You are on page 1of 693
QUANTUM ELECTRONICS Third Edition Amnon Yariv California Institute of Technology JOHN WILEY & SONS WILEY New York + Chichester + Brisbane * Toronto +¢ Singapore Copyright © 1967, 1975, 1989, by John Wiley & Sons, Inc, All rights reserved. Published simultaneously in Canada. Reproduction or translation of any part of this work beyond that permitted by Sections 107 and 108 of the 1976 United States Copyright Act without the permission of the copyright owner is unlawful. Requests for permission or further information should be addressed to the Permissions Department, John Wiley & Sons. LIBRARY OF CONGRESS Library of Congress Cataloging-in-Publication Data: Yariv, Amnon. Quantum electronics / Amnon Yariv.—3rd ed. pom. Bibliography: p. Includes index. ISBN 0-471-60997-8 {. Quantum electronics, L. Title, QC688.¥37 1988 537.5—delo Printed in the United States of America 1098765432 To The Memory of My Father Preface The thirteen years that have intervened since the appearance of the second edition of this book have witnessed some important developments in the field of lasers and quantum electronics. Foremost among them are: phase-conju- gate optics and its myriad applications, the long wavelength quaternary semi- conductor laser, and the deepened understanding of the physics of semicon- ductor lasers—especially that applying to their current modulation and limiting linewidth, laser arrays and the related concept of supermodes, quan- tum well semiconductor lasers, the role of phase amplitude coupling in laser noise, and free-electron lasers. The present edition retains nearly all the mate- ial of the second edition. There are four new chapters on semiconductor Tasers, quantum well lasers, free-electron lasers, and phase-conjugate optics. In addition, the chapters on laser noise and third-order nonlinear effects have been extensively revised. 1 benefited from teaching the material and from feedback at Caltech by colleagues and students who aided me in tightening and improving numer- ous points throughout the book. To these individuals special thanks are due. I am grateful to my administrative assistant, Jana Mercado, for her assistance in preparing this edition. I also take pleasure in thanking the talented group of students and ex-students, especially K. Lau, K, Vahala, C. Harder, T. Koch, M. Cronin-Golomb, S. Kwong, and the late C. Lindsey, whose research efforts are responsible for much of the new material added to this edition. The material on quantum well lasers benefited greatly from my association with Y. Arakawa while David Crouch contributed a major part of the treatment of squeezed states. Pasadena, California Amnon Yariv September 1987 Preface to the Second Edition This textbook introduces the main principles involved in the study and prac- tice of quantum electronics, which include the theory of laser oscillators, a wide range of optical phenomena, and devices that owe their existence to the intense and coherent optical fields made possible by the Jaser. The emphasis is almost exclusively on fundamental principles. An al- tempt is made, however, to bridge the gap between theory and practice through the use of numerical examples based on real situations. Approximately one-half of this edition is new. In addition, a number of topics related to microwave phenomena and magnetic resonance were omit- ted, The major changes are as follows. 1. The addition of treatments on Gaussian beam propagation in lenslike media, optical resonators, density matrix formulation of the interaction of light and matter, theory of laser oscillation, Van der Pol noise analysis of lasers, dye lasers, amplification in vibrational-rotational transitions, double heterojunction lasers, mode locking in homogeneously broadened lasers, Q-switching, saturated amplifiers and amplification of spontaneous emission, acoustooptic interactions, self-induced transparency, photon echoes, sponta- neous parametric fluorescence, distributed feedback lasers, and mode cou- pling in dielectric waveguides. 