced by a warning that in the

case

H = H1 ⊕H2, N =(

0 N12

N21 0

), N21 = −N †

12, (6.167)

equation (6.149) for G = 0 is not obtained using the Lagrange formalism, if we

take F for the coordinate field. For the two components of the vector F =(

F1

F2

),

separate equations are obtained and their connection is lost.It is recommended to use ξ = N−1 F as coordinate field. In the scalar wave

case, this means to define the coordinate field using solution of a generalization ofthe Poisson equation. In the Maxwell case, N cannot be inverted due to the zeroeigenvalue. Tip (1997) reminds one of projection upon the propagating modes. Inthe vacuum situation

(ξ 1ξ 2

)= N−1

0

(E⊥

B

)(6.168)

is introduced. The quantity−ξ 1 is the vector potential A in the Coulomb gauge. Thetheory comprises the relations

E⊥ = −∂A∂t

, B = ∇ × A,

E‖ = −∇Φ, Φ(x) =∫

ρ(x′)4π |x− x′| d3x′, ρ(x) = ∇ · E(x), (6.169)

where ρ(x) is the external charge density.In general case, we let P mean the projector upon the null space N = N (N ) of

N and Q = 1− P . It holds that

P =(

P1 00 P2

), Q =

(Q1 00 Q2

), Q j = 1− Pj , (6.170)

where Pj acts in H j , j = 1, 2. We choose G =(

G1

G2

). As

(0 N12

N21 0

)−1

=(

0 −N−121

−N−112 0

), (6.171)

a quantity

ξ = N−121 Q2 F2 (6.172)

Page 354: Quantum Aspects of Light Propagation

346 6 Periodic and Disordered Media

is introduced satisfying the generalization of the Coulomb gauge condition

P1ξ = 0. (6.173)

In terms of this quantity, propagating components can be expressed,

Q1 F1 = −∂ξ

∂t− N−1

21 Q2G2,

Q2 F2 = −N21ξ . (6.174)

Tip (1997) also considers a generalization of a gauge transformation

ξ = ξ + P1η, η ∈ H1. (6.175)

We make a replacement N−121 Q2G2 → N−1

21 Q2G2 − P1∂η

∂t ,

Q1 F1 = −∂ξ

∂t− N−1

21 Q2G2 + P1∂η

∂t(6.176)

and

Q2 F2 = −N21ξ. (6.177)

With respect to the application to the Maxwell equations, we assume that

G =(

G1

0

), P2 F2|t=0 = 0. (6.178)

Equations (6.176) and (6.177) simplify,

Q1 F1 = −∂ξ

∂t+ P1

∂η

∂t, (6.179)

F2 = −N21ξ. (6.180)

Tip (1997) introduces a generalization ξ0 of the scalar potential Φ of the Maxwelltheory. He considers another real Hilbert space H3 and an invertible operator Mfrom H3 into P1H1. Then

ξ0 = −M−1 P1

(F1 + ∂η

∂t

). (6.181)

He introduces a generalization ρ of the charge density,

ρ = −M†P1 F1. (6.182)

Page 355: Quantum Aspects of Light Propagation

6.1 Quantization in Periodic Media 347

He lets (. , .) j mean the inner product in H j and presents the Lagrangian

L = 1

2

(∂ξ

∂t+ Mξ0,

∂ξ

∂t+ Mξ0

)

1

− 1

2(N21ξ, N21ξ )2

+ (G1, ξ )1 − (ρ, ξ0)3 . (6.183)

Tip (1997) considers generalizations of the Coulomb, Lorentz, and temporalgauges. The field η has specific properties in each particular gauge. Expoundingthe C gauge, he writes the condition

P1ξ = 0, (6.184)

which leads to equations

∂2ξ

∂t2− N12 N21ξ = Q1G1,

M†Mξ0 = ρ. (6.185)

He eliminates ξ0 by expressing it in terms of ρ and presents a Lagrangian, the canon-ical momentum field associated with ξ , π = ξ , and a Hamiltonian.

Expounding the L gauge, he needs the condition

∂ξ0

∂t− M†P1ξ = 0, (6.186)

which leads to equations

∂2ξ0

∂t2+ M†Mξ0 = ρ, (6.187)

∂2ξ

∂t2− N12 N21ξ + M M†P1ξ = G1. (6.188)

He writes a Lagrangian, the momentum field π0 = ξ0, and a Hamiltonian.Expounding the T gauge, the author writes the condition (and the equation)

ξ0 = 0, (6.189)

which leads to an equation

∂2ξ

∂t2− N12 N21ξ = G1. (6.190)

He presents a Lagrangian and a Hamiltonian.At this level, Tip (1997) reminds one of the familiar method of canonical quan-

tization and concentrates himself on the C gauge case and the L gauge case. The C

Page 356: Quantum Aspects of Light Propagation

348 6 Periodic and Disordered Media

gauge case is discussed in full detail. He chooses an orthonormal basis {u j } in thesubspace Q1H1 ⊂H1 and decomposes

ξ =∑

j

ξ j u j , π =∑

j

π j u j , (6.191)

where ξ j = ξ (u j ) = (ξ, u j )1, π j = π (u j ) = (π, u j )1. He expresses the Hamiltonianin terms of ξ j , π j , which obey the Poisson brackets

{ξ j , πk} = δ jk . (6.192)

A quantization is accomplished by replacing the Poisson brackets by the commutators

[ξ (u j ), π (uk)] = iδ jk 1. (6.193)

We utilize hats here, although Tip (1997) does not write them, or lets them meansomething else. He defines the complexifications of the Hilbert spaces and the Fockspace F(H) over any Hilbert space H. He introduces the annihilation (creation)operator a(ϕ) (a†(ϕ)), which acts in F(H), where ϕ is the wave function of anannihilated (created) boson. In a rather complicated manner, it must be shown thatthe Hamiltonian can be expressed in terms of a(u j ), a†(u j ), which act in F(H),where H is the complexification of H1.

In the comment on the L gauge case, Tip (1997) chooses, in addition, anorthonormal basis {v j } in the subspace P1H and such a basis {w j } in the complexifi-cation H′ of the real space H3. He expounds that the Hamiltonian can be expressedin terms of a(u j ), a(v j ), b(w j ) and their Hermitian conjugates. Here b(w j ) is theannihilation operator, which acts in F(H′).

Applying this formalism to the Maxwell equations, he determines that

ξ = −√εA,π = −√ε∂A∂t

, ξ0 = Φ, M = √ε∇. (6.194)

Moreover, it holds that M† =−∇√ε, M† ·M=−∇√ε ·√ε∇, MM† =−√ε∇∇√ε.In the C gauge, the condition (6.184) becomes

∇ · εA = 0, (6.195)

leading to the equations

∂2(εA)

∂t2+ ∇ × 1

μ(∇ × A) = Q1 · J, (6.196)

∇√ε · √ε∇Φ = −ρ, (6.197)

Page 357: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 349

where J is the external current density. Tip (1997) also presents a Lagrangian and aHamiltonian. The Poisson brackets have the form

{ξ (x),π (y)} = Q1(x, y), (6.198)

where Q1(x, y) = δ(x− y)− P1(x, y), P1(x, y) are kernels associated with the pro-jectors Q1 and P1, respectively.

In the L gauge, the condition (6.186) becomes

∂Φ

∂t+ ∇ · εA = 0, (6.199)

leading to the equations of motion

∂2Φ

∂t2−∇ · ε∇Φ = ρ, (6.200)

∂2A∂t2

+ 1

ε∇ × 1

μ(∇ × A)−∇2εA = 1

εJ. (6.201)

A Lagrangian, the momentum field π0 = ∂Φ∂t , and a Hamiltonian are also presented.

The Poisson brackets have the form

{Φ(x),π0(y)} = δ(x, y),

{ξ (x),π (y)} = 1δ(x, y). (6.202)

Tip (1997) devotes an appropriate amount of place to the application to the atomicradiative decay in dielectrics.

Under the usual assumptions on the dielectric permittivity, a quantization ofthe Hamiltonian formalism of the electromagnetic field using a method close tothe microscopic approach was performed by Tip (1998). A proper definition ofband gaps in the periodic case and a new continuity equation for energy flow wasobtained, and an S-matrix formalism for scattering from absorbing objects wasworked out. In this way, the generation of Cerenkov and transition radiation havebeen investigated.

6.2 Corrugated Waveguides

The use of dielectric optical waveguides (Marcuse 1974, Yariv and Yeh 1984) andthe coupled-mode theory appropriate in the case of corrugated waveguides lead nat-urally to the task of describing the propagation of a general quantum state in thesedevices. The question of possibility and impossibility of quantizing the classicaldescription is not posed usually. The apology for it is that the copropagation doesnot make difficulties. The description can be analyzed in the framework of the the-oretical mechanics with the time variable replaced by the propagation distance. Wework easily with the Heisenberg and Schrodinger pictures and formulate respective

Page 358: Quantum Aspects of Light Propagation

350 6 Periodic and Disordered Media

equations. The quantum momentum operator can be derived with respect to themodal orthonormalization property on a cross section of an optical waveguide(Linares and Nistal 2003). The difficulties due to counterpropagation are regularlydisregarded. In fact, the classical description is free of difficulties, and the theoreticalmechanics that has been already modified by the replacement of the time variablewith the space variable can be extended to involve the counterpropagation (Luis andPerina 1996b). It has been concluded that the quantization will not be successfulin the case of counterpropagation in a nonlinear medium (except optical parametricprocesses).

The propagation in a linear dielectric, and in the devices based on such materi-als, can be quantized in time (Dalton et al. 1996, Dalton et al. 1999b, Dalton andKnight 1999a,b). Here we shall not criticize the use of operators in situations whichare classical essentially, since the literature abounds with this (Janszky et al. 1988,Perina 1995a,b, Perina and Perina, Jr. 1995b,c, Korolkova and Perina 1997c). Thedependence of outputs on inputs can be formulated in the spatial Heisenberg andSchrodinger pictures without respective differential equations. This restriction isdue to the peculiar nature of quantization. In the copropagation, the spatio-temporaldescription (cf. (Luks and Perinova 2002)) has not been used and therefore theunusual description has not been used by anybody even in the counterpropagation,where it is hardly dispensable.

The simplified quantization enables one to employ knowledge of quantum mecha-nical descriptions as follows. An amplifier should be described in the Schrodingerpicture, if we use quasidistributions and the antinormal ordering (the Husimi func-tions) for the expression of the input–output dependence. An attenuator ought to bedescribed in the Schrodinger picture, if we employ the quasidistributions and thenormal ordering (the Glauber diagonal representation) to a similar goal. To showthis, we present a formal definition of an amplifier and that of an attenuator. Thesedefinitions operate with the integrated quantum-noise terms which are needed forthe input–output relations to preserve the commutators. The integrated quantum-noise terms can be formally decomposed into creation operators in amplifier caseand into annihilation operators in the attenuator case. We believe that we can providemore than such a formal expansion considering fields of modes or mode densitiescoupled to the counterpropagating modes. These fields are well-known quantumreservoirs, e.g., (Louisell 1973), whose frequency dependence has been replaced bythe position dependence.

In a more complicated case when in the mode a (let us say) an attenuation pro-ceeds and in the mode b an amplification occurs (Severini et al. 2004), a morecomplicated behaviour in the Schrodinger picture can be expected, there is not aguide to choose the ordering. To describe the amplification and attenuation, oneneeds a full Heisenberg–Langevin approach.

Distributed feedback laser has attracted attention as a device, which in the frame-work of coupled-wave theory, deserves quantization (Toren and Ben Aryeh 1994).The quantization produced by the authors seems to be very complicated. Unfortu-nately, Toren and Ben Aryeh (1994) have not developed an overarching quantizationof the analysis of amplification and that of the contradirectional coupling.

Page 359: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 351

6.2.1 Lossless Propagation in a Waveguide Structure

Perina Jr., et al. (2004) study an optical parametric process, namely a second-orderprocess. They had studied photonic band-gap structures and continue the work(Tricca et al. 2004) for the second-harmonic generation in a planar nonlinear corru-gated waveguide. They consider both the influence of the corrugation of the waveg-uide on the longitudinal confinement of the signal and idler modes (a modificationof (Tricca 2004)) and this influence on the phase matching of the process.

They present the decomposition of the electric-field amplitude related to photonsof the respective modes

E(x, y, z, t) = i∑

m

√�ωm

2ε0εrLem

×{Am(z) fm(x, y) exp[i(kmzz − ωmt)]− c. c.} , (6.203)

where Am is the amplitude of the mth mode, fm means the transverse eigenfunctionof the mth mode, em stands for the polarization vector, ωm denotes the frequency,and km is the wave vector of the mth mode. The mean permittivity of the waveguideis denoted as εr, ε0 stands for the vacuum permittivity, � is the reduced Planckconstant, and L is the length of the structure. The abbreviation c. c. stands forcomplex-conjugated terms.

The function fm(x, y) is a solution of the equation

∇2T fm(x, y)− k2

mz fm(x, y)+ με0εrω2m fm(x, y) = 0. (6.204)

At present, one has not yet realized a perfect (or an imperfect) quantization. Thefunction fm(x, y) is normalized, has the property

∫ ∫| fm(x, y)|2 dx dy = 1. (6.205)

The electric-field amplitude E ≡ E(r, t) satisfies the wave equation inside thewaveguide

∇2E− με0εr∂2E∂t2

= μ∂2Pnl

∂t2, (6.206)

where ε0εr stands for the permittivity of the waveguide, μ denotes the vacuum per-meability, and Pnl describes the nonlinear polarization of the medium. The relativepermittivity εr(r) can be written as follows

εr(x, y, z) = εr(x, y)+Δεr(x, y, z). (6.207)

Page 360: Quantum Aspects of Light Propagation

352 6 Periodic and Disordered Media

Here Δεr(x, y, z) are small variations of the permittivity related to the corrugation.These variations decompose into harmonic functions

Δεr(x, y, z) =∞∑

q=−∞εq (x, y) exp

(iq

Λlz

), ε0(x, y) = 0, (6.208)

εq (x, y) are coefficients of the decomposition and Λl is the spatial period of thegrating. The polarization Pnl of the medium is determined using the second-ordersusceptibility tensor χ ,

Pnl = ε0χ : EE, (6.209)

where : denotes a contraction, i.e. a double sum to be carried out after the tensorsare replaced by their components and products of the corresponding componentsare formed. On the assumption of three monochromatic components, a substitutionof the amplitude (6.203) into the wave equation (6.206) provides three coupledHelmholtz equations for these components. We consider two directions of propa-gation for each of the monochromatic components. We have six modes: the signalforward-propagating mode (with amplitude AsF ), the signal backward-propagatingmode (AsB ), the idler forward-propagating mode (AiF ), the idler backward-propagating mode (AiB ), the pump forward-propagating mode (ApF ) and, finally, thepump backward-propagating mode (ApB ). In the framework of the coupled-modetheory, we represent each of the three Helmholtz equations with two ordinary dif-ferential equations for the amplitudes

dAsF

dz= iKs exp(−iδsz)AsB + KF exp(iδFz)ApF A∗iF ,

dAiF

dz= iKi exp(−iδiz)AiB + KF exp(iδFz)ApF A∗sF

,

dAsB

dz= −iK ∗

s exp(iδsz)AsF − KB exp(−iδBz)ApB A∗iB ,

dAiB

dz= −iK ∗

i exp(iδiz)AiF − KB exp(−iδBz)ApB A∗sB,

dApF

dz= iKp exp(−iδpz)ApB − K ∗

F exp(−iδFz)AsF AiF ,

dApB

dz= −iK ∗

p exp(iδpz)ApF + K ∗B exp(iδBz)AsB AiB , (6.210)

where Kp = 0 and

δa = |kaFz| + |kaBz| − δl, a = s, i,

δl = 2π

Λl,

δb = |kpbz| − |ksbz| − |kibz|, b = F, B. (6.211)

Page 361: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 353

The linear coupling constants Ks and Ki are given as

Ka = με0ω2a

2|kaFz|∫ ∫

ε1(x, y) f ∗aF(x, y) faB (x, y) dx dy, a = s, i, (6.212)

where we have assumed that |kaFz| = |kaBz|. The expressions for the nonlinear cou-pling constants KF and KB are

Kb = με0ωs

|ksbz|

√�ωpωiωs

2ε0εrLχ

...epeies

∫ ∫fpb (x, y) f ∗ib (x, y) f ∗sb

(x, y) dx dy

= με0ωi

|kibz|

√�ωpωsωi

2ε0εrLχ

...epesei

∫ ∫fpb (x, y) f ∗sb

(x, y) f ∗ib (x, y) dx dy

= με0ωp

|kpbz|

√�ωsωiωp

2ε0εrLχ

...eseiep

∫ ∫f ∗sb

(x, y) f ∗ib (x, y) fpb (x, y) dx dy, (6.213)

where... denotes a contraction, i.e. a treble sum to be carried out after the tensors are

replaced by their components and products of the corresponding components areformed. We have assumed that ωs

|ksb z | ≈ωi|kib z | ≈

ωp

|kpb z | .The dependencies of the solutions to equations (6.210) on the boundary data

AaF (0), AaB (L), a = s, i, p (of the solutions to the boundary-value problem) can beconsidered as classical input–output relations. These are significant for investigationof the effect of stochastic boundary data.

Approximate results can be obtained by considering variations of the classicalsolutions δAab , a = s, i, p, b = F, B. The variations verify linear equations

dδAsF

dz= iKs exp(−iδsz)δAsB + KF exp(iδFz)δ(ApF A∗iF ),

dδAiF

dz= iKi exp(−iδiz)δAiB + KF exp(iδFz)δ(ApF A∗sF

),

dδAsB

dz= −iK ∗

s exp(iδsz)δAsF − KB exp(−iδBz)δ(ApB A∗iB ),

dδAiB

dz= −iK ∗

i exp(iδiz)δAiF − KB exp(−iδBz)δ(ApB A∗sB),

dδApF

dz= iKp exp(−iδpz)δApB − K ∗

F exp(−iδFz)δ(AsF AiF ),

dδApB

dz= −iK ∗

p exp(iδpz)δApF + K ∗B exp(iδBz)δ(AsB AiB ), (6.214)

where the variations δ(XY ) are to be further transformed using the Leibniz formulaδ(XY ) = Y δX + XδY . These or the resulting equations do not depend on whetheror not solutions to the boundary-value problem for (6.210) or those to the problemwith the initial data AaF (0), AaB (0), which seems to be easy, are the case.

Page 362: Quantum Aspects of Light Propagation

354 6 Periodic and Disordered Media

The consideration of variations leads to an approximate quantization. The input–output relations for variations are linear and, on certain conditions, they may beinterpreted even as input–output relations for quantum corrections δ Aab , a = s, i, p,b = F, B.

Let us suppose that we first express a solution of the initial-value problem for(6.214). On introducing the notation

δAb(z) =

⎜⎜⎜⎜⎜⎜⎝

δAsb (z)δA∗sb

(z)δAib (z)δA∗ib (z)δApb (z)δA∗pb

(z)

⎟⎟⎟⎟⎟⎟⎠, b = F, B, (6.215)

this solution is(

δAF(L)δAB(L)

)=(VFF(L) VFB(L)VBF(L) VBB(L)

)(δAF(0)δAB(0)

). (6.216)

Then only we determine a solution of the boundary-value problem as

(δAF(L)δAB(0)

)= U

(δAF(0)δAB(L)

), (6.217)

where

U =(VFF(L)− VFB(L)V−1

BB(L)VBF(L) VFB(L)V−1BB(L)

−V−1BB(L)VBF(L) V−1

BB(L)

). (6.218)

The input–output relations for quantum corrections are

(δAF(L)δAB(0)

)= U

(δAF(0)δAB(L)

), (6.219)

where δAb(z), b = F, B, are given by relation (6.215), but with operators insteadof the classical variables. The complex conjugates are replaced with the Hermitianones. Nonclassical properties of the device are assessed by the operators

Aab (z) = Aab (z)1+ δ Aab (z), z = 0, L . (6.220)

Let the input operators, the components of the vectors δAF(0) and δAB(L), fulfilthe usual boson commutation relations. In other words, the first, the third, and thefifth component are annihilation operators, the second, the fourth, and the sixth oneare creation operators. Then also the output operators, the components of the vectorsδAF(L) and δAB(0), obey the boson commutation relations.

Page 363: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 355

A sufficient condition for it to be possible to carry out this approximate quanti-zation is the existence of a suitable momentum function G int(z) and the expressionof the equations (6.210) in the form

dX

dz= i

�[X, G int]. (6.221)

Here X (as well as Y in what follows) is any function of the variables Aab , A∗ab, and

G int(z) is a real function of these variables and of the coordinate z. With respect toa complicated expression of the bracket [X, Y ], operators are written about and thebracket has the usual meaning of a commutator in the paper (Perina, Jr., et al. 2004).We will prefer the classical variables and

[X, Y ] =∑

a=s,i,p

(∂ X

∂ AaF

∂Y

∂ A∗aF

− ∂Y

∂ AaF

∂ X

∂ A∗aF

)

−∑

a=s,i,p

(∂ X

∂ AaB

∂Y

∂ A∗aB

− ∂Y

∂ AaB

∂ X

∂ A∗aB

). (6.222)

The appropriate momentum function G int(z) is

G int(z) = �[Ks exp(iδsz)A∗sF

AsB + Ki exp(iδiz)A∗iF AiB

+ Kp exp(iδpz)A∗pFApB + c. c.

]

− �[iKF exp(iδFz)ApF A∗sF

A∗iF + iKB exp(−iδBz)ApB A∗sBA∗iB + c. c.

].

(6.223)

In such a case, a search for conservation laws is facilitated. Equations (6.210) havethe properties

d

dz

(|AsF |2 + |ApF |2 − |AsB |2 − |ApB |2) = 0,

d

dz

(|AiF |2 + |ApF |2 − |AiB |2 − |ApB |2) = 0. (6.224)

Perina, Jr., et al. (2005) pay attention also to the production of the longitudinalconfinement of the pump mode(s) through the corrugation. The equations (6.210)are utilized in full generality; the restriction Kp = 0 has been lifted. The explana-tions (6.211) are completed with another one,

δp = |kpFz| + |kpBz| − 2δl. (6.225)

The linear coupling constant Kp is expressed according to (6.212), which is com-pleted with a = p, and we have assumed that |kpFz| = |kpBz|.

Perina, Jr., et al. (2007) study degenerate optical parametric processes, namelysecond-harmonic and second-subharmonic generation. They have considered aniso-tropy of the waveguide.

Page 364: Quantum Aspects of Light Propagation

356 6 Periodic and Disordered Media

We present the decomposition of the electric-field amplitude related to photonsof the respective modes

E(x, y, z, t) = i∑

m

√�ωm

2ε0L

×{

Am(z)ε−12 · fm(x, y) exp[i(kmzz − ωmt)]− c. c.

}, (6.226)

where fm is the vector-valued transverse eigenfunction of the mth mode. The meanpermittivity tensor of the waveguide is denoted as ε.

The vector-valued function fm(x, y) is a solution of the equation

(I∇2T − ∇T∇T) · fm(x, y)− iβm∇Tez · fm(x, y)− iβmez∇T · fm(x, y)

−β2m(I− ezez) · fm(x, y)+ με0εω

2m · fm(x, y) = 0. (6.227)

The function fm(x, y) is normalized; it has the property

∫ ∫f∗m(x, y) · fm(x, y) dx dy = 1. (6.228)

The electric-field amplitude E(r, t) satisfies the wave equation inside the wave-guide

− ∇ × (∇ × E)− με0ε · ∂2E∂t2

= μ∂2Pnl

∂t2, (6.229)

where ε stands for the permittivity tensor of the waveguide. Every spectral compo-nent of the permittivity tensor ε(r, ω) can be written as follows

ε(x, y, z, ω) = ε(x, y, ω)+Δε(x, y, z, ω). (6.230)

Here Δε(x, y, z, ω) are small variations of the permittivity tensor related to thecorrugation. These variations decompose into tensor-valued harmonic functions

Δε(x, y, z, ω) =∞∑

q=−∞εq (x, y, ω) exp

(iq

Λlz

), ε0(x, y, ω) = 0(1), (6.231)

where εq (x, y, ω) are coefficients of the decomposition, and 0(1) is the second-rankzero tensor. The polarization Pnl(r, t) of the medium is determined using the second-order susceptibility tensor χ (r),

Pnl(r, t) = ε0χ (r) : E(r, t)E(r, t). (6.232)

Page 365: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 357

A spectral component χ (r,−2ω; ω,ω) of the second-order susceptibility can beexpressed as

χ (x, y, z,−2ω; ω,ω) =∞∑

q=−∞χq (x, y,−2ω; ω,ω) exp

(iq

Λnlz

), (6.233)

where Λnl describes the period of a possible periodical poling of the nonlinear mate-rial. On the assumption of two monochromatic components, a substitution of theamplitude (6.226) into the wave equation (6.229) provides two coupled Helmholtzequations for these components. We consider two directions of propagation foreach of the monochromatic components. We have four modes: the signal forward-propagating mode (with amplitude AsF ), the signal backward-propagating mode(AsB ), the pump forward-propagating mode (ApF ), and, finally, the pump backward-propagating mode (ApB ). In the framework of the coupled-mode theory, we representeach of the two Helmholtz equations with two ordinary differential equations for theamplitudes

dAsF

dz= iKs exp(−iδsz)AsB + 2KF,q exp(iδFz)ApF A∗sF

,

dAsB

dz= −iK ∗

s exp(iδsz)AsF − 2KB,q exp(−iδBz)ApB A∗sB,

dApF

dz= iKp exp(−iδpz)ApB − K ∗

F,q exp(−iδFz)A2sF,

dApB

dz= −iK ∗

p exp(iδpz)ApF + K ∗B,q exp(iδBz)A2

sB, (6.234)

where

δa = |βaF | + |βaB | − δl, a = p, s,

δb,q = |βpb | − 2|βsb | + q2π

Λnl, b = F, B. (6.235)

The linear coupling constants Kp and Ks are given as

Ka = με0ω2a

2|β ′aF|∫ ∫

f∗aF(x, y) · ε1(x, y, ωa) · faB (x, y) dx dy, a = p, s, (6.236)

where

β ′aF= βaF − βaF

∫ ∫|ez · faF (x, y)|2 dx dy

+ Im∫ ∫

[∇T · f∗aF(x, y)][ez · faF (x, y)] dx dy, (6.237)

Page 366: Quantum Aspects of Light Propagation

358 6 Periodic and Disordered Media

and we have assumed that |β ′aFz| = |β ′aBz|. The expressions for the nonlinear cou-pling constants KF and KB are

Kb,q = με0ωs

2|β ′sb|

√�ωpωsωs

2ε0L

∫ ∫[ε−

12 (ωp) · fpb (x, y)] · χq (x, y,−ωp; ωs, ωs)

:[ε12 (ωs) · f∗sb

(x, y)][ε−12 (ωs) · f∗sb

(x, y)] dx dy

= με0ωp

2|β ′pb|

√�ωsωsωp

2ε0L

∫ ∫[ε

12 (ωp) · fpb (x, y)] · χq (x, y,−ωp; ωs, ωs)

:[ε−12 (ωs) · f∗sb

(x, y)][ε−12 (ωs) · f∗sb

(x, y)] dx dy, (6.238)

where we have assumed that ωs|β ′sb

| ≈ωp

|β ′pb| .

The dependencies of the solutions to the equations (6.234) on the boundary dataAaF (0), AaB (L), a = s, p (of the solutions to the boundary-value problem) can beconsidered as classical input–output relations. These are significant for the study ofstochastic boundary data.

Approximate results can be obtained by considering variations of the classicalsolutions δAab , a = s, p, b = F, B. The variations satisfy linear equations

dδAsF

dz= iKs exp(−iδsz)δAsB + 2KF,q exp(iδFz)δ(ApF A∗sF

),

dδAsB

dz= −iK ∗

s exp(iδsz)δAsF − 2KB,q exp(−iδBz)δ(ApB A∗sB),

dδApF

dz= iKp exp(−iδpz)δApB − K ∗

F,q exp(−iδFz)δ(A2sF

),

dδApB

dz= −iK ∗

p exp(iδpz)δApF + K ∗B,q exp(iδBz)δ(A2

sB). (6.239)

The consideration of variations leads to an approximate quantization. The input–output relations for variations are linear and, as in the previous case, they may leadto input–output relations for quantum corrections δ Aab , a = s, p, b = F, B.

Let us suppose that we first express a solution of the initial-value problem for(6.239). On introducing notation

δAb(z) =

⎜⎜⎝

δAsb (z)δA∗sb

(z)δApb (z)δA∗pb

(z)

⎟⎟⎠ , b = F, B, (6.240)

this solution becomes (6.216). Thereafter we determine a solution of the boundary-value problem as (6.217), with (6.218). The input–output relations for quantumcorrections are (6.219), where δAb(z), b = F, B, are given by relation (6.240),but with operators instead of the classical variables. The complex conjugates arereplaced with the Hermitian ones. Nonclassical behaviour of a process is assessedby the operators (6.220).

Page 367: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 359

In this case, the equations (6.234) have the form (6.221), with

[X, Y ] =∑

a=s,p

(∂ X

∂ AaF

∂Y

∂ A∗aF

− ∂Y

∂ AaF

∂ X

∂ A∗aF

)

−∑

a=s,p

(∂ X

∂ AaB

∂Y

∂ A∗aB

− ∂Y

∂ AaB

∂ X

∂ A∗aB

). (6.241)

The appropriate momentum function G int(z) is

G int(z) = �

[Ks exp(iδsz)A∗sF

AsB + Kp exp(iδpz)A∗pFApB + c. c.

]

− �[iKF exp(iδFz)ApF A∗2

sF+ iKB exp(−iδBz)ApB A∗2

sB+ c. c.

]. (6.242)

Equations (6.234) have the property

d

dz

(|AsF |2 + 2|ApF |2 − |AsB |2 − 2|ApB |2) = 0. (6.243)

6.2.2 Coupled-Mode Theory Including Gain or Losses

We assume a monochromatic wave propagating in a waveguide in the form

E(x, y, z, t) =∑

m

Am(z)Em(x, y) exp[i(ωt − kmzz)], (6.244)

where Am(z), ddz Am(z) = 0, is the amplitude of the mth mode, ω is a frequency,

and kmz are propagation constants (components along the direction of propagationof the wave vector of each mode). Let us note the difference from relation (6.203),where the wave is real. These eigenmodes have the electric vectors and the magneticvectors of the form

Em(x, y, z, t) = Em(x, y) exp[i(ωt − kmzz)],

Hm(x, y, z, t) = Hm(x, y) exp[i(ωt − kmzz)], (6.245)

respectively. In fact, the fields are real, and they must be recovered as 12 [Em(x, y, z, t)

+E∗m(x, y, z, t)], etc.It is understood that counterpropagating modes are orthogonal. The normaliza-

tion and the orthogonality property of copropagating modes are expressed by therelation

1

2

∫ ∫(Ek ×H∗

m) · ez dx dy = vgm

L �ωδkm, (6.246)

where ez is the unit vector in the direction of the z-axis, vgm is the group velocity,vgm = ∂ω

∂kz|kz=kmz , L is a quantization length, and � is the reduced Planck constant.

The arguments of Ek and H∗m have been omitted for convenience. The treatment

Page 368: Quantum Aspects of Light Propagation

360 6 Periodic and Disordered Media

will be restricted to dielectric structures, which consist of pieces of homogeneousand isotropic materials, or those with a small gradient of the refractive index. Thenin (6.245), Em(x, y) and Hm(x, y) obey the vectorial wave equation

[∂2

∂x2+ ∂2

∂y2+ ω2με(x, y)− k2

mz

]Um(x, y) = 0, (6.247)

where μ is the magnetic permeability and ε(x, y) = ε0εr(x, y). In the case of apiecewise homogeneous dielectric structure, equation (6.247) is valid separately ineach of homogeneous domains. So the field must be determined separately in everydomain and, then, the tangential components of the fields must be joined at eachof the interfaces. Another important boundary condition for waveguide modes isthe vanishing of the field amplitudes at infinity. For the boundary conditions to besatisfied at all the points of the interfaces between homogeneous media, the parax-ial propagation constant kmz must be the same in the whole waveguide structure(Yariv and Yeh 1984). For definiteness, we will describe a planar waveguide and putUm = Em .

The solutions should be determined in the core (guiding region), which we desig-nate as D. We let C denote the boundary of D, which are two straight lines parallelto the y-axis. The quantity εr(x, y) is independent of y, and the solutions that areindependent of y are looked for. The transverse electric (TE) modes and the trans-verse magnetic (TM) modes are distinguished. The TE modes have

Emx (x, y) = Emz(x, y) = Hmy(x, y) = 0, (6.248)

where the first component is included by the definition of these modes, the secondone is present here due to the condition ∇ · E = 0, and the third one follows from aMaxwell equation. The TM modes have

Hmx (x, y) = Hmz(x, y) = Emy(x, y) = 0, (6.249)

where the first component is included by the definition of these modes, the secondone is present here due to the condition ∇ · H = 0, and the third one follows froma Maxwell equation. For suitable kmz , the TE mode boundary conditions are thecontinuity requirements on Emy(x, y)|C and Hmz(x, y)|C . The TM mode boundaryconditions are related to Hmy(x, y)|C and Emz(x, y)|C . Equation (6.247) is solved inthe whole x–y plane excepting the point set C perhaps. We have neglected a term∇ (∇ ·E), which is justified if the change of the quantity ε0(x, y) over a wavelengthis small. In the case of TE modes of planar dielectric waveguides ∇ · E = 0, since∇ · (εE) = ε∇ · E + E · ∇ε = 0, where E · ∇ε = 0. In fact, ε has a jump in thex-direction, whereas Ex vanishes. This implies that equation (6.247) is exact at C .In the case of TM modes of a planar waveguide, equation (6.247) does not hold atthe interface C .

In some cases, it is useful to consider complex dielectric permittivity. In this gen-eralization, the definition of unperturbed modes is based on Re{εr(x, y)}. Respecting

Page 369: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 361

it, the replacement of ε(x, y) by Re{εr(x, y)} should be made, where it is appropri-ate. Particularly, relation (6.207) becomes

εr(x, y, z) = Re{εr(x, y)} +Δεr(x, y, z). (6.250)

Now

Re{Δεr(x, y, z)} =∞∑

q=−∞ε(Re)

q (x, y) exp

(iq

Λlz

), ε

(Re)0 (x, y) = 0, (6.251)

Im{Δεr(x, y, z)} =∞∑

q=−∞ε(Im)

q (x, y) exp

(iq

Λlz

),

ε(Im)0 (x, y) = Im{εr(x, y)}. (6.252)

In the following, we restrict ourselves to the TE modes. Here E(x, y, z, t)(Em(x, y), Hm(x, y)) will be a shorthand for the component Ey(x, y, z, t) (Emy(x, y),Hmy(x, y)) along the direction of y. The orthogonality condition of the modes reads

⟨E∗m |E∗k

⟩ =∫ ∫

E∗m(x, y) · Ek(x, y) dx dy = |vgm |L

2�ω2μ

|kmz| δkm . (6.253)

With respect to the quantum treatment, we introduce the negative- and positive-frequency parts

E (−)(x, y, z, t) = 1

2E(x, y, z, t),

E (+)(x, y, z, t) = [E (−)(x, y, z, t)]∗, (6.254)

the corresponding envelope

Am(z) = 1

2A∗m(z), (6.255)

and we note that relation (6.244) can be rewritten as

E (+)(x, y, z, t) =∑

m

Am(z)E∗m(x, y) exp[−i(ωt − kmzz)]. (6.256)

This classical field is an eigenfunction (rather, eigenvalue that depends onparameters),

E (+)(x, y, z, t)

∣∣∣∣

{Am

(0

L

)}⟩= E (+)(x, y, z, t)

∣∣∣∣

{Am

(0

L

)}⟩, (6.257)

where E (+)(x, y, z, t) is the positive-frequency part of the electric strength operator,∣∣∣{

Am(0

L

)}⟩is a coherent state, 0 (L) corresponds to kmz > 0 (kmz < 0), and L is

the length of the optical device. We consider expansion (6.256) with E (+)(x, y, z, t)replaced by E (+)(x, y, z, t) and Am(z) replaced by Am(z).

Page 370: Quantum Aspects of Light Propagation

362 6 Periodic and Disordered Media

From the viewpoint of the integrated optics, the interaction of modes or couplingof modes has been interesting and also the possibility of quantization is attractive.Here we have touched even the impossibility of quantization. Whenever a linearrelation between pairs of complex amplitudes is appropriate (a linear canonicaltransformation), quantization is possible and the complex amplitudes can be inter-preted as operators. In the literature (Milburn et al. 1984) concerning the nonlinearprocesses, it has been shown that such operators do not obey the usual commutationrelations. In this case, we assert that the quantization has not been accomplished.Nevertheless, in this study, we interpret the input–output relations as quantal.

A special use of quantum-mechanical descriptions has been mentioned inSection 3.1.4 already. The full Heisenberg–Langevin approach has been utilized,which is based on the concept of a quantum reservoir.

We classify reservoirs as forward- and backward-propagating ones, even thoughit is not quite usual to combine these terms. But we do not utilize the concept of aquantum reservoir. We rely on the concept of quasi-continuous measurements (Ban1994). It is sufficient to know that the quasi-continuous measurement is realizedusing a system of lossless beam splitters or using a system of parametric amplifiers.A transition to a continuous limit is possible, and the differential equation of itsdescription coincides with the master equation for the description of a single modeobtained on an elimination of the quantum reservoir (Perinova and Luks 2000).Independent of this, a return to the classical description is possible provided that thequantum measurements are not studied.

The system of lossless beam splitters represents an “attenuator”, since the energyof the light mode is by parts reflected by the beam splitters to detectors. The systemof parametric amplifiers represents an “amplifier”, since photons of pump beam areconverted to photon-twin pairs. One of each pair is directed to the detector, and theother supplies energy to the light mode going through aligned axes of nonlinearcrystals. Similarly, repeated nondemolition measurements can be implemented.

Another idea, which however requires the replacement of the time variable bythe space variable, is the measurement of an observable of a cavity mode usingRydberg atoms. Literature is devoted to the useful case of repeated nondemolitionmeasurements, but we image easily also an attenuator similar to the system of loss-less beam splitters and an amplifier similar to the system of parametric amplifiers.This idea allows the transition to a continuous limit as well. Different frequenciesof a quantum reservoir are replaced with different times when the atoms interactwith the cavity mode. On the change of the time variable by the space variable, wedescribe the physical fact that the guided mode is in short but densely distributedsegments of a waveguide coupled to the sources or sinks of the energy.

We will approach the distributed feedback laser (Yariv and Yeh 1984, Torenand Ben Aryeh 1994) as a quantum amplifier. In the usual coupled-mode theory,it is assumed that the perturbation Δεr(x, y, z) of the dielectric permittivity is real,but the presence of a small gain can be also considered a perturbation and thenΔεr(x, y, z) is to be held for a complex quantity. Describing a lossy medium, oneassumes that the imaginary part of Δεr(x, y, z) is negative, but the gain mediumexhibits a positive imaginary part of Δεr(x, y, z). We assume that modes 1 and 2 are

Page 371: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 363

coupled, and we let k1z , k2z denote the z-components of the respective wave vectors.We will treat the particular case when k1z ≈ lπ

λ, i.e., mode 1 is strongly coupled with

the backward-propagating mode 2, k2z = −k1z . The classical description is basedon the differential equations. We complex conjugate the differential equations ofclassical description (Yariv and Yeh 1984) and replace A∗j (z) by A j (z). This resultsin the differential equations

d

dzA1(z) = iκ∗A2(z) exp(2iδz)+ γ

2A1(z),

d

dzA2(z) = −iκA1(z) exp(−2iδz)− γ

2A2(z), (6.258)

where

κ = ε0L4�|vg|

∫ ∫E∗1(x, y) · ε(Re)

−l (x, y)E1(x, y) dx dy,

δ = 1

2(k1z − k2z)− lπ

λ,

γ = ε0L2�|vg|

∫ ∫E∗1(x, y) · Im{εr(x, y)}E1(x, y) dx dy, (6.259)

with vg the group velocity of light. The input–output relations are given as

A1(L) = u11(L)A1(0)+ u12(L)A2(L),

A2(0) = u21(L)A1(0)+ u22(L)A2(L). (6.260)

Here u jk(L), j, k = 1, 2, are given by generalized relations from (Perinova et al.1991),

u11(L) = exp (iδL)

[cosh(DL)+ i

Δ

Dsinh(DL)

]−1

,

u12(L) = iκ∗

Dexp (i2δL) sinh(DL)

×[

cosh(DL)+ iΔ

Dsinh(DL)

]−1

,

u21(L) = iκ

Dsinh(DL)

[cosh(DL)+ i

Δ

Dsinh(DL)

]−1

,

u22(L) = u11(L), (6.261)

with

D =√|κ|2 −Δ2, Δ = δ + i

γ

2. (6.262)

The quantization of the classical description in the case of no gain can be accom-plished in the spatial Heisenberg picture using equations (6.260). The complex

Page 372: Quantum Aspects of Light Propagation

364 6 Periodic and Disordered Media

amplitudes A j (z) are replaced by the annihilation operators A j (z), j = 1, 2,z = 0, L . In general, the spatial Schrodinger picture is appropriate. Any inputstatistical operator ρ(0) evolves to the output operator ρ(L). In contrast, the oper-ators A1(0) and A2(L) do not change and that is why we abbreviate A1 ≡ A1(0),A2 ≡ A2(L) here.

The statistical properties of the input and output fields are described equivalentlyby the characteristic functions. The definition of these functions depends on theordering of field operators. We choose the antinormal ordering, which is suitable forthe amplifier. The antinormal characteristic functions of the input and output state,respectively, are

CA(β1, β2, 0)

= Tr{ρ(0) exp(−β∗1 A1 − β∗2 A2) exp(β1 A†

1 + β2 A†2)},

CA(β1, β2, L)

= Tr{ρ(L) exp(−β∗1 A1 − β∗2 A2) exp(β1 A†

1 + β2 A†2)}. (6.263)

The antinormal characteristic function for the output can be defined taking intoaccount the coefficients u jk(L), j, k = 1, 2, in (6.261),

CA(β1, β2, L) = Tr{ρ(0)

× exp[−(β∗1 u11(L)+ β∗2 u21(L)

)A1 −

(β∗1 u12(L)+ β∗2 u22(L)

)A2

]

× exp[(

β1u∗11(L)+ β2u∗21(L))

A†1 +

(β1u∗12(L)+ β2u∗22(L)

)A†

2

]}

= CA(β1u∗11(L)+ β2u∗21(L), β1u∗12(L)+ β2u∗22(L), 0

). (6.264)

Here we have reduced an alternative definition to a substitution into the charac-teristic function for the input. We may conclude with the inversion of the secondequation in (6.263),

ρ(L) = 1

π2

∫ ∫CA(β1, β2, L) exp(β1 A†

1 + β2 A†2)

× exp(−β∗1 A1 − β∗2 A2) d2β1 d2β2, (6.265)

where d2β j=d(Re {β j })d(Im {β j }), j = 1, 2. The above procedure is a completelypositive map. We will not present a proof, which can be established similarly asbelow in the case of attenuator. With the characteristic functions, the quasidistribu-tions are associated related to the same ordering

ΦA(A1, A2, 0) = 1

π4

∫ ∫CA(β1, β2, 0) exp(A1β

∗1 + A2β

∗2 − c. c.) d2β1 d2β2,

ΦA(A1, A2, L) = 1

π4

∫ ∫CA(β1, β2, L) exp(A1β

∗1 + A2β

∗2 − c. c.) d2β1 d2β2.