2. The deletion of chapters dealing with microwave masers, magnetism, magnetic resonance, and microwave parametric oscillators. 3. An exclusive use of the meter-kilogram-second (MKS) unit system. This text is primarily for the graduate student in physics and applied physics. The latter category often includes students in departments of electri- cal engineering and material science. The typical Caltech student taking the course from which this book was developed has a background of a one-year rigorous course in quantum me- chanics and one course in electrornagnetic theory. These are courses taken by the more advanced students in their senior year but often in the first year of graduate school. A good familiarity with these two topics is assumed, al- though most of the prerequisite background material is included here. The book can be used as a basis for a one-year course in quantum electronics or, alternatively, for these one-semester courses: X PREFACE TO THE SECOND EDINON 1. Lasers: Chapters 5-13. 2. Nonlinear Optical Effects and Stimulated Scattering Phenomena: Chapters 14 (part dealing with acoustooptics), 15-18. 3, Optical Modes and Propagation Phenomena: Chapters 5-7, 14, 19. Course 1 makes heavy use of quantum mechanics. In course 2 quantum mechanics is needed only in Chapter 15, while in course 3 it is not used at all. An electromagnetic background is needed in all three courses. I apologize to any of my colleagues whose work has not been acknow]- edged or adequately represented in this book. Since this is primarily a text- book, the material was chosen mainly because of pedagogic considerations rather than chronological precedence. I thank Ruth Stratton and Dian Rapchak for typing and proofreading the original manuscript and Paula Samazan for assisting with the references. Thanks are due to Dr. Jack Comly who made important contributions to Chapters 15 and 18 and to Mr. H. W. Yen who has gone over the whole manuscript rederiving the results and checking the internal consistency of the text. Amnon Yariv About the Author Amnon Yariv, a native of Israel, received his higher education at the Univer- sity-of California at Berkeley. After four years at the Bell Telephone Laborato- ties, he joined the California Institute of Technology (Caltech) where he is the Thomas G. Myers Professor of Electrical Engineering and Applied Physics. At Caltech he studies with a team of doctoral students and postdoctoral fellows, a number of research topics in laser physics, nonlinear optics, and optoelec- tronics. Some of his major contributions include the invention and co-inven- tion of the fields of integrated optoelectronic circuits, phase conjugate optics, and the authoring of the first papers on the theory of mode locking and nonlinear quantum optics. Dr, Yariv is also the founder and chairman of the board of ORTEL Corporation, a semiconductor laser company in Alhambra, California. When not working, which is rare, he spends time with his family (wife— Fran and three daughters} as well as with a piano, a tennis racket, and a windsurfer. Contents CHAPTERI Basic Theorems and Postulates of Quantum Mechanics 1.0 Introduction 1,1 The Schrédinger Wave Equation 1,2 The Time-Independent Schrédinger Wave Equation CHAPTER2 Some Solutions of the Time-Independent Schridinger Equation 2.0 Introduction 2.1 Parity 2.2 The Harmonic Oscillator 2.3 The Schrédinger Equation in Spherically Symmetric Potential Fields 2.4 The Angular Momentum Operators and Their Figenfunctions CHAPTER 3 Matrix Formulation of Quantum Mechanics 3.0 Introduction 3.1 Some Basic Matrix Properties 3.2 Transformation of a Square Matrix 3.3. Matrix Diagonalization 3.4 Representations of Operators as Matrices 3.5 Transformation of Operator Representations 3.6 Deriving the Eigenfunctions and Eigenvalues of an Operator by the Matrix Method 3.7 The Heisenberg Equations of Motion 3.8 Matrix Elements of the Angular Momentum Operators 3.9 Spin Angular Momenta 3.10 Addition of Angular Momentum 3.11 Time-Independent Perturbation Theory 18 18 18 27 30 xiv CONTENTS 3.12 Time-Dependent Perturbation Theory—Relation to Line Broadening 3.13 Density Matrices—Iniroduction 3.14 The Density Matrix 3.15 The Ensemble Average 3.16 Time Evolution of the Density Matrix 3.17 The Time Evolution Operator-Feynman Diagrams CHAPTER 4 Lattice Vibrations and Their Quantization 4.0 Introduction 4.1 Motion of Homogeneous Line 4.2 Wave Motion of a Line of Similar Atoms 4.3 A Line with Two Different Atoms 4.4 Lattice Sums 4.5 Quantization of the Acoustic Branch of Lattice Vibrations 4.6 Average Thermal Excitation of Lattice Modes CHAPTER 5 Electromagnetic Fields and Their Quantization 5.0 Introduction 5.1 Power Transport, Storage, and Dissipation in Electromagnetic Fields 5.2 Propagation of Electromagnetic Waves in Anisotropic Crystals 5.3 The Index Eliipsoid 5.4 Propagation in Uniaxial Crystals 5.5 Normal Mode Expansion of the Electromagnetic Field in a Resonator 5.6 The Quantization of the Radiation Field 5.7 Mode Density and Blackbody Radiation 5.8 The Coherent State CHAPTER 6 The Propagation of Optical Beams in Homogeneous and Lenslike Media 6.0 Introduction 6.1 The Lens Waveguide 6.2 The Identical-Lens Waveguide 6.3 The Propagation of Rays Between Mirrors 6.4 Rays in Lenslike Media 6.5 The Wave Equation in Quadratic Index Media 6.6 The Gaussian Beam in a Homogeneous Medium 68 638 68 7 74 76 83 83 83 106 106 106 111 111 112 115 116 CONTENTS 6.7 The Fundamental Gaussian Beam in a Lenslike Medium—The ABCD Law 6.8 A Gaussian Beam in a Lens Waveguide 6.9 High-Order Gaussian Beam Modes in a Homogeneous Medium. 6.10 High-Order Gaussian Beam Modes in Quadratic Index Media 6.11 Propagation in Media with a Quadratic Gain Profile 6.12 Elliptic Gaussian Beams CHAPTER7ZT Optical Resonators 7.0 Introduction 7.1 Spherical Mirror Resonators 7.2 Mode Stability (Confinement) Criteria and the Self-Consistent Resonator Solutions 7.3 The Resonance Frequencies 7.4 Losses in Optical Resonators 7.5 Unstable Optical Resonators CHAPTER 8& Interaction of Radiation and Atomic Systems 8.0 Introduction 8.1 Density Matrix Derivation of the Atomic Susceptibility 8.2 The Significance of x(v) 8.3 Spontaneous and Induced Transitions 8.4 The Gain Coefficient 8.5 The Einstein Treatment of Induced and Spontaneous Transitions 8.6 Homogeneous and Inhomogeneous Broadening 8.7 Gain Saturation in Systems with Homogeneous and Inhomogeneous Broadening CHAPTER? . Laser Oscillation 9.0 Introduction 9.1 The Laser Oscillation Condition 9.2 Laser Oscillation—Gencral Treatment 9.3 Power Output from Lasers CHAPTER 10 Some Specific Laser Systems 10.0 Introduction 120 123 124 125 127 129 136 136 136 141 145 147 149 155 155 155 162 164 169 171 173 176 183 183 183 189 191 202 202 xvi CONTENTS 10.1 Pumping and Laser Efficiency 10.2 The Ruby Laser 10.3 The Nd?