(6.266)

Page 373: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 365

The relationships between the characteristic functions become the relationshipsbetween the quasidistributions. The latter are more complicated (one needs elementsof the inverse to the matrix U (L) consisting of the elements u jk(L), j, k = 1, 2, andits determinant |U (L)|).

According to our statement below the relation (6.262), when γ = 0, the quan-tization can be accomplished in the Heisenberg picture. Relationships between sta-tistical operators may be intricate sometimes. Then we can adopt the Heisenberg–Langevin approach. We can formally define an amplifier to be a device which canbe described with input–output relations

A1(L) = u11(L) A1(0)+ u12(L) A2(L)+ M1(L),

A2(0) = u21(L) A1(0)+ u22(L) A2(L)+ M2(L), (6.267)

where M j (L), j = 1, 2, are integrated quantum-noise terms. A characteristic prop-erty of the amplifier by the definition is that the commutator matrix

([M1(L), M†

1(L)] [M1(L), M†2(L)]

[M2(L), M†1(L)] [M2(L), M†

2(L)]

)≤(

0 00 0

)1, (6.268)

i.e., both its eigenvalues, when 1 is factored out, are nonpositive. It means that wecan find numbers u jk(L) and creation operators A†

k(L), j = 1, 2, k = 3, 4, such that

M1(L) = u13(L) A†3(0)+ u14(L) A†

4(L),

M2(L) = u23(L) A†3(0)+ u24(L) A†

4(L). (6.269)

It is required that the commutation relations between input annihilation and creationoperators (6.305) below and those between such operators related to the expression(6.269),

[ A3(0), A†3(0)] = [ A4(L), A†

4(L)] = 1, (6.270)

could be rewritten as those between annihilation and creation output operators in(6.305). It is assumed that the operators of distinct modes mutually commute.

Instead of proving that the procedure for the derivation of the output statisticaloperator from the input one is a completely positive map, we can find expressions forintegrated quantum-noise terms and prove the characteristic property of the ampli-fier (6.268). In Yariv and Yeh (1984), much attention is paid to the regime of lightgeneration, defined by the condition

cosh(DL)+ iΔ

Dsinh(DL) = 0. (6.271)

In the explicit expression of the coefficients u jk(L), j = 1, 2, in (6.261), the divi-sion by zero occurs. It is obvious that the model is only tentative; for instance, theeffect of saturation can be described by a nonlinear model, while the model underdiscussion is linear.

Page 374: Quantum Aspects of Light Propagation

366 6 Periodic and Disordered Media

As the reciprocity theorem is derived on the assumption of the lossless medium,the attenuator has not been elaborated in the coupled-mode theory, contrary to theamplifier. But it is rather reasonable to assume that a perturbation Δεr(x, y, z) is acomplex quantity with a negative imaginary part. Using the same procedure as inthe case of amplifier, we arrive at the differential equations

d

dzA1(z) = iκ∗A2(z) exp(2iδz)− γ

2A1(z),

d

dzA2(z) = −iκA1(z) exp(−2iδz)+ γ

2A2(z), (6.272)

where κ is defined in (6.259) and

γ = − ε0L2�|vg|

∫ ∫E∗1(x, y) · Im{εr(x, y)}E1(x, y) dx dy. (6.273)

Here we expound the usual approach (the coherent-state technique), which has beenimplicitly used also in the previous exposition. The solution of an initial problemfor (6.272) can be obtained in the form

A1(z) = v11(z)A1(0)+ v12(z)A2(0),

A2(z) = v21(z)A1(0)+ v22(z)A2(0), (6.274)

where v jk(L), j, k = 1, 2, are given by generalized relations from Perinova et al.(1991),

v11(z) = exp (iδz)

[cosh(Dz)− i

Δ

Dsinh(Dz)

],

v12(z) = iκ∗

Dexp (iδz) sinh(Dz),

v21(z) = −iκ

Dexp (−iδz) sinh(Dz),

v22(z) = exp (−iδz)

[cosh(Dz)+ i

Δ

Dsinh(Dz)

], (6.275)

with

D =√|κ|2 −Δ2, Δ = δ − i

γ

2. (6.276)

The input–output relations have the form (6.260), where u jk(L), j, k = 1, 2, aregiven in (6.261) except Δ, Δ = δ − i γ

2 .The case of no losses coincides with the case of no gain, when the quantization

is possible as mentioned in the previous exposition. In general, the Schrodingerpicture can be recommended. With respect to (6.263), the input and output operatorsare denoted as ρ(0) and ρ(L), respectively. In the description, the normal orderingis used to simplify the form. The normal characteristic functions of the input andoutput states, respectively, are

Page 375: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 367

CN (β1, β2, 0) = Tr{ρ(0) exp(β1 A†

1 + β2 A†2) exp(−β∗1 A1 − β∗2 A2)

},

CN (β1, β2, L) = Tr{ρ(L) exp(β1 A†

1 + β2 A†2) exp(−β∗1 A1 − β∗2 A2)

}.(6.277)

Intuitively, the normal characteristic function is

CN (β1, β2, L) = Tr{ρ(0)

× exp[(

β1u∗11(L)+ β2u∗21(L))

A†1 +

(β1u∗12(L)+ β2u∗22(L)

)A†

2

]

× exp[−(β∗1 u11(L)+ β∗2 u21(L)

)A1 −

(β∗1 u12(L)+ β∗2 u22(L)

)A2

]}

= CN(β1u∗11(L)+ β2u∗21(L), β1u∗12(L)+ β2u∗22(L), 0

). (6.278)

Here we have expressed the statistical properties of the output through the normalcharacteristic function for the inputs. The procedure concludes with the inversion ofthe second equation in (6.277),

ρ(L) = 1

π2

∫ ∫CN (β1, β2, L) exp(−β∗1 A1 − β∗2 A2)

× exp(β1 A†1 + β2 A†

2) d2β1 d2β2. (6.279)

We assert that we have defined a completely positive map. We will provide a proof ofthis proposition in what follows. Meanwhile, we remark that for some states, therealso exist quasidistributions related to the normal ordering as ordinary functions

ΦN (A1, A2, 0) = 1

π4

×∫ ∫

CN (β1, β2, 0) exp(A1β∗1 + A2β

∗2 − c. c.) d2β1 d2β2,

ΦN (A1, A2, L) = 1

π4

×∫ ∫

CN (β1, β2, L) exp(A1β∗1 + A2β

∗2 − c. c.) d2β1 d2β2.

(6.280)

The relationships between the statistical operators may be involved sometimes.In response, we can adopt the Heisenberg–Langevin approach. We can formallydefine an attenuator to be a device which can be described with the input–outputrelations (6.267). A characteristic property of the attenuator by the definition is thatthe commutator matrix

([M1(L), M†

1(L)] [M1(L), M†2(L)]

[M2(L), M†1(L)] [M2(L), M†

2(L)]

)≥(

0 00 0

)1, (6.281)

Page 376: Quantum Aspects of Light Propagation

368 6 Periodic and Disordered Media

i.e., both its eigenvalues (let 1 be factored out) are nonnegative. It means that we canfind numbers u jk(L) and annihilation operators Ak(L), j = 1, 2, k = 3, 4, such that

M1(L) = u13(L) A3(0)+ u14(L) A4(L),

M2(L) = u23(L) A3(0)+ u24(L) A4(L). (6.282)

It is required that the commutation relations (6.305) between input annihilation andcreation operators and the relations (6.270) related to the expression (6.282) couldbe rewritten as those between annihilation and creation output operators in (6.311).The operators of different modes mutually commute, the input operators with theinput ones and the output operators with the output ones.

In place of proving that the procedure for derivation of the output statistical oper-ator from the input one is a completely positive map, we can find expressions forintegrated quantum noise terms and demonstrate the characteristic property of theattenuator (6.281).

(i) Expressions for the integrated quantum-noise terms

In order to take into account losses, we use the extended differential equations

d

dzA1(z) = iκ∗A2(z) exp(2iδz)− iG1(z, z),

d

dzA2(z) = −iκA1(z) exp(−2iδz)+ iG2(z, z), (6.283)

d

dzG1(ζ, z) = −iγ δ(ζ − z)A1(z),

d

dzG2(ζ, z) = iγ δ(ζ − z)A2(z), (6.284)

where G j (ζ, z), j = 1, 2, is a continuum of modes, which are right- and left-goingin dependence on j = 1 and j = 2, respectively. In terms of these modes, lossesare modelled. The detail that the losses of each mode A j (z), j = 1, 2, are modelledby coupling with modes propagating in the same direction is not essential, but it hasbeen chosen for simplicity. Quite a novel thing in this description is that the couplingof the mode G j (ζ, z) is concentrated into the position z = ζ . Solving (6.284) forcomplex amplitudes G j (ζ, z), j = 1, 2, we obtain that

G1(ζ, z) =⎧⎨

G1(ζ, 0) for z < ζ,

G1(ζ, 0)− i γ

2 A1(ζ ) for z = ζ,

G1(ζ, 0)− iγ A1(ζ ) for z > ζ,

G2(ζ, z) =⎧⎨

G2(ζ, 0)+ iγ A2(ζ ) for z > ζ,

G2(ζ, 0)+ i γ

2 A2(ζ ) for z = ζ,

G2(ζ, 0) for z < ζ.

(6.285)

Page 377: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 369

As the contradirectional coupling presents more difficulties than the codirectionalcoupling, we may use the solutions of (6.285) to create apparent paradoxes. Onsubstituting relation (6.285) into (6.283), we obtain differential equations

d

dzA1(z) = iκ∗A2(z) exp(2iδz)− γ

2A1(z)− iG1(z, 0),

d

dzA2(z) = −iκA1(z) exp(−2iδz)− γ

2A2(z)+ iG2(z, 0). (6.286)

These equations will contradict Equations (6.272) after one omits the Langevinterms as is usual in the treatment of the co-propagation, where such a manipula-tion is correct. The solutions of the initial-value problem for equations (6.286) and(6.284) are of the form

A1(z) = v11f(z)A1(0)+ v12f(z)A2(0)

− i∫ z

0v11f(z|ζ ′)G1(ζ ′, 0) dζ ′ + i

∫ z

0v12f(z|ζ ′)G2(ζ ′, 0) dζ ′, (6.287)

A2(z) = v21f(z)A1(0)+ v22f(z)A2(0)

− i∫ z

0v21f(z|ζ ′)G1(ζ ′, 0) dζ ′ + i

∫ z

0v22f(z|ζ ′)G2(ζ ′, 0) dζ ′, (6.288)

G1(ζ, z) = −γ θ(z − ζ )[iv11f(ζ )A1(0)+ iv12f(ζ )A2(0)

+∫ ζ

0v11f(ζ |ζ ′)G1(ζ ′, 0) dζ ′

−∫ ζ

0v12f(ζ |ζ ′)G2(ζ ′, 0) dζ ′

]+ G1(ζ, 0), (6.289)

G2(ζ, z) = γ θ(z − ζ )[iv21f(ζ )A1(0)+ iv22f(ζ )A2(0)

+∫ ζ

0v21f(ζ |ζ ′)G1(ζ ′, 0) dζ ′

−∫ ζ

0v22f(ζ |ζ ′)G2(ζ ′, 0) dζ ′

]+ G2(ζ, 0), (6.290)

where θ (z−ζ ) is the Heaviside (unit-step) function, v jkf(z) = v jkf(z|0), j, k = 1, 2,with

v jkf(z|z′) = exp[−γ

2(z − z′)

]v jk(z|z′)∣∣

γ=0 , (6.291)

where

v11(z|z′)∣∣γ=0 = exp[iδ(z − z′)]

×{

cosh[

D0(z − z′)]− i

δ

D0sinh

[D0(z − z′)

]},

Page 378: Quantum Aspects of Light Propagation

370 6 Periodic and Disordered Media

v12(z|z′)∣∣γ=0 = i

κ∗

D0exp[iδ(z + z′)] sinh

[D0(z − z′)

],

v21(z|z′)∣∣γ=0 = −i

κ

D0exp[−iδ(z + z′)] sinh

[D0(z − z′)

],

v22(z|z′)∣∣γ=0 = exp[−iδ(z − z′)]

×{

cosh[

D0(z − z′)]+ i

δ

D0sinh

[D0(z − z′)

]}, (6.292)

with

D0 =√|κ|2 − δ2 . (6.293)

Taking into account the relation

G2(ζ, z) =⎧⎨

G2(ζ, L) for z > ζ,

G2(ζ, L)− i γ

2 A2(ζ ) for z = ζ,

G2(ζ, L)− iγ A2(ζ ) for z < ζ,

(6.294)

we replace equations (6.286) by the following equations

d

dzA1(z) = iκ∗A2(z) exp(2iδz)− γ

2A1(z)− iG1(z, 0),

d

dzA2(z) = −iκA1(z) exp(−2iδz)+ γ

2A2(z)+ iG2(z, L). (6.295)

The solutions of the boundary-value problem

A j (z)∣∣z=0 = A j (0), j = 1, 2,

G1(ζ, z)|z=0 = G1(ζ, 0), G2(ζ, z)|z=L = G2(ζ, L), (6.296)

for equations (6.283) and (6.284) transformed to the form (6.295) and (6.296) arefor z = L as follows

A1(L) = v11g(L)A1(0)+ v12g(L)A2(0)− i∫ L

0v11g(L|ζ ′)G1(ζ ′, 0) dζ ′

+ i∫ L

0v12g(L|ζ ′)G2(ζ ′, L) dζ ′, (6.297)

A2(L) = v21g(L)A1(0)+ v22g(L)A2(0)

− i∫ L

0v21g(L|ζ ′)G1(ζ ′, 0) dζ ′ + i

∫ L

0v22g(L|ζ ′)G2(ζ ′, L) dζ ′,

(6.298)

G1(ζ, L) = −iγ v11g(ζ )A1(0)− iγ v12g(ζ )A2(0)

− γ

∫ ζ

0v11g(ζ |ζ ′)G1(ζ ′, 0) dζ ′ + G1(ζ, 0)

+ γ

∫ ζ

0v12g(ζ |ζ ′)G2(ζ ′, L) dζ ′ + G2(ζ, L), (6.299)

Page 379: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 371

G2(ζ, 0) = −iγ v21g(ζ )A1(0)− iγ v22g(ζ )A2(0)

− γ

∫ ζ

0v21g(ζ |ζ ′)G1(ζ ′, 0) dζ ′ + G1(ζ, 0)

+ γ

∫ ζ

0v22g(ζ |ζ ′)G2(ζ ′, L) dζ ′ + G2(ζ, L), (6.300)

where v jkg(z) = v jkg(z|0) = v jk(z), j, k = 1, 2, with

v11g(z|z′) = exp[iδ(z − z′)]{

cosh[

D(z − z′)]− i

Δ

Dsinh

[D(z − z′)

]},

v12g(z|z′) = iκ∗

Dexp[iδ(z + z′)] sinh

[D(z − z′)

],

v21g(z|z′) = −iκ

Dexp[−iδ(z + z′)] sinh

[D(z − z′)

],

v22g(z|z′) = exp[−iδ(z − z′)]

×{

cosh[

D(z − z′)]+ i

Δ

Dsinh

[D(z − z′)

]}. (6.301)

The solutions of differential equations (6.286) and (6.284) conserve pseudonorms(Perinova et al. 2006)

|A1(L)|2 − |A2(L)|2 + 1

γ

∫ L

0|G1(ζ, L)|2 dζ − 1

γ

∫ L

0|G2(ζ, L)|2 dζ

= |A1(0)|2 − |A2(0)|2 + 1

γ

∫ L

0|G1(ζ, 0)|2 dζ − 1

γ

∫ L

0|G2(ζ, 0)|2 dζ, (6.302)

|A1(L)|2 − |A2(L)|2 + 1

γ

∫ L

0|G1(ζ, L)|2 dζ + 1

γ

∫ L

0|G2(ζ, 0)|2 dζ

= |A1(0)|2 − |A2(0)|2 + 1

γ

∫ L

0|G1(ζ, 0)|2 dζ + 1

γ

∫ L

0|G2(ζ, L)|2 dζ. (6.303)

On the inversion of equations (6.297) through (6.300), we see that the input andoutput complex amplitudes conserve the norm

|A1(L)|2 + |A2(0)|2 + 1

γ

∫ L

0|G1(ζ, L)|2 dζ + 1

γ

∫ L

0|G2(ζ, 0)|2 dζ

= |A1(0)|2 + |A2(L)|2 + 1

γ

∫ L

0|G1(ζ, 0)|2 dζ + 1

γ

∫ L

0|G2(ζ, L)|2 dζ (6.304)

and the corresponding scalar product. We can establish the Heisenberg picture in thefollowing sense. We assume that the input annihilation and creation operators obeythe usual commutation relations

[ A1(0), A†1(0)] = [ A2(L), A†

2(L)] = 1, (6.305)

Page 380: Quantum Aspects of Light Propagation

372 6 Periodic and Disordered Media

where the operators in different modes commute. With respect to the noise opera-tors, we make similar assumptions

[G1(ζ, 0), G†1(ζ ′, 0)] = [G2(ζ, L), G†

2(ζ ′, L)] = γ δ(ζ − ζ ′)1. (6.306)

The output operators are

A1(L) = u11g(L) A1(0)+ u12g(L) A2(L)

− i∫ L

0w22g(L|ζ ′)G1(ζ ′, 0) dζ ′ − i

∫ L

0w12g(L|ζ ′)G2(ζ ′, L) dζ ′,

(6.307)

A2(0) = u21g(L) A1(0)+ u22g(L) A2(L)

− i∫ L

0w21g(L|ζ ′)G1(ζ ′, 0) dζ ′ − i

∫ L

0w11g(L|ζ ′)G2(ζ ′, L) dζ ′,

(6.308)

where

u jkg(L) = u jk(L), j, k = 1, 2, (6.309)

w22g(L|ζ ′) = exp[iδ(L − ζ ′)][

cosh(Dζ ′)+ iΔ

Dsinh(Dζ ′)

]

×[

cosh(DL)+ iΔ

Dsinh(DL)

]−1

,

w21g(L|ζ ′) = iκ

Dexp(−iδζ ′) sinh[D(L − ζ ′)]

×[

cosh(DL)+ iΔ

Dsinh(DL)

]−1

,

w12g(L|ζ ) = iκ∗

Dexp[iδ(L + ζ ′)] sinh(Dζ ′)

×[

cosh(DL)+ iΔ

Dsinh(DL)

]−1

,

w11g(L|ζ ′) = exp(iδζ ′){

cosh[D(L − ζ ′)]+ iΔ

Dsinh[D(L − ζ ′)]

}

×[

cosh(DL)+ iΔ

Dsinh(DL)

]−1

. (6.310)

The output operators obey the same commutation relations as (6.305)

[ A1(L), A†1(L)] = [ A2(0), A†

2(0)] = 1. (6.311)

Page 381: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 373

Now we see that the relations (6.307) and (6.308) have the form (6.267), wherethe integrated quantum noise terms are

M1(L) = −i∫ L

0w22g(L|ζ ′)G1(ζ ′|0) dζ ′

−i∫ L

0w12g(L|ζ ′)G2(ζ ′|L) dζ ′, (6.312)

M2(L) = −i∫ L

0w21g(L|ζ ′)G1(ζ ′|0) dζ ′

−i∫ L

0w11g(L|ζ ′)G2(ζ ′|L) dζ ′. (6.313)

Their commutators are

[M1(L), M†1(L)] =

∫ L

0

[|w22g(L|ζ ′)|2 + |w12g(L|ζ ′)|2] dζ ′ 1,

[M1(L), M†2(L)]

=∫ L

0

[w22g(L|ζ ′)w∗

21g(L|ζ ′)+ w12g(L|ζ ′)w∗11g(L|ζ ′)

]dζ ′ 1,

[M2(L), M†1(L)]

=∫ L

0

[w21g(L|ζ ′)w∗

22g(L|ζ ′)+ w11g(L|ζ ′)w∗12g(L|ζ ′)

]dζ ′ 1,

[M2(L), M†2(L)] =

∫ L

0

[|w21g(L|ζ ′)|2 + |w11g(L|ζ ′)|2] dζ ′ 1. (6.314)

The commutator matrix is

([M1(L), M†

1(L)] [M1(L), M†2(L)]

[M2(L), M†1(L)] [M2(L), M†

2(L)]

)

=∫ L

0

( |w22g(L|ζ ′)|2 w22g(L|ζ ′)w∗21g(L|ζ ′)

w21g(L|ζ ′)w∗22g(L|ζ ′) |w21g(L|ζ ′)|2

)dζ ′ 1

+∫ L

0

( |w12g(L|ζ ′)|2 w12g(L|ζ ′)w∗11g(L|ζ ′)

w11g(L|ζ ′)w∗12g(L|ζ ′) |w11g(L|ζ ′)|2

)dζ ′ 1

≥(

0 00 0

)1. (6.315)

We have shown that the device under investigation is an attenuator according to theformal definition. It has the attenuator characteristic property (6.281).

(ii) Derivation of normal characteristic function

Page 382: Quantum Aspects of Light Propagation

374 6 Periodic and Disordered Media

We can pass from the Heisenberg picture to the Schrodinger picture using charac-teristic functionals

CN (β1, β2, β1(z), β2(z), 0) = Tr

{ρe(0)

× exp

[β1 A†

1 + β2 A†2 +

∫ L

0β1(z′)G†

1(z′) dz′ +∫ L

0β2(z′)G†

2(z′) dz′]

× exp

[−β∗1 A1 − β∗2 A2 −

∫ L

0β∗1 (z′)G1(z′) dz′ −

∫ L

0β∗2 (z′)G2(z′) dz′

]}, (6.316)

CN (β1, β2, β1(z), β2(z), L) = Tr

{ρe(L)

× exp

[β1 A†

1 + β2 A†2 +

∫ L

0β1(z′)G†

1(z′) dz′ +∫ L

0β2(z′)G†

2(z′) dz′]

× exp

[−β∗1 A1 − β∗2 A2 −

∫ L

0β∗1 (z′)G1(z′) dz′ −

∫ L

0β∗2 (z′)G2(z′) dz′

]}. (6.317)

Here we abbreviate G1(z) ≡ G1(z, 0), G2(z) ≡ G2(z, L). Choosing the input quan-tum noise fields in the vacuum state, we have

CN (β1, β2, β1(z), β2(z), 0) = CN (β1, β2, 0). (6.318)

We do not describe the output quantum noise fields and so we are interested in thenormal characteristic function

CN (β1, β2, L) = CN (β1, β2, 0, 0, L)

= CN(β1u∗11(L)+ β2u∗21(L), β1u∗12(L)+ β2u∗22(L),

iβ1w∗22g(L , z)+ iβ2w

∗21g(L , z),

iβ1w∗12g(L , z)+ iβ2w

∗11g(L , z), 0

)

= CN(β1u∗11(L)+ β2u∗21(L), β1u∗12(L)+ β2u∗22(L), 0

). (6.319)

This already suffices for the proof that the procedure yields a completely positivemap.

In Nielsen and Chuang (2000), the notion of quantum fidelity of two states ρ, ρ ′

is introduced,

F = Tr

{√ρ

12 ρ ′ρ

12

}. (6.320)

Here we put ρ = ρ(0), ρ ′ = ρ(L). For pure states, ρ(0) = |ψ(0)〉〈ψ(0)|, ρ(L) =|ψ(L)〉〈ψ(L)|, and the formula for the fidelity simplifies

F = |〈ψ(L)|ψ(0)〉|. (6.321)

Page 383: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 375

Attenuated coherent states are coherent again,

|ψ(0)〉 = |A1(0), A2(L)〉 = |A1in, A2in〉,|ψ(L)〉 = |A1(L), A2(0)〉 = |A1out, A2out〉, (6.322)

where (see (6.260)) A1in = A1(0), A2in = A2(L), A1out=A1(L), A2out = A2(0). It isknown that for the coherent states (Perina 1991)

|〈A1out, A2out|A1in, A2in〉|

= exp

(−1

2|A1out − A1in|2 − 1

2|A2out − A2in|2

). (6.323)

The quantum fidelity should be applied in the time or space Schrodinger picture.In Severini et al. (2004), with focusing on mode 1, a transmission coefficient has

been introduced

T = 〈 A†1(L) A1(L)〉

〈 A†1(0) A1(0)〉

. (6.324)

The quantum averages are calculated in the state |ψ(0)〉. The transmission coeffi-cient ought to be utilized in the space Heisenberg picture. Both in this and in theSchrodinger picture the numerical analysis simplifies for |ψ(0)〉 = |Ain, 0〉. Thetransmission spectrum is a function of a mismatch coefficient δ (cf. Figs. 6.1, 6.2),

T (δ) = |u11(L)|2. (6.325)

The quantum fidelity spectrum is a function of the same coefficient

F(δ) = exp

{−1

2

[|u11(L)− 1|2 + |u21(L)|2] |Ain|2}

, (6.326)

but we restrict ourselves to |Ain|2 = 1 in Fig. 6.3.

Fig. 6.1 Transmissionspectrum for a corrugatedLiNbO3 planar waveguide asa function of the mismatchcoefficient δ, for κ = 3π

Page 384: Quantum Aspects of Light Propagation

376 6 Periodic and Disordered Media

Fig. 6.2 The same as inFig. 6.1, but for κ = 4π

L

µ

Fig. 6.3 Fidelity spectrumfor a corrugated LiNbO3

planar waveguide as afunction of the mismatchcoefficient δ. Here κ = 4π

L ,but other parameters are thesame as in Fig. 6.1. For|κ| = 3π

L , F(δ) = 1 is notobtained

µ

We have calculated the semiclassical and quantum transmission and fidelity spec-tra for a corrugated LiNbO3 planar waveguide as functions of the mismatch coef-ficient δ between the spatial corrugation of the refractive index of the guide andthe wavenumber of the propagating mode. We have used units [δ] = μm−1. Thespatial corrugation has caused a coupling, which is characterized by the couplingconstant κ , κ = 3π

L (Fig. 6.1) or κ = 4πL (Figs. 6.2, 6.3). The waveguide length is

L = 1824.598 μm. We assume that

−√

9+ 6.52

Lπ ≤ δ ≤

√9+ 6.52

Lπ. (6.327)

Here 6 is the number of maxima of the spectrum plotted in Fig. 6.1, and 6.5 is a valuewhich makes the curves end with the next minimum approximately. The losses arecharacterized by the damping constant γ ≥ 0, which is chosen as 0 (in the casewithout losses) and 0.01κ

√5. Line a (b) corresponds to the damping coefficient

γ = 0 (γ = 0.01κ√

5).

Page 385: Quantum Aspects of Light Propagation

6.2 Corrugated Waveguides 377

The output two-mode state differs from the input state only by an inessentialphase factor when either

δ ≡ 0 mod2π

L,√

δ2 − |κ|2 ≡ 0 mod2π

L, (6.328)

or

δ ≡ π

Lmod

L,√

δ2 − |κ|2 ≡ π

Lmod

L. (6.329)

These conditions cannot be fulfilled for |κ| = 3πL , more generally, for |κ| =

πL mod 2π

L . It can be satisfied for |κ| = 4πL , more generally, |κ| = 0 mod 2π

L . In

Fig. 6.3 this condition is met for δ = 5πL , |κ| = 4π

L ,√

δ2 − |κ|2 = 3πL (one of the

Pythagorean triangles).In (Severini et al. 2004), a contradirectional coupler has been described by the

following equations

d

dzA1(z) = iKLA2(z) exp(2iδz)− γ

2A1(z),

d

dzA2(z) = −iKLA1(z) exp(−2iδz)− γ

2A2(z), (6.330)

where KL ≡ κ ≥ 0. Although the authors specify that α ≡ γ

2 characterizes leak-age phenomena, they do not write signs conformable to the attenuator case, cf.equations (6.272). The second equation describes rather amplification in mode 2.The amplification has perhaps led to the derivation of “a state whose quantum prop-erties are preserved”.

Let us assume that in a medium, an attenuator in mode 1 and an amplifier in mode2 occur, although we do not know of such a case. We could work with semiclassicaland quantum noises as in the previous sections, but neither the normal ordering northe antinormal one lead to a simple Schrodinger picture. With respect to the quantumnoise of the amplifiers, one can assert that there is no state in which all the quantumproperties are preserved.

We have dealt with the problem of quantum description of light modes whichpropagate in a periodic medium in opposite directions (Perinova et al. 2006).Although we believe that partial differential equations comprising both time andspace derivatives would be appropriate for the description, we have neglected thetime ones and retained the space ones. As the coupled-mode theory is a classicaldescription by means of ordinary differential equations involving a space derivativewhich leads to a quantum description of copropagating modes, it is also widelyused for such a simple quantum description of counterpropagating modes. We haveincluded also the gain and losses which have not been described to our knowledgeyet. As application we have treated the conditions that can be imposed on a waveg-uide for the output state to be the same as the incoming one.

Bozhevolnyi et al. (2005) study propagation of long-range surface plasmonpolaritons along periodically modulated medium both theoretically and

Page 386: Quantum Aspects of Light Propagation

378 6 Periodic and Disordered Media

experimentally. Surface plasmon polaritons are quasi-two-dimensional electromag-netic excitations that propagate along a dielectric–metal interface. Their applicationprospects are narrow. More complex excitations are an exception that are createdin the configuration of two similar and very close metal–dielectric interfaces, suchas surfaces of a thin metal film embedded in a dielectric. Then it is appropriate tospeak of long-range surface plasmon polaritons.

Similarity to dielectric symmetric waveguides suggests to realize the band-gapeffect for the long-range surface plasmon polaritons. The metal films are periodi-cally thickness modulated. This is achieved with a periodic array of metal ridges.Provided that we know the electric field E0(r) propagating along the metal film andthe electric field Green tensor G(r, r′) for the same structure, we can obtain the totalelectric field E(r) resulting in the process of multiple scattering by the ridges bysolving the equation

E(r) = E0(r)+ k20

∫G(r, r′)

[ε(r′)− εref(r′)

]E(r′) d2r′. (6.331)

Here k0 is the free-space wave number, ε is the dielectric constant of the total struc-ture inclusive of the metal ridges, and εref is the dielectric constant of the referencestructure (only a metal film embedded in a dielectric).

The gap in transmission and the peak in reflection are centred at λg ≈ 2nΛ,where n is the refractive index of the dielectric and Λ is the grating period. For lowridges, the gap and the peak improve with the increase of the ridge height. For largerheights, the band-gap effect was not achieved. For n = 1.543 and Λ = 500 nm,we obtain λg = 1543 nm. Band gaps centred at 1550 nm and 20 nm wide havebeen simulated and experimentally investigated. The lengths of the structures wereL = 20, 40, 80, 160 μm. The band-gap effect has been utilized for design andfabrication of a compact wavelength add-drop filter.

Deng et al. (2006) have reported second-harmonic generation in a sample madeof lithium niobate. Near the surface, a waveguide was fabricated applying theproton-exchange technique. Ultraviolet laser lithography was applied to make pho-tonic band-gap gratings on the sample. On the sample, two different gratings areinscribed. The first one couples the pump into the waveguide and the pump wavemay come at an angle of around 45o. The second one is the photonic band-gap grat-ing. A numerical model utilizes the coupled-mode theory. The corrugation couplesTM to TM modes. The authors have measured that the second-harmonic generationin a waveguide mode is very weak compared with the second harmonic radiated intothe substrate from the Cherenkov condition.

6.3 Photonic Crystals

The integral equation for quantum mechanical Green operator is a pattern for otherintegral equations of the field theory, in particular for the relation comprising theinput and retarded electromagnetic fields (Białynicki-Birula and Białynicka-Birula

Page 387: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 379

1975). A treatment of two- and three-dimensional photonic crystals may use a sim-ilar approach.

(i) Quadrature-phase squeezing in photonic crystals

The Green function method has been used for classical fields in Sakoda andOhtaka (1996a,b). Sakoda (2002) has obtained the enhancement of a quantum opti-cal process by use of a perturbation theory based on a Green function formalism.The results are related to degenerate optical parametric amplification, but the per-turbation theory is not limited to this process and can be applied to other quantum-optical processes in the photonic crystals.

The quantization proceeds according to Glauber and Lewenstein (1991). Theeigenmode of the electric field is denoted as Ekn(r), where k is a wave vector inthe first Brillouin zone and n is a band index. The eigenmodes are normalized withthe condition

Vε(r)E∗kn(r)Ek′n′(r) d3r = V δkk′δnn′ , (6.332)

where V means the volume of the photonic crystal, and ε is a spatially periodicdielectric constant. The volume of unit cell is denoted as V0.

On expressing the electric-field operator in the form

E(r, t) =∑

k,n

i

√�ωkn

2ε0V[akn(t)Ekn(r)− a†

kn(t)E∗kn(r)], (6.333)

where akn and a†kn are the usual photon annihilation and creation operators, respec-

tively, and writing the magnetic-induction operator similarly, the total electromag-netic energy (in the volume) is reduced to a quantum-mechanical Hamiltonian

H =∑

k,n

�ωkn

[a†

kn(t)akn(t)+ 1

21

]. (6.334)

The nonlinear medium is described by a second-order susceptibility tensorχ (2)(r) that has the same spatial periodicity as ε(r). But χ (2)(r) is nonzero onlyin the region 0 ≤ z ≤ l = anz , where a is the lattice constant of the crystal and nz

is a positive integer.It is assumed that a pump wave denoted Ep(r, t) and a signal wave denoted

Es(r, t) propagate along the z-axis. Both waves are single mode. We let kpz and ksz

denote the z-components of their wave vectors, and we introduce a phase mismatchΔkz , Δkz = kpz − 2ksz . The frequency of the pump wave, ωp, is twice that of thesignal wave, ωs. We let vg denote the group velocity of the signal wave.

Since

Ep(r, t) = iA[Ep(r) exp(−iωpt + iθ )− E∗p(r) exp(iωst − iθ )] (6.335)

Page 388: Quantum Aspects of Light Propagation

380 6 Periodic and Disordered Media

is a classical quantity and

Es(r, t) = i

√�ωs

2ε0V[as(0)Es(r) exp(−iωst)− a†

s (0)E∗s (r) exp(iωst)], (6.336)

where Ep(r) and Es(r) are eigenmodes of the electric field, A is the amplitude ofthe pump wave, θ the shift of its phase, and as(0) and a†

s (0) are photon annihilationand creation operators, respectively, can be considered to be an output electric-fieldoperator, the integral equation for this operator is formulated,

ε0ε(r)E(r, t)+ P(r, t) = ε0ε(r)Es(r, t)+ ε(r)

V

×∑

k,n

ωknEkn(r)∫

V

∫ t

−∞E∗kn(r′) · P(r′, t ′) sin

[ωkn(t − t ′)

]d3r′ dt ′, (6.337)

where P(r, t) is the output nonlinear polarization operator,

P(r, t) ≈ 2χ (2)(r) : Ep(r, t)E(r, t). (6.338)

The solution of the integral equation (6.337) is of the form

E(r, t) ≈ i

√�ωs

2ε0V[bEs(r) exp(−iωst)− b†E∗s (r) exp(iωst)], (6.339)

where

b = as(0) cosh(|β ′|l)+ exp[i(θ + φ′)]a†s (0) sinh(|β ′|l),

b† = a†s (0) cosh(|β ′|l)+ exp[−i(θ + φ′)]as(0) sinh(|β ′|l), (6.340)

with φ′ being the phase of the effective coupling constant (inclusive of the ampli-tude A),

β ′ = βηξ, β = ωs AFs,p,s

ε0vg,

η =sin(

lΔkz

2

)

nz sin(

aΔkz

2

) , ξ = exp

[ia(nz − 1)Δkz

2

],

Fs,p,s = 1

V0

V0

E∗s (r) · χ (2)(r) : Ep(r)E∗s (r) d3r. (6.341)

The nonlinear properties of photonic crystals were reviewed in Slusher and Eggleton(2003).

Page 389: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 381

(ii) Parametric down conversion in a multilayered structure simply described

A description of spontaneous parametric down conversion in finite-length multilayerstructure has been developed using semiclassical and quantum approaches (Centiniet al. 2005). The semiclassical model has allowed one to find the criterion for design-ing and optimizing the structure. The quantum model is related to the properties ofemitted entangled photon pairs.

One considers a one-dimensional dispersive lossless inhomogeneous medium,where both the dielectric constant, ε(z, ω), and the nonlinear susceptibility, d (2)(z),are functions of a single spatial coordinate z. The study is limited to s-polarizedplane monochromatic waves that fall onto the interfaces in the normal direction. Theplanes z = 0 and z = L are the first and last interfaces of the structure embedded inair. The classical treatment begins with the following nonlinear, coupled Helmholtzequations:

d2 Es

dz2+ ω2

s εs(z)

c2Es = −2

ω2s

c2d (2)(z)E∗

i Ep,

d2 Ei

dz2+ ω2

i εi(z)

c2Ei = −2

ω2i

c2d (2)(z)E∗

s Ep,

d2 Ep

dz2+ ω2

pεp(z)

c2Ep = −2

ω2p

c2d (2)(z)E∗

i Es, (6.342)

where εn(z) ≡ ε(z, ωn), ωn is the angular frequency of the field n, n = s, i, p,and s, i, and p stand for signal, idler, and pump, respectively. It is assumed thatωp = ωs + ωi. The treatment has followed (D’Aguanno et al. 2002).

The solutions of the corresponding linear equations E can easily be decomposedinto forward- and backward-propagating waves E = EF + EB, where F (B) meansthe forward (backward) propagation. The solutions to these equations are intro-duced, Θ(+)

n and Θ(−)n , which fulfil the following boundary conditions,

Θ(+)nF (−0) = 1, Θ

(+)nB (L + 0) = 0,

Θ(−)nF (−0) = 0, Θ

(−)nB (L + 0) = 1. (6.343)

These solutions have the familiar properties

Θ(+)nF (L + 0) = t(+)

a , Θ(+)nB (−0) = r(+)

a ,

Θ(−)nB (−0) = t(−)

n , Θ(−)nF (L + 0) = r(−)

n , (6.344)

where r(+)n and t(+)

n (r(−)n and t(−)

n ) are the linear reflection and transmission com-plex coefficients for left-to-right (right-to-left) propagation, cf. (Born and Wolf1999, Yeh 1988). The solutions of the nonlinear equations are decomposed as

En = A(+)n (z)Θ(+)

n (z)+ A(−)n (z)Θ(−)

n (z), (6.345)

Page 390: Quantum Aspects of Light Propagation

382 6 Periodic and Disordered Media

where A(+)n (z) and A(−)

n (z) are slowly varying complex envelope functions. Espe-cially,

EnB(z) = [A(+)n (0)r(+)

n + A(−)n (0)t(−)

n

]exp

(−i

ωnz

c

)for z ≤ 0,

EnF(z) = [A(−)n (L)r(−)

n + A(+)n (L)t(+)

n

]

× exp

(iωn(z − L)

c

)for z ≥ L . (6.346)

As usual with the semiclassical approach, a scalar product is introduced,

〈 f |g〉 = 1

L

∫ L

0f ∗(z)g(z) dz. (6.347)

A description has been developed using, e.g., the integrals

Γ(k,l)(s, j) =

⟨Θ( j)

s |d (2)Θ(k)p Θ

(l)∗i

⟩, (6.348)

where j, k, l = +,−.In contrast, we present a quantum description using the solutions of linear equa-

tions Θ(+)a and Θ(−)

a , a = s, i, which fulfil the boundary conditions

Θ(+)aB (−0) = 0, Θ

(+)aF (L + 0) = 1,

Θ(−)aB (−0) = 1, Θ

(−)aF (L + 0) = 0. (6.349)

The connecting relations are

Θ(+)a = Θ(+)

a t(+)∗a +Θ(−)

a r(−)∗a ,

Θ(−)a = Θ(+)

a r(+)∗a +Θ(−)

a t(−)∗a . (6.350)

We introduce the overlap integrals

Γ(k,l)(s, j) =

⟨Θ( j)

s |d (2)Θ(k)p Θ

(l)∗i

⟩. (6.351)

The nonlinear interaction in the entire photonic band-gap structure is describedby an interaction Hamiltonian H (t) given as a sum of operators H (l)(t) (l =1, . . . , N ) that characterize every layer of the structure,

H (t) =N∑

l=1

H (l)(t), (6.352)

where N is the number of layers of the structure. Strong pump positive-frequencyelectric-field amplitudes E (l,+)

p (z, t), weak signal and idler positive-frequency

electric-field operator amplitudes E (l,+)a (z, t), a = s, i, and their Hermitian-

conjugated expressions E (l,−)a (z, t) are introduced. Then

H (l)(t) = d (l)∫ zl

zl−1

[E (l,+)

p (z, t)E (l,−)s (z, t)E (l,−)

i (z, t)+ H.c.]

dz. (6.353)

Page 391: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 383

Here d (l) = �

c

√ωsωid (l), and d (l) means the second-order susceptibility of the lth

layer.A monochromatic pump electric field has the positive-frequency amplitude

E (l,+)p (z, t) =

{B(l)

pF exp[ik(l)

p (z − zl−1)]

+ B(l)pB exp

[−ik(l)

p (z − zl−1)]}

exp(−iωpt); (6.354)

k(l)p means the wave vector of the pump field in the lth layer and B(l)

pF (B(l)pB) is a

complex coefficient. Down-converted fields are polychromatic, and it is convenient

to express their amplitudes in units of√

2�ωaε0V , where V is the quantization volume.

Then the positive-frequency operator amplitudes are

E (l,+)a (z, t) =

∫b(l)

a

{b(l)

aF(ωa) exp[ik(l)

a (ωa)(z − zl−1)]

+ b(l)aB(ωa) exp

[−ik(l)a (ωa)(z − zl−1)

]}exp(−iωat) dωa . (6.355)

Here b(l)a = 1√

ε(l)a

, ε(l)a stands for the relative permittivity in the lth layer for the

field a.It holds that

B(l)pF = A(+)

p (0)Θ(+)pF (zl−1)+ A(−)

p (L)Θ(−)pF (zl−1),

B(l)pB = A(+)

p (0)Θ(+)pB (zl−1)+ A(−)

p (L)Θ(−)pB (zl−1), l = 1, . . . , N + 1; (6.356)

particularly, B(N+1)pB = A(−)

p (L), but B(0)pF = A(+)

p (0). Similarly,

b(l)a b(l)

aF(ωa) = b(N+1)aF (ωa)Θ(+)

aF (zl−1)+ b(0)aB(ωa)Θ(−)

aF (zl−1),

b(l)a b(l)

aB(ωa) = b(N+1)aF (ωa)Θ(+)

aB (zl−1)+ b(0)aB(ωa)Θ(−)

aB (zl−1), l = 1, . . . , N . (6.357)

Here b(N+1)aF (ωa) and b(0)

aB(ωa) are output operators.The solution |ψ〉s,i of the Schrodinger equation correct up to first order on the

assumption of the initial vacuum state |vac〉s,i for the down-converted fields reads as

|ψ〉s,i = |vac〉s,i − i

�lim

T→∞

∫ T

−TH (t)|vac〉s,i dt. (6.358)

It can be written in the form

|ψ〉s,i = |vac〉s,i +∫ ∫ [

ΦFF(ωs, ωi)b(N+1)†sF b(N+1)†

iF |vac〉s,i+ΦFB(ωs, ωi)b

(N+1)†sF b(0)†

iB |vac〉s,i +ΦBF(ωs, ωi)b(0)†sB b(N+1)†

iF |vac〉s,i+ΦBB(ωs, ωi )b

(0)†sB b(0)†

iB |vac〉s,i dωs dωi

]. (6.359)

Page 392: Quantum Aspects of Light Propagation

384 6 Periodic and Disordered Media

Here

ΦFF(ωs, ωi) = 2π√

ωsωiL

icδ(ωp − ωs − ωi)

[A(+)

p Γ(+,+)(s,+) + A(−)

p Γ(−,+)(s,+)

],

ΦFB(ωs, ωi) = 2π√

ωsωiL

icδ(ωp − ωs − ωi)

[A(+)

p Γ(+,−)(s,+) + A(−)

p Γ(−,−)(s,+)

],

ΦBF(ωs, ωi) = 2π√

ωsωiL

icδ(ωp − ωs − ωi)

[A(+)

p Γ(+,+)(s,−) + A(−)

p Γ(−,+)(s,−)

],

ΦBB(ωs, ωi) = 2π√

ωsωiL

icδ(ωp − ωs − ωi)

[A(+)

p Γ(+,−)(s,−) + A(−)

p Γ(−,−)(s,−)

]. (6.360)

Even though these functions are not probability amplitudes, they can be combined,e.g., with parameters of a finite spatial region to give such amplitudes.

(iii) Parametric down conversion in a multilayered structure including polarization

The foregoing description has been generalized and revised in part in (Perina Jr.,et al. 2006). For instance, the structure may be embedded in a medium with therelative permittivity ε(0)

n (ε(N+1)n ) in front of (beyond) the sample. The linear indices

of refraction are introduced, n(l)m =

√ε

(l)m , l = 0, . . . , N + 1, m = s, i, p.

The treatment has been restricted to plane waves with wave vectors parallelto the yz-plane. The forward-propagating fields have the wave vectors k(l)

mF =eyk(l)

m sin(ϑ (l)m )+ ezk(l)

m cos(ϑ (l)m ), where

k(l)m = ωm

cn(l)

m . (6.361)

The angles ϑ (l)m fulfil the Snell law:

n(l)m sin(ϑ (l)

m ) = constant, l = 0, . . . , N + 1. (6.362)

The wave-vectors k(l)mB = eyk(l)

m sin(ϑ (l)m )−ezk(l)

m cos(ϑ (l)m ) characterize the backward-

propagating fields. For simplicity, k(l)m ≡ k(l)

mF, k(l)m,x = 0, k(l)

m,y = k(l)m sin(ϑ (l)

m ), k(l)m,z =

k(l)m cos(ϑ (l)

m ).At this moment, we may still use classical concepts and restrict ourselves to

monochromatic waves. We distinguish the TE- and TM-waves. These have the elec-tric fields of the forms

Em,TE = Em,TEex ,

Em,TM = Em,TMey cos [ϑm(z)]− Em,TMez sin [ϑm(z)] , (6.363)

where ϑm(z) = ϑ (l)m for z in the lth layer and

Em,TM = 1

ikm,y(z)

∂yEm,TM = 1

ikm,z(z)

∂zEm,TM, (6.364)

Page 393: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 385

with km,y(z) = k(l)m,y and km,z(z) = k(l)

m,z (in the lth layer).The linear Helmholtz equations read

∂2 Em,α

∂z2+ k2

m,z Em,α = 0, m = s, i, p, α = TE, TM. (6.365)

The prolongation conditions depend on the polarization. The case α = TE is veryattractive, since the conditions require that Em,α and ∂

∂z Em,α be continuous at thepoints z = zl . Let x and y be zero for definiteness. The case α = TM includes theconditions that Em,α cos[ϑm(z)] and 1

cos[ϑm (z)]∂∂z Em,α are continuous at z = zl .

We introduce solutions to these equations, Θ(+)m,α and Θ(−)

m,α , which satisfy thefollowing boundary conditions:

Θ(+)mF,α(−0, ωm) = 1, Θ

(+)mB,α(L + 0, ωm) = 0,

Θ(−)mF,α(−0, ωm) = 0, Θ

(−)mB,α(L + 0, ωm) = 1. (6.366)

Similarly, we will need some of the solutions Θ(+)m,α and Θ(−)

m,α , which fulfil the bound-ary conditions

Θ(+)mB,α(−0, ωm) = 0, Θ

(+)mF,α(L + 0, ωm) = 1,

Θ(−)mB,α(−0, ωm) = 1, Θ

(−)mF,α(L + 0, ωm) = 0. (6.367)

In quantum physics, we will use E(+)p,α(z, ωp) instead of Ep,α(z) for the positive-

frequency electric-field amplitude of a monochromatic component at frequencyωp with polarization α. The positive-frequency electric-field amplitude E(+)

p (z, t)

is decomposed into the TE- and TM-wave contributions E(+)p,TE(z, t) and E(+)

p,TM(z, t)and expressed in the forms

E(+)p (z, t) = E(+)

p,TE(z, t)+ E(+)p,TM(z, t)

= 1√2π

∫ ∞

0E(+)

p (z, ωp) dωp

= 1√2π

∫ ∞

0

[E(+)

p,TE(z, ωp)+ E(+)p,TM(z, ωp)

]dωp. (6.368)

Here

E(+)p,α(z, ωp) = A(l)

pF,α(ωp)e(l)pF,α(ωp) exp[ik(l)

p,z(z − zl−1)]

+ A(l)pB,α(ωp)e(l)

pB,α(ωp) exp[−ik(l)p,z(z − zl−1)], (6.369)

with

A(l)pF,α(ωp) = A(0)

pF,α(ωp)Θ(+)pF,α(zl−1, ωp)+ A(N+1)

pB,α (ωp)Θ(−)pF,α(zl−1, ωp),

A(l)pB,α(ωp) = A(0)

pF,α(ωp)Θ(+)pB,α(zl−1, ωp)

+ A(N+1)pB,α (ωp)Θ(−)

pB,α(zl−1, ωp), l = 1, . . . , N + 1. (6.370)

Page 394: Quantum Aspects of Light Propagation

386 6 Periodic and Disordered Media

For l = 0, we use this formula on replacement of zl−1 by z0. Further,

e(l)mF,TE(ωm) = e(l)

mB,TE(ωm) = ex ,

e(l)mF,TM(ωm) = ey cos(ϑ (l)

m )− ez sin(ϑ (l)m ),

e(l)mB,TM(ωm) = ey cos(ϑ (l)

m )+ ez sin(ϑ (l)m ), (6.371)

where m = p. The positive-frequency electric-field operators E(+)s (z, t) and E(+)

i (z, t)for the signal and idler fields can be decomposed into the TE- and TM-wave contri-butions E(+)

a,TE(z, t) and E(+)a,TM(z, t) and expressed as follows (Vogel et al. 2001)

E(+)a (z, t) = E(+)

a,TE(z, t)+ E(+)a,TM(z, t)

= 1√2π

∫ ∞

0E(+)

a (z, ωa) dωa

= 1√2π

∫ ∞

0

[E(+)

a,TE(z, ωa)+ E(+)a,TM(z, ωa)

]dωa , a = s, i. (6.372)

Here

E(+)a,α(z, ωa) =

√�ωa

2ε0cB{

a(l)aF,α(ωa)e(l)

aF,α(ωa) exp[ik(l)a,z(z − zl−1)]

+ a(l)aB,α(ωa)e(l)

aB,α(ωa) exp[−ik(l)a,z(z − zl−1)]

}, (6.373)

with B the area of the transverse profile of a beam and

b(l)a a(l)

aF,α(ωa) = b(N+1)a a(N+1)

aF,α (ωa)Θ(+)aF,α(zl−1, ωa)

+ b(0)a a(0)

aB,α(ωa)Θ(−)aF,α(zl−1, ωa),

b(l)a a(l)

aB,α(ωa) = b(N+1)a a(N+1)

aF,α (ωa)Θ(+)aB,α(zl−1, ωa)

+ b(0)a a(N+1)

aB,α (ωa)Θ(−)aB,α(zl−1, ωa), l = 1, . . . , N . (6.374)

Further, e(l)aF,α(ωa) and e(l)

aB,α(ωa) are defined by relation (6.371), where m = a.

The operators a(N+1)aF,α (ωa) and a(0)

aB,α(ωa) obey the following commutation rela-tions

[a(N+1)

aF,α (ωa), a(N+1)†a′F,α′ (ω′a)

]= δα,α′δa,a′δ(ωa − ω′a)1,

[a(N+1)

aF,α (ωa), a(N+1)a′F,α′ (ω′a)

]= 0,

[a(0)

aB,α(ωa), a(0)†a′B,α′ (ω

′a)]= δα,α′δa,a′δ(ωa − ω′a)1,

[a(0)

aB,α(ωa), a(0)a′B,α′ (ω

′a)]= 0,

Page 395: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 387

[a(N+1)

aF,α (ωa), a(0)†a′B,α′(ω

′a)]= 0,

[a(N+1)

aF,α (ωa), a(0)a′B,α′(ω

′a)]= 0. (6.375)

The interaction Hamiltonian H (t) describing spontaneous parametric down-conversion can be written as

H (t) = ε0B(2π )

32

∫ zN

z0

∫ ∞

0

∫ ∞

0

∫ ∞

0

α,β,γ=TE,TM

d(z)

...[E(+)

p,α(z, ωp)E(−)s,β (z, ωs)E

(−)i,γ (z, ωi)+ H. c.

]dz dωp dωs dωi, (6.376)

where d(z) means a third-order tensor of nonlinear susceptibility and... denotes a

contraction, i.e., treble sum after the tensors are replaced by their components, andproducts of the corresponding components are formed.

The solution |ψ〉s,i of the Schrodinger equation correct up to first order on theassumption of the initial vacuum state |vac〉s,i for the down-converted fields is givenby the relation

|ψ〉s,i = |vac〉s,i +∫ ∞

0

∫ ∞

0

β,γ=TE,TM

[ΦFβFγ (ωs, ωi)

× b(N+1)s a(N+1)†

sF,β (ωs)b(N+1)i a(N+1)†

iF,γ (ωi)|vac〉s,i+ΦFβBγ (ωs, ωi)b

(N+1)s a(N+1)†

sF,β (ωs)b(0)i a(0)†

iB,γ (ωi)|vac〉s,i+ΦBβFγ (ωs, ωi)b

(0)s a(0)†

sB,β(ωs)b(N+1)i a(N+1)†

iF,γ (ωi)|vac〉s,i+ΦBβBγ (ωs, ωi)b

(0)s a(0)†

sB,β(ωs)b(0)i a(0)†

iB,γ (ωi)|vac〉s,i]

dωs dωi. (6.377)

Here

ΦFβFγ (ωs, ωi) = − i

∫ ∞

0

α=TE,TM

√ωsωiδ(ωp − ωs − ωi)

× A(0)pF,α(ωp)

m,n,o=F,B

∫ zN

z0

Θ(+)pm,α(z, ωp)Θ(+)∗

sn,β (z, ωs)Θ(+)∗io,γ (z, ωi)

× d(z)... epm,α(z, ωp)esn,β(z, ωs)eio,γ (z, ωi) dz dωp, (6.378)

ΦFβBγ (ωs, ωi) = − i

∫ ∞

0

α=TE,TM

√ωsωiδ(ωp − ωs − ωi)

× A(0)pF,α(ωp)

m,n,o=F,B

∫ zN

z0

Θ(+)pm,α(z, ωp)Θ(+)∗

sn,β (z, ωs)Θ(−)∗io,γ (z, ωi)

× d(z)... epm,α(z, ωp)esn,β(z, ωs)eio,γ (z, ωi) dz dωp, (6.379)

Page 396: Quantum Aspects of Light Propagation

388 6 Periodic and Disordered Media

ΦBβFγ (ωs, ωi) = − i

∫ ∞

0

α=TE,TM

√ωsωiδ(ωp − ωs − ωi)

× A(0)pF,α(ωp)

m,n,o=F,B

∫ zN

z0

Θ(+)pm,α(z, ωp)Θ(−)∗

sn,β (z, ωs)Θ(+)∗io,γ (z, ωi)

× d(z)... epm,α(z, ωp)esn,β(z, ωs)eio,γ (z, ωi) dz dωp, (6.380)

ΦBβBγ (ωs, ωi) = − i

∫ ∞

0

α=TE,TM

√ωsωiδ(ωp − ωs − ωi)

× A(0)pF,α(ωp)

m,n,o=F,B

∫ zN

z0

Θ(+)pm,α(z, ωp)Θ(−)∗

sn,β (z, ωs)Θ(−)∗io,γ (z, ωi)d(z)

... epm,α(z, ωp)esn,β(z, ωs)eio,γ (z, ωi) dz dωp, (6.381)

with epm,α(z, ωp) = e(l)pm,α(ωp), esn,β(z, ωs) = e(l)

sn,β (ωs), eio,γ (z, ωi) = e(l)io,γ (ωi) (in

the lth layer) when we restrict ourselves to the case A(N+1)pB,α (ωp) = 0.

Corona and U’Ren (2007) study type-II, frequency degenerate, collinear para-metric down-conversion in a χ (2) material with uniaxial birefringence. The type-II interaction means that the pump is an extraordinary wave, and the signal andidler are extraordinary and ordinary. For this operation, signal and idler photons areorthogonally polarized. The material is characterized by a spatial periodicity in itslinear optical properties. Introducing μ = o for the ordinary ray and μ = e for theextraordinary ray, the index of refraction will be

nμ(ω, z) ={

nμ1(ω), 0 < z < a,

nμ2(ω), a < z < Λ.(6.382)

The authors assume a = Λ2 in a numerical analysis. The Bloch waves are written in

the form

E(z, t) = EK (z, ω) exp{i[K (ω)z − ωt]}. (6.383)

Here EK (z, ω) is the Bloch envelope, which has the same period, Λ, as the material,and K (ω) stands for the Bloch wave number. The Bloch waves are described in thevicinity of K = mπ

Λ. Further m = 1.

A standard perturbative approach to the quantum description has been adopted byCorona and U’Ren (2007). There Eμ(r, t) (μ = p, s, i) represents the electric-fieldoperators related to each of the interacting fields. They assume that the pump fieldis classical or that the replacement

E (+)p (r, t) →

∫αp(ω)EKp (z, ω) exp{i[Kp(ω)z − ωt]} dω, (6.384)

where Kp(ω) is the Bloch wave number, EKp (z, ω) is the Bloch envelope, and αp(ω)is the spectral amplitude, may be done in the interaction Hamiltonian. The Bloch

Page 397: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 389

envelope may be expressed as a Fourier series,

EKp (z, ω) =∑

l

εpl(ω)eiGl z, (6.385)

in terms of the spatial harmonics Gl = 2πlΛ

. The authors touch the quantization ofthe signal and idler fields when they present the operator

E (+)μ (r, t) = i

∫ ∑

l

εμl(ω)lμ(ω)aμ

[Kμ(ω)+ Gl

]

× exp(

i{[

Kμ(ω)+ Gl]

z − ωt} )

dω, (6.386)

where Kμ(ω) is the Bloch wave number, εμl are the Bloch envelope Fourier seriescoefficients, and aμ (K (ω)) is the annihilation operator for the signal (s) or idler (i)mode. The normalization constant is

lμ(ω) =√

�ωK ′μ(ω)

2εμ(ω)S, (6.387)

where K ′μ(ω) is the first frequency derivative of Kμ, εμ(ω) is the permittivity in the

nonlinear medium, and S is the transverse beam area.The approach to quantization is macroscopic. The joint spectral amplitude has

been defined, which depends on the length of the crystal, L . The authors prove with anumerical calculation that each of the three interacting fields propagates essentiallyas a plane wave. They may utilize this approximation.

To obtain conditions for factorizability, the authors expand the mismatch LΔK (ω),where ΔK (ω) = Kp(ω) − Ks(ω) − Ki(ω). They let ωo denote the degenerate fre-quency and introduce the mismatch τ

( j)μ , μ = s, i, j = 2, 3, 4, in the j th frequency

derivatives between the wave numbers of the pump and the signal (idler) wavepackets

τ ( j)μ = L(K ( j)

p − K ( j)μ ), (6.388)

where

K ( j)μ = 1

j!

d j Kμ(ω)

dω j

∣∣∣∣ω=ωo

, K ( j)p = 1

j!

d j Kp(ω)

dω j

∣∣∣∣ω=2ωo

, (6.389)

and the frequency detunings νs,i = ωs,i − ωo. Whereas the zeroth-order approxima-tion is

LΔK (ω) ≈ LΔK (0) + O(1), (6.390)

where

ΔK (0) = Kp(2ωo)− Ks(ωo)− Ki(ωo), (6.391)

Page 398: Quantum Aspects of Light Propagation

390 6 Periodic and Disordered Media

and O(J + 1) represents (J + 1)th- and higher-order terms in the detunings; theauthors present a fourth-order approximation, J = 4.

They then choose αp(ωs + ωi) in the relation f000(ωs, ωi) = ωp(ωs + ωi)φ000(ωs, ωi), where φ000(ωs, ωi) is given by relation (15) in Corona and U’Ren(2007), as a Gaussian function with the parameter σ 2 and, in that defining relation,they replace sinc( LΔK (ω)

2 ) by another Gaussian (with the parameter 1γ

). The approx-imation of the function f (ωs, ωi) ≡ f000(ωs, ωi) comprises the quantity Φsi(νs, νi)given in (25) in the cited paper.

The expression is complicated. The factorability occurs if Φsi(νs, νi) = 0. Pro-vided that τ (1)

s = τ(1)i = 0, the expression simplifies. Then relations for | f (νs, νi)|

and arg[ f (νs, νi)] can be written. Further the authors assume that the signal and idlerphotons undergo much stronger dispersion than the pump. Particularly, they assumethat |τ ( j)

p |�|τ ( j)s |, |τ ( j)

i |, j = 2, 3, 4, τ ( j)p =L K ( j)

p . At last, they assume that the pumpis broad-band,

σ � 2 4

√4

γ

1√τ

(2)s + τ

(2)i

. (6.392)

While the phase of the joint spectral amplitude has been factorable without the con-dition (6.392), only now the modulus of the joint spectral amplitude reduces to

| f (νs, νi)| ≈ exp

4

(τ (2)

s ν2s + τ

(2)i ν2

i

)2]

. (6.393)

In Corona and U’Ren (2007), it has been shown also that this modulus can describea nearly factorable two-photon state.

The authors first evidence that in nonlinear photonic crystals, complete groupvelocity matching, K ′

p=K ′s=K ′

i , or τ (1)s =τ

(1)i =0, can be achieved. Here it is assumed

that a = Λ2 .

Then one may search for the lattice period Λ, the permittivity contrast

α = 2[εμ1(ω)− εμ2(ω)

]

εμ1(ω)+ εμ2(ω), (6.394)

in which an independence of the frequency and of μ = e, o is actually assumed,and the crystal propagation angle θpm such that ΔK (0) = 0, K ′

p = K ′s, K ′

p = K ′i .

Here ωp = 2ωo, ωs = ωi = ωo as in relation (6.389). The fact that the producedtwo-photon state is nearly factorable has been shown in Law et al. (2000).

(iv) Further principles and effects

Diao and Blair (2007) have paid attention to the use of multilayer thin film structuresfor optical bistability and multistability. They have considered single-cavity andcoupled-cavity structures. In the case of a single-cavity structure, mirrors consistof M quarter-wave low-index layers and M−1 quarter-wave high-index layers. Thecavity is based on L quarter-wave high-index layers.

Page 399: Quantum Aspects of Light Propagation

6.3 Photonic Crystals 391

The coupled-cavity structures have N cavities and N + 1 mirrors. It is assumedthat the low-index layers are constructed from silicon dioxide. They have a nonlinearcoefficient of the refractive index change n2,silica ≈ 3 × 10−16 cm2W−1. The high-index layers have a coefficient n2.

For these structures, linear transmission (in magnitude and phase) and groupdelay may be calculated. The optical bistability and multistability are analyzedmainly by dependence of nonlinear transmission (in magnitude and phase) on nor-malized input intensity. As it is a multivalued function of the input, also the non-linear transmission (in magnitude only) is plotted versus the intensity within thecavity.

Photonic crystal fibres (PCFs) are dielectric optical fibres with an array of air-holes running along the fibre. Usually, the fibres employ a single dielectric mate-rial. Mortensen (2005) notes that other base materials have been studied besidesthe silica. Typically, the airholes are arranged in a triangular lattice with a pitchΛ. A waveguide is formed as a cladding and a core using the core defect, i.e., byremoval of a single airhole. From this, the author has realized a call for a theoryof photonic crystal fibres with an arbitrary base material. Solving a scalar two-dimensional Schrodinger-like equation, geometrical eigenvalues γ 2 have been cal-culated, dependent on the normalized airhole diameter d

Λ. If d

Λis below a critical

value, only the eigenvalue for the fundamental core mode γ 2c,1 and that for the fun-

damental cladding mode γ 2cl are seen. If d

Λis above the critical value, an eigenvalue

for the second-order core mode γ 2c,2 diverges from that for the fundamental cladding

mode. Then an abbreviation, γ 2c ≡ γ 2

c,1, is used and, for dΛ≤ 0.8, a third-order

polynomial is presented, which fits the eigenvalues for the fundamental core modewell. One considers the V parameter of the form

VPCF =√

γ 2cl − γ 2

c , (6.395)

and the endlessly single-mode regime is associated with the condition VPCF < π

(Mortensen et al. 2003).The dependence of the effective index neff = cβ

ωon the normalized free-space

wavelength λΛ

is expressed with a second-order polynomial, which agrees with fullyvectorial plane-wave simulations in the short-wavelength limit λ � Λ. The com-parison has been performed for a single slightly subcritical value of the normalizedairhole diameter d

Λand both for the fundamental core mode and for the fundamental

cladding mode.Della Villa et al. (2005) study formation of band gaps in photonic quasicrystals.

They have considered a photonic quasicrystal with a Penrose-type lattice. They haveinferred a band gap from the normalized local density of states

ρ(r0, ω) = Im {G(r0, r0, ω)} , (6.396)

where G(r, r0, ω) is the Green function. Numerical calculations were performedfor finite-size quasicrystals made of hundreds of rods. The normalized local densitywas determined at the centre of the quasicrystal, and the choice of the central point

Page 400: Quantum Aspects of Light Propagation

392 6 Periodic and Disordered Media

did not affect results. In comparison with the photonic crystal with square lattice,the quasicrystal exhibits small additional band gaps. In the central band gap of thephotonic quasicrystal, the normalized local density of states exhibits the exponentialdecay similar to that of the photonic crystal. The central band gap seems to stemfrom relatively short-distance interactions. Lateral band gaps stem from long-rangeinteractions.

The Fourier spectrum of the permittivity profile for the quasicrystal, togetherwith the usual Bragg condition, predicts the central and upper band gap, and a lowercontrast of the permittivities is more advantageous. The frequency, at which thelower band gap occurs, may not be explained using single scattering and should soinclude multiple scattering.

The use of two-dimensional photonic crystals instead of the conventional one-dimensional feedback grating can lower the lasing threshold of distributed feedbacklasers. Such photonic-crystal-based organic lasers have been studied (Harbers et al.2005).

The photonic crystals can change the spontaneous emission dynamics of excitedatoms (Yablonovich 1987, John 1987). The photonic crystals modify also sponta-neous emission of quantum dots (Lodahl et al. 2004, Yoshie et al. 2004).

Hughes (2005a) introduces a scheme that enables one to study quantum corre-lations between two quantum dots in a planar-photonic-crystal nanocavity. He con-siders the fundamental cavity mode ec(r), which fulfils the normalization condition

all spaceεc(r)|ec(r)|2 d3r = 1, (6.397)

where εc(r) is the permittivity of the nanocavity structure. He considers a tensor-valued Green function

Gb(r, r′; ω) = Gt(r, r′; ω)−( c

ω

)2 δ(r− r′)1|ec(r)| , (6.398)

where for simplicity

Gt(r, r′; ω) =( c

ω

)2 ω2c ec(r)e∗c (r)

ω2c − ω2 − iωΓc

, (6.399)

where ωc is the cavity resonance frequency and Γc = ωcQ is the cavity linewidth.

Here we have digressed from Hughes (2005b), who has made some modification ofthe function.

The quantum dots are modelled as two-level atoms (Dung et al. 2002b). We intro-duce a quantum mechanical basis for the quantum dots a and b and the cavity modeas |A〉 ⊗ |B〉 ⊗ |k〉 = |ABk〉, where each variable can assume a value of 0 or 1. Weconcentrate on wave functions of the form

|ψ(t)〉e = Ca(t)|100〉 + Cb(t)|010〉 + Cp(t)|001〉, (6.400)

where Ca(t), Cb(t), and Cp(t) are complex amplitudes. We consider the reducedwave function of the form

Page 401: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 393

|ψ(t)〉 = 1√|Ca(t)|2 + |Cb(t)|2

[Ca(t)|10〉 + Cb(t)|01〉], (6.401)

where Ca(t) and Cb(t) obey integro-differential equations (Hughes 2005c). The timedependence of the entanglement (E(t)) is calculated for simple initial conditionsfrom the concurrence (C(t)) using (Hughes 2005c)

E(t) = −x log2(x)− (1− x) log2(1− x), (6.402)

where

x = 1

2+ 1

2

√1− C(t), C(t) = 4|Ca(t)|2|Cb(t)|2. (6.403)

As a continuation of the paper Sakoda and Haus (2003), Sibilia et al. (2005) havestudied the properties of super-radiant emission from a two-level atomic systemembedded in a one-dimensional photonic band gap structure. The description by areduced system of equations has further been reduced. Attention has been paid tothe Rabi splitting. The effect of the location of the atoms has been obtained using aclassical model of an amplifier.

In rotating Bose–Einstein condensates vortices develop. Mustecaplıoglu andOktel (2005) have shown that a vortex lattice can act as a photonic crystal andgenerate photonic band gaps. They have considered a two-dimensional triangularlattice. A numerical simulation of the propagation of an electromagnetic wave ina finite lattice has indicated that tens of vortices are enough for the infinite latticeproperties to occur. Those authors have proposed a method to measure the rotationfrequency of the condensate using a directional band gap.

6.4 Quantization in Disordered Media

Quantization of the electromagnetic field in disordered media may be realizedby any of the expounded or mentioned approaches. In particular it is likely thatan approach as that in Section 2.2.4 could be chosen. Many explanations fromSection 6.1 remain valid even on the assumption of a disordered medium.

With respect to application to random lasers, a great emphasis is laid on thenotion of a mode. As the electric permittivity is a scalar random field on the usualassumption of an isotropic dielectric, eigenfrequencies and modal functions of adevice are random. We form an idea on properties of the eigenenergies and modalfunctions by the condensed-matter theory and even by nuclear physics as well. Itmay lead to the restriction that the vectorial character is neglected in the modalfunctions. Certain models are mathematically very demanding to the contrary. Manystudies encounter the laser dynamics, which forces one to determine pseudomodes(after Garraway and Knight 1996), lossy modes, quasimodes otherwise).

Patra (2002), e.g., assumes a random medium closed in a cavity to be allowed tosuppose orthogonal modal functions. He neglects the vectorial character of modalfunctions. Whereas he may assume that, in an opening from the cavity, the value of

Page 402: Quantum Aspects of Light Propagation

394 6 Periodic and Disordered Media

a modal function has a Gaussian distribution, he encounters a difficulty in the wholecavity. He mentions that the modal functions are Gaussian random fields accordingto the literature, but he does not use, in fact, this assumption. Patra (2002) restrictshimself to values of a finite set of modal functions in a finite number of points andimplements orthogonality using columns of a random unitary matrix.

Loudon (1999) has expounded the polariton dispersion relation in the frameworkof the classical Lorentz theory. The linear response theory for a perfect crystal hasbeen generalized to include a randomly diluted crystal. The dependence of polaritonradiative damping rates on the occupation probability with admixture atoms hasbeen expressed.

The quantum, semiclassical, and classical theories of spontaneous emission havebeen characterized. The Glauber–Lewenstein model and more quantum theory havebeen applied to the radiative decay of dilute active atoms in lossy homogeneousand inhomogeneous dielectrics. The phase, group, and energy velocities for opticalpulse propagation through a lossy dielectric have been defined and their physicalmeanings have been illustrated.

6.4.1 Quantization in Chaotic Cavity

Recently (Patra 2002) a model of laser has been adopted that ignores the phase of thefield and, on including suitable Langevin terms like in Mishchenko and Beenakker(1999), provides the photon statistics. Perinova et al. (2004) hope that the calcu-lations can be refined, when the open systems theory is invoked (see, for examplePerinova and Luks (2000) and references therein).

An optical cavity is considered which is coupled to the environment by a smallopening of a diameter d . We concentrate on Np cavity eigenmodes described by thechaotic cavity modal functions Θi (r), each with an eigenfrequency ωi . The quantumdescription is reduced to only considering the number ni (t) of photons in each modei . Photons in mode i escape through the opening with rate γi .

The cavity is filled with an amplifying medium. The medium can be a four-levellaser dye, in which the lasing transition is directed from the third to the second level,the transition’s resonance frequency being Ω. In Patra (2002), the density of excitedatoms has been considered at every point r in the cavity. The coupling of mode i tothe medium at the point r has been given by Ki (r) = w(ωi )|Θi (r)|2, i = 1, . . . , Np,where w(ωi ) has been the transition matrix element of the atomic transition 3 → 2(de-excitation). At the level of a numerical solution, a linearization and a discretiza-tion have been performed. We discretize the space by introducing the Borel mea-surable neighbourhoods U (0 j ) of “uniformly” located centres 0 j , j = 1, . . . , Ns,which exhaust all the space and have the same volume ΔV each, NsΔV = V , thecavity volume. The description is reduced to only considering the density of excitedatoms N j in each neighbourhood U (0 j ). Excitations are created by pumping witha rate Pj and are lost nonradiatively with a rate a j . The coupling of mode i to themedium in the neighbourhood U (0 j ) is given by Ki j , Ki j ≡ Ki (0 j ). The original

Page 403: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 395

notation due to (Patra 2002) has been changed as follows: gi → γi , Ki j → Ki jΔV ,N j → N jΔV , Pj → PjΔV , a j → a jΔV .

We have utilized the annihilation (creation) operators ai (t) (a†i (t)), i = 1, . . . , Np,

which are assigned to the cavity modes and obey the commutation relations

[ai (t), a†i ′(t)] = δi i ′ 1, [ai (t), ai ′(t)] = 0. (6.404)

Using them, we can reinterpret the photon numbers ni (t) as the operators ni (t) =a†

i (t)ai (t).In fact, we still have to consider the matter field operators A j (t), A†

j (t), whichobey the commutation relations

[ A j (t), A†j ′(t)] =

1

ΔVδ j j ′ 1, [ A j (t), A j ′(t)] = 0. (6.405)

Using these operators, we can reinterpret the densities of excited atoms N j (t) as theoperators N j (t) = A†

j (t) A j (t).In analogy with exponential phase operators (Perinova et al. 1998)

exp[−iϕi (t)] = a†i (t)

[ni (t)+ 1

]− 12 ,

exp[iϕi (t)] =[ni (t)+ 1

]− 12 ai (t), (6.406)

we introduce the quantum phase operators

exp[−iΦ j (t)] = A†j (t)

[N j (t)+ 1

ΔV

]− 12

,

exp[iΦ j (t)] =[

N j (t)+ 1

ΔV

]− 12

A j (t). (6.407)

In this exposition, we restrict ourselves to the quantum description in the frame-work of the Schrodinger picture. The time dependence of the operators was relatedto the Heisenberg picture, is not present in the Schrodinger picture, and is dropped.We adopt the master equation approach, which concerns the temporal evolution ofthe statistical operator ρ(t) normalized such that

Tr{ρ(t)} = 1 (6.408)

and leads to the rate equations (6.418) straightforwardly. We propose that the masterequation has the form

∂tρ(t) = ˆLattρ(t)+ ˆLampρ(t)+ ˆLAttρ(t)+ ˆLnlnρ(t), (6.409)

where ˆLatt,ˆLamp,

ˆLAtt,ˆLnln are the Liouvillian superoperators with respective

properties

Page 404: Quantum Aspects of Light Propagation

396 6 Periodic and Disordered Media

ˆLattρ(t) =Np∑

i=1

γi

[ai ρ(t)a†

i −1

2ni ρ(t)− 1

2ρ(t)ni

], (6.410)

ˆLampρ(t) = ΔVNs∑

j=1

Pj[exp(−iΦ j )ρ(t)exp(iΦ j )− ρ(t)

], (6.411)

ˆLAttρ(t) = ΔVNs∑

j=1

a j

[A j ρ(t) A†

j −1

2N j ρ(t)− 1

2ρ(t)N j

], (6.412)

ˆLnlnρ(t) = ΔV

Np∑

i=1

Ns∑

j=1

Ki j

[a†

i A j ρ(t) A†j ai − 1

2(n + 1)N j ρ(t)

−1

2ρ(t)N j (n + 1)

]. (6.413)

The subscript “att” denotes the attenuation (escape of photons), the subscript “amp”denotes the amplification (pumping), the subscript “Att” another attenuation (relax-ation of the medium), and the subscript “nln” denotes a nonlinear process. The termˆLattρ(t) is well known from the quantum theory of damping at zero temperature

(Haken 1970). In analogy with it, the term ˆLAttρ(t) has been proposed. This isequivalent to treating the excited atoms as bosons. All the terms have the Lindblad

form Lindblad (1976). For instance, the term ˆLnlnρ(t) has such a form with the

Lindblad operators Oi jnln = a†i A j . The Lindblad form of the term ˆLampρ(t) has

been gained at the cost of considering the Lindblad operators O jamp = exp(−iΦ j ).The unusual operators may be related either to the discretization of the space or to atreatment of the excitations as bosons. To our knowledge, the phenomenology addedhas not been proved to be wrong. The master equation must be completed with aninitial condition that gives the statistical operator ρ(t0) which commutes with all theoperators ni , i = 1, . . . , Np, N j , j = 1, . . . , Ns.

It is natural to ask whether a limit of Ns → ∞ may be taken to absolve thequantum description from the parameters Ns and ΔV . This is not excluded, but theemerging model has too much in common with a quantum field theory. A need fora renormalization would bring us farther than an appropriate choice of ΔV .

We let |n1, . . . , nNp , N1, . . . , NNs〉 denote the (normalized) simultaneous eigen-kets of the operators ni , i = 1, . . . , Np, N j , j = 1, . . . , Ns. Using properties of theoperators which underlie to averaging, we obtain the rate equations for probabilities

p(n1, . . . , nNp , N1, . . . , NNs , t)

= 〈n1, . . . , nNp , N1, . . . , NNs |ρ(t)|n1, . . . , nNp , N1, . . . , NNs〉, (6.414)

normalized such that∞∑

n1=0

. . .

∞∑

nNp=0

∞∑′

N1=0

. . .

∞∑′

NNs=0

p(n1, . . . , nNp , N1, . . . , NNs , t) = 1, (6.415)

Page 405: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 397

where each prime means that the summation proceeds with the step 1ΔV . We get also

an initial condition related to the time t0.A derivation of rate equations is made easier when we write

ai (t) = exp[iϕi (t)][ni (t)]12 , A j (t) = exp[iΦ j (t)][N j (t)]

12 , (6.416)

a†i (t) A j (t) = exp[−iϕi (t)]exp[iΦ j (t)]

[ni (t)+ 1

] 12 N j (t)

12 . (6.417)

They have the form

∂tp(n1, . . . , nNp , N1, . . . , NNs , t) = Latt p(n1, . . . , nNp , N1, . . . , NNs , t)

+ Lamp p(n1, . . . , nNp , N1, . . . , NNs , t)

+ LAtt p(n1, . . . , nNp , N1, . . . , NNs , t)

+ Lnln p(n1, . . . , nNp , N1, . . . , NNs , t),(6.418)

where Latt, Lamp, LAtt, Lnln are operators with the respective properties

Latt p(n1, . . . , nNp , N1, . . . , NNs , t)

=Np∑

i=1

γi [(ni + 1)p(n1, . . . , ni + 1, . . . , nNp , N1, . . . , NNs , t)

− ni p(n1, . . . , nNp , N1, . . . , NNs , t)], (6.419)

Lamp p(n1, . . . , nNp , N1, . . . , NNs , t)

= ΔVNs∑

j=1

Pj

[p

(n1, . . . , nNp , N1, . . . , N j − 1

ΔV, . . . , NNs , t

)

− p(n1, . . . , nNp , N1, . . . , NNs , t)

], (6.420)

LAtt p(n1, . . . , nNp , N1, . . . , NNs , t)

= ΔVNs∑

j=1

a j

[(N j + 1

ΔV

)

× p

(n1, . . . , nNp , N1, . . . , N j + 1

ΔV, . . . , NNs , t

)

− N j p(n1, . . . , nNp , N1, . . . , NNs , t)

], (6.421)

Page 406: Quantum Aspects of Light Propagation

398 6 Periodic and Disordered Media

Lnln p(n1, . . . , nNp , N1, . . . , NNs , t)

= ΔV

Np∑

i=1

Ns∑

j=1

Ki j

[ni

(N j + 1

ΔV

)

× p

(n1, . . . , ni − 1, . . . , nNp , N1, . . . , N j + 1

ΔV, . . . , NNs , t

)

− (ni + 1)N j p(n1, . . . , nNp , N1, . . . , NNs , t)

]. (6.422)

The relation (6.419) indicates that a photon escapes from the i th mode with theprobability γi niΔt within a period of duration Δt . Similarly, the term (6.421) sug-gests that an excited atom near 0 j decays with the probability a j N jΔV Δt within aperiod of Δt . But the term (6.420) expresses that an excited atom near 0 j emergeswith probability PjΔV Δt within such a period. Finally, the term (6.422) tells thatan excited atom near 0 j emits a photon into the i th mode with the probabilityKi j (ni + 1)N jΔV Δt within such a period. The initial condition related to the timet0 gives the probabilities p(n1, . . ., nNp , N1, . . ., NNs , t0).

6.4.2 Open Systems Approach

The open systems approach enables us to “unravel” the master equation (6.409) with

respect to one, two, three, or all of the Liouvillian superoperators ˆLatt,ˆLamp, ˆLAtt,

ˆLnln. The unravelling is interesting in the steady state too. This state evolves froman old initial datum after a long time period since t0 ≤ 0 and it is considered to bepart of a new initial condition at the time t = 0.

The new description presents a stochastic process ρc(t) whose values are sta-

tistical operators ρc(t). Here the subscript c stands for condition, and it will beexplained in the following. We suppose that such a process has a Markovian prop-erty. The event of a continuous change or maybe conservation and the event of aninstantaneous discontinuous change of the statistical operator whose probability isasymptotically proportional to Δt for Δt → 0 are statistically independent of suchpast events. The continuous change could be described with a master equation

∂tρ

c(t) = ˆL∓,attρc

(t)+ ˆL∓,ampρc(t)+ ˆL∓,Attρc

(t)+ ˆL∓,nlnρc(t), (6.423)

where

ˆL−,att = ˆLatt,

ˆL+,attρc(t) =

Np∑

i=1

γi

[〈ni 〉c(t)ρ

c(t)− 1

2ni ρc

(t)− 1

c(t)ni

], (6.424)

Page 407: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 399

ˆL−,amp = ˆLatt,

ˆL+,ampρc(t) = ΔV

Ns∑

j=1

Pj

c(t)− ρ

c(t)]= 0, (6.425)

ˆL−,Att = ˆLAtt,

ˆL+,Attρc(t) = ΔV

Ns∑

j=1

a j

[〈N j 〉c(t)ρ

c(t)− 1

2N j ρc

(t)− 1

c(t)N j

], (6.426)

ˆL−,nln = ˆLnln,

ˆL+,nlnρc(t) = ΔV

Np∑

i=1

Ns∑

j=1

Ki j

[〈(n + 1)N j 〉c(t)a†

i A j ρc(t)

− 1

2(n + 1)N j ρc

(t)− 1

c(t)N j (n + 1)

]. (6.427)

Having introduced the subscripts ∓, we have alluded to the choice of 24 − 1 unrav-ellings, + means unravelled and − means unmodified.

After the discontinuous change, the new statistical operator

ρ ′c(t) = ai ρc

(t)a†i

〈ni 〉c(t),

exp(−iΦ j )ρc(t)exp(iΦ j )

1,

A j ρc(t) A†

j

〈N j 〉c(t),

a†i A j ρc

(t) A†j ai

〈(ni + 1)N j 〉c(t)(6.428)

and the factors of asymptotical proportionality are γi 〈ni 〉c(t), Pj ,a j 〈N j 〉c(t)ΔV , Ki j 〈(ni+ 1)N j 〉c(t)ΔV . We call them intensities. The discontinuous

changes and intensities are related to the components of superoperators ˆLatt (Np

components), ˆLamp (Ns components), ˆLAtt (Ns components), ˆLnln (Np Ns).The application of the new description is perspicuous and it does not contra-

dict a physical intuition when it is related only to components of the superopera-

tor ˆLatt. The new description becomes more transparent after introducing variablesm1(t), . . . , m Np

(t), which obey the initial condition m1(0) = . . . = m Np(0) = 0.

On the continuous change of the statistical operator ρc(t), all of these variables are

conserved. In the discontinuous change of the statistical operator related to the i th

component of ˆLatt, mi (t) is increased by unity. Even if from this it follows onlythat the variables m1(t), . . . , m Np

(t) do not burden the description, we state thatthe expression of ρ

c(t) in dependence on m1(t ′), . . . , m Np

(t ′), 0 ≤ t ′ ≤ t , and thetreatment of m1(t), . . . , m Np

(t) as classical stochastic processes with the intensi-ties, which are given as quantum averages, is lucid. Moreover, it is appropriate tothe physical intuition, when we identify mi (t) with numbers of photons (or photo-counts), which have been registered with a detector since the time t = 0.

Page 408: Quantum Aspects of Light Propagation

400 6 Periodic and Disordered Media

The unravelling can be most easily understood as the equality

ρ(t) = E[ρc(t)], (6.429)

where E stands for the expectation value. We introduce the notation

ρm1,...,m Np(t) = E[ρ

c(t)|m1(t) = m1, . . . , m Np

(t) = m Np ]

× p(m1, . . . , m Np , t), (6.430)

where E[ρc(t)|A] is the conditioned expectation value of the random operator ρ

c(t)

conditioned on the event A and

p(m1, . . . , m Np , t) = Pr[m1(t) = m1, . . . , m Np(t) = m Np ], (6.431)

with Pr denoting the probability. Invoking the probability theory, we note that

ρ(t) =∞∑

m1=0

. . .

∞∑

m Np=0

ρm1,...,m Np(t). (6.432)

On introducing the Hilbert space with a complete orthonormal basis |m1, . . . , m Np〉det

and considering this space in a tensor product with the original Hilbert space, wecan define a statistical operator

ρe(t) =∞∑

m1=0

. . .

∞∑

m Np=0

ρm1,...,m Np(t)

⊗ |m1, . . . , m Np〉det det〈m1, . . . , m Np |, (6.433)

where “e” means extended. Letting Trdet denote the partial trace over the extendingHilbert-space factor, Trdet ≡ Trd1 . . . TrdNp

, we see easily that

ρ(t) = Trdet ρe(t). (6.434)

The master equation has the form

∂tρe(t) = ˆLe,attρe(t)+ ˆLe,ampρe(t)+ ˆLe,Attρe(t)+ ˆLe,nlnρe(t), (6.435)

where

ˆLe,attρe(t) =Np∑

i=1

γi

[ai exp(−iθi )ρe(t)exp(iθi )a

†i −

1

2ni ρe(t)− 1

2ρe(t)ni

],

(6.436)

ˆLe,ampρe(t) = ˆLampρe(t),

ˆLe,Attρe(t) = ˆLAttρe(t),

ˆLe,nlnρe(t) = ˆLnlnρe(t), (6.437)

Page 409: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 401

exp(iθi ) = 1⊗∞∑

m1=0

. . .

∞∑

mi=0

. . .

∞∑

m Np=0

1 (6.438)

× |m1, . . . , mi , . . . , m Np〉det det〈m1, . . . , mi + 1, . . . , m Np |,

exp(−iθi ) =[exp(iθi )

]†,

ρe(0) = ρ(0)⊗ |0, . . . , 0〉det det〈0, . . . , 0|. (6.439)

Now we extend the joint probability distribution of photon numbers and the den-sities of excited atoms by numbers of emitted photons and introduce the probabi-lities

pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

= 〈n1, . . . , nNp , N1, . . . , NNs |ρm1,...,m Np(t)

× |n1, . . . , nNp , N1, . . . , NNs〉. (6.440)

The rate equations for these probabilities have the form

∂tpe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

= Le,att pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

+ Le,amp pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

+ Le,Att pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

+ Le,nln pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t), (6.441)

where

Le,att pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

=Np∑

i=1

γi [(ni + 1)pe(n1, . . . , ni + 1, . . . , nNp , N1, . . . , NNs ,

m1, . . . , mi − 1, . . . , m Np , t)

− ni pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)], (6.442)

Le,amp pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t) = ΔVNs∑

j=1

Pj

×[

pe

(n1, . . . , nNp , N1, . . . , N j − 1

ΔV, . . . , NNs , m1, . . . , m Np , t

)

− pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

], (6.443)

Page 410: Quantum Aspects of Light Propagation

402 6 Periodic and Disordered Media

Le,Att pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

= ΔVNs∑

j=1

a j

[(N j + 1

ΔV

)

× pe

(n1, . . . , nNp , N1, . . . , N j + 1

ΔV, . . . , NNs , m1, . . . , m Np , t

)

− N j pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

], (6.444)

Le,nln pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

= ΔV

Np∑

i=1

Ns∑

j=1

Ki j

[ni

(N j + 1

ΔV

)pe

(n1, . . . , ni − 1, . . . , nNp , N1, . . . , N j

+ 1

ΔV, . . . , NNs , m1, . . . , m Np , t

)

− (ni + 1)N j pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)

]. (6.445)

The initial condition related to the time origin is

pe(n1, . . . , nNp , N1, . . . , NNs , m1, . . . , m Np , t)∣∣t=0

= p(n1, . . . , nNp , N1, . . . , NNs , 0)δm10 . . . δm Np 0. (6.446)

Only the term (6.442) means some added phenomenology. It is assumed that anarray of ideal detectors is available and, whenever a photon escapes from the i thmode, it is absorbed by the i th detector.

By analogy with (6.438), we introduce the operators

mi = 1e ⊗∞∑

m1=0

. . .

∞∑

mi=0

. . .

∞∑

m Np=0

|m1, . . . , m Np〉det det〈m1, . . . , m Np |. (6.447)

We will introduce a shorthand notation

nk ≡ nk11 . . . n

kNp

Np,

N l ≡ N l11 . . . N

lNsNs

,

mr ≡ mr11 . . . m

rNp

Np. (6.448)

Considering the moments 〈nk N l mr 〉(t), we can rewrite equation (6.441) in the formof a hierarchy of equations

d

dt〈nk N l mr 〉(t) = 〈 ˆL†

e,attnk N l mr 〉(t)+ 〈 ˆL†

e,ampnk N l mr 〉(t)

+ 〈 ˆL†e,Attn

k N l mr 〉(t)+ 〈 ˆL†e,nlnnk N l mr 〉(t), (6.449)

Page 411: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 403

where

ˆL†e,attn

k N l mr = −Np∑

i=1

γi ni

[nk1

1 . . . nkNp

NpN l mr1

1 . . . mrNp

Np

− nk11 . . . (ni − 1e)ki . . . n

kNp

NpN l mr1

1 . . . (mi + 1e)ri . . . mrNp

Np

],

(6.450)

ˆL†e,ampnk N l mr = ΔV

Ns∑

j=1

Pj

[nk N l1

1 . . .

(N j + 1e

ΔV

)l j

. . . NlNsNs

mr

− nk N l11 . . . N

lNsNs

mr

], (6.451)

ˆL†e,Attn

k N l mr = −ΔVNs∑

j=1

a j N j

[nk N l1

1 . . . NlNsNs

mr

− nk N l11 . . .

(N j − 1e

ΔV

)l j

. . . NlNsNs

mr

], (6.452)

ˆL†e,nlnnk N l mr = ±ΔV

Np∑

i=1

Ns∑

j=1

Ki j (ni + 1e)N j

[± nk1

1

. . . (ni + 1e)ki . . . nkNp

NpN l1

1 . . .

(N j − 1e

ΔV

)l j

. . . NlNsNs

mr

∓ nk11 . . . n

kNp

NpN l1

1 . . . NlNsNs

mr

], (6.453)

with † denoting the Hermitian conjugation and 1e denoting the identity operator. Themoment equations (6.449) can be decoupled by various assumptions and approxi-mations. First, we observe that the equations for the means 〈ni 〉(t), 〈N j 〉(t) needthose for the second moments 〈ni N j 〉(t)

d

dt〈ni 〉(t) = −γi 〈ni 〉(t)+ΔV

Ns∑

j=1

Ki j 〈(ni + 1)N j 〉(t), (6.454)

d

dt〈N j 〉(t) = Pj − a j 〈N j 〉(t)−

Np∑

i=1

Ki j 〈(ni + 1)N j 〉(t). (6.455)

We neglect such a coupling by the factorizing approximation (Patra 2002, Rice andCarmichael 1994)

〈ni N j 〉(t) ≈ 〈ni 〉(t)〈N j 〉(t). (6.456)

Page 412: Quantum Aspects of Light Propagation

404 6 Periodic and Disordered Media

We introduce the shorthand ni (t) ≡ 〈ni 〉(t), N j (t) ≡ 〈N j 〉(t), and mi (t) ≡ 〈mi 〉(t),and, from now on, we consider the differential equations

d

dtni (t) = −γi ni (t)+ΔV

Ns∑

j=1

[ni (t)+ 1]Ki j N j (t), (6.457)

d

dtN j (t) = Pj − a j N j (t)−

Np∑

i=1

[ni (t)+ 1]Ki j N j (t), (6.458)

d

dtmi (t) = γi ni (t), (6.459)

and the appropriate initial conditions related to the time t0 and to the time origin,which give averages of the initial data of the unlinearized model if possible.

Further, we introduce the variations δni (t)= ni−ni (t)1e, δ N j (t)= N j−N j (t)1e,δmi (t) = mi − mi (t)1e, i = 1, . . . , Np, j = 1, . . . , Ns, of zero expectation val-ues. Assuming that the equations (6.457), (6.458), (6.459) have been solved for themeans ni (t), N j (t), mi (t), i = 1, . . . , Np, j = 1, . . . , Ns, and neglecting higher(third) moments, we derive approximate at most linear equations for the variances

〈[δni (t)]2〉(t), 〈[δ N j (t)]

2〉(t), 〈[δmi (t)]2〉(t) (6.460)

and the covariances

〈δni (t)δn j (t)〉(t), 〈δni (t)δ N j (t)〉(t), 〈δ N j (t)δ N j ′(t)〉(t),〈δni (t)δm j (t)〉(t), 〈δ N j (t)δmi (t)〉(t), 〈δmi (t)δmi ′(t)〉(t) (6.461)

from the hierarchy of moment equations. Also in the relations (6.460) and (6.461),we have written the argument t to the right of the angular brackets to indicate thatthe quantum description is still being carried out in the Schrodinger picture, eventhough some time dependence has been created by the subtracted means.

6.4.3 Semiclassical Approach

Semiclassical approach consists in the equations of motion for the photon numbersin the eigenmodes 1, . . . , Np and the densities of excited atoms in the neighbour-hoods U (0 j ), j = 1, . . . , Ns. These equations may be supplemented with those forthe number of counts taken by the detectors.

Stationary values of the means ni (t), N j (t) can be characteristic of stationaryprocesses ni (t), N j (t), which can be obtained in the limit t0 → −∞. We alsoobtain the time independence of the variances 〈[δni (t)]2〉(t), 〈[δN j (t)]2〉(t) and thecovariances 〈δni (t)δn j (t)〉, 〈δni (t)δN j (t)〉, 〈δN j (t)δN j ′(t)〉, and mi (t)=γi tni (0). Inthis case, the number of modes above the threshold Nl is interesting, mode i beingabove the threshold if and only if ni (t) ≥ 2.

Page 413: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 405

We will restrict ourselves to the Fano factor (Perina 1991) which can be calcu-lated as

Fsugg =

Np∑i=1

Np∑j=1〈δmi (T )δm j (T )〉

Np∑i=1

mi (T )

, (6.462)

where the subscript sugg stands for suggested, T is the time needed to explore theentire space inside the cavity, T = Ω2V 2

π2c3 is chosen as a detection time, and c is thespeed of light. As a short-T approximation, we obtain the usual formula for the Fanofactor, in fact

d

dT(Fsugg − 1)

∣∣∣∣T=0

= 1

T(Fcorr − 1), (6.463)

where Fcorr means the Fano factor, the subscript corr stands for correlated, which isgiven as

Fcorr =

Np∑i=1

Np∑j=1

TiT j 〈δni (0)δn j (0)〉Np∑

i=1Ti ni (0)

, (6.464)

with the transmission probability Ti = γi T , 0 ≤ Ti ≤ 1.With respect to reduced information on an individual random laser, the prob-

lem becomes a stochastic problem, an ensemble of cavities with small variationsin a shape or scatterer positions being considered. The coefficients γi and Ki j thusbecome random quantities. The coefficients γi are independent and identically dis-tributed, the probability density is

P({γi }) =Np∏

i=1

P(γi ), (6.465)

where

P(γi ) = 1√2πγiγ

e−γi2γ , (6.466)

which is the gamma or the Porter–Thomas distribution (Porter 1965) with the meanloss rate of a cavity γ = T

T , T being the mean transmissivity of a cavity, T =16π2d6Ω6

c6 . The expectation values γ i ≡ 〈γi 〉, i = 1, . . . , Np, are equal to the meanloss rate of the cavity, γ i = γ . Ki j , i = 1, . . . , Np, j = 1, . . . , Ns, are independentof γi , and they are distributed as squared moduli of elements of a random unitarymatrix. We hope to be consistent with Patra (2002), even after the correction of itsformula (16).

Page 414: Quantum Aspects of Light Propagation

406 6 Periodic and Disordered Media

As an illustration, a laser with a cavity supporting ten modes has been considered,where one mode is coupled out much less than the others, i.e., γ1 = 0.01 = g,γ2 = · · · = γ10 = 0.1, Ki j = 0.1 for all i , a1 = · · · = a10 = 1.

It can be accepted that some characteristics of a random laser fluctuate not verymuch about the ensemble means. Such characteristics are the scaled Fano factorF−1

g of the lasing mode (γi = g)

F − 1

g= T

( 〈(δni )2〉ni

− 1

)(6.467)

and the number of modes above threshold Nl. In computer simulations, the scaledFano factor may and may not fluctuate very much about the conditional mean〈F−1

g 〉|Nl when the probability of Nl, p(Nl), is small or not as can be seen fromFig. 6.4.

Fig. 6.4 The conditionalmean of the scaled Fanofactor in the lasing mode inthe dependence on thenumber of modes above thethreshold (curve a) and theprobability of the number ofmodes above the threshold(curve b) for cavities with afixed number Np = Ns modes

In this figure, the conditional mean increases in dependence on Nl (curve a). Theirregularity at 7–10 is likely due to the simulation. The expectation value of Nl isabout 3 (curve b). In the simulation Np = Ns = 10, γ = 0.125, and P1 = · · · =PNs = P , P = 30.

The simulation study of the conditional mean 〈F−1g 〉| g

γfor the lasing mode is

more difficult than the previous one, because gγ

is a continuous variable. Denotingthe probability density of g with P(1)(g), we note that g

γhas the probability density

γ P(1)(g). The calculated values of the conditional means are plotted in Fig. 6.5 asthe curve a and the probabilities

P (1)

(g

γ

)=∫ g+0.025 γ

gP(1)(g

′) dg′ (6.468)

are depicted as the curve b. The distribution of gγ

indicates that the conditional meanmay be calculated well only for the smallest values of g

γ, and it is estimated worse

otherwise.

Page 415: Quantum Aspects of Light Propagation

6.4 Quantization in Disordered Media 407

Fig. 6.5 The conditionalmean of the scaled Fanofactor in the lasing mode inthe dependence on g

γ(curve

a) and the probability of gγ

inbins of length of 0.025 (curveb) for cavities with a fixednumber Np = Ns modes

In Fig. 6.6 the scaled Fano factors Fcorr−1g and Fsugg−1

g are plotted (curves a, c, e and

b, d, f, respectively) in dependence on T ∈ [0, 5] for P = g, 100.2g, 100.4g (pairs(a, b), (c, d), (e, f)). The straight lines depicting Fcorr−1

g are tangent to the respective

curves Fsugg−1g at the origin.

Fig. 6.6 The scaled Fanofactors Fcorr−1

g and Fsugg−1g

(curves a, c, e and b, d, f,respectively) in thedependence on T ∈ [0, 5] forP = g, 100.2g, 100.4g (pairs(a, b), (c, d), (e, f)),respectively

Considering the master equation in the Lindblad form, we have derived rateequations for the probability distributions describing “classical” state of the ran-dom laser. From this, using standard approximations, we have rederived well-knownequations for the means and linear equations for the correlators. Both for traditionaland random lasers, the Fano factor has been proposed based on open systems theory.Comparison of this proposal with the usual Fano factor has been made in the tradi-tional laser. In the random laser, the scaled Fano factor in the lasing mode has beenaveraged in dependence on the number of modes above threshold and, alternatively,in dependence on the scaled loss rate.

Fu and Berman (2005) have complemented the Green function approach tothe spontaneous emission from an atom embedded inside a disordered dielectric(Fleischhauer 1999). The virtual cavity result has been reproduced using an ampli-tude approach which has been extended to second order.

Page 416: Quantum Aspects of Light Propagation

408 6 Periodic and Disordered Media

6.5 Propagation in Amplifying Random Media

The light propagation in disordered media is contiguous with the concept of lighttransport analogous to the electron transport in the condensed-matter physics, andits description is of importance for the experimental study of random lasers.

Quantum statistical properties of light are determined in disordered media.Models of a random laser with incoherent and coherent feedback are mentionedand it is stated that, in the framework of such models, photon statistics of opticalmodes were determined.

Both localization and laser theories which were developed in the 1960s havebeen jointly utilized in the study of a random laser. They have been used in studiesof strongly scattering gain media. Lasing in disordered media has been a subjectof intense theoretical and experimental studies. Random lasers have been classifiedinto incoherent and coherent random lasers. Research works on both types of ran-dom lasers have been surveyed in the monographic chapter (Cao 2003). In order toexplain quantum-statistical properties of random lasers, quantum theory is needed.Standard quantum theory for lasers applies only to quasidiscrete modes and cannotaccount for lasing in the presence of overlapping modes. In a random medium, thecharacter of lasing modes depends on the amount of disorder. Weak disorder leadsto a poor confinement of light and to strongly overlapping modes. Statistics natu-rally belongs to the theory of amplifying random media (Beenakker 1998, Patra andBeenakker 1999, 2000, Mishchenko et al. 2001), which is restricted to linear mediaand has not been used for the description of random lasers above lasing threshold(Cao 2003). The random laser model of Patra (2002, 2003), who calculated morethan only statistics of the photon number, has been completed (Luks and Perinova2003). In the framework of the open systems theory, the equations of motion involvethose for numbers of photons absorbed by detector. This extension corrects thephoton-number statistics. Hackenbroich et al. (2002) have developed a quantizationscheme for optical resonators with overlapping (nonorthogonal) modes. Cheng andSiegman (2003) have derived a generalized formalism of radiation-field quantiza-tion which need not rely on a set of orthogonal eigenmodes. True eigenmodes ofa system will be nonorthogonal and the method is intended for quantization of anopen system which contains a gain or loss medium.

6.5.1 Strongly Scattering Media

John (1984) has proposed a range of wavelengths or frequencies, in which electro-magnetic waves in a strongly scattering disordered medium undergo the Andersonlocalization (Anderson 1958, Abrahams et al. 1979). Although the derivation is con-ducted in d = 2 + ε dimensions, in consequence it holds that the photon mobilityedge ω∗ is as (ω∗l)2 � 1

2π(d = 3), where l is the photon elastic mean free path.

The range of wavelengths has the property λ ∼ l.In the case where λ � l, early experiments showed that the intensity of light

scattered from a concentrated suspension of latex microspheres in water presented

Page 417: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 409

a sharp peak centred at the backscattering direction (Kuga and Ishimaru 1984,van Albada and Lagendijk 1985, Wolf and Maret 1985). This peak is a coherenceeffect, which is present in disordered media. Akkermans et al. (1986) have analyzedthe multiple scattering of light to explain the peak line shape. The explanation isbased on the constructive interferences between time-reversed paths of light in asemi-infinite medium. The intensity reflected in directions distant by more than onedegree is almost constant and is incoherent. The analysis is not restricted to scalarwaves. It respects that polarization was analyzed parallel and perpendicular to theincident one.

John et al. (1996) have contributed to the field of optical tomography (see, e.g.,Huang et al. (1991) and references in John et al. (1996)). They assume a wave offrequency ω and velocity c. They recall that a simplified view of a photon as aquantum mechanical particle leads to the use of the Wigner coherence function,which is

I (R, k) =∫

exp(ik · r)⟨E∗(

R+ r2

)E(

R− r2

)⟩

ensembled3r. (6.469)

The authors let I0(R, k) denote the source coherence function. On choosing R′,k′, the source I0(R, k) = δ(R − R′)δ(k − k′) “radiates” a coherence functionΓ(R − R′; k, k′), which is called a propagator. The Wigner coherence function canbe measured (Raymer et al. 1994).

Here we must also refer to John and Stephen (1983) and McKintosh and John(1989) for reviews of the theory. In what follows we mention only the description ofa homogeneously disordered dielectric material. Its statistical properties are givenby the ensemble-averaged autocorrelation function

Bh(r− r′) = k40〈ε∗h (r)εh(r′)〉ensemble, (6.470)

where h stands for homogeneous and k0 = ωc . The Fourier transform is

Bh(q) =∫

exp(−iq · r)Bh(r) d3r. (6.471)

They define

Γh(Q; k′, k) = 1

k40

∫exp(−iQ · R)Γh(R; k′, k) d3R. (6.472)

In the field of optical tomography, conventional radiative transfer theory has beenapplied (Case and Zweifel 1976). In this theory, the (time-independent) specificlight intensity I c(R, k) (c means conventional) of a homogeneous medium withoutabsorption obeys the following phenomenological Boltzmann transport equation(Case and Zweifel 1976)

(k · ∇R)I c(R, k)+ kσ (k)I c(R, k)

= I c0 (R, k)+

∫k ′σ (k′ → k)I c(R, k′)

d3k′

(2π )3, (6.473)

Page 418: Quantum Aspects of Light Propagation

410 6 Periodic and Disordered Media

where σ (k) and σ (k′ → k) are the total and angular scattering coefficients, respec-tively, that satisfy the relation

∫σ (k′ → k)

d3k′

(2π )3= σ (k). (6.474)

In (6.473), I c0 (R, k) is the source specific intensity. Replacing k′ by k′′, I c

0 (R, k) by(2π )3δ(R−R′)δ(k−k′), and I c(R, k) by Γc

h(R−R′; k, k′), we obtain the equation forthe Green function for the specific light intensity. Its Fourier transform Γc

h(Q; k, k′)satisfies the equation

ik ·QΓch(Q; k, k′) = (2π )3δ(k− k′)

+∫

k1σ (k1 → k)Γch(Q; k1, k′)

d3k1

(2π )3− kσ (k)Γc

h(Q; k, k′).

(6.475)

From the optical coherence theory, it follows that

2k ·QΓh(Q; k, k′) = ΔGk(Q)(2π )3δ(k− k′)

+∫

ΔGk(Q)Bh(k− k1)Γh(Q; k1, k′)d3k1

(2π )3

−∫

ΔGk1 (Q)Bh(k− k1)d3k1

(2π )3Γh(Q; k, k′), (6.476)

where

ΔGk(Q) = G+(

k+ Q2

)− G−

(k− Q

2

),

G± (k) = 1

k20 − k2 −Σ±(k)

, (6.477)

with

Σ±(k) =∫

Bh(k− q)

k20 − q2 −Σ±(q)

d3q(2π )3

. (6.478)

A comparison of (6.475) with (6.476) may be made in a thorough and expert manner.Even though the coherent backscattering effects have not been included in John et al.(1996), they may be described using the results of MacKintosh and John (1989).

On considering multiple light scattering near an inhomogeneity, the analysis ofthe propagation and measurement of the Wigner distribution function may enhancethe resolving power of optical tomography. A formal analogy between wave prop-agation in a multiple scattering medium with a statistical inhomogeneity and thequantum mechanical scattering of a particle by a localized potential has beenexplored.

Page 419: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 411

Bruce and Chalker (1996) have generalized the treatment of quasi-one-dimensional systems due to Dorokhov (1982) and Mello et al. (1988b) for it toinclude absorption. Interestingly, they recall that it is not easy to obtain transmissionproperties. They consider a waveguide or optical fibre, along which N modes canpropagate in each direction.

Slightly deviating from them, we let a(L) and b(L) mean vectors of wave ampli-tudes (N×1 matrices) for the right-hand propagation and the left-hand propagation,respectively. Here L stands for a propagation distance. With respect to the disor-dered medium, the coupled modes are described by stochastic differential equations,which we write in the matrix form

(a(L + δl)b(L + δl)

)−(

a(L)b(L)

)

=[

(x y−y∗ −x∗

)− μ2a

(1 00 −1

)

− 1

2μ2

(x y−y∗ −x∗

)2](

a(L)b(L)

). (6.479)

Here μ = √δl, with δl > 0 an infinitesimal length, a parametrizes the strength of

absorption, 1 is an N × N unit matrix, x ≡ x[L ,L+δl] is a random N × N Hermitianmatrix, and y ≡ y[L ,L+δl] is a random N × N symmetric matrix. The two matricesx[L1,L2], y[L1,L2] are statistically independent of the two matrices x[L3,L4], y[L3,L4],when the intervals [L1, L2] and [L3, L4] do not overlap. Considering again x ≡x[L ,L+δl] and y ≡ y[L ,L+δl], the elements xαβ , yαβ are random variables, which arespecified in terms of the first and second moments,

〈xαβ〉 = 〈yαβ〉 = 0, 〈xαβ yγ δ〉 = 0, 〈yαβ yγ δ〉 = 0,

〈xαβ x∗γ δ〉 =δαγ δβδ

N, 〈yαβ y∗γ δ〉 =

δαγ δβδ + δαδδβγ

N + 1. (6.480)

A general solution of equations (6.479) has the form(

a(L)b(L)

)=(

vFFf(L|L0) vFBf(L|L0)vBFf(L|L0) vBBf(L|L0)

)(a(L0)b(L0)

), (6.481)

where L0, L0 ≤ L , is another propagation length and subscripts F and B meanforward (to the right) and backward (to the left), cf. Section 6.2.

Now we introduce the matrix(

τ ρ ′

ρ τ ′

), (6.482)

with

τ = vFFf(L1|L)− vFBf(L1|L)v−1BBf(L1|L)vBFf(L1|L),

ρ ′ = vFBf(L1|L)v−1BBf(L1|L),

Page 420: Quantum Aspects of Light Propagation

412 6 Periodic and Disordered Media

ρ = −v−1BBf(L1|L)vBFf(L1|L),

τ ′ = v−1BBf(L1|L), (6.483)

where L1 ≡ L + δl and a reflection matrix for waves incident from the right

r′(L) ≡ vFBf(L|0)v−1BBf(L|0). (6.484)

From the stochastic differential equation (6.479), it follows that(

τ ρ ′

ρ τ ′

)=(

1 00 1

)+ iμ

(x yy∗ x∗

)

− μ2a

(1 00 1

)− 1

2μ2

(x yy∗ x∗

)2

. (6.485)

The authors assume the increase of the system length L by δl. As the new reflec-tion matrix r′1 is given by a relatively simple relation

r′1 = ρ ′ + τ (r′ + r′ρr′ + . . .)τ ′, (6.486)

where r′1 ≡ r′(L1), r′ ≡ r′(L), the authors derive that

r′1 = r′ + iμ(y + xr′ + r′x∗ + r′y∗r′)

+ μ2

[−2ar′ − 1

2(yy∗ + x2)r′ − 1

2r′(y∗y + x∗2)− xr′x∗

]. (6.487)

On subtracting r′ from each side and dividing by μ2, we obtain a usual idea of astochastic differential equation. The mathematical notation does not require eventhe division by μ2.

It is relatively easy to derive the stochastic differential equation for Λ, the diag-onal matrix of eigenvalues Λα of the matrix r′†r′. The authors introduce λα asλα = Λα

1−Λα. The joint probability distribution of the set {λα}, W (λ1, λ2, . . . , λN , L),

evolves with the system length L according to the Fokker–Planck equation

∂W

∂L= 2

N + 1

N∑

α=1

∂λα

{λα(1+ λα)

×⎡

⎣∑

β,β �=α

1

λβ − λα

W + 2a(N + 1)W + ∂W

∂λα

⎦}. (6.488)

This equation has a stationary solution in the limit of long samples (L → ∞).Without absorption, this limit is trivial:

r ′mn =∑

k

UmkUnk, (6.489)

Page 421: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 413

where Umn are elements of the matrix U that has a uniform distribution on the unitarygroup U(N ). They present a result equivalent to relation (6.503) by Beenakker et al.(1996).

Brouwer and Beenakker (1996) have expounded a diagrammatic method foraveraging over the circular ensemble of random-matrix theory. The role of thecircular ensemble of unitary matrices in the scattering matrix approach has beenrespected. The method has been modified to the ensemble of uniformly distributedunitary symmetric matrices, which is referred to as the circular orthogonal ensem-ble. Even though these matrices are of the form U = VVT , with the matrix Vuniformly distributed over the unitary group, this efficient method is available.The results have been extended to unitary matrices of quaternions. Brouwer andBeenakker (1996) have applied the method to two types of mesoscopic systems inthe condensed-matter physics.

In the article (Beenakker 1997), the author concentrates himself on two types ofmesoscopic systems in the condensed-matter theory. In conclusions, he mentionsthat the propagation of electromagnetic waves through a waveguide is the opticalanalogue of conduction through a wire. In the problem of localization by disorder,the analogy is incomplete. The (relative) dielectric constant ε(x, y) not only fluctu-ates around unity but is always positive. Potentials V greater than the Fermi energyhave no optical analogue.

One new aspect of the optical problem is the behaviour in the case that the dielec-tric constant has a nonzero imaginary part. The intensity of radiation which haspropagated over a distance L is multiplied by a factor exp(σ L), with σ negative(positive) for absorption (amplification). Here the growth rate σ is related to thedielectric constant by the relation σ = −2k Im

√ε. The Dorokhov–Mello–Pereyra–

Kumar equation, which applies to the conduction through a wire, is accordinglygeneralized. Another new aspect of the optical problem is the frequency dependenceof the term

(ωc

)2ε (k2

0ε) in the Helmholtz equation, whereas an energy-dependentpotential does not occur in the mesoscopic systems.

van Rossum and Nieuwenhuizen (1999) have provided a discussion of the prop-agation of waves in random media. The description of radiation transport respectsthree length scales: macroscopic, mesoscopic, and microscopic. The diffusion the-ory presents the diffusion approximation at the macroscopic level. Important cor-rections are calculated with the radiative transfer equation, which describes intensitytransport at the mesoscopic level and is derived from the microscopic wave equation.A precise treatment of the diffuse intensity includes the effects of boundary lay-ers. Situations such as the enhanced backscattering cone and imaging of objects inopaque media are also discussed in the cited work. Mesoscopic correlations betweenmultiple scattered intensities are introduced. The correlation functions and intensitydistribution functions are derived.

Guo (2002) has modelled the propagation of a plane scalar wave through auniform dielectric slab using the multiple scattering method. This approach canbe used to model light propagation in stratified media, which represents alsoone-dimensional photonic crystals. The multiple scattering method can easily be

Page 422: Quantum Aspects of Light Propagation

414 6 Periodic and Disordered Media

generalized to treat light propagation in nonuniform media, such as light propaga-tion in random media.

The scalar approximation is based on the assumption that the transverse electricwaves are propagated. An incident plane wave, Einc(r) = eik0·r, with a wave vectork0, may be a component of an incident pulse. The corresponding total field E(r) maybe a component of the corresponding diffracted, transmitted, and reflected pulses.The total field E(r) obeys the following integral equation (in Gaussian units),

E(r) = Einc(r)+ 4πk20

σ

G0(r, r1)χ (r1)E(r1) d3r1, (6.490)

where

G0(r, r1) = 1

4π |r− r1|eik0|r−r1|. (6.491)

The subscript σ indicates exclusion of a small volume surrounding the position at r.Guo (2002) has assumed that the slab is formed by uniformly distributed dipoles.

He has concentrated on the resonant scattering case, where the complex electricpermittivity is pure imaginary and energy is lost.

In mathematics, random motions in random media are treated (Bolthausen andSznitman 2002). Lectures are restricted to discrete models. Then a one-dimensionalmodel of diffusion in a constant medium is the nearest neighbour random walk xn ,xn assuming integer values. Here are two ways to introduce the medium randomnessin a simple random walk.

(i) The probabilities of jumping to the right neighbour are chosen as independentidentically distributed random variables p(x), 0 ≤ p(x) ≤ 1, x being an integer.

(ii) The probabilities of jumping to the right neighbour are given by the relations

p(x) = cx,x+1

cx−1,x + cx,x+1, (6.492)

where cx,x+1 are independent identically distributed random variables,cx,x+1 > 0. In disordered media physics, rather, this model occurs (Bolthausenand Sznitman 2002, Hughes 1996).

6.5.2 Incoherent and Coherent Random Lasers

Gouedard et al. (1993) have developed mirrorless light sources based on heavilydoped neodymium materials pumped by nanosecond laser pulses. These sourcesgenerate quasimonochromatic short pulses and present characteristics of spatialand temporal incoherence. Such devices may find applications in domains such asholography, transport of energy in fibres for medical applications, and laser inertialconfinement fusion. A result of their study is the existence of a threshold belowwhich amplified spontaneous emission occurred in one of two compounds men-tioned by the authors. Above the threshold the specific behaviour has been found.This behaviour was reported by Ter-Gabrielyan et al. (1991).

Page 423: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 415

The authors have considered two schemes of the origin:

1. Many different microcrystallites are lasing sequentially in very short pulses.2. The grains of the powder emit collectively due to distributed feedback provided

by multiple scattering.

Scheme 2 may be related to the photon localization effect, but this has not beenproved or disproved in Gouedard et al. (1993). Second-harmonic generation instrongly scattering media has been investigated by Kravtsov et al. (1991).

Lawandy et al. (1994) have opined that composite systems may have spectraland temporal properties characteristic of a multimode laser oscillator, even thoughthese systems do not comprise an external cavity. They have investigated a laser dyedispersed in a strongly scattering medium. This medium was a colloidal suspensionof titanium dioxide particles. Colloidal solutions were optically pumped by linearlypolarized 532-nm radiation. Either single 7-ns-long pulses or a 125-ns-long train ofnine 80-ps-long pulses were used.

Most experiments were performed using the long pulses. The presence of theTiO2 nanoparticles led first to a larger emission linewidth, but when the energy ofthe excitation pulses was increased, the emission at 617 nm grew rapidly and theline narrowed. The emission at the peak wavelength was studied as a function ofthe pump pulse energy for four different TiO2 nanoparticle densities. A threshold ofthe pump energy has been observed. This threshold is obvious in plots of emissionlinewidth as a function of the pump pulse energy when the TiO2 nanoparticle con-centration is varied. Excitation with a train of 80-ps pulses demonstrated a thresh-old behaviour in the temporal characteristics of the colloid. When the pump pulseenergy was increased beyond the threshold, the response was a sharp peak.

The authors also concluded that a theory for this process did not exist. The liter-ature cited by them required that every dimension of the sample be large comparedto the optical scattering length.

Sha et al. (1994) performed similar experiments. Single 3-ns-long laser pulseswere used. A series of spectral experiments were carried out with a density of5 × 1011 cm−3 of TiO2 nanoparticles. When the pump pulse energy was varied,a threshold at 620 nm and a possible one at 650 nm were demonstrated.

The highest dye concentration exhibited a reduction of the lasing threshold, whenthe density of scattering particles was increased. For high particle density from5 × 1011 to 2.5 × 1012 cm−3 the lowest dye concentration revealed the thresholdof 0.17 mJ, higher than 0.07 mJ for this concentration in the neat solvent. Temporalcharacteristics of the response have been investigated as well. Above the threshold,the bandwidth of emission and the temporal width of the emitted pulses are nar-rowed. Also these authors stated that the exact mechanism for this process had notyet been explained.

Wiersma et al. (1995a) declared that Lawandy et al. (1994) had prepared onlyamplified spontaneous emission. For the pump geometry used the amplified spon-taneous emission proceeds perpendicular to the propagation direction of the pumppulse (side emission). It can be achieved that the amplification is negligible paral-lel to this direction. The paper under criticism does not record the side emission.

Page 424: Quantum Aspects of Light Propagation

416 6 Periodic and Disordered Media

Addition of the TiO2 particles brings about scattering or the detection of someamplified spontaneous emission light.

Lawandy and Balachandran (1995) have mainly presented data that clearly showthat for a fixed pump pulse energy there exists a threshold scatterer concentrationfor both side and front emissions. For the front emission it can be demonstrated thatthe threshold pump pulse energy decreases with increasing particle density.

Wiersma et al. (1995b) performed coherent backscattering measurements fromamplifying random media. They used optically pumped Ti:sapphire powders. Theyhave found that the light intensity as a function of the angular distance from theexact backscattering direction exhibits a top, which sharpens with increasing gain.They have solved the stationary diffusion equation with gain (Davison and Sykes1958, Letokhov 1968) and have used an approach by Akkermans et al. (1986) toobtain the coherent component of the backscattered intensity.

Beenakker et al. (1996) have recognized the notion of a random laser. Letokhov(1968) called it a “laser with incoherent feedback”, but randomness admits the “lastcoherence effect that survives” (Akkermans et al. 1986). It is appropriate when themodel includes an illuminated area, S. The authors associate the number of modesN � S

λ2 with it, where λ is a wavelength of the incident light. The authors were infact motivated by a paper of Pradhan and Kumar (1994) on the case N = 1 and havegeneralized it.

The reflection of a monochromatic plane wave (frequency ω, wavelength λ) bya slab (thickness L , transverse dimension W ) is considered, which represents a dis-ordered medium (mean free path l). This medium either amplifies or absorbs theradiation. For a statistical description an ensemble of slabs with different configura-tions of scatterers is considered. The authors let σ mean the amplification per unitlength. A negative value of σ indicates absorption. The parameter γ = σ l is theamplification per mean free path. The treatment is limited to scalar waves.

It is assumed that the slab is embedded in an optically passive waveguide withoutdisorder. We let N be the number of modes which can propagate in the waveguide atfrequency ω. The modal functions are normalized such that each mode carries unitpower. The N×N reflection matrix r has the elements rmn , rmn means the amplitudeof a wave reflected into mode m from an incident mode n.

The matrix r is symmetric, rT = r, and its singular-value decomposition is aproduct of U, the diagonal form of the matrix r, and UT . Here U is a unitary matrix,which has elements Umn (Mello et al. 1988a). It is specific that they assume thereflection eigenvalues to be nonnegative. They let

√Rn , n = 1, 2, . . . , N mean the

singular values. Then

rmn =∑

k

UmkU (∗)nk

√Rk . (6.493)

The difference between the symmetric matrix r and the Hermitian one r is signed bythe omission and the use of an asterisk, respectively.

The unit power of the mode n is amplified (or reduced) to the value

an =∑

m

|rmn|2. (6.494)

Page 425: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 417

The statistical calculation uses the assumption that W � L and U is uniformlydistributed in the unitary group. As a consequence, an has a distribution independentof n. Further it is assumed that λ � l, λ � 1

σ. On introducing

μn = 1

Rn − 1, (6.495)

the distribution P(μ1, μ2, . . . , μN , L) obeys the Fokker–Planck equation

l∂ P

∂L= 2

N + 1

N∑

i=1

∂μi

⎧⎨

⎩μi (1+ μi )

⎣∑

j, j �=i

1

μ j − μi+ γ (N + 1)

⎦ P

⎫⎬

+ 2

N + 1

N∑

i=1

∂μi

[μi (1+ μi )

∂ P

∂μi

], (6.496)

with the initial condition P|L=0 = N∏

i δ(μi + 1).On introducing the notation a ≡ 〈an〉, var a ≡ 〈(an − a)2〉, it can be derived that

ld

dLa = (a − 1)2 + 2γ a, (6.497)

ld

dLvar a = 4(a − 1+ γ ) var a + 2

Na(a − 1)2. (6.498)

Equation (6.498) has been derived for large N . Equation (6.497) is in agreementwith Selden (1974). The initial conditions for Equations (6.497) and (6.498) area(0)=0, var a(0)=0, respectively. In the case of absorption (γ < 0) and in the limitL →∞ the solution, which we do not quote here, yields

a∞ = 1− γ −√

γ 2 − 2γ , (6.499)

var a∞ = 1

2N

a∞(1− a∞)2

1− γ − a∞, (6.500)

which is a stationary solution of equations (6.497) and (6.498) obviously.In the case of amplification (γ > 0), the condition L < Lc must be fulfilled,

where Lc is a critical length,

Lc = larccos(γ − 1)√

2γ − γ 2. (6.501)

At this length a and var a diverge. It does not imply that a probability distributionof an independent of n, which has the density

P(a, L) =⟨δ

(a − 1−

k

UnkU ∗nk

1

μk

)⟩, (6.502)

does not exist for L ≥ Lc. It can be characterized by a modal value amax.

Page 426: Quantum Aspects of Light Propagation

418 6 Periodic and Disordered Media

The stationary solution of equation (6.496) is the Laguerre ensemble (Bronk1965)

P(μ1, μ2, . . . , μN ,∞) = C∏

i

exp[−γ (N + 1)μi ]∏

i< j

|μ j − μi |, (6.503)

where C is a normalization constant. The density

ρ(μ) =⟨∑

i

δ(μ− μi )

⟩(6.504)

is introduced, which in the large-N limit yields

ρ(μ) = N

π

√2γ

μ− γ 2, 0 < μ <

2

γ. (6.505)

The average in (6.502) consists in the average of Unk’s over the unitary group fol-lowed by the average of the μk’s over the Laguerre ensemble (6.503). The pertinentresult is known (Dyson 1962), even though only for large N , and it is an inverseLaplace transform. It has been found that the modal value of the distribution of an

independent of n is amax=1 + 0.8γ N . Predictions of random-matrix theory havebeen compared with numerical simulations of the analogous electronic Andersonmodel with a complex scattering potential.

John and Pang (1996) have determined the emission intensity properties of amodel dye system, which is immersed in a multiply scattering medium with trans-port mean free path l∗. Since they considered a rhodamine 640 dye solution basedon the literature, they respected the emission at 620 and 650 nm. They assumedthat the dye solution with scattering titanium particles fills the whole sample regionbetween the two planes z = 0 and z = L . They have used a generic dye laser scheme(Svelto and Hanna 1977, 1989) that explains bichromatic emission from the singletstates and the triplet states. The description comprises laser rate equations for thesinglet states and those for the triplet states. The intensity of the pumping beam is

of the form Iinc exp(−zlz

), where Iinc is the intensity at z = 0 and lz =√

l∗ la3 , with

la the absorption length. Those equations are completed with a diffusion equationfor the photon flux. A nonlinear diffusion equation for a dimensionless intensity ispresented, from which populations of dye molecules in singlet and triplet states areeliminated.

The emission spectra at nine different pump intensities agree with experiments(Sha et al. 1994). The linewidth and emission intensity at 620 nm as a function ofpump intensity for different values of transport mean free path l∗ are consistent withobservations (Balachandran and Lawandy 1995, Lawandy et al. 1994).

Wiersma and Lagendijk (1996) have confirmed that they set high standards uponrandom lasers. In their paper they have presented calculations on light diffusion withamplification and have reported on experiments.

In a random medium light is multiply scattered. The relevant length scales thatdescribe the scattering process are the scattering mean free path ls defined as the

Page 427: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 419

average distance between two scattering events and transport mean free path ldefined as the average distance a wave travels before its propagation direction ran-domizes. In an amplifying random medium it is necessary to define two more lengthscales: the gain length lg and amplification length lamp. The gain length is definedas the path length over which the intensity is multiplied by a factor e = exp(1).The amplification length is defined as the (rms) distance between the beginning

and ending points for paths of length lg, lamp =√

llg

3 . A sample of an amplifyingrandom medium in the form of a slab has been studied. Light and the amplify-ing medium become unstable if the thickness L is above the critical thickness Lcr,Lcr = πlamp.

Wiersma and Lagendijk (1996) have assumed the laser material to be a four-levelsystem. They have considered an incident pump and probe pulse and spontaneousemission. They have described their system with coupled differential equations. Theset is formed by three diffusion equations and the rate equation for the concentrationN1(r, t) of laser particles in a metastable state. The first two diffusion equations arelinear relative to the energy densities of the pump light and the probe light. Externalfields in those equations are the intensities of the incoming pump and probe pulses,respectively.

The boundary condition is the vanishing of the energy densities outside the slabat the distance z0≈0.7l from the surfaces of the slab. The authors have been ableto calculate the outgoing flux at either the front or rear surface of the slab, which isdetermined by the gradient of the energy density at the sample surface. Experimentshave demonstrated that it is possible to realize an amplifying random medium.

Wiersma and Lagendijk (1996) have distinguished three regimes depending onthe amount of scattering.

1. Weak scattering and gain. If l is of the order of the sample size, one says that thescattering is very weak. Addition of scatterers lifts a directionality in the outputof the system, cf. (Wiersma et al. 1995a). It is assumed that l continues to be ofthe order of the sample size.

2. Modest scattering and gain. If λ � l � L , one says that the scattering is tem-perate. The calculations have shown that modest scattering with gain can lead toa pulsed output, cf. (Gouedard et al. 1993).

3. Strong scattering and gain. If l is smaller than or equal to the wavelength λ, onesays that the scattering is strong. In this regime the Anderson localization of lightis expected to occur. Around 1996 there was no experimental evidence for thephotonic Anderson localization.

Paasschens et al. (1996) have studied the propagation of radiation through adisordered waveguide with a complex dielectric constant ε. They have called thesystems dual, which differ only in the sign of the imaginary part of ε. In the case ofthe scattering matrix

S =(

r t′

t r′

), (6.506)

Page 428: Quantum Aspects of Light Propagation

420 6 Periodic and Disordered Media

they have introduced transmittances T , T ′ and reflectances R, R′

T = 1

NTr{tt†}, R = 1

NTr{rr†},

T ′ = 1

NTr{t′t′†}, R′ = 1

NTr{r′r′†}. (6.507)

They have let Tn mean the eigenvalues of the matrix tt† (all of them depend on L)and have introduced their localization lengths ξn with the properties

1

ξn= − lim

L→∞1

Lln[Tn(L)]. (6.508)

The decay length ξ is defined as ξ = max(ξ1, ξ2, . . . , ξN ). For L →∞ it holds thatξ (σ ) = ξ (−σ ), where a dependence on σ is indicated, σ = −2k Im

√1+ i Im ε,

with k the free-space wave number of the radiation.In the case N = 1, the authors have presented also the Fokker–Planck equa-

tion for the joint probability distribution P(R, T, L) of the reflectance R and trans-mittance T . They have included the familiar relation for P(R, L = ∞) and haveintroduced γ = σ l.

Since 〈T 〉 diverges for γ > 0 at the lasing threshold Lc ≈ lc(γ )|γ | , where

c(γ ) = C+ (ln 2)γ − e2γ Ei(−2γ ), (6.509)

where C is Euler’s constant and Ei(x) = ∫ x−∞

et

t dt is the exponential integral, 〈ln T 〉(N ≥ 1) and var ln T (N = 1) are studied. The sum 〈ln T 〉+ L

ξ0, where ξ0 = (N + 1) l

2is the localization length for σ = 0, is mostly negative in consequence of the approx-imate equality 〈ln T 〉 + L

ξ≈ 0 and the observation that ξ ≤ ξ0.

Beenakker (1998) has generalized results concerning amplified spontaneousemission from a random medium for them to include thermal radiation. He hasapplied the method of random-matrix theory (Mehta 1991) to quantum optics. Hethinks of a linear amplifier as a system in thermal equilibrium at a negative temper-ature (Jeffers 1993, Matloob et al. 1997).

It is assumed that radiation, maybe no photons, comes into a random mediumvia an N -mode waveguide. The radiation is transformed in the random medium andgoes out from it via the same waveguide. Annihilation operators ˜ain

n (ω), ˜aoutn (ω),

˜bn(ω), ˜cn(ω), n = 1, 2, . . . , N (ω), are introduced. They satisfy the commutationrelations

[ ˜an(ω), ˜a†m(ω′)] = δnmδ(ω − ω′)1, [ ˜an(ω), ˜am(ω′)] = 0, (6.510)

for ˜a = ˜ain, ˜aout, ˜b, ˜c. The operators ˜ainn (ω) are related to the incoming modes and

the operators ˜aoutn (ω) describe the outgoing modes, ˜bn(ω) and ˜cn(ω) are quantum

noises for absorption and amplification, respectively.

Page 429: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 421

On introducing

˜ain(ω) =

⎜⎝

˜ain1 (ω)

...˜ain

N (ω)

⎟⎠ , ˜aout(ω) =

⎜⎝

˜aout1 (ω)

...˜aout

N (ω)

⎟⎠ ,

˜b(ω) =

⎜⎜⎝

˜b1(ω)...

˜bN (ω)

⎟⎟⎠ , ˜c†(ω) =

⎜⎝

˜c†1(ω)...

˜c†N (ω)

⎟⎠ , (6.511)

the input–output relations take the form

˜aout(ω) = S ˜ain(ω)+ Q ˜b(ω)+ V ˜c†(ω), (6.512)

where S, Q, V are matrices, which satisfy the conditions

QQ† − VV† = 1− SS†, (6.513)

where 1 is the unit matrix. Besides zero mean values, the quantum noises have theproperties

〈 ˜b†n(ω) ˜bm(ω′)〉 = δnmδ(ω − ω′) f (ω, T ),

〈 ˜c†n(ω)˜cm(ω′)〉 = δnmδ(ω − ω′) f (ω, T ), (6.514)

where T is the temperature and

f (ω, T ) = 1

exp(

�ωkBT

)− 1

. (6.515)

The negative temperature is obtained according to the relation

− [ f (ω, T )+ 1] = f (ω,−T ). (6.516)

The author also mentions the photodetection theory: The probability that n photonsare counted in a time t is given by the relation

p(n) = 1

n!〈: W n exp(−W ) :〉, (6.517)

where : : means the normal ordering of the operators inside and

W (t) =∫ t

0aout†(t ′)aout(t ′) dt ′, (6.518)

with

aout(t) = 1√2π

∫ ∞

0e−iωt ˜aout(ω) dω. (6.519)

Page 430: Quantum Aspects of Light Propagation

422 6 Periodic and Disordered Media

The generating function F(ξ ) of factorial cumulants

F(ξ ) = ln

[ ∞∑

n=0

(1+ ξ )n p(n)

]= ln〈: exp[ξ W (t)] :〉 (6.520)

is introduced. Here factorial means that the connection of usual cumulants withmoments of the photon–number distribution is applied to the factorial momentsof the distribution p(n) and W (t) stands for the integrated intensity. The factorialcumulants are

κp = dp F(ξ )

dξ p

∣∣∣∣ξ=0

, p = 1, 2, . . . ,∞. (6.521)

If ωct � 1, where ωc is the frequency interval within which SS† does not varysignificantly, then

F(ξ ) = −t∫ ∞

0ln{det[1− (1− SS†)ξ f (ω, T )]

} dω

2π, (6.522)

where det indicates the determinant. If Ωct � 1, where Ωc is the frequency rangeover which SS† differs appreciably from the unit matrix, then

F(ξ ) = − ln

{det

[1− t

∫ ∞

0(1− SS†)ξ f (ω, T )

]dω

}. (6.523)

The long-time limit depends only on the set of eigenvalues σ1, σ2, . . . , σN ofSS† (Beenakker 1998). These eigenvalues are called scattering strengths. As afrequency-resolved measurement leads to

F(ξ ) = − tδω

∞∑

n=1

ln [1− (1− σn)ξ f (ω, T )]

= − tδω

2πln{det[1− (1− SS†)ξ f (ω, T )

]}, (6.524)

where δω is a frequency interval, the factorial cumulants are

κp = (p − 1)!tδω

2π[ f (ω, T )]p

N∑

n=1

(1− σn)p, p = 1, 2, . . . ,∞. (6.525)

Particularly,

n = κ1 = tδω

2πf (ω, T )

N∑

n=1

(1− σn),

var n = κ2 = tδω

2π[ f (ω, T )]2

N∑

n=1

(1− σn)2. (6.526)

Page 431: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 423

But the blackbody radiation has

n = ν f, var n = ν f (ω, T )[1+ f (ω, T )], (6.527)

where

ν = Ntδω

2π. (6.528)

It has the property

var n = n + 1

νn2, (6.529)

whence it is advantageous to introduce

νeff = n2

var n − n, (6.530)

an effective number of degrees of freedom. Hence,

νeff = ν

[∑Nn=1(1− σn)

]2

N∑N

n=1

[(1− σn)2 − 1

f (ω,T )

∑Nn=1(1− σn)

] . (6.531)

As f (ω, T ) →∞ for T →∞, relation

νeff = ν

[∑Nn=1(1− σn)

]2

N∑N

n=1(1− σn)2(6.532)

is obtained for T →∞. Turning to applications, Beenakker (1998) concentrates onthe long-time regime with N � 1. For random media, a scattering-strength densityρ(σ ) is considered. Relations (6.525) and (6.532) are written using

N∑

n=1

M(σn) →∫

ρ(σ )M(σ ) dσ, (6.533)

where M(σn)=(1− σn)p.Beenakker (1998) presents further results for a semi-infinite random medium. Let

τs denote the inverse scattering rate and τa the inverse absorption or amplificationrate. Let us introduce γ = 16

3τsτa

. In the regime γ N 2 � 1,

ρ(σ ) = N

π

√γ

1

(1− σ )2

√1

σ− 1− γ

4, for 0 < σ <

1

1+ γ

2

, (6.534)

ρ(σ ) = 0 elsewhere. From this,

n = 1

2ν f (ω, T )γ

(√

1+ 4

γ− 1

). (6.535)

Beenakker (1998) has arrived at the following conclusions. For strong absorption,γ � 1, νeff = ν is obtained as for the blackbody radiation. For weak absorption,

Page 432: Quantum Aspects of Light Propagation

424 6 Periodic and Disordered Media

γ � 1, it is found that νeff = 2ν√

γ . It has been recognized that it is Glauber’sresult for the Lorentzian spectrum (Glauber 1963).

Specific results are added for an optical cavity coupled to a photodetector viaan N -mode waveguide. Here N � 1 is assumed and the modes overlap. Let usintroduce γ = τdwell

τa, where τdwell is the mean dwell time of a photon in the cavity

without absorption. In the limit of weak absorption, γ � 1,

ρ(σ ) = N

1

(1− σ )2

√(σ − σ−)(σ+ − σ ) for σ− < σ < σ+, (6.536)

where σ± = 1 − 3γ ± 2γ√

2. In the limit of strong absorption, γ � 1, relation(6.534) holds. From this,

n = ν fγ

1+ γ. (6.537)

It holds that νeff = ν for γ � 1. For γ � 1, we find that νeff = ν2 , which is finite

(nonvanishing) for γ → 0.The general formulae can also be applied to amplified spontaneous emission.

Beenakker (1998) investigates only the random laser below the laser threshold. Thesemi-infinite medium is above the laser threshold for an arbitrarily small amplifica-tion, but the cavity is below the threshold, as long as γ < 1.

Cao et al. (1999) observed random laser action with coherent feedback in semi-conductor powder. They found that the scattering mean free path is less than theemission wavelength. A comparison with the random laser theory (Wiersma andLagendijk 1996, Zhang 1995) was realized. The laser emission from the powdercould be observed in all directions. Their work is very different from the work onpowder laser (Markushev et al. 1986, Ter-Gabrielyan et al. 1991). When the particlesize was much larger than the wavelength, a single particle could serve as a res-onator. When the particle size was less than this wavelength, a single particle wastoo small to serve as a laser resonator. Laser resonators were formed by recurrentlight scattering.

Patra and Beenakker (1999) have continued the study of an amplifying disorderedcavity (Beenakker 1998). Besides the cavity they have investigated an amplifyingdisordered waveguide. The disordered medium is illuminated by monochromaticradiation of a single propagating mode in a coherent state.

First they consider an amplifying disordered medium embedded in a waveguidethat supports N (ω) propagating modes at frequency ω. The incoming radiation inmode n is described by an annihilation operator ˜ain

n (ω), where n = 1, 2, . . . , Nfor a mode on the left-hand side of the medium and n = N + 1, N + 2, . . . , 2Nfor a mode on the right-hand side. The outgoing radiation in mode n is describedby an annihilation operator ˜aout

n (ω), where again n = 1, 2, . . . , N for a mode onthe left-hand side of the medium and n = N + 1, N + 2, . . . , 2N for a mode onthe right-hand side. These two sets of operators are connected by the input–outputrelations

˜aout(ω) = S(ω)˜ain(ω)+ V(ω)˜c†(ω). (6.538)

Page 433: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 425

Here S(ω) is a 2N × 2N scattering matrix, V(ω) is a 2N × 2N matrix, and ˜c†(ω) isa vector of 2N creation operators ˜c†1(ω), ˜c†2(ω),. . . , ˜c†2N (ω). The scattering matrix Shas the form

S(ω) =(

r′ t′

t r

), (6.539)

where r ≡ r(ω) and r′ ≡ r′(ω) are N × N reflection matrices and t ≡ t(ω) andt′ ≡ t′(ω) are N×N transmission matrices. In the case under consideration t′ = tT ,r = rT , and r′ = r′T .

It is assumed that the operators ˜cn(ω) and ˜c†n(ω) commute with the operators˜am(ω) and ˜a†

m(ω). This implies that

V(ω)V†(ω) = S(ω)S†(ω)− 1. (6.540)

The state of the incoming radiation is

|ψ〉 =2N⊗

m=1m �=m0

|vac〉m ⊗ |ψ(ω)〉m0 , (6.541)

where |vac〉m is a vacuum state of mode m of the incoming radiation and |ψ(ω)〉m isa coherent state related to an unnormalized wave function ψ(ω), |ψ(ω)|2 → I0δ(ω−ω0). The coherent state is defined in terms of the usual unnormalized single-photonstates |ω〉n with the property

n〈ω|ω′〉m = δnmδ(ω − ω′). (6.542)

The counting of photons is studied using the integrated intensity

W (τ ) =∫ τ

0I (t) dt, (6.543)

where

I (t) = ηaout†(t)Paout(t), (6.544)

with η the efficiency of the photodetector,

P =(

0 00 1

), (6.545)

being a 2N × 2N matrix divided into four N × N matrices, and

aoutn (t) = 1√

∫ ∞

0e−iωt ˜aout

n (ω) dω. (6.546)

The generating function is

F(ξ ) = ln〈: exp(ξ W ) :〉. (6.547)

Page 434: Quantum Aspects of Light Propagation

426 6 Periodic and Disordered Media

The result for the detection in the long-time regime ωcτ � 1 is relatively simple. Itis found that

F(ξ ) = Fex(ξ )− τ

∫ ∞

0ln{det[1− ηξ f (ω, T )(1− rr† − tt†)]

}dω, (6.548)

where

Fex(ξ ) = ηξτ

{t†0

[1− ηξ f (ω0, T )(1− r0r

†0 − t0t

†0)]−1

t0

}

m0m0

, (6.549)

with {. . .}m0m0 denoting the matrix element located in m0th row and m0th column.In relation (6.549) t0 ≡ t(ω0), r0 ≡ r(ω0).

The general description may be applied also to an optical cavity filled with anamplifying random medium. It holds that t = 0 because there is no transmis-sion. Patra and Beenakker (1999) have studied how the noise figure F increaseson approaching the laser threshold. Near the laser threshold the noise figure has adivergent ensemble average. Its modal value is of the order of the number N ofpropagating modes in the medium.

The noise power is increased by

Pex = 2η2 f I0

[t†0(1− r0r

†0 − t0t

†0)t0

]

m0m0

. (6.550)

It has been found that Pex increases monotonically with increasing amplificationrate, but it has a maximum as a function of absorption rate for certain geometries.

Mishchenko and Beenakker (1999) say clearly that they borrow from the fieldof electronic conduction in disordered metals. Besides, they take into account theabsorption and emission of photons. They let fk(r, t) denote the density of photonnumber at the position r and such that

∫fk(r, t) d3r is the total photon number in

the mode k. The authors consider the random field fk(r, t) and its mean fk(r). Thesystem of wave vectors k is discrete. This has been adopted for ease of notation.

It is appropriate that this notion is independent of the time. But this does notfacilitate reading of a Boltzmann equation

cs · ∇r fk(r, t) = Ik(r, t), (6.551)

where s = k|k| and

Ik(r, t) =∑

k′

[Jkk′(r, t)− Jk′k(r, t)

]+ I+k (r, t), (6.552)

where

Jkk′(r, t) ≡ Jkk′(

fk(r, t), fk′ (r, t)),

Jk′k(r, t) ≡ Jk′k(

fk′(r, t), fk(r, t)). (6.553)

On writing it in the form

0 = −cs · ∇r fk(r, t)+ Ik(r, t), (6.554)

Page 435: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 427

we see that a continuity equation was first generalized

∂ fk(r, t)

∂t= −cs · ∇r fk(r, t)+ Ik(r, t). (6.555)

Here Ik(r, t) has to mean gain and loss terms.Let us note the form of the gain term due to amplification

I+k (r, t) = w+k

[1+ fk(r, t)

], (6.556)

with w+k the amplification rate. The unity enables a zero density to be amplified.

The gain term due to scattering from the mode with the wave vector k′,

Jkk′ (r, t) = wkk′ fk′ (r, t)[1+ fk(r, t)], (6.557)

is similar. Obviously, it is nonlinear in the described field. This does not mean thatEquation (6.551) is not linear. It holds that

Jkk′ (r, t)− Jk′k(r, t) = wkk′[ fk′(r, t)− fk(r, t)]

= wk′k[ fk′(r, t)− fk(r, t)]. (6.558)

Mishchenko and Beenakker (1999) continue the results such as those in Kogan(1996). Consistently with (6.551), the Boltzmann–Langevin equation for the ran-dom field itself is presented

cs · ∇r fk(r, t)+ ∂ fk(r, t)

∂t=∑

k′[Jkk′(r, t)− Jk′k(r, t)]

+ I+k (r, t)− I−k (r, t)+ Lk(r, t), (6.559)

where

Jkk′ (r, t) ≡ Jkk′(

fk(r, t), fk′ (r, t)),

Jk′k(r, t) ≡ Jk′k(

fk′ (r, t), fk(r, t)). (6.560)

Here I+k (r, t) is a gain term due to amplification,

I+k (r, t) = w+k [1+ Ik(r, t)] (6.561)

and I−k (r, t) is a loss term due to absorption,

I−k (r, t) = w−k fk(r, t), (6.562)

with w−k the absorption rate. In (6.559), Lk(r, t) is a Langevin term,

Lk(r, t) =∑

k′[δ Jkk′(r, t)− δ Jk′k(r, t)]+ δ I+k (r, t)− δ I−k (r, t), (6.563)

which is remarkable for copying the form of the previous terms. The elementarystochastic processes δ Jkk′ (r, t), δ I±k (r, t) have zero means,

δ Jkk′ (r, t) = 0, δ I+k (r, t) = 0, δ I−k (r, t) = 0, (6.564)

Page 436: Quantum Aspects of Light Propagation

428 6 Periodic and Disordered Media

and they have properties

δ Jkk′ (r, t)δ Jqq′(r′, t ′) = Δ(r, t, r′, t ′)δkqδk′q′ Jkk′(r, t),

δ I±k (r, t)δ I±k′ (r′, t ′) = Δ(r, t, r′, t ′)δkk′ I±k (r, t), (6.565)

where

Δ(r, t, r′, t ′) = δ(r− r′)δ(t − t ′). (6.566)

They are also characterized by the stochastic independence between δ Jkk′ (r, t) andδ I±q (r′, t ′) and by the same relationship between δ I+k (r, t) and δ I−q (r′, t ′).

The authors make a diffusion approximation, which is related to the expansionof the random field with respect to s. Then they consider the propagation throughan absorbing or amplifying disordered waveguide (of length L). The noise poweris decomposed into the fluctuations in the transmitted radiation, those in the ther-mal radiation, and the excess noise, which is characterized in Henry and Kazarinov(1996). The expressions for the thermal fluctuations and the excess noise agree withBeenakker (1999).

As a contribution the noise power of the thermal radiation emitted by a sphereis given (per unit surface area). Numerical results both for the waveguide geometryand for the sphere geometry are presented.

Beenakker (1999) has presented the statistics of thermal radiation in dependenceon the deviation 1− SS† from the unitarity of the scattering matrix S of the system.He has recovered the familiar results for black-body radiation in the limit S → 0.A simple expression for the mean photocount has been identified as Kirchhoff’slaw. A generalization of the Kirchhoff law has been derived. For the extension ofthe Kirchhoff law to the statistics of quanta, which exists in the case of single-modedetection, reference is made to (Bekenstein and Schiffer 1994). Due to a similarity,the theory has been easily applied to a random amplifying medium (or a “randomlaser”) below the laser threshold.

Zacharakis et al. (2000) measured photon–number distributions of fluorescenceof an organic dye. The dye was mixed with poly(methyl methacrylate), which fixedscatterers. They were able to measure the photon–number distributions for differenttime delays and at different wavelengths. The source of excitations was a frequency-doubled 200-fs pulsed laser emitting at 400 nm.

The photon-number distribution from the sample when it was pumped abovethreshold had different character for different time delays. For a small time delaythis distribution was Poisson-shaped with an imperfect vacuum value. When thetime delay increased, a Bose–Einstein distribution was appropriate. In contrast tothe high-energy case, when the excitation energy is below threshold, the photonnumber has the Bose–Einstein distribution, which is independent of the time delay.

In Patra and Beenakker (2000), a continuation of Patra and Beenakker (1999)and Beenakker (1998) is contained. The treatment includes a noise characteris-tic averaged over an ensemble of random media with different positions of thescatterers. The authors assume a waveguide with N (ω) propagating modes at fre-quency ω. Modes 1, 2, . . . , N are on the left-hand side of the medium and modes

Page 437: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 429

N + 1, N + 2, . . . , 2N are on the right-hand side. The outgoing radiation in mode nis described by an annihilation operator ˜aout

n (ω). A vector ˜aout(ω) is defined. Similarnotation is introduced for incoming radiation.

The operators, which belong to the same stage, fulfil the commutation relationsbetween the annihilation and creation ones,

[ ˜an(ω), ˜a†m(ω′)] = δnmδ(ω − ω′)1, [ ˜an(ω), ˜am(ω′)] = 0, (6.567)

where ˜a= ˜ain, ˜aout.In the input–output relations, also the vectors b and c occur, each of which has 2N

annihilation operators for its elements. Their correlation functions are determineddependent on the temperature of the medium.

The usual quantization of discrete frequencies is obtained by considering thosein the relations for a frequency step Δ, the frequencies ωp = pΔ, and subscripts asthose in the relations

˜aoutnp =

1√Δ

∫ ωp+1

ωp

˜aoutn (ω) dω,

Snp,n′ p′ = Snn′ (ωp)δpp′ . (6.568)

Patra and Beenakker (2000) have considered a useful modification

˜a′outnp = 1√

Δ˜a

outnp . (6.569)

We have used the prime to distinguish this modification from the usual annihilationoperator.

A characteristic function is

χ (β,Δ) =⟨: exp

12

(˜a′†β − β† ˜a

′)]:⟩. (6.570)

The vector β has the elements βnp = βn(ωp).The statistical properties of the bath are

χabs(β,Δ) = exp(−β†fβ) (6.571)

for an absorbing medium and

χamp(β,Δ) = exp(β†fβ) (6.572)

for an amplifying medium. In these relations f means a matrix with the elementsfnp,n′ p′ = δnn′δpp′ f (ωp, T ), where f (ωp, T ) is given in (6.515). The characteristicfunction of the outgoing state is

χout(β,Δ) = exp[−β†(1− SS†)fβ

]χin(S†β,Δ). (6.573)

The photocount distribution is the probability P(n, τ ) that n photons are absorbedby a photodetector within a time τ . The appropriate generating function is denotedby F(ξ, τ ). The integrated intensity W (τ ) is defined by the relation

Page 438: Quantum Aspects of Light Propagation

430 6 Periodic and Disordered Media

W (τ ) =∫ τ

0

2N∑

n=1

ηnaout†n (t)aout

n (t) dt. (6.574)

Here ηn ∈ [0, 1] is the detection efficiency of the nth mode. We let η denote a2N × 2N diagonal matrix containing the detection efficiencies ηn on the diag-onal (ηnm = ηnδnm). The discretization of frequencies will lead to the integratedintensity W (τ,Δ). It can be written using the matrix η = η ⊗ 1frequencies, with theelements ηnp,n′ p′=ηnn′δpp′ . Here 1frequencies is the unit matrix with the elements δpp′ .The respective generating function is denoted by F(ξ, τ, Δ).

The generating function F(ξ, τ,Δ) may be determined fromexp [F(ξ, τ, Δ)], which in turn is a linear integral transform of the characteristicfunction χout(β,Δ). On respecting the input–output relation (6.573), we obtain therelation

exp [F(ξ, τ, Δ)] = 1

det(−ξπη)

×∫

χin(S†β,Δ) exp

[1

ξβ†(η)−1β − β†(1− SS†)fβ

]dβ,

(6.575)

where Δ = 2πτ

is chosen. The thermal fluctuations can be separated and we obtainthe relations

F(ξ, τ,Δ) = Fth(ξ, τ,Δ)

+ ln

[1

det(πM)

∫χin(β,Δ) exp(−β†M−1β) dβ

], (6.576)

Fth(ξ, τ, Δ) = − ln{det[1− ξ η(1− SS†)f]}, (6.577)

and

M = −ξS†[1− ξ η(1− SS†)f]−1ηS. (6.578)

Returning to the continuous frequency, relation (6.577) can be written as

Fth(ξ, τ ) = − τ

∫ ∞

0ln{

det(1− ξη[1− S(ω)S†(ω)] f (ω, T )

)}dω, (6.579)

where f (ω, T ) is given in (6.515). It is stated that all factorial cumulants dependlinearly on the detection time in the long-time limit. The notion of detection effi-ciency may be utilized also to treatment of particular cases, such as the detection atone side of the waveguide (Beenakker 1999).

Patra and Beenakker (2000) assume that the incident radiation is in the idealsqueezed state |ζ, α〉 = C S|0〉, where

S = exp

[1

2Δ(a

′in T ζ † ˜a′in − ˜a

′in †ζ ˜a

′in ∗)

], (6.580)

Page 439: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 431

is the squeezing operator and

C = exp[Δ

12 ( ˜a

′in †α − α† ˜a

′in)]

(6.581)

is the displacement operator.In relation (6.580), T means the transposition, ζ is the diagonal matrix with

the elements ζnp,n′ p′ = ζn(ωp)δnn′δpp′ , and α is the vector with the elementsαnp = αn(ωp). Useful are the real parameters ρn(ωp), φn(ωp) such that ζn(ω) =ρn(ω) exp (iφn(ω)). Here ˜a

′in †np is the column with the elements ˜a

′innp.

The characteristic function of the incident radiation in the case where this radia-tion is in the ideal squeezed state is

χin(β,Δ) = exp{α†β − β†α − 1

4βT [e−iφ sinh(2ρ)]β

− 1

4β†[eiφ sinh(2ρ)]β∗ − β†(sinh ρ)2β

}. (6.582)

On substitution into relation (6.573), we obtain that the characteristic function ofthe outgoing radiation is

χout(β,Δ) = exp{α†S†β − β†Sα − 1

4βT S∗[e−iφ sinh(2ρ)]S†β

− 1

4β†S[eiφ sinh(2ρ)]ST β∗

− β†[f − S(f − (sinh ρ)2

)S†]β

}. (6.583)

Using an integral transformation, we obtain the generating function F(ξ, τ, Δ),

F(ξ, τ, Δ) = Fth(ξ, τ,Δ)− 1

2ln(det X)− 1

2

(α∗

α

)T

X−1

(Mα

M∗α∗

), (6.584)

where the matrix X is defined in terms of the matrix M,

X = 1+(

M sinh ρ −Meiφ cosh ρ

−M∗e−iφ cosh ρ M∗ sinh ρ

)(sinh ρ 0

0 sinh ρ

). (6.585)

Further Patra and Beenakker (2000) consider the case, where only the mode m0 issqueezed.

The Fano factor is the ratio

F = P

I, (6.586)

where P = 1τ

(κ2 + κ1) is the noise power and I = 1τκ1

is the mean current. Forsimplicity this characteristic is considered in the limit of τ → ∞. The detectionefficiency does not depend on the mode. The thermal contributions may be neglectedin the considered case, since they are spread out over a wide range of frequencies.

Page 440: Quantum Aspects of Light Propagation

432 6 Periodic and Disordered Media

In the exposition, η ceases to be a matrix and it means the common value of thedetector efficiencies ηn . For (6.582), one has

Fin = 1+ |α cosh ρ − α∗eiφ sinh ρ|2 − |α|2 + 12 (sinh ρ)2 cosh(2ρ)

|α|2 + (sinh ρ)2. (6.587)

Patra and Beenakker (2000) consider the Fano factor both for the direct detection(Fdirect) and for the homodyne detection (Fhomo). In the case of direct detection, theauthors find that

Fdirect − 1 = η(t†0t0)m0m0 (Fin − 1)

+ 2η f (ω0, T )[t†0(1− r0r

†0 − t0t

†0)t0]m0m0

(t†0t0)m0m0

. (6.588)

The Fano factors for the direct and homodyne detections depend on the reflectionand transmission matrices of the waveguide. These matrices depend on the positionsof the scatterers inside the waveguide. The distribution of these matrices can bechosen by random-matrix theory (Beenakker 1997). The details may be found inBrouwer (1998). It is examined how the distribution depends on the mean free path land the amplification (absorption) length ξa =

√Dτa, where D = cl

3 is the diffusionconstant and 1

τais the amplification (absorption) rate.

On neglect of the correlation between numerator and denominator for an absorb-ing disordered waveguide it is obtained that

Fdirect = 1+ 4lη

3ξa sinh s(Fin − 1)

+ η

2f (ω0, T )

[3− 2s + coth s

sinh s− s coth s − 1

(sinh s)2+ s

(sinh s)3

]. (6.589)

Here s ≡ Lξa

. In the limit of strong absorption, the Fano factor Fdirect = 1 +3

2ηf (ω0, T ). The Fano factor Fin may be given by equation (6.587), but Equation

(6.589) is valid even for any state of the incident radiation.For an amplifying disordered waveguide it is found that

Fdirect = 1+ 4lη

3ξa sin s(Fin − 1)

+ η

2f (ω0, T )

[3− 2s − cot s

sin s+ s cot s − 1

(sin s)2− s

(sin s)3

]. (6.590)

The laser threshold is s = π . In Patra and Beenakker (2000) also the average Fanofactors for the homodyne detection are presented. The effect of absorption on quasi-modes of a random waveguide has been studied in Sebbah et al. (2007).

Cao et al. (2000) have mentioned the concept of the Anderson localization.Optical absorption counteracts photon localization. Also optical gain reduces photonlocalization length in a one-dimensional random medium. After the previous exper-iment on coherent feedback for lasing, the authors have arrived at enhancement of

Page 441: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 433

the scattering strength and at spatial confinement of laser light in the disorderedmedium. They opine that the optical gain enhances the photon localization at leastin a three-dimensional medium. The coherent amplification of the scattered lightenhances the interference effect and helps the spatial confinement.

ZnO particles of the average size about 50 nm and a ZnO powder film of thick-ness about 30 μm were prepared. The scattering mean free path l in the ZnO powderhas been estimated l ∼ 0.5λ. The ZnO powder film is photoluminescent, when it ispumped at 266 nm. The pump beam falls in the normal direction on an about 20 μmspot of the film. The spectrum of emission from the powder film is measured. Atthe same time, the spatial distribution of the emitted light intensity is imaged in theultraviolet.

Figure 1, which we do not reproduce here, presents the measured spectra andspatial distribution of emission in a ZnO powder film at two different pump powers.For the lower pump level, the spectrum consists of a single broad spontaneous emis-sion peak. The spatial distribution of the spontaneous emission intensity is almostuniform across the excitation area. It depends on the pump intensity distribution.

For the higher pump level, when the pump intensity exceeds a threshold, sharppeaks emerge in the emission spectrum. Bright tiny spots appear in the spatial distri-bution of the emission. When the pump intensity increases further, additional sharppeaks emerge in the emission spectrum. Also more bright spots appear in the emittedlight pattern. Above the threshold, the total emission intensity begins to approachthe pump power.

The fact that the bright spots in the emission pattern and the lasing modes in theemission spectrum always occur simultaneously could mean that the bright spots areefficient scatterers. Then the bright spots should scatter the spontaneously emittedlight below the lasing threshold. As the bright spots do not exist below the lasingthreshold, the laser light intensity is high at their locations.

The authors have measured the short scattering mean free path, i.e. very stronglight scattering on average. They expect small regions of higher disorder andstronger scattering. In other words, they assume many resonant cavities formed bymultiple scattering and interference. Every cavity has its lasing threshold. The lasingpeaks in the emission spectrum illustrate cavity resonant frequencies and the brightspots in the spatial light pattern give positions and shapes of the cavities.

To verify this hypothesis, the authors have reduced the size of the randommedium to a cluster of ZnO nanoparticles. The existence of the lasing thresholdis related to nonradiative and radiative recombination of the excited carriers. Theauthors have calculated the electromagnetic-field distribution in a random mediumusing the finite-difference time-domain method according to the book (Taflove1995). In the calculations the assumption that the ZnO particles are surrounded byair has been used. Optical gain has been introduced to the Maxwell equations by thenegative conductance σ .

The simulation has shown that, when the optical gain is just above the lasingthreshold, the emission spectrum consists of a single peak. When the optical gainincreases further, additional lasing modes appear. The authors have concluded thatoptical gain helps spatial confinement of light in a random medium.

Page 442: Quantum Aspects of Light Propagation

434 6 Periodic and Disordered Media

The authors have paid more attention to the Anderson localization. In spite ofthe fact that there is no criterion for the Anderson localization in an active randommedium, they have determined the Thouless number δ = 0.75 < 1 in favour of theAnderson localization in the lasing mode.

Thareja and Mitra (2000) have reported on an experiment on optically pumpedZnO powder. In this medium a random laser has been demonstrated. The theoreticexplanation of this effect has been based on the paper (Cao et al. 1999).

Jiang and Soukoulis (2000) emphasize that, in contrast to the paper (Lawandyet al. 1994), where discrete lasing peaks were not observed, a new interesting prop-erty was reported, e.g. in Cao et al. (1999).

The authors provide references, e.g. in (Zyuzin 1995, John and Pang 1996), butthey have pointed out a limitation of the diffusion approach. They see a limitation,though mild, also in the approach as in (Paasschens et al. 1996, Jiang and Soukoulis1999, Jiang et al. 1999). Essentially, they return to the semiclassical laser theory,e.g., (Siegman 1986).

The authors have combined the equations for electron densities with Maxwell’sequations and have used the finite-difference time-domain method according tothe book (Taflove 1995). After a long relaxation time stationary solutions can beobtained. The time dependence of the electric field inside the system and in itsvicinity is examined. The emission spectra and the modes inside the system canbe obtained after the Fourier transformation.

The system is a one-dimensional simplification of the reported experiments. Itconsists of many dielectric layers of fixed thickness bestowed between two surfaces,with the space among the dielectric layers filled with a gain medium. The distancebetween the neighbouring dielectric layers is assumed to be a random variable. Thetotal length of the system is L .

The numerical simulations have shown the following:

(i) In periodic and short (L < ξ ) random systems an extended mode dominates inthe field and the spectrum.

(ii) For either strong disorder or a long (L � ξ ) system a low threshold value forlasing is obtained. By increasing the length or the gain more peaks appear inthe spectrum. The peaks are coming from localized modes.

(iii) The saturation can be observed. The number of the peaks is proportional to thelength of the system.

(iv) The emission spectra are not the same for various output directions. In thethree-dimensional case lasing peaks need not be so sharp as in the one-dimensional case.

The alternating layers are made of dielectric materials with dielectric constantsε1 = ε0 and ε2 = 4ε0. The thickness of the first layer, which simulates the gainmedium, is a random variable an = a0(1 + Wγ ), where a0 = 300 nm, W is thestrength of randomness, and γ is a random value in the range [−0.5, 0.5]. The thick-ness of the second layer, which simulates the scatterers, is a constant b = 180 nm.In the layers, which represent the gain medium, a four-level electronic material

Page 443: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 435

is admixed. The electron densities at a ground, first, second, and third levels areN0(x, t),. . . ,N3(x, t), respectively. An external mechanism pumps electrons fromthe ground level to the third one at a certain pumping rate Pr. Nonradiative transi-tions occur from the higher level to the lower one with the lifetimes of the upperlevels τ32, τ21, τ10. The radiative transition from the second level to the first one, orback has the centre frequency ωa.

According to the monograph (Siegman 1986), the polarization density P(x, t)depends nonlinearly on the population inversion ΔN (x, t) = N1(x, t) − N2(x, t)and on the electric field E(x, t). An equation for the polarization density P(x, t)can be written. Equations for electron densities at every level can be utilized.

One must introduce sources into the system. The distance between the twosources Ls must be smaller than the localization length ξ . The sources simulate thespontaneous emission. They have a Lorentzian spectrum centred around ωa and theiramplitudes depend on N2. Two leads are assumed at the left-hand and right-handsides of the system. A numerical method for solving the mentioned equations withan absorbing-boundary condition is described.

Jiang and Soukoulis (2000) have performed the numerical simulations for peri-odic and random systems. They associate a lasing threshold with each of the sys-tems. With the increase of the randomness, the threshold intensity decreases. It hasbeen found that, in the case of a periodic system and a short (L < ξ ) random system,one mode dominates, even if the gain increases far above the threshold. In the caseof a long (L � ξ ) random system, the stationary behaviour is marked with beats.There are more than one localized mode and each one has its specific frequency.In a figure, which is not reproduced here, it is shown how, for the pumping ratesPr = 104, 106, 1010 s−1, one lasing mode appears and then more lasing modes.So more than one mode can exist together and each mode seems to repel othersto reserve itself some space. There exists a saturated number of lasing modes Nm,which is proportional to the length of the system L . There exists an average modelength Lm = L

Nm, which is proportional to the localization length. The emission

spectra at the right-hand and left-hand sides of the system are different. In the realthree-dimensional experiments, Jiang and Soukoulis (2000) assume that every local-ized mode has its direction, strength, and position.

Cao et al. (2001) have measured the photon statistics of random lasers withresonant feedback. They have found that, when the pump intensity increases, thephoton–number distribution changes continuously from the Bose–Einstein distri-bution at the threshold to the Poisson distribution well above the threshold. Thenormalized second factorial moment G2 = 〈n(n−1)〉

〈n〉2 decreases correspondingly from2 to 1. By comparing the photon statistics of a random laser with resonant feedbackand this statistics of a random laser with nonresonant feedback, the authors haveformed the idea about two lasing mechanisms.

For a random laser with nonresonant feedback, the fluctuation of the total numberof photons in all modes of laser emission is smaller than the fluctuation of this num-ber in blackbody radiation with the same number of modes (Zacharakis et al. 2000).However, the photon–number distribution in a single mode remains the Bose–Einstein distribution even well above the threshold. The quasimodes correspond to

Page 444: Quantum Aspects of Light Propagation

436 6 Periodic and Disordered Media

eigenfrequencies, whose imaginary parts represent the decay rates, in fact they arepseudomodes. When kl > 1 (k is a wave number and l is the transport mean freepath), the quasimodes overlap spectrally and the emission spectrum is continuous.

In other words, in the case of weak scattering, the quasimodes decay fast andthey are strongly coupled. So the loss of quasimodes is much lower than the lossof a single quasimode. In an active random medium, when the optical gain forinteracting quasimodes reaches the loss of these quasimodes, lasing with nonres-onant feedback emerges. A significant spectral narrowing is observed. Well abovethe threshold, the total photon–number fluctuation decreases due to gain saturation.However, strong coupling of quasimodes excludes stabilization of the lasing in asingle quasimode.

When the amount of optical scattering increases, the decay rates of the quasi-modes decrease and the mixing of the quasimodes weakens. When the opticalgain increases, lasing with nonresonant feedback occurs first. As the optical gainincreases further, it exceeds the loss of a quasimode that has a long lifetime. Thisresembles a traditional laser. A further increase of optical gain leads to lasing inmore low-loss quasimodes. Laser emission manifests itself by discrete peaks. Thisprocess is lasing with resonant feedback.

When the scattering strength increases further, the lasing threshold in individuallow-loss quasimodes drops below the threshold for lasing in coupled quasimodes.So the mentioned stage of lasing with nonresonant feedback is absent. Well abovethe threshold, the fluctuations of individual photon numbers decrease due to gainsaturation.

Vanneste and Sebbah (2001) also perceived that the studies of the dependenceof laser action on the strength of the disorder (the randomness) and, for highlyscattering media, of the Anderson localization on laser gain were not finished. TheAnderson localization of electronic waves has later been extended to electromag-netic waves (John 1984). The average localization length ξ characterizes an expo-nential decrease of the envelope of a localized mode. Also properties of the elec-tronic or photonic transport depend on this parameter. The localized eigenmodesare microcavities in fact and they can serve as the feedback cavities of the laser.The reports of experiments and theoretic interpretations admit only nonresonantfeedback of spontaneous emission amplified along open scattering paths (Vannesteand Sebbah 2001).

Only recently laser action in a random medium with resonant feedback has beenreported, e.g. (Cao et al. 1999). In the experiment, a semiconductor powder wasused, which played simultaneously the roles of the random and active media. Pos-sible connection with the Anderson localization was merely mentioned (Cao et al.2000). In theory it is assumed that the localized modes of the passive random systemare preserved in the presence of gain, but one may ask, how much they are modifiedby the gain. The doubt whether localization is enhanced or inhibited by the gainended in a low quotation index in one of the papers (Zhang 1995, Paasschens et al.1996, Jiang 1999).

Vanneste and Sebbah (2001) examine the role of strong localization in the lasingaction process. The numerical model describes the full dynamics of the field and the

Page 445: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 437

levels’ populations in a two-dimensional active random medium. First they choosea window of modes strongly localized in the spectrum of the passive medium andexamine the spatial and spectral characteristics of these modes. Next the gain isactivated and the passive modes are compared with the laser modes. It results thatthe active medium is described by the modes of the passive system. The amplifyingmedium has only a small effect on the frequencies. The two-dimensional spatialprofile of the localized wave functions is reproduced without distortion.

They consider two-dimensional disordered medium of size L2 made of circularscatterers with radius r , refractive index n2, and surface filling fraction φ, imbeddedin a matrix of index n1. This system is equivalent to an array of dielectric cylindersparallel to the z-axis. The matrix also plays the role of an active medium. Theyutilize the rate equations of a four-level atomic system and Maxwell’s equationswith a polarization term including atomic population inversion. It is a generalizationof the paper (Jiang and Soukoulis 2000). A TM field defined by the components Ez ,Hx , and Hy is considered.

The modes of the passive system have been studied as follows. The time responseto a short pulse is recorded and Fourier transformed. It is damped. The first andsecond half of the time record can be Fourier transformed. It leads to a conclu-sion that the leaky modes (quasimodes) with shorter lifetimes have not survivedin the second half of the time record. The modes with longer lifetimes are exam-ined by a monochromatic source on their spatial pattern and time evolution. It canbe concluded that the regime of the Anderson localization has been attained. Theinvestigation has continued with introducing gain by uniform pumping of the atomsin the whole system. Above threshold a stationary regime is attained after a tran-sient exponential growth of the field amplitude. The structure of mode has beenpreserved.

At higher pump levels, the laser emission is multimode. After a transient regimethe field becomes stationary in beats between several excited modes. The choice ofindividual localized modes is possible by pumping locally. So it is meaningful toconsider local thresholds for lasing. Vanneste and Sebbah (2001) have concludedwith a question, if the introduction of gain contributes somehow to the discrimina-tion between the diffusive and localized regimes in actual experiments.

Burin et al. (2001) have based their model for a random laser on a planar systemof resonant scatterers pumped by an external laser. At the beginning they expoundalso the following classification: Random lasing with nonresonant feedback appearsas the remarkable narrowing of the luminescence spectrum to a single peak of widthabout several nanometres. The coherent feedback lasing is identified as the seriesof high and narrow peaks having the width decreasing with the increase of pumppower to at least the tenth nanometre scale. They distinguish two different theoreti-cal approaches to the description of random lasing:

(i) The diffusion model with coherent backscattering corrections, e.g. (Wiersmaand Lagendijk 1996) appropriate for describing the regime of nonresonant feed-back, but which fails to predict the lasing threshold behaviour for the laser oper-ation. The criticism is directed to the disability of prediction of the formation

Page 446: Quantum Aspects of Light Propagation

438 6 Periodic and Disordered Media

of high-quality random cavities. It seems to comprise even the rejection of theAnderson localization (considered in the paper (Jiang and Soukoulis 1999)).

(ii) In another approach, the criticism is contrasted with the intended comparison ofthe model with the random matrix approach (Frahm et al. 2000, Misirpashaevand Beenakker 1998).

The authors have analysed an experiment on ZnO disk-shaped powder samplesusing a classical microscopic model. The medium is represented by a set of randomscatterers. The number of these particles is denoted by N . Each of them is consid-ered as an electric dipole oscillator. The resonant frequency of the kth particle isdenoted by ωk and the transition dipole moment of the kth particle is denoted bydk , its length is |dk |=dk . The position vector of a particle k relative to a centre jis denoted by Rk j . It is assumed that the polarization component pk , |pk |=pk , isparallel to the transition dipole moment,

pk = pkdk

dk. (6.591)

We suppose that

pk = pkeizt , Ek = Ekeizt , Ek j = Ek j eizt . (6.592)

If it may be put � = 1, then the equations for the collective eigenfrequencies z andthe collective eigenvectors {pk} of the system have the form

− z2pk = −(ωk − ig)2pk + 2ωkdk(dk · Ek), (6.593)

where g is a gain rate and

Ek =∑

jj �=k

Ek j + i2

3q3pk (6.594)

are the electric fields of other particles and the damping term, with

Ek j = eiq Rk jp j − 3n(n · p j )

R3k j

(1− iq Rk j

)

+q2eiq Rk jp j − n(n · p j )

Rk j, (6.595)

where

n = Rk j

Rk j, q = z

c. (6.596)

The equations are linear, but z enters in a relatively complicated manner. It is acomplex number, the imaginary part of which is the decay rate.

The iteration method they have used calculates the lasing threshold. They couldnot treat a very large system with the number of particles exceeding N = 1000.

Page 447: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 439

They have restricted the analysis to a two-dimensional system of scatterers. Theyhave ignored the difference of dielectric constants of substrate and air. They reportthat they have studied the case ωk = ω0. The positions of all N particles weregenerated randomly within the circle of the radius R = √

N ηcω0

. Three differ-ent values of the parameter η, η = 0.3, 1, 3, have been probed. In compari-son with the random matrix approach, they have seen that the high-quality col-lective modes, in fact, occupy a few scatterers (e. g., from 5 to 10 for N =100).

Ling et al. (2001) have presented a detailed experimental study of random laserswith resonant feedback. One of two materials was a poly(methyl methacrylate) filmthat contained dye and titanium dioxide particles. The other was zinc oxide poly-crystalline film on a sapphire substrate. The dependences of the incident pump-pulseenergy at the lasing threshold and the number of lasing modes for a fixed pumpintensity on the transport mean free path have been measured. The effects of thepump area and the sample size have been determined too. The idea published inCao et al. (2001) has been developed to an analytical model and the theoreticalpredictions have agreed with the experimental results.

Burin et al. (2002) have announced an analytical approach to random lasing ina one-dimensional medium. They have dealt with the lasing threshold. They havediscussed application to the regime of strong three-dimensional localization of light.They have derived that the lasing threshold has strong fluctuations from sample tosample.

The original approach (Letokhov 1968) is based on a diffusion formalism. Itpredicts the lasing instability, when the length of the diffusion path L2

lt, with lt being

the mean-free path length of light, attains the gain length lg. Not even John (1984)modifies this criterion. But experimental and numerical studies have shown a differ-ent value of gain rate at which lasing sets in.

The authors consider a different physical mechanism from diffusive motion.They study a one-dimensional medium as studied by (Jiang and Soukoulis 1999).They identify relevant channels responsible for lasing with the quasimodes. Theresults may be applied to higher dimension in the strong localization regime. Thisphenomenon has been reported in the studies (Chabanov et al. 2000, Wiersma et al.1997).

It is assumed that a one-dimensional scattering medium is situated between theplanes x = 0 and x = L . A gain medium is assumed. The description is basedon an imaginary correction of the frequency ω → ω + ig

2 , where g is the gainrate. The lasing threshold is associated with the singularity in the transmissionthrough the sample ( 0

0 ) as in papers (Jiang and Soukoulis 1999, Beenakker 1998).The authors relate the threshold with the intensity of the field near the source point.They show the equivalence convincingly. In the strong localization regime the lasingthreshold is very small.

We will consider a source in the middle of the structure. We let rl, rr denotethe reflection coefficients from the left-hand and right-hand halves, respectively.Provided that lt � L , |rl| ≈ 1, |rr| ≈ 1, it is essential that the reflected wavesinterfere constructively and the sum of reflection phases (r (ω) = |r (ω)|eiΦ(ω)) is

Page 448: Quantum Aspects of Light Propagation

440 6 Periodic and Disordered Media

a multiple of 2π . The authors call this resonance. The resonances approach eigen-modes of the whole system closed between mirrors.

But in the passive medium the equality is not reached, |rl| < 1, |rr| < 1. Theequation

rl(ω)rr(ω) = 1 (6.597)

is rewritten as

|rlrr| exp{

i[Φl

(ω + i

g

2

)+Φr

(ω + i

g

2

)]}= 1, (6.598)

or

|rlrr| exp

[−g

2

(dΦl

dω+ dΦr

)]= 1. (6.599)

Letting gc denote the solution of equation (6.599), we can express it, approximately,as

gc ≈ −|tl|2 + |tr|2

dΦ1dω+ dΦr

. (6.600)

We differ in the sign. The validity or invalidity of the sign could be ultimately deter-mined according to examples of Φl (ω), Φr (ω). The exposition comprises otherminor errors. For example, the Green function

G(x, xs) ={

cr[eik(x−xs) + rre−ik(x−xs)

], x ≥ xs,

cl[e−ik(x−xs) + rleik(x−xs)

], x ≤ xs,

(6.601)

is written without brackets, but also without cr, cl, which would have been obtainedon a possible removal of the brackets. Here k = ω

c and xs is the source point. TheGreen function should be continuous at xs, but ∂

∂x G(x, xs) should have a jump 2c at

x = xs. The authors set the jump to unity dubiously.In fact, we solve the following equations:

cr(1+ rr)− cl(1+ rl) = 0,

cr(1− rr)− cl(−1+ rl) = 2

iω. (6.602)

We find

cr = 1

1+ rl

1− rlrr, cl = 1

1+ rr

1− rlrr. (6.603)

A consideration of gain rate may be restricted to the range (ω0, ω0 + δω). Thedensity of modes ρ(ω) per unit energy and length can be introduced. In the regimeof the strong localization, the transmission t decreases exponentially with the size ofthe sample,

t ≈ exp

(− L

l1

), (6.604)

Page 449: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 441

where l1 is the localization length. The fluctuations are described by the logarith-mically normal distribution, with the variance of the transmission logarithm L

l2,

l1, l2 ≈ lt.A resonance state is described by the coordinate of its centre and two transmis-

sion coefficients

tl(x) ≈ e−xl1 , tr(x) ≈ e−

L−xl1 . (6.605)

The nominator of (6.600) is in fact denoted by τ0. A study of dwell times has beenmentioned in Burin et al. (2002). But the fact that the nominator comprises a sumhas not been touched on

g(x) ≈ 1

τ0

{exp

(−2x

l1

)+ exp

[−2(L − x)

l1

]}, (6.606)

g ≡ g

(L

2

)≈ 2

τ0exp

(− L

l1

). (6.607)

Here we have completed a coefficient which is not present in the cited paper, how-ever.

The authors remark that the exponential dependence has not been obtained in theliterature as, e.g. in (Jiang and Soukoulis 1999). The lasing threshold strongly fluctu-ates from mode to mode depending on the fluctuation of the centre x , transmission,and dwell time. Strong fluctuations of lasing threshold are also demonstrated by thenumerical simulations.

Patra (2002) considers an optical cavity, which is coupled to the outside byan opening that is small compared to the wavelength of the emitted radiation. Asthe opening is small, it is possible to work with the true modes of the cavity.These modes are distinguished by their eigenfrequencies ωi , i = 1, 2, . . . , Np,as is asserted with respect to the irregular shape of the cavity. A finite number ofmodes can be connected with the size of the cavity, but the dependence has not beenexpressed. Each mode i can be described by the number ni of photons in it. Photonsin mode i can escape through the opening with the rate gi .

The cavity is filled with an amplifying medium. This medium is modelled as afour-level laser dye. The density of excited atoms at the point r in the cavity is N (r).Excitations are created by pumping with the rate P(r) and are lost nonradiativelywith the rate a(r). The coupling of mode i to the medium at the point r is quantifiedby a coefficient Ki (r) ≡ w(ωi )|Θi (r)|2, where w(ω) is the transition matrix elementand Θi (r) is an eigenfunction of the mode i .

The semiclassical equations of motion for ni and N (r) respect the quantumphysics of the matter only. Here ni cannot mean the photon number. The Langevinterms in the equations are in fact to correct the “continuous variation” of the photonnumbers, which is not appropriate, to jumps of the photon numbers. Similarly alsothe medium can be treated, when the cavity volume is divided into Ns cells (Luksand Perinova 2003).

Page 450: Quantum Aspects of Light Propagation

442 6 Periodic and Disordered Media

To linearize the description, we write ni = ni + δni and N (r) = N (r) + δN (r),where ni and N (r) are the average solutions, ni ≡ 〈ni 〉, N (r) ≡ 〈N (r)〉, and δni andδN (r) are new variables. Patra (2002) discretizes the equations in space as follows.In the cavity, he picks Ns points r j , j = 1, 2, . . . , Ns. Then Ki j ≡ Ki (r j ), N j ≡N (r j ), and analogues for other quantities are introduced. For simplicity, Pj doesnot depend on j , Pj ≡ P . The stationary mean numbers ni and mean densities N j

are the solution of equations

0 = −gi ni + V

Ns

Ns∑

j=1

(ni + 1)Ki j N j , i = 1, 2, . . . , Np, (6.608)

0 = Pj − a j N j −Np∑

i=1

(ni + 1)Ki j N j , j = 1, 2, . . . , Ns. (6.609)

The factor VNs

is set equal to unity. On the stationary average variables the cor-relations 〈δni (t)δni ′(t)〉, 〈δni (t)δN j ′(t)〉, 〈δN j (t)δni ′(t)〉, and 〈δN j (t)δN j ′(t)〉 aredetermined.

The author defines the Fano factor Fi of the radiation emitted from the cavityin the mode i and the Fano factor F for a measurement where the photons emit-ted from the cavity in all modes are detected simultaneously. He discusses threesimple cases of lasing regimes. He characterizes the single-mode laser. He treatsthe dependence of the Fano factor of the total radiation F and the dependence ofthe Fano factor of the radiation of one of lasing modes Fi on the pumping rate P .On several simplifying assumptions he compares the laser with Ns = Np = 10and g1 = 1

100 , gi = 110 for i = 2, . . . , 10 with the laser with Ns = Np = 10

and all gi = 1100 . In the second case, the mode with the smallest gi becomes the

lasing mode and the behaviour is similar to a single-mode laser. In the third case,the mode competition occurs. Even though the total radiation behaves similarly asin the case of a single-mode laser, the Fano factor of any of lasing modes growsexcessively.

Patra (2002) intends to study the mode-competition noise. He holds a spatialoverlap of the modes for the necessary condition of the mode competition. Herecalls the ensemble of cavities with small variations in shape or scatterer positions.Neither the volume V nor the hole diameter d should vary. Then, for instance, in thedescriptions (6.608) and (6.609) oriented to the average variables, the coefficientsbegin to be interpreted as random variables. A parameter of their distribution is, forinstance, the mean loss rate g that depends on the frequency ω. The expression wedo not reproduce here is related to the assumption that neither the volume V nor thehole diameter d changes.

It is assumed that the loss rates gi are independent and identically distributed.For simplicity, Ki j = |Θi j |2 is computed, where Θi j are elements of an Ns × Ns

random unitary matrix. The author, for instance, has calculated the Fano factor Fof the radiation emitted from the primary lasing mode depending on the pumpingrate P . Along with the quantity F he has presented the number of modes above

Page 451: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 443

the lasing threshold Nl. Actually, the axes are denoted as Pg and F−1

g . Whereasthese quantities have been determined for a particular sample, a number of furtherquantities have been found using 9×105 samples, e.g. the average of the Fano factordepending on the number of modes above the threshold Nl.

Trying to interpret experiments, the author recalls the spatial confinement oflight, which is assumed in the experiments with the random lasers and has beenconfirmed, as we know, by simulations. The author holds these modes somewhatcontradictorily for virtual multimode chaotic cavities. The deliberateness of thisinterpretation is underlined by considering the difference between these totally openchaotic cavities and the chaotic cavities with a small opening, which are studiedin Patra (2002). Again he reminds of the fact that his paper does not include themodes which overlap. Therefore a comparison between the papers by (Zacharakis2000, Cao 2001), in which the author is interested, cannot be based on the model.Also the concept of a “more or less” resonant feedback only glints in the exposi-tion.

Cao (2003) has contributed a chapter on lasing in disordered media. Lasing islight amplification by stimulated emission with feedback. The author has distin-guished two kinds of feedback: intensity or energy feedback and field or amplitudefeedback. These mechanisms are also called incoherent (or nonresonant) feedbackand coherent (or resonant) feedback, respectively. She has let lt denote the transportmean free path.

Many experiments have been classified as random lasers with incoherent feed-back. Some experiments have been reviewed as random lasers with coherent feed-back. Transition between the two types of random lasers has been characterized as,e.g. in Cao et al. (2001). Classical and quantum theories have been reviewed in short.It has been admitted that the characteristics of lasing modes in three-dimensionalrandom media are not fully understood. A previous theory has predicted that in athree-dimensional random medium with lt > λ, the quasimodes extend over theentire sample. The experiments have shown that the size of individual lasing modesis much smaller than that of the entire sample even when lt � λ, cf. (Apalkov et al.2002).

Patra (2003) has noted that, in the theoretical treatment of disordered materi-als, two particular geometries are of great importance, the disordered slab and thechaotic cavity. Random-matrix theory began at the moment, when Wigner (1956)formulated the eigenvalue distribution for a closed chaotic cavity. For chaotic cavi-ties with broken time-reversal symmetry, the distribution of imaginary part of eigen-value (the decay rate distribution) is known analytically for an opening of any size(Fyodorov and Sommers 1997). The open systems in optics are not attributed bro-ken time-reversal symmetry, but only “approximability” by a cavity with brokensymmetry and an opening of half the real size.

Patra (2003) has studied a disordered slab of the length L and the width N . Hehas added an argument that N can be identified with the number of propagatingmodes for an ideal coupling at the centre of the conduction band. From the resultsof the numerical simulations, he has found mainly that the mean free path l is invery good approximation given by the relation

Page 452: Quantum Aspects of Light Propagation

444 6 Periodic and Disordered Media

l = 6

w32

, (6.610)

where w is an amount of randomness, 0.3 ≤ w ≤ 1.0. He has distinguished thediffusive and localized regimes. In both cases, he has also been able to find simpleanalytical formulae, which describe the decay rate distribution well. It has beenproposed that these distributions depend on a scaling factor γ0,

γ0 = 2l

L2(6.611)

in the diffusive regime and

γ0 = a

N 2exp

(− L

), (6.612)

with a = 1.12, ξ = N+12 l, in the localized regime. The results can be applied to

both electronic and photonic systems. It has been shown that, on certain conditions,the lasing threshold of a random laser is close to γ0 in both the diffusive and thelocalized regime.

Florescu and John (2004) have considered the probability distribution for photonsin the vicinity of position r, travelling with a wave vector k, and a master equationfor the diffusion model with the gain describing this distribution. The steady-statephoton distribution function has been characterized. The statistical properties of theemitted laser light have been studied. The dependence of the photon statistics onscatterer density, gain concentration, and position within a sample has been dis-cussed.

van der Molen et al. (2007) have listed five optical and material properties ofthe sample, seven experimental details, and seven experimental data of the randomlaser needed for a quantitative comparison between different experiments. Com-parisons between different theories and those between experiments and theory areforeseen too. Their recent experiment has been advertised, in fact, by a comparisonwith 8 perfectly reported ones from among 160 papers. An analysis of the model in(Mujumdar et al. 2004) has been added.

6.5.3 Modal Decomposition in Optical Resonators

Cheng and Siegman (2003) have reapproached the quantum description of a laser.They have mentioned the paper (Deutsch 1991), where the quantization of quasi-modes (let us call them according to other work) is expounded, and the paper (Ho1998), where rather the pseudomodes (let us call them with respect to other liter-ature) are paid attention to. Cheng and Siegman (2003) contribute with a kind ofgeneralized quasimode approach.

The authors introduce nonorthogonal modes un and nonstandard annihilation(creation) operators aun (a†

un ). On using these quantities, the electric-field operatoris expressed. A kind of normalization is represented just by a formal identity of the

Page 453: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 445

proposed expansion with that of the electric-field operator in terms of orthogonalmodes.

Next, adjoint modes φn are introduced, which satisfy the biorthogonal relation

(φn|um) ≡∫

φ∗n um d3x = δnm . (6.613)

The authors recall that (φn|φn) > 1 (Siegman 1989). A description of the laserusing quantum Langevin equations is provided, where the correlation relations ofthe atomic noise operators can be obtained from the Einstein relation (Sargent III1977). We would remark that these correlation relations involve reservoir aver-ages on the right-hand sides, which may suggest other commutation relations. Theauthors assume that the laser is in a single mode oscillation and label the oscillationmode as u0. Finally, they derive the well-known quantum laser line-width equations

Δωl ≈ (φ0|φ0)�ωlγ

2c

2Pl, (6.614)

where ωl is the centre frequency, γl is the cavity decay rate, and Pl = |A0|2�ωlγc,with A0 = 〈au0〉R being the reservoir averaged steady-state solutions, which arequantum corrected.

6.5.4 Chaotic Resonators

Graf et al. (1992) studied the distribution of eigenfrequencies in a microwave sta-dium billiard with chaotic dynamics. Alt et al. (1995) paid attention to wave func-tions or decay amplitudes (widths). The cavity was coupled to three antennas. Thewidths of the resonances are proportional to the squares of the eigenfunction at therespective locations of the antennas. Tests were in perfect agreement with predic-tions of Gaussian orthogonal ensemble for the statistics of eigenfunctions.

A resonance labelled μ has a frequency ωμ and decay widths Γμj , j = 1, 2, 3,where j labels an antenna. The total width, Γμ, is

Γμ =3∑

j=1

Γμj + Γμ,wall, (6.615)

where Γμ,wall represents dissipation in the walls of a superconducting cavity. TheGaussian orthogonal ensemble predicts a Gaussian distribution for the decay ampli-tudes ±√Γμj , or a gamma distribution for Γμj with fixed j . The prediction canbe tested that the decay amplitudes for different channels are uncorrelated. Thisholds provided that the distances between neighbouring antennas are larger thanthe maximum wavelength used in the experiment. The Gaussian orthogonal ensem-ble predicts that the resonance widths are statistically independent of the resonanceenergies.

Nockel and Stone (1997) have focused on the effectively two-dimensional case ofa deformed cylindrical resonator. Any resonator is characterized by a set of modes,each of which has a resonant frequency ω and resonance width δω = 1

τ, where τ

Page 454: Quantum Aspects of Light Propagation

446 6 Periodic and Disordered Media

is the lifetime of a photon in the mode. A cylindrical dielectric resonator has long-lived resonances due to “whispering gallery” modes in which light is trapped bytotal internal reflection. The lifetime is finite in consequence of evanescent leakage,the optical analogue of quantum tunnelling. Asymmetric resonant cavities exhibitpartially chaotic ray dynamics. A deformation of the cavity gives rise to a broad-ening of the whispering-gallery resonances. For small deformations the evanescentleakage increases. For large deformations the lifetime is shortened by a processcalled “chaos-assisted tunnelling” (Doron and Frischat 1995).

The predictions of ray optics are compared with wave solutions. Rays escapingnear the critical angle are emitted tangent to the surface. In the framework of thetheory of two-dimensional “billiards” (Gutzwiller 1990), it can be derived that thepoints of escape are the points of maximum curvature of the asymmetric resonantcavities. This implies strong emission maxima in the far field in directions tan-gent to the points of maximum curvature for appropriate values of the refractiveindex. For smaller indices the points of highest curvature may be screened by stableislands.

Gornik (1998) reminds of the interest in lasing not only at extremely low currentsbut also in directionality control expected from micrometre-size lasers. A circularsymmetric cylinder or disk laser may be based on a whispering-gallery resonance.Light only weakly leaks out from the cavity. A power increase and output direc-tionality is achieved by deformation, at which a circular cross section becomesquadrupolar. At small deformations chaotic whispering-gallery resonances occur.At larger deformations the lasers operate on bow-tie-shaped modes.

Gmachl et al. (1998) have reported on semiconductor microcylinder lasers, whichhave a high power and high directionality due to the deformation of cross sectionof an optical resonator. They have focused on semiconductor lasers that have aneffective index of n ≈ 3.3. The deformation has been called a flattened quadrupole.The quadrupolar shape has been derived from a resist pattern, which has a stadiumshape. They chose quantum cascade lasers to emit light of wavelength λ = 5.2 μm(mid-infrared).

While in circular and weakly deformed lasers whispering-gallery type modesoccur, in the higher range of deformations a bow-tie shaped resonance is born,which is related to the improved performance of the lasers. Figures in the citedwork have been obtained by the numerical solution of the Helmholtz equation. Thisanalysis has been supported by a study of the short-wavelength limit of the problemby (Gutzwiller 1990, Reichl 1992). Poincare’s surface of section has been completedby the Husimi function for a particular deformation (LeBoeuf and Saraceno 1990).

Patra et al. (2000) have presented a random-matrix theory for the linewidth of alaser cavity in which the radiation propagates chaotically. They continue the paper(Misirpashaev and Beenakker 1998), which contains results on the lasing thresh-old and mode competition. They investigate the Petermann factor K , or excessnoise factor. The factor K is expressed in terms of a non-Hermitian matrix H,H = H0 − iπWWT . Here H0 is an M × M Hermitian random matrix and W isa real, nonrandom M × N coupling matrix, M is the number of considered modes,and N is the number of the output channels. At the end of the calculation the limit

Page 455: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 447

M → ∞ is taken. The distribution of the factor K is calculated exactly on theassumption that N = 1. It is surprising that the mean Petermann factor convergesnonanalytically to unity for the transmission probability T of the output channelgoing to zero.

Frahm et al. (2000) relate the Petermann factor with the general theory of res-onant scattering. They concentrate themselves on the Petermann factor in chaoticcavities in the case of broken time-reversal symmetry. They have calculated that theaverage 〈K 〉 scales with the square root of the number of scattering channels N .The exact result is compared with the result of a numerical simulation. Only froma numerical simulation they have found that the sublinear increase holds also in thecase of preserved time-reversal symmetry.

Hackenbroich et al. (2001) recall the facts known since the development of thequantum theory of lasing in the 1960s. Below the laser threshold, the photon statis-tics is well represented by a thermal distribution. Above the threshold, nonlinearinteractions stabilize the field intensity, and the photon statistics approaches thePoissonian distribution. The authors mention the renewal of interest in the photonstatistics of amplifying media. This interest goes hand in hand with the attentionpaid to random media. While linear random amplifiers were intensively studied,in about the same time, there were few results for an amplifying random mediumabove the lasing threshold. The authors concentrate themselves on a chaotic laserresonator, which is considered to be a random medium. They restrict themselves tothe single-mode lasing.

The randomness means that an experiment is performed on an ensemble ofmodes. The variation of parameters of a single resonator and the variation of shapeof different resonators are typical, manners how to reach the randomness.

A chaotic resonator is filled with an active medium. The radiation is allowed togo out through a single opening, a partially transmitting mirror. This mirror reflectsback a photon with the mean probability R. It transmits a photon with the meanprobability T . The cavity is connected to a waveguide and a frequency selectivephotodetector. The waveguide supports M modes. It is assumed that MT � 1. Themean cavity escape rate Γ and the counting time t are introduced. The frequencyresolution requires a longer counting time t . The active medium is described bythree parameters A, B, and C, which characterize the linear gain, the nonlinear sat-uration, and the total loss, respectively. The coefficients A and B are considered tobe nonrandom, but C is a random quantity, since C = Γ+ κ , where Γ is the photonescape rate and κ is the absorption rate, which is nonrandom, but Γ is a randomvariable.

The authors apply the input–output theory (Gardiner and Collet 1985), mainly alinear coupling between the waveguide and the cavity field, which is described interms of constants γp, p = 1, 2, . . . , M . In the linear regime below the threshold,the input–output theory yields the theory according to (Beenakker 1998).

Then the authors consider the nonlinear regime near and above the threshold. Thephotocount statistics is given in dependence on statistical properties of the integratedcavity field intensity. Considering an ensemble of modes, Γ has the χ2

ν distribution

Page 456: Quantum Aspects of Light Propagation

448 6 Periodic and Disordered Media

P(Γ) = AνΓν2−1 exp

(−νΓ

), (6.616)

where ν = βM , with β = 1, 2, and Aν is a normalization constant. Here β = 1, ifthe system is time-reversal invariant and β = 2 otherwise. The special case M =β = 1 is known as the Porter–Thomas distribution. The factorial moments

μr =∞∑

m=0

m!

(m − r )!p(m) (6.617)

in the dependence on Γ become random quantities μr |Γ and their probability den-sity is

P(μr ) =∫

P(Γ)δ (μr − μr |Γ) dΓ, (6.618)

where

μr |Γ =⟨: W r :

⟩ = Γr⟨: I r :

⟩. (6.619)

The case M � 1 can be analyzed from the viewpoint of small relative fluctuationsΔΓ

Γ→ 0 for M → 0. In the weak-coupling limit Γ → 0, the moments have a

power-law behaviour.For short counting intervals, the photocount distribution is connected simply

with the stationary photon-number distribution Pn in the lasing mode. This distri-bution depends on the parameters of lasing A, B, and C. The authors have foundthe probability density P(μ1). The existence of μmax ≡ supΓ {μ1|Γ} may be uti-lized. They discuss the dependence of μ1 on Γ, which in turn depends on A,B, κ . Above the threshold, it means that κ < A; below the threshold, it meansthat κ > A. For example, (a) κ = 0.7 < 1 = A is above the threshold and(b) κ = 2.0 > 1 = A is below the threshold. The mean photocount μ1 in depen-dence on Γ, above the threshold, has a maximum, which can be approximated bya quadratic polynomial, and so its probability density has a peak in this maximum.Below the threshold, a supremum is rather the right term, and the density vanishestherein.

In the limit Γ →∞, μ1 > 0, which is an amplified spontaneous emission. Mono-tonic decrease (above the threshold) and monotonic increase (below the threshold)contribute with a broad peak in the right-hand neighbourhood μ1(∞) (above thethreshold) and in the left-hand neighbourhood μ1(∞) (below the threshold), if Γ

is not small. The mean photocount can be expressed in terms of parameters of thestationary distribution of the laser, but μr , with r ≥ 2, already depends on otherparameters.

Related fluctuation phenomena occur in neutron waves and in semiconductorquantum dots. Chaotic lasers feature not only wave chaos but also the nonlineardynamics of the laser. The distribution of Γ may not be the Porter–Thomas distribu-tion then.

Yang and Kellman (2000, 2002) have reported on a method of calculation of thesemiclassical wave function for an Einstein–Brillouin–Keller quantizing torus. The

Page 457: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 449

semiclassical wave function of the “primitive type”, which diverges at the classicalcaustics, is computed directly from running a classical trajectory.

Hackenbroich et al. (2002) recall the laser action in amplifying random media.While (Beenakker 1998) is devoted to the subthreshold radiation, the nonlinearlasing regime has been treated only modestly. Random laser modes have unusualproperties. The modal functions and modal frequencies in the random lasers areanalyzed in a statistical manner. The character of modes depends on the amountof disorder. For strong disorder, the localization of light may occur, i.e., the riseof well-separated modes in different regions of space. Weak disorder leads to animperfect confinement of light and to strongly overlapping modes.

The authors state the lack of satisfactory scheme for the field quantization inrandom media. They pay attention to optical resonators with overlapping modes anddevelop a quantization scheme for such resonators. This scheme does not neglectthe case of a chaotic resonator. The frequency spectrum of the resonator modes withappropriate correlations may be utilized. The mode coupling by dissipation has astochastic character as well and its description has been conceived.

The authors consider a two-dimensional optical resonator for the sake of sim-plicity. It is actually a waveguide, which does not guide any fields. It is assumedthat the electric field contained in the resonator is polarized in the direction of thetranslational symmetry of the resonator, which is the direction of the z-axis. Theauthors call these fields TM fields in conformity with the waveguide terminology.The terminology of the reflection and refraction at a planar interface, which is stilla little related, would be “TE fields” on the contrary.

It is proper to perform an exact quantum description of the total system compris-ing the resonator and the external radiation field, since there are openings of the totalwidth W in the boundary of the resonator. Hackenbroich et al. (2002) have providedan outline of a two-dimensional generalization of the approach to which we allude inSection 3.3. By the authors’ exposition, this approach consists in the finding of theexact eigenfunctions ψm(ω, r) of the Helmholtz equation. This equation describesa harmonic wave in the free space along with boundary conditions of significance.The eigenfunctions are labelled by the continuous frequency ω and the integer m.The latter can be interpreted as a channel number. It is important that the conjugateof the field A(r, t) = A(r, t)ez is the field Π(r, t) = Π(r, t)ez , where

A(r, t) = c∫

ψT (ω, r)q(ω, t) dω,

Π(r, t) = 1

c

∫ψT (ω, r)p(ω, t) dω, (6.620)

with ψ(ω, r) being an M-component vector with components ψm(ω, r), q(ω, t)being an M-component vector with components qm(ω, t), and p(ω, t) is an M-component vector with components pm(ω, t). To quantize the fields, we imposethe canonical commutation relations [qm(ω, t), pn(ω′, t)] = i�δmnδ(ω − ω′)1.

The authors opine that the modes-of-the-universe approach does not provide anyinformation about the resonator itself. Evoking Feshbach’s projector formalism, the

Page 458: Quantum Aspects of Light Propagation

450 6 Periodic and Disordered Media

authors assume the single-photon Hamiltonian of the form

H =∑

λ

Eλ|φλ〉〈φλ| +∑

m

∫E |χm(E)〉〈χm(E)| dE

+∑

λ,m

∫[Wλm(E)|φλ〉〈χm(E)| + H. c.] dE, (6.621)

where |φλ〉 are photon states with the energy Eλ in the decoupled resonator and|χm(E)〉 are photon states with the energy E in the decoupled channel m.

The resonator wave functions φλ(r) are nonzero only within the resonator, andthe channel wave functions χ (E, r) are nonzero only outside. These functions mustsatisfy the Dirichlet conditions along the boundary of the total system except theopenings. The boundary condition along the ideal surface bridging the openings isarbitrary with the exception that the total Hamiltonian must be self-adjoint. Then thecoupling amplitudes Wλm(E) are given by surface integrals of products of resonatorand channel wave functions. The diagonalization of the Hamiltonian is performedaccording to the papers by Viviescas and Hackenbroich (2003) and Dittes (2000).It leads to a transformation of the relations (6.620). Particularly, it can be derivedthat the intracavity fields have the same representation as in the case of the closedcavity (and the Gaussian system of units). The field Hamiltonian is rather similarto the well-known system-bath Hamiltonian of quantum optics. As the nonresonantterms can be neglected, provided that the broadening of the resonator modes is muchsmaller than their frequency, this similarity is remarkable.

The equations for mode operators have the form

˙aλ(t) = −iωλaλ(t)− π∑

λ′(WW†)λλ′ aλ′ (t)+ Fλ(t), (6.622)

where Fλ(t) is a well-defined noise operator and W is the coupling matrix with theelements Wλm . The mode operators aλ are coupled by the damping matrix WW†.The noise operators Fλ are appropriately correlated, 〈Fλ(t)F†

λ (t ′)〉 ∝ (WW†)λλ′δ(t−t ′). Hackenbroich et al. (2002) formulate the condition, on which equation (6.622)simplifies to the standard equation of motion for nonoverlapping modes. From therelation (6.622), they conclude that for the interesting case of wave chaos, the modedynamics will be governed by a non-Hermitian random matrix (Fyodorov and Som-mers 1997, Chalker and Mehlig 1998), eventually. They study also an open resonatorabove the laser threshold. They announce that they have derived the linewidth

δω = K δωST, (6.623)

where δωST is the fundamental (Shawlow–Townes) linewidth and K is the Peter-mann factor

K = 〈l|l〉〈r |r〉. (6.624)

Page 459: Quantum Aspects of Light Propagation

6.5 Propagation in Amplifying Random Media 451

Here |r〉 and 〈l| are, respectively, the right and left eigenvectors of the non-Hermitianmatrix H, whose anti-Hermitian part expresses the damping and gain.

An experimental evidence that random-matrix theory is helpful has been pro-vided in Mendez-Sanchez et al. (2003). In many microwave experiments withchaotic and disordered billiards, the distribution of reflection coefficients was inves-tigated. For all regimes, the agreement between experimental and theoretical pre-dictions has been almost perfect.

Page 460: Quantum Aspects of Light Propagation

Chapter 7Conclusions

In this book we have mainly reviewed canonical quantum descriptions of light prop-agation in a nonlinear dispersionless dielectric medium and in linear and nonlineardispersive dielectric media. These descriptions have regularly been simplified by atransition to one-dimensional propagation, which has been illustrated also by someoriginal simple descriptions. Besides this we have reported criticisms of descriptionof light propagation in a nonlinear medium using a spatial variable instead of andsimilarly as the quantum-mechanical time parameter.

We have adopted the standpoint that macroscopic quantum electrodynamicsarises both through reduction of a microscopic description and through immediateapplication of a quantization scheme to macroscopic fields. The origins of immedi-ate macroscopic theories are connected with the possibility of determining the elec-trical displacement field as the canonical momentum (up to the sign) to the vectorpotential. One may also consider as a member of this class the possibility of startingfrom the assumption that the canonical momentum is the magnetic-induction fieldwith the dual vector potential whose curl is the electric displacement field.

Due to a relatively large number of various fields and their components and therelative complexity of commutators, discrete and continuous expansions in terms ofannihilation and creation operators must be presented. In the papers reviewed here,the ground (vacuum) state of the electromagnetic field has mostly been assumed.The Heisenberg picture of time evolution has been preferred. In one of these papersthe opinion has been formulated that the quantum description of sources is to betreated separately, but a detailed treatment cannot be found there. As a criterion forthe correctness of the theories, equivalence between the Heisenberg equations ofmotion for the electric displacement and magnetic-induction fields and the Maxwellequations is required.

Not always it is necessary to utilize the formalism of the electromagnetic fieldin the matter. For description of experiments with correlated photons it suffices todescribe the electromagnetic field between optical devices and to know the input–output relations for the optical elements, both passive and active, with which theradiation is transformed.

The role of the momentum operator has been analyzed: its limited domain ofapplication, one-dimensional propagation, is still appealing for its simplicity. Theintegral expression for the momentum operator has been given. It has been stressed

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 7, C© Springer Science+Business Media, LLC 2009

453

Page 461: Quantum Aspects of Light Propagation

454 7 Conclusions

that the use of this operator as the generator of spatial progression depends on theknowledge of appropriate equal-space commutators between field operators. Resultson the connection with the slowly-varying-envelope approximation have been pre-sented, including some important applications.

The possibility has been examined of determining a phenomenological Hamil-tonian, and the quantization scheme with dual potential has been expounded andappropriately modified, so that the real relative permittivity of the lossless mediumis expressed at least locally (in terms of a quadratic Taylor polynomial). In simpli-fying to the one-dimensional case, an application to a nonlinear dispersive (Kerr-like) medium and quantum solitons has been presented. Various notions of theslowly-varying-envelope approximation and the quantum-paraxial approximationhave been presented, along with applications of these concepts in both linear andnonlinear cases.

Multimode consideration of nonclassical effects is needed when one wants toimprove the processing of images and of parallel signals. The quantum fluctuationsof light at different spatial points in the plane perpendicular to the propagation direc-tion of the light beam have been investigated.

A new application to a nonlinear dispersionless (Kerr-like) medium has beenconcerned with the propagation of arbitrary pulses. The need for renormalizationhas been declared even in this, one-dimensional case. As an example, the nonlinearprocess has been presented, whose description includes the renormalization. Finally,the application of one of the macroscopic approaches has lead to the description ofseveral linear optical devices and to the study of radiating atoms in a linear medium,which is a recurrent theme by the way.

Derivation of a macroscopic description from a microscopic model has beenperformed; it is not possible – in the prevailing Heisenberg picture – to eliminatethe matter fields completely. The argument has been provided not only with thedescription of the dispersion up to any order or accuracy, but also with losses, asfollows from the Kramers–Kronig relations.

The exposition of this microscopic model has been delivered from the viewpointof the macroscopic variables and the operator-valued noise current, which is a formof the matter field. A nonlinear modification of this description of the linear disper-sive absorbing medium has been performed for the Kerr medium. The application toa two-level atom positioned in the centre of a spherical cavity has been mentioned.

The magnetic properties are usually neglected, but they must be included in thephenomenological quantum description of negative-index materials. Even thoughthe Casimir effect is not regularly connected with the propagation, an expression ofthe noise which is quantal in essence fits in the framework of the electromagnetic-field quantization.

A different notion of the macroscopic fields has been mentioned as well: Theseare not comprised in a microscopic model in advance, but approximately equal oper-ators of measurable modes of the dressed matter fields.

The periodic and disordered media have been paid attention to, respecting theirimportance in photonics. The macroscopic approach to quantization of the elec-tromagnetic field in a periodic medium has mainly been spoken of, but also the

Page 462: Quantum Aspects of Light Propagation

7 Conclusions 455

papers have been mentioned, which have utilized specific approaches. Finite peri-odic media can couple free-space modes. In applications, the corrugated waveguidesare important. Photonic crystals are infinite or finite periodic media. The literatureon one-dimensional periodic media dominates, since both theory and fabrication ofsuch media are simpler.

We have mentioned the quantization of the electromagnetic field in a disor-dered medium mainly in connection with various physical studies. The macroscopicapproach to the quantization of the electromagnetic field suffices usually, but adescription of the disordered or random medium is not easy. We have cited thepapers, whose authors have restricted themselves to the quantum input–output rela-tions and the detection theory. This is also related to application of results of furtherfields of the quantum physics. Many papers have reported the random lasers. Eventhough we have intended to review rather the theory, we see that we have been ableto only provide a partial review. In the essence of the matter is that much theoreticalwork is not published as the optical physics.

Page 463: Quantum Aspects of Light Propagation

References

Abraham, M. (1909). Rendiconti del Circolo Matematico di Palermo, 28:1.Abraham, M. (1910). Rendiconti del Circolo Matematico di Palermo, 30:33.Abrahams, E., Anderson, P. W., Licciardello, D. C., and Ramakrishnan, T. V. (1979). Phys. Rev.

Lett., 42:673.Abram, I. (1987). Phys. Rev. A, 35:4661.Abram, I. and Cohen, E. (1991). Phys. Rev. A, 44:500.Abram, I. and Cohen, E. (1994). J. Mod. Opt., 41:847.Agarwal, G. S. (1975). Phys. Rev. A, 11:230.Agarwal, G. S. and Wolf, E. (1970). Phys. Rev. D, 2:2206.Agrawal, G. P. (1989). Nonlinear Fiber Optics. Academic Press, New York.Agrawal, G. P. (1995). Nonlinear Fiber Optics. Academic Press, San Diego, CA.Ahmedov, H. and Duru, I. H. (2003). J. Math. Phys., 44:5487.Aiello, A. (2000). Phys. Rev. A, 62:063813.Aiello, A. and Woerdman, J. P. (2005). Phys. Rev. A, 72:060101.Akkermans, E., Wolf, P. E., and Maynard, R. (1986). Phys. Rev. Lett., 56:1471.Allen, L., Barnett, S. M., and Padgett, M. J. (2003). Optical Angular Momentum. Institute of

Physics Publishing (IOP), Bristol.Allen, L. and Stenholm, S. (1992). Opt. Commun., 93:253.Alt, H., Graf, H.-D., Harney, H. L., Hofferbert, R., Lengeler, H., Richter, A., Schardt, P., and

Weidenmuller, H. A. (1995). Phys. Rev. Lett., 74:62.Altman, A. R., Koprulu, K. G., Corndorf, E., Kumar, P., and Barbosa, G. A. (2005). Phys. Rev.

Lett., 94:123601.Anderson, P. W. (1958). Phys. Rev., 109:1492.Antezza, M., Pitaevskii, L. P., Stringari, S., and Svetovoy, V. B. (2006). Phys. Rev. Lett., 97:223203.Apalkov, V. M., Raikh, M. E., and Shapiro, B. (2002). Phys. Rev. Lett., 89:016802.Artoni, M. and Loudon, R. (1997). Phys. Rev. A, 55:1347.Artoni, M. and Loudon, R. (1998). Phys. Rev. A, 57:622.Artoni, M. and Loudon, R. (1999). Phys. Rev. A, 59:2279.Assanto, G., Laureti-Palma, A., Sibilia, C., and Bertolotti, M. (1994). Opt. Commun., 110:599.Babiker, M. and Loudon, R. (1983). Philos. Trans. R. Soc. Lond. Ser. A, 385:439.Bache, M., Brambilla, E., Gatti, A., Lugiato, L. A. (2004). Phys. Rev. A, 70:023823.Balachandran, R. M. and Lawandy, N. M. (1995). Opt. Lett., 20:1271.Ban, M. (1994). Phys. Rev. A, 49:5078.Barash, Yu. S. and Ginzburg, V. L. (1975). Sov. Phys. Usp., 18:305.Barnett, S. M., Gilson, C. R., Huttner, B., and Imoto, N. (1996). Phys. Rev. Lett., 77:1739.Barnett, S. M., Jeffers, J., Gatti, A., and Loudon, R. (1998). Phys. Rev. A, 57:2134.Barnett, S. M., Matloob, R., and Loudon, R. (1995). J. Mod. Opt., 42:1165.Barnett, S. M. and Radmore, P. M. (1988). Opt. Commun., 68:364.Barton, G. (1987). Proc. R. Soc. London, Ser. A, 410:175.

A. Luks, V. Perinova, Quantum Aspects of Light Propagation,DOI 10.1007/b101766 BM2, C© Springer Science+Business Media, LLC 2009

457

Page 464: Quantum Aspects of Light Propagation

458 References

Barton, G. (1991a). J. Phys. A, 24:991.Barton, G. (1991b). J. Phys. A, 24:5533.Barton, G. (2001). J. Phys. A: Math. Gen., 34:4083.Bechler, A. (1999). J. Mod. Opt., 46:901.Beenakker, C. W. J. (1997). Rev. Mod. Phys., 69:731.Beenakker, C. W. J. (1998). Phys. Rev. Lett., 81:1829.Beenakker, C. W. J. (1999). In Fouque, J.-P., editor, Diffuse Waves in Complex Media, NATO

Science Series C531, pp. 137–164. Kluwer Academic Publishers, Dordrecht.Beenakker, C. W. J., Paasschens, J. C. J., and Brouwer, P. W. (1996). Phys. Rev. Lett., 76:1368.Bekenstein, J. D. and Schiffer, M. (1994). Phys. Rev. Lett., 72:2512.Ben-Aryeh, Y., Luks, A., and Perinova, V. (1992). Phys. Lett. A, 165:19.Ben-Aryeh, Y. and Serulnik, S. (1991). Phys. Lett. A, 155:473.Bennink, R. S., Bentley, S. J., Boyd, R. W., and Howell, J. C. (2004). Phys. Rev. Lett., 94:033601.Bergman, K. and Haus, H. A. (1991). Optics Lett., 16:663.Berman, P. R. (2004). Phys. Rev. A, 69:022101.Berman, P. R. (2007). Phys. Rev. A, 76:042106.Bespalov, V. G., Kozlov, S. A., Shpolyanskiy, Yu. A., and Walmsley, I. A. (2002). Phys. Rev. A,

66:013811.Białynicka-Birula, Z. and Białynicki-Birula, I. (1987). J. Opt. Soc. Am., 4:1621.Białynicki-Birula, I. (1996a). In Wolf, E., editor, Progress in Optics, Vol. 36, p. 245. North-

Holland, Amsterdam.Białynicki-Birula, I. (1996b). In Eberly, J. H., Mandel, L., and Wolf, E., editors, Coherence and

Quantum Optics VII, p. 313. Plenum Press, New York.Białynicki-Birula, I. (1998). Phys. Rev. Lett., 80:5247.Białynicki-Birula, I. (2000). Opt. Commun., 179:237.Białynicki-Birula, I. and Białynicka-Birula, Z. (1975). Quantum Electrodynamics. Pergamon,

Oxford.Bjork, G., Soderholm, J., and Sanchez-Soto, L. L. (2004). J. Opt. B: Quantum Semiclass. Opt.,

6:S478.Bjorken, J. D. and Drell, S. D. (1965). Relativistic Quantum Fields. McGraw-Hill, New York.Bleany, B. I. and Bleany, B. (1985). Electricity and Magnetism. Oxford University Press, Oxford.Bloembergen, N. (1965). Nonlinear Optics. Benjamin, New York.Bloembergen, N. and Pershan, P. S. (1962). Phys. Rev., 128:606.Blow, K. J., Loudon, R., and Phoenix, S. J. D. (1991). J. Opt. Soc. Am. B, 8:1750.Blow, K. J., Loudon, R., and Phoenix, S. J. D. (1992). Phys. Rev. A, 45:8064.Blow, K. J., Loudon, R., Phoenix, S. J. D., and Sheperd, T. J. (1990). Phys. Rev. A, 42:4102.Bolthausen, E. and Sznitman, A.-S. (2002). Ten Lectures on Random Media. Birkhauser, Basel.Bordag, M., Geyer, B., Klimchitskaya, G. L., and Mostepanenko, V. M. (2000). Phys. Rev. Lett.,

85:503.Bordag, M., Mohideen, U., and Mostepanenko, V. M. (2001). Phys. Rep., 353:1.Born, M. and Infeld, L. (1934). Proc. Roy. Soc. Lond. A, 147:522.Born, M. and Infeld, L. (1935). Proc. Roy. Soc. Lond. A, 150:141.Born, M. and Wolf, E. (1968). Principles of Optics. Pergamon Press, New York.Born, M. and Wolf, E. (1999). Principles of Optics, 7th ed. Cambridge University Press, Cam-

bridge, England.Boyd, R. W. (2003). Nonlinear Optics, 2nd ed. Academic Press, New York.Boyer, T. H. (1974). Phys. Rev. A, 9:2078.Boyer, T. H. (2003). Am. J. Phys., 71:990.Bowmeester, D., Ekert, A., and Zeilinger, A., editors (2000). The Physics of Quantum Information.

Springer-Verlag, Berlin.Bozhevolnyi, S. I., Boltasseva, A., Søndergaard, T., Nikolajsen, T., and Leosson, K. (2005). Opt.

Commun., 250:328.

Page 465: Quantum Aspects of Light Propagation

References 459

Bressi, G., Carugno, G., Onofrio, R., and Ruoso, G. (2002). Phys. Rev. Lett., 88:041804.Brevik, I., Aarseth, J. B., and Høye, J. S. (2002). Phys. Rev. E, 66:026119.Bronk, B. V. (1965). J. Math. Phys. (NY), 6:228.Brouwer, P. W. (1998). Phys. Rev. B, 57:10526.Brouwer, P. W. and Beenakker, C. W. J. (1996). J. Math. Phys., 37:4904.Brown, S. A. and Dalton, B. J. (2001a). J. Mod. Opt., 48:597.Brown, S. A. and Dalton, B. J. (2001b). J. Mod. Opt., 48:639.Brown, S. A. and Dalton, B. J. (2002). J. Mod. Opt., 49:1009.Brown, L. S. and Maclay, G. J. (1969). Phys. Rev., 184:1272.Bruce, N. A. and Chalker, J. T. (1996). J. Phys. A: Math. Gen., 29:3761.Brun, T. A. and Barnett, S. M. (1998). J. Mod. Opt., 45:777.Bruno, P. (2002). Phys. Rev. Lett., 88:240401.Burin, A. L., Ratner, M. A., Cao, H., and Chang, R. P. H. (2001). Phys. Rev. Lett., 87:215503.Burin, A. L., Ratner, M. A., Cao, H., and Chang, S. H. (2002). Phys. Rev. Lett., 88:093904.Burnham, D. C. and Weinberg, D. L. (1970). Phys. Rev. Lett., 25:84.Caetano, D. P. and Souto Ribeiro, P. H. (2004). Opt. Commun., 239:121.Cahill, K. E. and Glauber, R. J. (1999). Phys. Rev. A, 59:1538.Campos, R. A., Saleh, B. E. A., and Teich, M. A. (1990). Phys. Rev. A, 42:4127.Cao, H. (2003). In Wolf, E., editor, Progress in Optics, Vol. 45, pp. 317–370. Elsevier Science

(North-Holland), Amsterdam.Cao, H., Ling, Y., Xu, J. Y., Cao, C. Q., and Kumar, P. (2001). Phys. Rev. Lett., 86:4524.Cao, H., Xu, J. Y., Zhang, D. Z., Chang, S.-H., Ho, S. T., Seelig, E. W., Liu, S., and Chang, R. P.

H. (2000). Phys. Rev. Lett., 84:5584.Cao, H., Zhao, Y. G., Ho, S. T., Seelig, E. W., Wang, Q. H., and Chang, R. P. (1999). Phys. Rev.

Lett., 82:2278.Carmichael, H. J. (1987). J. Opt. Soc. Am. B, 4:1565.Carniglia, C. K. and Mandel, L. (1971). Phys. Rev. D, 3:280.Carter, S. J. and Drummond, P. D. (1991). Phys. Rev. Lett., 67:3757.Carter, S. J., Drummond, P. D., Reid, M. D., and Shelby, R. M. (1987). Phys. Rev. Lett., 58:1841.Casado, A., Fernandez-Rueda, A., Marshall, T., Risco-Delgado, R., and Santos, E. (1997a). Phys.

Rev. A, 55:3879.Casado, A., Marshall, T. W., and Santos, E. (1997b). J. Opt. Soc. Am. B, 14:494.Case, K. M. and Zweifel, P. F. (1976). Linear Transport Theory. Addison-Wesley,

Reading, MA.Casimir, H. B. G. and Polder, D. (1948). Phys. Rev., 73:360.Caticha, A. and Caticha, N. (1992). Phys. Rev. B, 46:479.Caves, C. M. (1981). Phys. Rev. D, 23:1693.Caves, C. M. (1982). Phys. Rev. D, 26:1817.Caves, C. M. and Crouch, D. D. (1987). J. Opt. Soc. Am. B, 4:1535.Caves, C. M. and Schumaker, B. L. (1985). Phys. Rev. A, 31:3068.Centini, M., Perina, Jr., J., Sciscione, L., Sibilia, C., Scalora, M., Bloemer, M. J., and Bertolotti,

M. (2005). Phys. Rev. A, 72:033806.Cerboneschi, E., Renzoni, F., and Arimondo, E. (2002). J. Opt. B: Quantum Semiclass. Opt.,

4:S267.Chabanov, A. A., Stoytchev, M., and Genack, A. Z. (2000). Nature, 404:850.Chalker, J. T. and Mehlig, B. (1998). Phys. Rev. Lett., 81:3367.Chan, H. B., Aksyuk, V. A., Kleiman, R. N., Bishop, D. J., and Capasso, F. (2001a). Science,

291:1941.Chan, H. B., Aksyuk, V. A., Kleiman, R. N., Bishop, D. J., and Capasso, F. (2001b). Phys. Rev.

Lett., 87:211801.Chefles, A. and Barnett, S. M. (1996). J. Mod. Opt., 43:709.Chen, F., Klimchitskaya, G. L., Mohideen, U., and Mostepanenko, V. M. (2003). Phys. Rev. Lett.,

90:160404.

Page 466: Quantum Aspects of Light Propagation

460 References

Chen, F., Klimchitskaya, G. L., Mostepanenko, V. M., and Mohideen, U. (2006). Phys. Rev. Lett.,97:170402.

Chen, F., Mohideen, U., Klimchitskaya, G. L., and Mostepanenko, V. M. (2002a). Phys. Rev. Lett.,88:101801.

Chen, F., Mohideen, U., Klimchitskaya, G. L., and Mostepanenko, V. M. (2002b). Phys. Rev. A,66:032113.

Chen, F., Mohideen, U., Klimtchitskaya, G. L., and Mostepanenko, V. M. (2005). Phys. Rev. A,72:020101.

Chen, F., Mohideen, U., Klimtchitskaya, G. L., and Mostepanenko, V. M. (2006). Phys. Rev. A,74:022103.

Chen, F., Mohideen, U., Klimchitskaya, G. L., and Mostepanenko, V. M. (2007). Phys. Rev. Lett.,98:068901.

Cheng, Y.-J. and Siegman, A. E. (2003). Phys. Rev. A, 68:043808.Chew, W. C. (1995). Waves and Fields in Inhomogeneous Media. IEEE Press, New York.Chizhov, A. V., Nazmitdinov, R. G., and Shumovsky, A. S. (1991). Quantum Opt., 3:1.Chui, S. T. and Hu, L. B. (2002). Phys. Rev. B, 65:144407.Clemmow, P. C. (1996). The Plane Wave Spectrum Representation of Electromagnetic Fields.

Oxford University Press/IEEE Press, Oxford, UK.Cohen-Tannoudji, C., Dupont-Roc, J., and Grynberg, G. (1989). Photons and Atoms: Introduction

to Quantum Electrodynamics. Wiley, New York.Collett, M. J. and Gardiner, C. W. (1984). Phys. Rev. A, 30:1386.Collett, M. J., Loudon, R., and Gardiner, C. W. (1987). J. Mod. Opt., 34:881.Collett, M. J. and Walls, D. F. (1985). Phys. Rev. A, 35:2887.Corona, M. and U’Ren, A. B. (2007). Phys. Rev. A, 76:043829.Coulson, C. A. (1952). Valence. Clarendon Press, Oxford.Crenshaw, M. E. (2003). Phys. Rev. A, 67:033805.Crenshaw, M. E. (2004). Opt. Commun., 235:153.Crenshaw, M. E. and Bowden, C. M. (2000a). Phys. Rev. Lett., 85:1851.Crenshaw, M. E. and Bowden, C. M. (2000b). Phys. Rev. A, 63:013801.Crenshaw, M. E. and Bowden, C. M. (2002). J. Mod. Opt., 49:511.Crouch, D. D. (1988). Phys. Rev. A, 38:508.Cubukcu, E., Aydin, K., Ozbay, E., Foteinopoulou, S., and Soukoulis, C. M. (2003). Phys. Rev.

Lett., 91:207401.Cui, T. J., Hao, Z.-C., Yin, X. X., Hong, W., and Kong, J. A. (2004). Phys. Lett. A, 323:484.Dalton, B. J. and Babiker, M. (1997). Phys. Rev. A, 56:905.Dalton, B. J., Barnett, S. M., and Garraway, B. M. (2003). In Bigelow, N. P., Eberly, J. H.,

Stroud, Jr., C. R., Walmsley, I. A., editors, Coherence and Quantum Optics VIII, Proceed-ings of the Eighth Rochester Conference on Coherence and Quantum Optics, p. 495. KluwerAcademic/Plenum Publishers, New York.

Dalton, B. J., Barnett, S. M., and Knight, P. L. (1999a). J. Mod. Opt., 46:1107.Dalton, B. J., Barnett, S. M., and Knight, P. L. (1999b). J. Mod. Opt., 46:1315.Dalton, B. J., Barnett, S. M., and Knight, P. L. (1999c). J. Mod. Opt., 46:1495.Dalton, B. J., Barnett, S. M., and Knight, P. L. (1999d). J. Mod. Opt., 46:1559.Dalton, B. J., Guerra, E. S., and Knight, P. L. (1996). Phys. Rev. A, 54:2292.Dalton, B. J. and Knight, P. L. (1999a). J. Mod. Opt., 46:1817.Dalton, B. J. and Knight, P. L. (1999b). J. Mod. Opt., 46:1839.D’Aguanno, G., Akozbek, N., Mattiucci, N., Scalora, M., Bloemer, M. J., and Zheltikov, A. M.

(2005). Opt. Lett., 30:1998.D’Aguanno, G., Centini, M., Scalora, M., Sibilia, C., Bertolotti, M., Bloemer, M. J., and Bowden,

C. M. (2002). J. Opt. Soc. Am. B, 19:211.da Silva, J. C., Khanna, F. C., Matos Neto, A., and Santana, A. E. (2002). Phys. Rev. A, 66:052101.Davison, B. and Sykes, J. B. (1958). Neutron Transport Theory. Oxford University Press, New

York.

Page 467: Quantum Aspects of Light Propagation

References 461

Dechoum, K., Marshall, T. W., and Santos, E. (2000). J. Mod. Opt., 47:1273.de Groot, S. R. (1969). In de Boer, G. E. U. J., editor, Studies in Statistical Mechanics, Vol. 4,

p. 11. North-Holland, Amsterdam.de Lange, O. L. and Raab, R. E. (2006). Am. J. Phys., 74:301.Della Villa, A., Enoch, S., Tayeb, G., Pierro, V., Galdi, V., and Capolino, F. (2005). Phys. Rev. Lett.,

94:183903.den Dekker, A. J. and van den Bos, A. (1997). J. Opt. Soc. Am. B, 14:547.Deng, C., Haus, J. W., Sarangan, A., Mahfoud, A., Sibilia, C., Scalora, M., and Zheltikov, A.

(2006). Laser Physics, 16:927.Derjaguin, B. V. (1960). Sci. Am., 203:47.Derjaguin, B. V. and Abrikosova, I. I., (1957). Sov. Phys. JETP, 3:819.Deutsch, I. H. (1991). Am. J. Phys., 59:834.Deutsch, D. and Candelas, P. (1979). Phys. Rev. D, 20:3063.Deutsch, I. H. and Garrison, J. C. (1991a). Opt. Commun., 86:311.Deutsch, I. H. and Garrison, J. C. (1991b). Phys. Rev. A, 43:2498.Diao, L. and Blair, S. (2007). J. Opt. A: Pure Appl. Opt., 9:972.Di Stefano, O., Savasta, S., and Girlanda, R. (1999). Phys. Rev. A, 60:1614.Di Stefano, O., Savasta, S., and Girlanda, R. (2000). Phys. Rev. A, 61:023803.Di Stefano, O., Savasta, S., and Girlanda, R. (2001a). J. Mod. Opt., 48:67.Di Stefano, O., Savasta, S., and Girlanda, R. (2001b). J. Opt. B: Quantum Semiclass. Opt., 3:288.Dirac, P. A. M. (1927). Proc. Roy. Soc. Lond. A, 11:4243.Dirac, P. A. M. (1964). Lectures on Quantum Mechanics. Belfer Graduate School of Science, New

York.Dittes, F. M. (2000). Phys. Rep., 339:215.Dolling, G., Wegener, M., Soukoulis, C. M., and Linden, S. (2007). Opt. Lett., 32:53.Dorokhov, O. (1982). Pis. Zh. Eksp. Teor. Fiz., 36:259.Doron, E. and Frischat, S. D. (1995). Phys. Rev. Lett., 75:3661.Drummond, P. D. (1990). Phys. Rev. A, 42:6845.Drummond, P. D. (1994). In Evans, M. V. and Kielich, S., editors, Modern Nonlinear Optics, Part

3, p. 379, J. Wiley, New York.Drummond, P. D. and Carter, S. J. (1987). J. Opt. Soc. Am. B, 4:1565.Drummond, P. D., Carter, S. J., and Shelby, R. M. (1989). Opt. Lett., 14:373.Drummond, P. D. and Hardman, A. D. (1993). Europhys. Lett., 21:279.Drummond, P. D. and He, H. (1997). Phys. Rev. A, 56:R1107.Drummond, P. D. and Hillery, M. (1999). Phys. Rev. A, 59:691.Dung, H.-T., Buhmann, S. Y., Knoll, L., Welsch, D.-G., Scheel, S., and Kastel, J. (2003). Phys.

Rev. A, 68:043816.Dung, H.-T., Knoll, L., and Welsch, D.-G. (1998). Phys. Rev. A, 57:3931.Dung, H.-T., Knoll, L., and Welsch, D.-G. (2000). Phys. Rev. A, 62:053804.Dung H.-T., Knoll, L., and Welsch, D.-G. (2001). Phys. Rev. A, 64:013804.Dung, H.-T., Knoll, L., and Welsch, D.-G. (2002a). Phys. Rev. A, 65:043813.Dung, H.-T., Knoll, L., and Welsch, D.-G. (2002b). Phys. Rev. A, 66:063810.Dutra, S. M. and Furuya, K. (1997). Phys. Rev. A, 55:3832.Dutra, S. M. and Furuya, K. (1998a). Europhys. Lett., 43:13.Dutra, S. M. and Furuya, K. (1998b). Phys. Rev. A, 57:3050.Dutra, S. M. and Nienhuis, G. (2000). Phys. Rev. A, 62:063805.Dyson, F. J. (1962). J. Math. Phys. (NY), 3:157.Dzyaloshinskii, I. E., Lifshitz, E. M., and Pitaevskii, L. P. (1960). Sov. Phys. JETP, 9:1282.Dzyaloshinskii, I. E., Lifshitz, E. M., and Pitaevskii, L. P. (1961). Adv. Phys., 10:165.Dzyaloshinskii, I. E. and Pitaevskii, L. P. (1959). Sov. Phys. JETP, 9:1282.Emig, T. (2003). Europhys. Lett., 62:466.Emig, R. (2007). Phys. Rev. Lett., 98:060801.Emig, R., Graham, N., Jaffe, R. L., and Kardar, M. (2007). Phys. Rev. Lett., 99:170403.

Page 468: Quantum Aspects of Light Propagation

462 References

Engheta, N. (1998). Microwave Opt. Technolog. Lett., 17:86.Fano, U. (1956). Phys. Rev., 103:1202.Fano, U. (1961). Phys. Rev., 124:1866.Felbacq, D. and Bouchitte, G. (2005). Phys. Rev. Lett., 94:183902.Fermani, R., Scheel, S., and Knight, P. L. (2007). Phys. Rev. A, 75:062905.Finlayson, N., Banyai, W. C., Wright, E. M., Seaton, C. T., Stegeman, G. I., Cullen, T. J., and

Ironside, C. N. (1988). Appl. Phys. Lett., 53:1144.Fisher, R. A. (1983). Optical Phase Conjugation. Academic Press, New York.Fiurasek, J., Krepelka, J., and Perina, J. (1999a). Opt. Commun., 167:115.Fiurasek, J., Krepelka, J., and Perina, J. (1999b). Acta Phys. Slovaca, 49:689.Fiurasek, J. and Perina, J. (1998). Acta Phys. Slovaca, 48:361.Fiurasek, J. and Perina, J. (1999). Quantum statistics of light in Raman and Brillouin couplers with

losses. In Hrabovsky, M., Strba, A., and Urbanczyk, W., editors, SPIE, Vol. 3820, pp. 23–33.SPIE, Bellingham.

Fiurasek, J. and Perina, J. (2000a). J. Opt. B: Quantum Semiclass. Opt., 2:10.Fiurasek, J. and Perina, J. (2000b). J. Mod. Opt., 47:1399.Fleischhauer, M. (1999). Phys. Rev. A, 60:2534.Florescu, L. and John, S. (2004). Phys. Rev. E, 69:046603.Foteinopoulou, S., Economou, E. N., and Soukoulis, C. M. (2003). Phys. Rev. Lett., 90:107402.Foteinopoulou, S. and Soukoulis, C. M. (2003). Phys. Rev. B, 67:235107.Fox, A. G. and Li, T. (1961). Bell Sys. Tech. J., 40:453.Frahm, K. M., Schomerus, H., Patra, M., and Beenakker, C. W. J. (2000). Europhys. Lett., 49:48.Franson, J. D. (1989). Phys. Rev. Lett., 62:2205.Friberg, S. R., Machida, S., and Yamamoto, Y. (1992). Phys. Rev. Lett., 69:3165.Fu, H. and Berman, P. R. (2005). Phys. Rev. A, 72:022104.Fulling, S. A., Kaplan, L., and Wilson, J. H. (2007). Phys. Rev. A, 76:012118.Fyodorov, Y. V. and Sommers, H.-J. (1997). J. Math. Phys., 38:1918.Galvez, E. J., Holbrow, G. H., Pysher, M. J., Martin, J. W., Cortemanche, N., Heilig, L., and

Spencer, J. (2005). Am. J. Phys., 73:127.Garabedian, P. A. (1964). Partial Differential Equations. Chelsea, New York, Chap. 10.Gardiner, C. W. and Collett, M. J. (1985). Phys. Rev. A, 31:3761.Garraway, B. M. (1997a). Phys. Rev. A, 55:2290.Garraway, B. M. (1997b). Phys. Rev. A, 55:4636.Garraway, B. M. and Knight, P. L. (1996). Phys. Rev. A, 54:3592.Gatti, A. (2003). In Bigelow, N. P., Eberly, J. H., Stroud, Jr., C. R., Walmsley, I. A., editors, Coher-

ence and Quantum Optics VIII, Proceedings of the Eighth Rochester Conference on Coherenceand Quantum Optics, p. 129. Kluwer Academic/Plenum Publishers, New York.

Gatti, A., Brambilla, E., Bache, M., Lugiato, L. A. (2004). Phys. Rev. A, 70:013802.Gatti, A., Brambilla, E., and Lugiato, L. (2003). Phys. Rev. Lett., 90:133603.Gatti, A. and Lugiato, L. A. (1995). Phys. Rev. A, 52:1675.Gea-Banacloche, J., Lu, N., Pedrotti, L. M., Prasad, S., Scully, M. O., and Wodkiewicz, K. (1990a).

Phys. Rev. A, 41:369.Gea-Banacloche, J., Lu, N., Pedrotti, L. M., Prasad, S., Scully, M. O., and Wodkiewicz, K. (1990b).

Phys. Rev. A, 41:381.Genet, C., Lambrecht, A., Maia Neto, P., and Reynaud, S. (2003). Europhys. Lett., 62:484.Geyer, B., Klimchitskaya, G. L., and Mostepanenko, V. M. (2003). Phys. Rev. A, 67:062102.Ghosh, R., Hong, C. K., Ou, Z. Y., and Mandel, L. (1986). Phys. Rev. A, 34:3962.Ghosh, R. and Mandel, L. (1987). Phys. Rev. Lett., 59:1903.Ginzburg, V. L. (1940). J. Phys. USSR, 2:441.Glauber, R. J. (1963). Phys. Rev., 130:2529.Glauber, R. J. (1965). In Blandin, A., DeWitt, C., and Cohen-Tannoudji, C., editors, Quantum

Optics and Electronics, p. 144. Gordon and Breach, London.

Page 469: Quantum Aspects of Light Propagation

References 463

Glauber, R. J. and Lewenstein, M. (1989). In Tombesi, P. and Pike, E. R., editors, Squeezed andNonclassical Light, p. 203. Plenum Press, New York.

Glauber, R. J. and Lewenstein, M. (1991). Phys. Rev. A, 43:467.Gmachl, C., Capasso, F., Narimanov, E. E., Nockel, J. U., Stone, A. D., Faist, J., Sivco, D. L., and

Cho, A. Y. (1998). Science, 280:1556.Gordon, J. P. (1986). Opt. Lett., 11:662.Gornik, E. (1998). Science, 280:1544.Gouedard, C., Husson, D., Sauteret, C., Auzel, F., and Migus, A. (1993). J. Opt. Soc. Am. B,

10:2358.Graf, H.-D., Harney, H. L., Lengeler, H., Lewenkopf, C. H., Rangacharyulu, C., Richter, A.,

Schardt, P., and Weidenmuller, H. A. (1992). Phys. Rev. Lett., 69:1296.Graham, R. and Haken, H. (1968). Z. Phys., 213:420.Greenberger, D. M., Horne, M. A., and Zeilinger, A. (1993). Phys. Today, 46(8):22.Greiner, W. (1998). Classical Electrodynamics. Springer, New York.Grimes, C. A. and Grimes, D. M. (1991). Phys. Rev. B, 43:10780.Gruner, T. and Welsch, D.-G. (1995). Phys. Rev. A, 51:3246.Gruner, T. and Welsch, D.-G. (1996a). Phys. Rev. A, 53:1818.Gruner, T. and Welsch, D.-G. (1996b). Phys. Rev. A, 54:1661.Guo, W. (2002). Am. J. Phys., 70:1039.Guo, W. (2005). Opt. Commun., 250:137.Guo, W. (2007). Phys. Rev. A, 76:023834.Gutzwiller, M. C. (1990). Chaos in Classical and Quantum Mechanics. Springer, New York.Hackenbroich, G., Viviescas, C., Elattari, B., and Haake, F. (2001). Phys. Rev. Lett., 86:5262.Hackenbroich, G., Viviescas, C., and Haake, F. (2002). Phys. Rev. Lett., 89:083902.Haken, H. (1970). Laser Theory. Springer-Verlag, Berlin.Hansen, P. B., Kloch, A., Aakjer, T., and Rasmussen, T. (1995). Opt. Commun., 119:178.Harbers, R., Moll, N., Mahrt, R. F., Erni, D., and Bachtold, W. (2005). J. Opt. A: Pure Appl. Opt.,

7:S230.Hatami-Hanza, H. and Chu, P. L. (1995). Opt. Commun., 119:347.Haus, H. A. and Kartner, F. X. (1992). Phys. Rev. A, 46:R1175.Haus, H. A. and Lai, Y. (1990). J. Opt. Soc. Am. B, 7:386.Haus, H. A., Watanabe, K., and Yamamoto, Y. (1989) J. Opt. Soc. Am. B, 6:1138.Hawton, M. (1999). Phys. Rev. A, 59:3223.Healey, W. P. (1982). Nonrelativistic Quantum Electrodynamics. Academic Press, London.Heitler, W. (1954). The Quantum Theory of Radiation, 3rd edition. Oxford University

Press/Clarendon Press, Oxford.Henry, C. H. and Kazarinov, R. F. (1996). Rev. Mod. Phys., 68:801.Herec, J. (1999). Acta Phys. Slovaca, 49:731.Herzog, T. J., Rarity, J. G., Weinfurter, H., and Zeilinger, A. (1994). Phys. Rev. Lett., 72:629.Hillery, M. and Drummond, P. D. (2001). Phys. Rev. A, 64:013815.Hillery, M. and Mlodinow, L. D. (1984). Phys. Rev. A, 30:1860.Hillery, M. and Mlodinow, L. D. (1997). Phys. Rev. A, 55:678.Hinds, E. A. (1990). In Bates, D. and Bederson, B., editors, Advances in Atomic, Molecular and

Optical Physics, Vol. 28, p. 237. Academic Press, New York.Ho, K. C., Leung, P. T., Maassen van den Brink, A., and Young, K. (1998). Phys. Rev. E, 58:2965.Ho, S.-T. and Kumar, P. (1993). J. Opt. Soc. Am. B, 10:1620.Hopfield, J. J. (1958). Phys. Rev., 112:1555.Horne, M., Shimony, A., and Zeilinger, A. (1990). In Anandan, J. S., editor, Quantum Coherence,

pp. 356–372. World Scientific, Singapore.Høye, J. S., Brevik, I., and Aarseth, J. B. (2001). Phys. Rev. E, 63:051101.Hradil, Z. (1996). Phys. Rev. A, 53:3687.Huang, Y. and Gao, L. (2003). Phys. Lett. A, 318:592.Huang, X. and Schaich, W. L. (2004). Am. J. Phys., 72:1232.

Page 470: Quantum Aspects of Light Propagation

464 References

Huang, D., Swanson, E. A., Lin, C. P., Schuman, J. S., Stinson, W. G., Chang, W., Hee, M. R.,Flotte, T., Gregory, K., Puliafito, C. A., and Fujimoto, J. G. (1991). Science, 254:1178.

Hughes, B. D. (1996). Random Walks and Random Environments, Vol. 2. Clarendon Press, Oxford.Hughes, S. (2005a). Opt. Lett., 30:1393.Hughes, S. (2005b). Phys. Rev. Lett., 94:033903.Hughes, S. (2005c). Phys. Rev. Lett., 94:227402.Huttner, B. and Barnett, S. M. (1992a). Europhys. Lett., 18:487.Huttner, B. and Barnett, S. M. (1992b). Phys. Rev. A, 46:4306.Huttner, B., Baumberg, J. J., and Barnett, S. M. (1991). Europhys. Lett., 16:177.Huttner, B., Serulnik, S., and Ben-Aryeh, Y. (1990). Phys. Rev. A, 42:5594.Hynne, F. and Bullough, R. K. (1990). Phil. Trans. Roy. Soc. A (London), 330:253.Iannuzzi, D. and Capasso, F. (2003). Phys. Rev. Lett., 91:029101.Imoto, N. (1989). In Kobayashi, S., Egawa, H., Murayama, Y., and Nomura, S., editors, Proc. 3rd

Int. Symp. on Foundations of Quantum Mechanics (Tokyo), p. 306. Physical Society of Tokyo,Tokyo.

Inagaki, T. (1998). Phys. Rev. A, 57:2204.Inoue, T. and Hori, H. (2001). Phys. Rev. A, 63:063805.Intravaia, F., Henkel, C., and Lambrecht, A. (2007). Phys. Rev. A, 76:033820.Itzykson, C. and Zuber, J. B. (1980). Quantum Field Theory. McGraw-Hill, New York.Jaekel, M.-T. and Reynaud, S. (1991). J. Phys. I, 1:1395.Janowicz, M., Reddig, D., and Holthaus, M. (2003). Phys. Rev. A, 8:043823.Janszky, J., Sibilia, C., Bertolotti, M., Adam, P., and Petak, A. (1995). J. Opt. B: Quantum Semi-

class. Opt., 7:509.Janszky, J., Sibilia, C., Bertolotti, M., and Yushin, Y. (1988). J. Mod. Opt., 35:1757.Jauch, J. M. and Watson, K. M. (1948). Phys. Rev., 74:950.Javanainen, J., Ruostekoski, J., Vestergaard, B., and Francis, M. R. (1999). Phys. Rev. A, 64:1959.Jeffers, J. and Barnett, S. M. (1994). J. Mod. Opt., 41:1121.Jeffers, J., Barnett, S. M., Loudon, R., Matloob, R., and Artoni, M. (1996). Opt. Commun., 131:66.Jeffers, J. R., Imoto, N., and Loudon, R. (1993). Phys. Rev. A, 47:3346.Jensen, S. M. (1982). IEEE J. Quantum Electron., QE-18:1580.Jiang, X., Li, Q., and Soukoulis, C. M. (1999). Phys. Rev. B, 59:R9007.Jiang, X. and Soukoulis, C. M. (1999). Phys. Rev. B, 59:6159.Jiang, X. and Soukoulis, C. M. (2000). Phys. Rev. Lett., 85:70.Joannapoulos, J. D., Meade, R. D., and Winn, J. N. (1995). Photonic Crystals. Princeton University,

Princeton, NJ.John, S. (1984). Phys. Rev. Lett., 53:2169.John, S. (1987). Phys. Rev. Lett., 58:2486.John, S. and Pang, G. (1996). Phys. Rev. A, 54:3642.John, S., Pang, G., and Yang, Y. (1996). J. Biomed. Opt., 1:180.John, S. and Stephen, M. J. (1983). Phys. Rev. B, 28:6358.Joobeur, A., Saleh, B. E. A., and Teich, M. C. (1994). Phys. Rev. A, 50:3349.Karpati, A., Adam, P., Janszky, J., Bertolotti, M., and Sibilia, C. (2000). J. Opt. B: Quantum Semi-

class. Opt., 2:133.Kelley, P. L. and Kleiner, W. H. (1964). Phys. Rev., 136:316.Kennedy, T. A. B. and Wright, E. M. (1988). Phys. Rev. A, 38:212.Kenneth, O., Klich, I., Mann, A., and Revzen, M. (2002). Phys. Rev. Lett., 89:033001.Kenneth, O., Klich, I., Mann, A., and Revzen, M. (2003). Phys. Rev. Lett., 91:029102.Khanbekyan, M., Knoll, L., and Welsch, D.-G. (2003). Phys. Rev. A, 67:063812.Khosravi, H. and Loudon, R. (1991). Proc. Roy. Soc. Lond. A, 433:337.Khosravi, H. and Loudon, R. (1992). Proc. Roy. Soc. Lond. A, 436:373.Kitagawa, M. and Yamamoto, Y. (1986). Phys. Rev. A, 34:3974.Kittel, C. (1987). Quantum Theory of Solids, 2nd ed. J. Wiley, New York.Klimchitskaya, G. L. and Mostepanenko, V. M. (2001). Phys. Rev. A, 63:062108.

Page 471: Quantum Aspects of Light Propagation

References 465

Klingshirn, C. F. (1995). Semiconductor Optics. Springer-Verlag, Berlin, p. 57.Klyshko, D. N. (1988). Photons and Nonlinear Optics. Gordon and Breach, New York.Knoll, L. and Leonhardt, U. (1992). J. Mod. Opt., 39:1253.Knoll, L., Scheel, S., Schmidt, E., Welsch, D.-G., and Chizhov, A. V. (1999). Phys. Rev. A, 59:4716.Knoll, L., Scheel, S., and Welsch, D.-G. (2001). In Perina, J., editor, Coherence and Statistics of

Photons and Atoms, p. 1. J. Wiley, New York.Knoll, L., Vogel, W., and Welsch, D.-G. (1986). J. Opt. Soc. Am. B, 3:1315.Knoll, L., Vogel, W., and Welsch, D.-G. (1987). Phys. Rev. A, 36:3803.Knoll, L., Vogel, W., and Welsch, D.-G. (1990). Phys. Rev. A, 42:503.Knoll, L., Vogel, W., and Welsch, D.-G. (1991). Phys. Rev. A, 43:543.Knoll, L. and Welsch, D.-G. (1992). Prog. Quantum Electron., 16:135.Kobe, D. H. (1999). Phys. Lett. A, 253:7.Kogan, Sh. (1996). Electronic Noise and Fluctuations in Solids. Cambridge University Press, Cam-

bridge.Kolobov, M. I. (1999). Rev. Mod. Phys., 71:1539.Kolobov, M. I. (2007). In Kolobov, M. I., editor, Quantum Imaging, p. 316. Springer-Verlag, New

York.Kolobov, M. I. and Kumar, P. (1993). Opt. Lett., 18:849.Kolobov, M. I. and Lugiato, L. A. (1995). Phys. Rev. A, 52:4930.Kolobov, M. I. and Sokolov, I. V. (1989a). Zh. Eksp. Teor. Fiz., 96:1945.Kolobov, M. I. and Sokolov, I. V. (1989b). Sov. Phys. JETP, 69:1097.Kondrat’ev, I. G. and Smirnov, A. I. (2003). Phys. Rev. Lett., 91:249401.Korolkova, N., Loudon, R., Gardavsky, G., Hamilton, M. W., and Leuchs, G. (2001). J. Mod. Opt.,

48:1339.Korolkova, N. and Perina, J. (1997a). Opt. Commun., 137:263.Korolkova, N. and Perina, J. (1997b). Opt. Commun., 136:135.Korolkova, N. and Perina, J. (1997c). J. Mod. Opt., 44:1525.Kovalev, V. F., Bychenkov, V. Yu., and Tikhonchuk, V. T. (2000). Phys. Rev. A, 61:033809.Krause, D. E., Decca, R. S., Lopez, D., and Fischbach, E. (2007). Phys. Rev. Lett., 98:050403.Kravtsov, V. E., Agranovich, V. M., and Grigorishin, K. I. (1991). Phys. Rev. B, 44:4931.Kubo, R. (1962). J. Phys. Soc. Jpn., 17:1100.Kuchment, P. (2004). Waves in Random Media, 14:S107.Kuga, Y. and Ishimaru, A. (1984). J. Opt. Soc. Am. A, 8:831.Kumar, P. and Shapiro, J. H. (1984). Phys. Rev. A, 30:1568.Kupiszewska, D. (1992). Phys. Rev. A, 46:2286.Kupiszewska, D. and Mostowski, J. (1990). Phys. Rev. A, 41:4636.Kurizki, G. and Haus, J. W., editors. (1994). Journal of Modern Optics, Special Issue Photonic

Band Structures, Vol. 41, pp. 171–408. Taylor and Francis, London.Kurokawa, N. and Wakayama, M. (2002). Indagationes Mathematicae, Nova Series, 13:63.Kweon, G. and Lawandy, N. M. (1995). Opt. Commun., 118:388.Lai, W. K., Buzek, V., and Knight, P. L. (1991). Phys. Rev. A, 43:6323.Lai, Y. and Haus H. A. (1989). Phys. Rev. A, 40:844.Lamoreaux, S. K. (1997). Phys. Rev. Lett., 78:5.Lamoreaux, S. K. (1999). Am. J. Phys., 67:850.Lamoreaux, S. K. (2005). Rep. Prog. Phys., 68:201.Landau, L. D. and Lifshitz, E. M. (1960). Electrodynamics of Continuous Media. Pergamon Press,

Oxford.Landau, L. D. and Lifshitz, E. M. (1963). Electrodynamics of Continuous Media. Pergamon Press,

Oxford.Landau, L. and Lifshitz, E. (1980). Statistical Physics, Part II. Pergamon Press, Oxford.Landau, L. D., Lifshitz, E. M., and Pitaevskii, L. P. (1984). Electrodynamics of Continuous Media,

Pergamon Press, Oxford.Lang, R. and Scully, M. O. (1973). Opt. Commun., 9:331.

Page 472: Quantum Aspects of Light Propagation

466 References

Lang, R., Scully, M. O., and Lamb, Jr, W. E. (1973). Phys. Rev. A, 7:1788.Law, C. K., Walmsley, I. A., and Eberly, J. H. (2000). Phys. Rev. Lett., 84:5304.Lawandy, N. M. and Balachandran, R. M. (1995). Nature, 373:204.Lawandy, N. M., Balachandran, R. M., Gomes, A. S. L., and Sauvain, E. (1994). Nature, 368:436.Lax, M., Louisell, W. H., and McKnight, W. B. (1974). Phys. Rev. A, 11:1365.LeBoeuf, P. and Saraceno, M. (1990). J. Phys. A: Math. Gen., 23:1745.Ledinegg, E. (1966). Zeit. Phys., 191:177.Lenac, Z. and Tomas, M. S. (2007). Phys. Rev. A, 75:042101.Leonhardt, U. (2000). Phys. Rev. A, 62:012111.Leonhardt, U. and Philbin, T. G. (2007a). New J. Phys., 9:254.Leonhardt, U. and Philbin, T. G. (2007b). J. Opt. A: Pure Appl. Opt., 9:S289.Leonhardt, U. and Piwnicki, P. (2001). J. Mod. Opt., 48:977.Letokhov, V. S. (1968). Sov. Phys. JETP, 26:835.Leutheuser, V., Langbein, U., and Lederer, F. (1990). Opt. Commun., 75:251.Levenson, M. D., Shelby, R. M., Aspect, A., Reid, M., and Walls, D. F. (1985). Phys. Rev. A,

32:1550.Li, L. W., Kooi, P. S., Leong, M. S., and Yeo, T. S. (1994). IEEE Transactions Microwave Theory

Tech., 42:2302.Li, Y., Mikami, H., Wang, H., and Kobayashi, T. (2005). Phys. Rev. A, 72:063801.Lifshitz, E. M. (1955). Zh. Eksp. Teor. Fiz., 29:94.Lifshitz, E. M. (1956). Sov. Phys. JETP, 2:73.Lifshitz, E. M. and Pitaevskii, L. P. (1980). Statistical Physics: Part 2. Pergamon, Oxford.Linares, J. and Nistal, N. C. (2003). J. Mod. Opt., 50:781.Lindblad, G. (1976). Commun. Math. Phys., 48:119.Ling, Y., Cao, H., Burin, A. L., Ratner, M. A., Liu, X., and Chang, R. P. (2001). Phys. Rev. A,

64:063808.Liu, G. J., Liu, J., Liang, B. M., Li, Q., and Jin, G. L. (2003). Opt. Lett., 28:1347.Lodahl, P., van Driel, A. F., Nikolaev, I. S., Irman, A., Overgaag, K., Vanmaekelbergh, D., and Vos,

W. L. (2004). Nature (London), 430:654.Lombardi, A. M. (2002). J. Opt. B: Quantum Semiclass. Opt., 4:S473.London, F. (1930). Z. Phys., 63:245.Loudon, R. (1963). Proc. Phys. Soc., 83:393.Loudon, R. (1970). J. Phys. A, 3:233.Loudon, R. (1998). Phys. Rev. A, 57:4904.Loudon, R. (1999). In Fouque, J.-P., editor, Diffuse Waves in Complex Media, NATO Science Series

C531, pp. 213–247. Kluwer Academic Publishers, Dordrecht.Loudon, R. (2000). The Quantum Theory of Light, 3rd ed. Oxford University Press, New York.Loudon, R. (2002). J. Mod. Opt., 49:821.Loudon, R. (2003). Phys. Rev. A, 68:013806.Loudon, R., Barnett, S. M., and Baxter, C. (2005). Phys. Rev. A, 71:063802.Loudon, R. and Knight, P. L. (1987). J. Mod. Opt., 34:709.Louisell, W. H. (1973). Quantum Statistical Properties of Radiation. J. Wiley, New York.Lu, D., Hu, W., Zheng, Y., and Yang, Z. (2003). Opt. Commun., 228:217.Lugiato, L. A., Brambilla, M., and Gatti, A. (1999). In Bederson, B. and Walther, H., editors,

Advances in Atomic, Molecular, and Optical Physics, p. 229. Academic Press, Boston.Lugiato, L. A. and Gatti, A. (1993). Phys. Rev. Lett., 70:3868.Lugiato, L. A. and Marzoli, I. (1995). Phys. Rev. A, 52:4886.Luis, A. and Perina, J. (1996a). Phys. Rev. A, 53:1886.Luis, A. and Perina, J. (1996b). Quantum Semiclass. Opt., 8:39.Lukosz, W. (1971). Physica (Amsterdam), 56:109.Luks, A. and Perinova, V. (2002). In Wolf, E., editor, Progress in Optics, Vol. 45,

pp. 295–431. Elsevier Science, Amsterdam.

Page 473: Quantum Aspects of Light Propagation

References 467

Luks, A. and Perinova, V. (2003). In Nowak, J., Zajac, M., and Masajada, J., editors, Proceedingsof SPIE, Vol. 5259, pp. 22–29. SPIE, Bellingham, WA.

Maclay, J. (2000). Phys. Rev. A, 61:052110.Mandel, L. (1964). Phys. Rev., 136:B1221.Mandel, L. (1966). Phys. Rev., 144:1071.Mandel, L. and Wolf, E. (1995). Optical Coherence and Quantum Optics. Cambridge University

Press, New York.Marcuse, D. (1974). Theory of Dielectric Optical Waveguides. Academic Press, New York.Marinescu, N. (1992). Prog. Quantum Electron., 16:183.Markos, P. (2005). Cs. cas. fyz., 55:531 (in Slovak).Markos, P., Rousochatzakis, I., and Soukoulis, C. M. (2002). Phys. Rev. E, 66:045601.Markos, P. and Soukoulis, C. M. (2002a). Phys. Rev. B, 65:033401.Markos, P. and Soukoulis, C. M. (2002b). Phys. Rev. E, 65:036622.Markushev, V. M., Zolin, V. F., and Briskina, Ch. M. (1986). Sov. J. Quant. Electron., 16:281.Marques, R., Martel, J., Mesa, F., and Medina, F. (2002). Phys. Rev. Lett., 89:183901.Marques, R., Martel, J., Mesa, F., and Medina, F. (2003). Phys. Rev. Lett., 91:249402.Matloob, R. (1999a). Phys. Rev. A, 60:50.Matloob, R. (1999b). Phys. Rev. A, 60:3421.Matloob, R. (2001). Opt. Commun., 192:287.Matloob, R. (2004a). Phys. Rev. A, 69:052110.Matloob, R. (2004b). Phys. Rev. A, 70:022108.Matloob, R. (2005). Phys. Rev. A, 71:062105.Matloob, R. and Falinejad, H. (2001). Phys. Rev. A, 64:042102.Matloob, R., Keshavarz, A., and Sedighi, D. (1999). Phys. Rev. A, 60:3410.Matloob, R., Loudon, R., Artoni, M., Barnett, S. M., and Jeffers, J. (1997). Phys. Rev. A, 55:1623.Matloob, R., Loudon, R., Barnett, S. M., and Jeffers, J. (1995). Phys. Rev. A, 52:4823.Matloob, R. and Pooseh, G. (2000). Opt. Commun., 181:109.Matsko, A. B. and Kozlov, V. V. (2000). Phys. Rev. A, 62:033811.Mazzitelli, F. D., Sanchez, M. J., Scoccola, N. N., and von Stecher, J. (2003). Phys. Rev. A,

76:013807.McDonald, K. T. (2001). Am. J. Phys., 69:607.McKintosh, F. C. and John, S. (1989). Phys. Rev. B, 40:2383.Mehra, J. (1967). Physica, 37:145.Mehta, M. L. (1991). Random Matrices. Academic Press, New York.Mello, P. A., Akkermans, E., and Shapiro, B. (1988a). Phys. Rev. Lett., 61:459.Mello, P. A., Pereyra, P., and Kumar, N. (1988b). Ann. Phys., 181:290.Mendez-Sanchez, R. A., Kuhl, U., Barth, M., Lewenkopf, C. H., and Stockmann, H.-J. (2003).

Phys. Rev. Lett., 91:174102.Mendonca, J. T., Martins, A. M., and Guerreiro, A. (2000). Phys. Rev. E, 62:2989.Messina, R. and Passante, R. (2007a). Phys. Rev. A, 75:042113.Messina, R. and Passante, R. (2007b). Phys. Rev. A, 76:032107.Milburn, G. J., and Walls, D. F. (1981). Opt. Commun., 39:40.Milburn, G. J., Walls, D. F., and Levenson, M. D. (1984). J. Opt. Soc. Am. B, 1:390.Milonni, P. W. (1994). The Quantum Vacuum: An Introduction to Quantum Electrodynamics. Aca-

demic Press, San Diego.Milonni, P. W. (1995). J. Mod. Opt., 42:1991.Milonni, P. W., Cook, R. J., and Goggin, M. E. (1988). Phys. Rev. A, 38:1621.Milonni, P. W. and Maclay, G. J. (2003). Opt. Commun., 228:161.Milton, K. A. (2004). J. Phys. A: Math. Gen., 37:R209.Mishchenko, E. G. and Beenakker, C. W. J. (1999). Phys. Rev. Lett., 83:5475.Mishchenko, E. G., Patra, M., and Beenakker, C. W. J. (2001). Europhys. J. D, 13:289.Mishra, S. and Satpathy, S. (2003). Phys. Rev. B, 68:045121.

Page 474: Quantum Aspects of Light Propagation

468 References

Misirpashaev, T. Sh. and Beenakker, C. W. J. (1998). Phys. Rev. A, 57:2041.Misner, C. W., Thorne, K. C., and Wheeler, J. A. (1970). Gravitation. Freeman, San Francisco.Mista, Jr., L. (1999). Acta Phys. Slovaca, 49:737.Mista, Jr., L. and Perina, J. (1997). Czech. J. Phys., 47:629.Mogilevtsev, D. N., Korolkova, N., and Perina, J. (1996). J. Opt. B: Quantum Semiclass. Opt.,

8:1169.Mogilevtsev, D. N., Korolkova, N., and Perina, J. (1997). J. Mod. Opt., 44:1293.Mohideen, U. and Roy, A. (1998). Phys. Rev. Lett., 81:4549.Monken, C. H., Ribeiro, P. H. S., and Padua, S. (1998). Phys. Rev. A, 57:3123.Morris, G. M. (1989). In Arsenault, H. H., Szoplik, T., and Macukow, B., editors, Optical Process-

ing and Computing, p. 343. Academic Press, New York.Mortensen, N. A. (2005). Opt. Lett., 30:1455.Mortensen, N. A., Folkenberg, J. R., Nielsen, M. D., and Hansen, K. P. (2003). Opt. Lett., 28:1879.Mosset, A., Devaux, F., and Lantz, E. (2005). Phys. Rev. Lett., 94:223603.Mrowczynski, S. and Muller, B. (1994). Phys. Rev. D, 50:7542.Mujumdar, S., Ricci, M., Torre, R., and Wiersma, D. S. (2004). Phys. Rev. Lett., 93:53903.Munday, J. N. and Capasso, F. (2007). Phys. Rev. A, 75:060102.Mustecaplıoglu, O. E. and Oktel, M. O. (2005). Phys. Rev. Lett., 94:220404.Nagasako, E. M., Boyd, R. W., and Agarwal, G. S. (1997). Phys. Rev. A, 55:1412.Naqvi, Q. A. and Abbas, M. (2003). Opt. Commun., 227:143.Nesterenko, V. V. and Pirozhenko, I. G. (1997). Phys. Rev. D, 57:1284.Newton, R. G. (1966). Scattering Theory of Waves and Particles. McGraw-Hill, New York.Nielsen, M. A. and Chuang, I. L. (2000). Quantum Computation and Quantum Information. Cam-

bridge University Press, Cambridge.Nilsson, O., Yamamoto, Y., and Machida, S. (1986). IEEE J. Quant. Electr., QE-2:2043.Nockel, J. U. and Stone, D. (1997). Nature, 385:45.Notomi, M. (2000). Phys. Rev. B, 62:10696.Oughstun, K. E. and Cartwright, N. A. (2005). J. Mod. Opt., 52:1089.Ozbay, E., Guven, K., and Aydin, K. (2007). J. Opt. A: Pure Appl. Opt., 9:S301.Paasschens, J. C. J., Misirpashaev, T. S., and Beenakker, C. W. J. (1996). Phys. Rev. B, 54:11887.Padgett, M., Barnett, S. M., and Loudon, R. (2003). J. Mod. Opt., 50:1555.Paspalakis, E. and Kis, Z. (2002). Phys. Rev. A, 66:025802.Passante, R., Power, E. A., and Thirunamachandran, T. (1998). Phys. Lett. A, 249:77.Passante, R. and Spagnolo, S. (2007). Phys. Rev. A, 76:042112.Patra, M. (2002). Phys. Rev. A, 65:043809.Patra, M. (2003). Phys. Rev. E, 67:016603.Patra, M. and Beenakker, C. W. J. (1999). Phys. Rev. A, 60:4059.Patra, M. and Beenakker, C. W. J. (2000). Phys. Rev. A, 61:063805.Patra, M., Schomerus, H., and Beenakker, C. W. J. (2000). Phys. Rev. A, 61:023810.Pedrotti, F. L. and Pedrotti, L. S. (1993). Introduction to Optics. Prentice-Hall, Englewood Cliffs,

NJ.Peierls, R. (1976). Proc. R. Soc. A (London), 347:75.Peierls, R. (1985). In Bassani, F., Fumi, F., Loudon, R., and Tosi, M. P., editors, Highlights of

Condensed Matter Theory. North-Holland, Amsterdam.Pendry, J. B. (2000). Phys. Rev. Lett., 85:3966.Pendry, J. B. (2003). Nature, 423:22.Pendry, J. B., Holden, A. J., Robbins, D. J., and Stewart, W. J. (1999). IEEE Trans. Microwave

Theory Tech., 47:2075.Pendry, J. B., Holden, A. J., Stewart, W. J., and Youngs, I. (1996). Phys. Rev. Lett., 76:4773.Perina, J. (1981a). Optica Acta, 28:325.Perina, J. (1981b). Optica Acta, 28:1529.Perina, J. (1991). Quantum Statistics of Linear and Nonlinear Optical Phenomena, 2nd edn.

Kluwer Academic Publishers, Dordrecht.

Page 475: Quantum Aspects of Light Propagation

References 469

Perina, J. (1995a). J. Mod. Opt., 42:1517.Perina, J. (1995b). Acta Physica Slovaca, 45:279.Perina, J. and Bajer, J. (1995). J. Mod. Opt., 42:2337.Perina, J. and Krepelka, J. (2005). J. Opt. B: Quantum Semiclass. Opt., 7:246.Perina, J. and Krepelka, J. (2006). Opt. Commun., 265:632.Perina, J., Krepelka, J., Perina, Jr., J., Bondani, M., Allevi, A., and Andreoni, A. (2007). Phys.

Rev. A, 76:043806.Perina, J. and Perina, Jr., J. (1995a). J. Opt. B: Quantum Semiclass. Opt., 7:541.Perina, J. and Perina, Jr., J. (1995b). J. Opt. B: Quantum Semiclass. Opt., 7:849.Perina, J. and Perina, Jr., J. (1995c). J. Opt. B: Quantum Semiclass. Opt., 7:863.Perina, J. and Perina, Jr., J. (1996). J. Mod. Opt., 43:1951.Perina, Jr., J. (2008). Phys. Rev. A, 77:013803.Perina, Jr., J., Centini, M., Sibilia, C., Bertolotti, M., and Scalora, M. (2006). Phys. Rev. A,

73:033823.Perina, Jr., J., Haderka, O., Sibilia, C., Bertolotti, M., and Scalora, M. (2007). Phys. Rev. A,

76:033813.Perina, Jr., J. and Perina, J. (1997). J. Opt. B: Quantum Semiclass. Opt., 9:443.Perina, Jr., J. and Perina, J. (2000). In Wolf, E., editor, Progress in Optics, Vol. 41,

pp. 361–419. North-Holland, Amsterdam.Perina, Jr., J., Sibilia, C., Tricca, D., and Bertolotti, M. (2004). Phys. Rev. A, 70:043816.Perina, Jr., J., Sibilia, C., Tricca, D., and Bertolotti, M. (2005). Phys. Rev. A, 71:043813.Perinova, V. and Luks, A. (2000). In Wolf, E., editor, Progress in Optics, Vol. XL,

pp. 115–269. North-Holland, Amsterdam.Perinova, V. and Luks, A. (2003). Fort. Phys., 51:210.Perinova, V., Luks, A., and Krepelka, J. (2003). J. Opt. B: Quantum Semiclass. Opt., 5:254.Perinova, V., Luks, A., and Krepelka, J. (2004). J. Opt. B: Quantum Semiclass. Opt., 6:S104.Perinova, V., Luks, A., and Krepelka, J. (2006). J. Phys. B: At. Mol. Opt. Phys., 39:2267.Perinova, V., Luks, A., Krepelka, J., and Jelınek, V. (2000). J. Opt. B: Quantum Semiclass. Opt.,

2:746.Perinova, V., Luks, A., Krepelka, J., Sibilia, C., and Bertolotti, M. (1991). J. Mod. Opt., 38:2429.Perinova, V., Luks, A., and Perina, J. (1998). Phase in Optics. World Scientific, Singapore.Petersson, L. E. R. and Smith, G. S. (2003). J. Opt. Soc. Am. A, 20:2378.Pfeifer, R. N. C., Nieminen, T. A., Heckenberg, N. R., and Rubinsztein-Dunlop, H. (2007). Rev.

Mod. Phys., 79:1197.Pitaevskii, L. P. (2006). Phys. Rev. A, 73:047801.Plunien, G., Muller, B., and Greiner, W. (1986). Phys. Rep., 134:87.Pokrovsky, A. L. and Efros, A. L. (2002). Phys. Rev. Lett., 89:093901.Porter, C. E. (1965). Statistical Theory of Spectra: Fluctuations. Academic Press, New York.Potasek, M. J. and Yurke, B. (1987). Phys. Rev. A, 35:3974.Power, E. A. and Thirunamachandran, T. (1993). Phys. Rev. A, 48:4761.Power, E. A. and Zienau, S. (1959). Philos. Trans. R. Soc. Lond. Series A, 251:427.Pradhan, P. and Kumar, N. (1994). Phys. Rev. B, 50:9644.Raabe, C., Knoll, L., and Welsch, D.-G. (2003). Phys. Rev. A, 68:033810.Raabe, C., Knoll, L., and Welsch, D.-G. (2004). Phys. Rev. A, 69:019901.Raabe, C. and Welsch, D.-G. (2005). Phys. Rev. A, 71:013814.Raabe, C. and Welsch, D.-G. (2006). Phys. Rev. A, 73:047802.Rarity, J. G. and Tapster, P. R. (1990). Phys. Rev. Lett., 64:2495.Raymer, M. G., Beck, M., and McAlister, D. F. (1994). Phys. Rev. Lett., 72:1137.Raymond Ooi, C. H. and Scully, M. O. (2007). Phys. Rev. A, 76:043822.Reed, M. and Simon, B. (1979). Methods of Modern Mathematical Physics: III, Scattering Theory.

Academic Press, New York.Rehacek, J. and Perina, J. (1996). Opt. Commun., 132:549.

Page 476: Quantum Aspects of Light Propagation

470 References

Reichl, L. (1992). The Transition to Chaos in Conservative Classical Systems: Quantum Manifes-tations. Springer-Verlag, New York.

Reimann, P. (2002). Phys. Rep., 361:57.Rice, P. R. and Carmichael, H. J. (1994). Phys. Rev. A, 50:4318.Rizzuto, L. (2007). Phys. Rev. A, 76:062114.Robinson, F. N. H. (1971). Physica (Utrecht), 54:329.Robinson, F. N. H. (1973). Macroscopic Electromagnetism, 1st ed., Pergamon Press, Oxford.Rodriguez, A., Ibanescu, M., Iannuzzi, D., Capasso, F., Joannopoulos, J. D., and Johnson, S. G.

(2007a). Phys. Rev. Lett., 99:080401.Rodriguez, A., Ibanescu, M., Iannuzzi, D., Joannopoulos, J. D., and Johnson, S. G. (2007b). Phys.

Rev. A, 76:032106.Rodrigues, R. B., Maia Neto, P. A., Lambrecht, A., and Reynaud, S. (2006a). Europhys. Lett.,

76:822.Rodrigues, R. B., Maia Neto, P. A., Lambrecht, A., and Reynaud, S. (2006b). Phys. Rev. Lett.,

96:100402.Rodrigues, R. B., Maia Neto, P. A., Lambrecht, A., and Reynaud, S. (2007a). Phys. Rev. A,

75:062108.Rodrigues, R. B., Maia Neto, P. A., Lambrecht, A., and Reynaud, S. (2007b). Phys. Rev. Lett.,

98:068902.Roman, P. (1969). Introduction to Quantum Field Theory. J. Wiley, New York.Roppo, V., Centini, M., Sibilia, C., Bertolotti, M., de Ceglia, D., Scalora, M., Akozbek, N., Bloe-

mer, M. J., Haus, J. W., Kosareva, O. G., and Kandidov, V. P. (2007). Phys. Rev. A, 76:033829.Rosenau da Costa, M., Caldeira, A. O., Dutra, S. M., and Westfahl, Jr., H. (2000). Phys. Rev. A,

61:022107.Rosenbluh, M. and Shelby, R. M. (1991). Phys. Rev. Lett., 66:153.Rosewarne, D. M. (1991). Quantum Opt., 3:193.Roth, J.-P. (1985). In Mokobodzki, G. and Pinchon, D., editors, Theorie du Potentiel, Pro-

ceedings of the Colloquia Jacques Deny, Lecture Notes in Mathematics, Vol. 1096,pp. 521–539. Springer-Verlag, Berlin.

Ruostekoski, J. (2000). Phys. Rev. A, 61:033605.Ruppin, R. (2000). Solid State Commun., 116:411.Ruppin, R. (2002). Phys. Lett. A, 299:309.Sakoda, K. (2002). J. Opt. Soc. Am. B, 19:2060.Sakoda, K. and Haus, J. W. (2003). Phys. Rev. A, 68:053809.Sakoda, K. and Ohtaka, K. (1996a). Phys. Rev. B, 54:5732.Sakoda, K. and Ohtaka, K. (1996b). Phys. Rev. B, 54:5742.Saleh, B. E. A. and Teich, M. C. (1991). Fundamentals of Photonics. J. Wiley, New York.Sanz, M., Papageorgopoulos, A. C., Egelhoff, W. F., Nieto-Vesperinas, M., and Garcıa, N. (2003).

Phys. Rev. E, 67:067601.Sargent III, M., Scully, M. O., and Lamb Jr., W. E. (1977). Laser Physics. Addison-Wesley, Read-

ing, MA.Savage, G. M. and Walls, D. F. (1987). J. Opt. Soc. Am. B, 4:1514.Scardicchio, A. and Jaffe, R. L. (2006). Nucl. Phys., B743:249.Schaden, M. and Spruch, L. (1998). Phys. Rev. A, 58:935.Schaden, M. and Spruch, L. (2000). Phys. Rev. Lett., 84:459.Scheel, S., Knoll, L., and Welsch, D.-G. (1998). Phys. Rev. A, 58:700.Scheel, S., Knoll, L., and Welsch, D.-G. (1999a). Phys. Rev. A, 60:4094.Scheel, S., Knoll, L., and Welsch, D.-G. (2000). Phys. Rev. A, 61:069901.Scheel, S., Knoll, L., Welsch, D.-G., and Barnett, S. M. (1999b). Phys. Rev. A, 60:1590.Schmidt, E., Jeffers, J., Barnett, S. M., Knoll, L., and Welsch, D.-G. (1998). J. Mod. Opt., 45:377.Schmidt, E., Knoll, L., and Welsch, D.-G. (1996). Phys. Rev. A, 54:843.Schram, K. (1960). Physica (Utrecht), 26:1080.Schram, K. (1973). Phys. Lett., 43A:282.

Page 477: Quantum Aspects of Light Propagation

References 471

Schubert, M. and Wilhelmi, B. (1986). Nonlinear Optics and Quantum Electronics. J. Wiley, NewYork.

Schwinger, J., DeRaad, Jr., L. L., and Milton, K. A. (1978). Ann. Phys. (New York),115:1.

Scully, M. O. and Zubairy, M. S. (1997). Quantum Optics. Cambridge University Press, Cam-bridge.

Sczaniecki, L. (1983). Phys. Rev. A, 28:3493.Sebbah, P., Hu, B., Kopp, V. I., and Genack, A. Z. (2007). J. Opt. Soc. Am. B, 24:A77.Selden, A. C. (1974). Opt. Commun., 10:1.Serulnik, S. and Ben-Aryeh, Y. (1991). Quantum Opt., 3:63.Severini, S., Tricca, D., Sibilia, C., Bertolotti, M., and Perina, J. (2004). J. Opt. B: Quantum Semi-

class. Opt., 6:110.Sha, W. L., Liu, C.-H., and Alfano, R. R. (1994). Opt. Lett., 19:1922.Shelby, R. M., Drummond, P. D., and Carter, S. J. (1990). Phys. Rev. A, 2:2966.Shelby, R. A., Smith, D. R., and Schultz, S. (2001). Science, 292:77.Shen, Y. R. (1967). Phys. Rev., 155:921.Shen, Y. R. (1969). In Glauber, R. J., editor, Quantum Optics, p. 473. Academic Press,

New York.Shen, Y. R., (1984). The Principles of Nonlinear Optics. J. Wiley, New York.Shen, J.-Q. (2004). J. Opt. A: Pure Appl. Opt., 6:239.Shih, Y. (2003). Rep. Prog. Phys., 66:1009.Shirasaki, M. and Haus, H. A. (1990). J. Opt. Soc. Am. B, 7:30.Shirasaki, M., Haus, H. A., and Liu Wong, D. (1989). J. Opt. Soc. Am. B, 6:82.Shirkov, D. V. and Kovalev, V. F. (2001). Phys. Rep., 352:219.Shukla, P. K., Marklund, M., Brodin, G., and Stenflo, L. (2004). Phys. Lett. A, 330:131.Sibilia, C., Centini, M., Sakoda, K., Haus, J. W., and Bertolotti, M. (2005). J. Opt. A: Pure Appl.

Opt., 7:S198.Slusher, R. E. and Eggleton, B. J. (2003). Nonlinear Photonic Crystals. Springer-Verlag, Berlin.Siegman, A. E. (1986). Lasers. Mill Valley, California.Siegman, A. E. (1989). Phys. Rev. A, 39:1253.Slusher, R. E., Grangier, P., La Porta, A., Yurke, B., and Potasek, M. J. (1987). Phys. Rev. Lett.,

59:2566.Smith, G. S. (1997). An Introduction to Classical Electromagnetic Radiation. Cambridge Univer-

sity Press, Cambridge, UK.Smith, D. R., Schultz, S., Markos, P., and Soukoulis, C. M. (2002). Phys. Rev. B, 65:195104.Sokolov, I. V. (1991a). Zh. Eksp. Teor. Fiz., 100:780.Sokolov, I. V. (1991b). Sov. Phys. JETP, 73:431.Solimeno, S., Crosignani, B., and Di Porto, P. (1986). Guiding, Diffraction, and Confinement of

Optical Radiation. Academic Press, Orlando.Sparnaay, M. J. (1989). In Sarlemijn, A. and Sparnaay, M. J., editors, Physics in the Making, pp.

235–246. Elsevier, Amsterdam.Stegeman, G. I., Bowden, C. M., and Zheltikov, A. M., editors. (2002). Journal of the Opti-

cal Society of America B, Special Issue Nonlinear Optics of Photonic Crystals, Vol. 19,pp. 2046–2296. Optical Society of America, Washington, DC.

Steinberg, S. (1985). In Sanchez Mondragon, J. and Wolf, K. B., editors, Lie Methods in Optics.Springer-Verlag, Berlin.

Stenholm, S. (2000). Opt. Commun., 179:247.Strekalov, D. V., Sergienko, A. V., Klyshko, D. N., and Shih, Y. H. (1995). Phys. Rev. Lett., 74:3600.Sudarshan, E. C. G. (1963). Phys. Rev. Lett., 10:277.Suttorp, L. G. and Wubs, M. (2004). Phys. Rev. A, 70:013816.Svelto, O. and Hanna, D. C. (1977). In Schafer, F. P., editor, Dye Lasers, p. 78. Springer-Verlag,

Berlin.Svelto, O. and Hanna, D. C. (1989). Principles of Lasers. Plenum Press, New York.

Page 478: Quantum Aspects of Light Propagation

472 References

Svetovoy, V. B. (2007). Phys. Rev. A, 76:062102.Taflove, A. (1995). Computational Electrodynamics: The Finite-Difference Time-Domain Method.

Artech House, London.Tai, C.-T. (1994). Dyadic Green Functions in Electromagnetic Theory. IEEE Press,

New York.Tanaka, H., Niwa, H., Hayami, K., Furue, S., Nakayama, K., Kohmoto, T., Kunitomo, M., and

Fukuda, Y. (2003). Phys. Rev. A, 68:053801.Tapster, P. R., Rarity, J. G., and Owens, P. C. M. (1994). Phys. Rev. Lett., 73:1923.Tarasov, V. E. (2001). Phys. Lett. A, 288:173.Tatarinova, L. L. and Garcia, M. E. (2007). Phys. Rev. A, 76:043824.Taylor, J. R. (1972). Scattering Theory: The Quantum Theory of Nonrelativistic Collisions. Krieger,

Malabar, Florida.Ter-Gabrielyan, N. E., Markushev, V. M., Belan, V. R., Briskina, C. M., Dimitrova, O. V., Zolin, V.

F., and Lavrov, A. V. (1991a). Sov. J. Quantum Electron., 21:281.Ter-Gabrielyan, N. E., Markushev, V. M., Belan, V. R., Briskina, Ch. M., and Zolin, V. F. (1991b).

Sov. J. Quantum Electron., 21:32.Ter Haar, D. (1961). Elements of Hamiltonian Mechanics. Interscience, New York.Thareja, R. K. and Mitra, A. (2000). Appl. Phys. B, 71:181.Tip, A. (1997). Phys. Rev. A, 56:5022.Tip, A. (1998). Phys. Rev. A, 57:4818.Tip, A. (2004). Phys. Rev. A, 69:013804.Tip, A. (2007). Phys. Rev. A, 76:043835.Tip, A., Knoll, L., Scheel, S., and Welsch, D.-G. (2001). Phys. Rev. A, 63:043806.Toda, M., Adachi, S., and Ikeda, K. (1989). Prog. Theor. Phys. Suppl., 98:323.Tomas, M. S. (1995). Phys. Rev. A, 51:2545.Tomas, M. S. (2002). Phys. Rev. A, 66:052103.Toren, M. and Ben-Aryeh, Y. (1994). Quantum Opt., 6:425.Townsend, P. D., Baker, G. L., Jackel, J. L., Shelburne III, J. A., and Etemad, S. (1989). SPIE,

1147:256.Tricca, D., Sibilia, C., Severini, S., Bertolotti, M., Scalora, M., Bowden, M., and Sakoda, K. (2004).

J. Opt. Soc. Am. B, 21:671.Tucker, J. and Walls, D. F. (1969). Phys. Rev., 178:2036.Ujihara, K. (1975). Phys. Rev. A, 12:148.Ujihara, K. (1976). Jpn. J. Appl. Phys., 15:1529.Ujihara, K. (1977). Phys. Rev. A, 16:562.Valanju, P. M., Walser, R. M., and Valanju, A. P. (2002). Phys. Rev. Lett., 88:187401.Valencia, A., Scarcelli, G., D’Angelo, M., and Shih, Y. (2005). Phys. Rev. Lett., 94:063601.van Albada, M. P. and Lagendijk, A. (1985). Phys. Rev. Lett., 55:2692.van der Molen, K. L., Mosk, A. P., and Lagendijk, A. (2007). Opt. Commun., 278:110.van Kampen, N. G., Nijboer, B. R. A., and Schram, K. (1968). Phys. Lett. A, 26:307.van Kranendonk, J. and Sipe, J. E. (1977). In Wolf, E., editor, Progress in Optics, Vol. 15, p. 245.

North-Holland, Amsterdam.Vanneste, C. and Sebbah, P. (2001). Phys. Rev. Lett., 87:183903.van Rossum, M. C. W. and Nieuwenhuizen, T. M. (1999). Rev. Mod. Phys., 71:313.Veselago, V. G. (1967). Sov. Phys. Solid State, 8:2854.Veselago, V. G. (1968). Sov. Phys. Uspechi, 10:509.Veselago, V. G. (2002). Uspekhi Fiz. Nauk, 172:1215 (in Russian).Viviescas, C. and Hackenbroich, G. (2003). Phys. Rev. A, 67:013805.Vogel, W. and Welsch, D.-G. (1994). Lectures on Quantum Optics. Akademie Verlag, Berlin.Vogel, W., Welsch, D.-G., and Wallentowitz, S. (2001). Quantum Optics. Wiley-VCH, Weinheim.Walls, D. F. (1983). Nature (London), 306:141.Walls, D. F. and Milburn, G. J. (1994). Quantum Optics. Springer-Verlag, Berlin, p. 127.Wang, L. J., Zou, X. Y., and Mandel, L. (1991a). Phys. Rev. A, 44:4614.

Page 479: Quantum Aspects of Light Propagation

References 473

Wang, L. J., Zou, X. Y., and Mandel, L. J. (1991b). J. Opt. Soc. Am. B, 8:978.Watanabe, K., Nakano, H., Honold, A., and Yamamoto, Y. (1989). Phys. Rev. Lett., 62:2257.Weigert, S. (1996). Phys. Lett. A, 214:215.Weinert-Raczka, E. (1996). Pure Appl. Opt., 5:251.Weinert-Raczka, E. and Lederer, F. (1993). Opt. Commun., 102:478.Wennerstrom, H., Daicic, J., and Ninham, B. W. (1999). Phys. Rev. A, 60:2581.Wiersma, D. S., Bartolini, P., Lagendijk, A., and Righini, R. (1997). Nature, 390:671.Wiersma, D. S. and Lagendijk, A. (1996). Phys. Rev. E, 54:4256.Wiersma, D. S., van Albada, M. P., and Lagendijk, A. (1995a). Nature, 373:203.Wiersma, D. S., van Albada, M. P., and Lagendijk, A. (1995b). Phys. Rev. Lett., 75:1739.Wigner, E. P. (1932). Phys. Rev., 40:749.Wigner, E. P. (1956). In Proceedings of the Conference on Neutron Physics by Time-of-Flight,

pp. 59–70. Gatlinburg, Tennessee. (Oak Ridge National Laboratory Report No. ORNL 2309,unpublished).

Williamson, J. (1936). Am. J. Math., 58:141.Wiseman, H. M. and Mølmer, K. (2000). Phys. Lett. A, 270:245.Wolf, P. E. and Maret, G. (1985). Phys. Rev. Lett., 55:2696.Woolley, R. G. (1971). Philos. Trans. R. Soc. Lond. Series A, 321:557.Wright, E. M. (1990). J. Opt. Soc. Am. B, 7:1141.Wubs, M. and Suttorp, L. G. (2001). Phys. Rev. A, 63:043809.Yablonovich, E. (1987). Phys. Rev. Lett., 58:2059.Yablonovitch, E. and Gmitter, T. J. (1987). Phys. Rev. Lett., 58:2486.Yablonovitch, E. and Gmitter, T. J. (1989). Phys. Rev. Lett., 63:1950.Yamamoto, Y. and Imoto, N. (1986). IEEE J. Quant. Electr., QE–22:2032.Yang, C. N. (1967). Phys. Rev. Lett., 19:1312.Yang, C. N. (1968). Phys. Rev., 168:1920.Yang, S. and Kellman, M. E. (2000). Phys. Rev. A, 62:022105.Yang, S. and Kellman, M. E. (2002). Phys. Rev. A, 65:034103.Yanik, M. F. and Fan, S. (2005). Phys. Rev. A, 71:013803.Yariv, A. (1985). Optical Electronics, 3rd ed. J. Wiley, New York.Yariv, A. (1989). Quantum Electronics, 3rd ed. J. Wiley, New York, p. 136.Yariv, A. and Yeh, P. (1984). Optical Waves in Crystals. J. Wiley, New York.Yeh, P. (1988). Optical Waves in Layered Media. J. Wiley, New York.Yoshie, T., Scherer, A., Hendrickson, J., Khitrova, G., Gibbs, H. M., Rupper, G., Ell, C., Shchekin,

O. B., and Deppe, D. G. (2004). Nature (London), 432:200.Yuen, H. P. and Shapiro, J. H. (1979). Opt. Lett., 4:334.Yurke, B. (1984). Phys. Rev. A, 29:408.Yurke, B. (1985). Phys. Rev. A, 32:300.Yurke, B., Grangier, P., Slusher, R. F., and Potasek, M. J. (1987). Phys. Rev. A, 35:3586.Zacharakis, G., Papadogiannis, N. A., Filippidis, G., and Papazoglou, T. G. (2000). Opt. Lett.,

25:923.Zalesny, J. (2001). Phys. Rev. E, 63:026603.Zhang, Z.-Q. (1995). Phys. Rev. B, 52:7960.Zhang, J., Peng, K., and Braunstein, S. L. (2003). Phys. Rev. A, 68:013808.Zharov, A. A., Shadrivov, I. V., and Kivshar, Y. S. (2003). Phys. Rev. Lett., 91:037401.Zhou, F. and Spruch, L. (1995). Phys. Rev. A, 52:297.Zinn-Justin, J. (1989). Quantum Field Theory and Critical Phenomena. Oxford University Press,

Oxford.Ziolkowski, R. W. (2001). Phys. Rev. E, 63:046604.Ziolkowski, R. W. and Heyman, E. (2001). Phys. Rev. E, 64:056625.Zou, X. Y., Wang, L. J., and Mandel, L. (1991). Phys. Rev. Lett., 67:318.Zyuzin, A. Yu. (1995). Phys. Rev. E, 51:5274.

Page 480: Quantum Aspects of Light Propagation

Index

Aamplifying random medium, 415, 419Anderson localization, 408, 419, 432, 437anomalous dispersion, 317atom-field interaction, medium assisted, 207atom-field medium-assisted interaction, 258

Bblackbody radiation, 423Bloch envelope, 388Bloch wave number, 388Bloch waves, 388

CCasimir effect, 280Casimir–Polder force, 297, 299cavity quantum electrodynamics, standard

model, 213cavity, single-mode, 305chaotic cavity, 441, 443circular orthogonal ensemble, 413commutation relation, equal space, 103contradirectional, 105corrugated waveguide, 349, 351counterdirectional coupling, 108coupled-mode theory, 362

Ddamped-polariton model, 235disordered cavity, 420, 424, 426disordered medium, 393disordered waveguide, 424, 428, 432dispersive lossless dielectric, 32, 121dispersive lossy linear and nonlinear

dielectrics, 242

Eeffective description, 305, 317electric multipole moment, macroscopic

density, 309

electromagnetically induced transparency, 317ensemble of cavities, 405extra quantum noise, 307

FFabry–Perot cavity, 200Fabry–Perot cavity, artificial, 201Fabry–Perot cavity, true, 201Fano diagonalization, 229Fano factor, 431, 442Fano factor, suggested formula

for, 405Fano factor, usual formula

for, 405feedback, nonresonant, 435, 443feedback, resonant, 435, 439, 443Fermi’s golden rule, generalization, 342

Ggauge, generalized Coulomb, 347gauge, generalized Lorentz, 347gauge, generalized temporal, 347Gaussian orthogonal ensemble, 445Gaussian state, 20Gaussian units, 15generator of spatial progression, 105guest atom, 305

HHeaviside–Lorentz units, 8, 12Hopfield microscopic

model, 224host medium, 305Huttner–Barnett model, restriction, 312

Iinfinite-dimensional harmonic oscillator, 20input–output relations, 256interference experiment, 42

475

Page 481: Quantum Aspects of Light Propagation

476 Index

KKerr-like nonlinearity, 391Kirchhoff’s law, 428

LLaguerre ensemble, 418left-handed materials, 264light propagation, 7light propagation, in a medium, 8light propagation, steady state, 8linear coupler, 213

Mmacroscopic averages, 308macroscopic averaging, 314macroscopic canonical quantization, 186macroscopic quantization, 389macroscopic theory, 342magnetic multipole moment, macroscopic

density, 309master equation, 395microscopic approach, 349mode competition, 442mode in macroscopic quantization, artificial

medium, several ones, 213mode in macroscopic quantization, artificial

medium, single one, 188mode in macroscopic quantization, true

medium, 187model of laser, 394modes of universe, 133momentum operator, 8, 9, 17momentum operator, slowly-varying

amplitude, 97multiple scattering, 317multiple scattering method, 413

Nnarrow-band field, 121negative-index materials, 264noiseless amplification, 169non-Hermitian modes, 137nonlinear optical couplers, 110nonorthogonal modes, 444, 449

Ooperator, space dependent, 8operator, time dependent, 8optical tomography, 409

Pparametric down-conversion, 388parametric down-conversion experiment, 36paraxial photon, 148

periodic media, 321permittivity, frequency-dependent, 308Petermann factor, 446Petermann factor, mean, 447photodetection plane, 152photodetection theory, 421photodetection, extended Hilbert space, 400photon-number operator, 9Photonic crystal fibres, 391photonic crystal, one-dimensional, 393photonic crystal, two-dimensional, 392photonic crystals, 379, 392photonic crystals, one-dimensional, 413photonic crystals, vortex lattice, 393photonic quasicrystals, 391Poisson brackets, 348polariton field, 305polariton field, damped, 313

Qquantization of evolution equations, 342quantum chemistry, an analogy, 214quantum dots, 392quantum imaging, 152quantum scattering theory, 193quantum soliton, 132quasi-cavity, 201quasimode, 133

Rradiation pressure, 259random laser, 393, 416, 424, 437, 444random media, 414rate equations, 397reduced velocity, microscopic derivation, 319reflection coefficient,distribution, 451reflection matrix, 412, 416removing divergences, 173

Sscattering, strong, 408, 419scattering, weak , 419single photon, 317single-mode regime, condition, 391source atom, 317source-field operator, 25space–time displacement operators, 102spatial distribution, of steady-state field, 137spatial progression, 17spatial Schrodinger picture, 10squeezed state, 430stationary quantum field, 40stationary solutions, 434, 437super-radiance, 393

Page 482: Quantum Aspects of Light Propagation

Index 477

Ttemporal mode, 86threshold, 415, 428, 433, 435, 448threshold, laser, 424, 426, 432threshold, lasing, 435, 439translational generator, 9transport theory, 8transport theory, many-body, 8two-level atom, 137

Uunitary translation operator, 10unravelling, continuous change, 398unravelling, discontinuous change, 399

unravelling, of master equation, 398unstable cavity, 137

Vvector potential, dual, 12, 102, 126vector potential, primal, 23, 32, 35

Wwave functional, 20Wigner functional, 20

Zzero electric-susceptibility limit, 237