*: YAG Laser 10.4 The Neodymium-Glass Laser 10.5 The He-Ne Laser 10.6 The Carbon Dioxide Laser 10.7 Organic-Dye Lasers CHAPTER ILL Semiconductor Diode Lasers 11.0 Introduction 1L.L Some Semiconductor Background 11.2 Optically Induced Band-to-Band Transitions in Semiconductors 11,3 Diode Lasers 11.4 GalnAsP Lasers 11.5 Some Real Lasers 11.6 Direct-Current Modulation of Semiconductor Lasers CHAPTER 12 Quantum Well Lasers 12.0 Introduction 12.4 The Quantum Mechanics 12.2 Gain in Quantum Well Lasers 12.3 Some Numerical Considerations CHAPTER 13 The Free-Electron Laser 13.0 Introduction 13.1 The Kinematics of Free-Electron-Photon Interaction 13.2 Theory of Optical Gain in Free-Electron Lasers 13.3 The Pondermotive Potential CHAPTER 14 The Modulation of Optical Radiation 14.0 Introduction 14.1 The Electrooptic Effect 14.2 Electrooptic Retardation 14.3. Electrooptic Amplitude Modulation 14.4 Phase Modulation of Light 14.5 Transverse Electrooptic Modulators 202 202 208 211 214 216 224 232 232 232 236 243 251 251 255 264 264 269 271 277 277 277 283 289 298 298 298 307 310 313 315 ‘ CONTENTS 14.6 High-Frequency Modulation Considerations 14.7 Electrooptic Beam Deflection. 14.8 The Photoelastic Effect 14.9 Bragg Diffraction of Light by Acoustic Waves 14.10 Deflection of Light by Sound 14.11 Bragg Scattering in Naturally Birefringent Crystals CHAPTER 15 Coherent Interactions of a Radiation Field and An Atomic System 15.0 Introduction 15.1 Vector Representation of the Interaction of a Radiation Field with a Two-Level Atomic System 15.2 Superradiance 15.3 Photon Echoes 15.4 Self-Induced Transparency CHAPTER 16 Introduction to Nonlinear Optics-—-Second-Harmonic Generation 16.0 Introduction 16.1 The Nonlinear Optical Susceptibility Tensor 16.2 The Nonlinear Field Hamiltonian 16.3 On the Physical Origins of the Nonlinear Optical Coefficients 16.4 The Electromagnetic Formulation of the Nonlinear interaction 16.5 Optical Second-Harmonic Generation 16.6 Second-Harmonic Generation with a Depleted Input 16.7 Second-Harmonic Generation with Gaussian Beams 16.8 Internal Second-Harmonic Generation CHAPTERI7 Parametric Amplification, Oscillation, and Fluorescence 17.0 Introduction and Lumped Circuit Analog 17.1 The Basic Equations of Parametric Amplification 17.2 Parametric Oscillation 17.3 Power Output and Pump Saturation in Parametric Oscillators 17.4 Frequency Tuming in Parametric Oscillation 17.5 Quantum Mechanical Treatment of Parametric Interactions 17.6 Frequency Up-Conversion 17.7 Spontaneous Parametric Fluorescence xvii 318 323 325 327 335 337 342 342 342 352 355 357 378 378 379 383 384 389 392 398 400 402 407 407 409 411 418 419 42) 425 430 xviii | CONTENTS 17.8 Backward Parametric Amplification and Oscillation 17.9 Squeezed States of the Electromagnetic Field CHAPTER i8 Third-Order Optical Nonlinearities—Stimulated Raman and Brillouin Scattering 18.0 Tntroduction 18.1 The Nonlinear Constants 18,2 molecular Raman Scattering 18.3 Stimulated Molecular Raman Scattering 18.4 Electromagnetic Treatment of Stimulated Raman Scattering 18.5 Anti-Stokes Scattering 18.6 Stimulated Brillouin Scattering 18.7 A Classical Treaument of Brillouin Scattering 18.8 Self-Focusing of Optical Beams CHAPTER 19 Phase-Conjugate-Optics and Photorefractive Beam Coupling 19.0 introduction 19.1 Propagation Through a Distorting Medium 19.2 Image Transmission in Fibers 19.3 Theory of Phase Conjugation by Four-Wave Mixing 19.4 Optical Resonators with Phase-Conjugate Reflectors 19.5 The ABCD Formalism of Phase-Conjugate Optical Resonators 19.6 Some Practical Applications of Phase Conjugation 19.7 Optical Phase Conjugation by Stimulated Nonlinear Scattering 19.8 Beam Coupling and Phase Conjugation by the Photorefractive Effect CHAPTER 20 Q-Switching and Mode Locking of Lasers 20.0 Introduction 20.1 Q-Switching 20.2 Mode Locking in Inhomegencously Broadened Laser Systems. 20.3 Mode Locking in Homogeneously Broadened Laser Systems 20.4 Relaxation Oscillation in Lasers 20.5 Passive Mode Locking 435 437 A453 453 453 457 465 469 473 475 475 482 495 495 495 A97 498 506 507 310 513 516 534 534 534 542 553 560 565 CONTENTS. CHAPTER 21 Noise and Spectra.of Laser Amplifiers and Oscillators 21.0 Introduction 21.1 Noise in Laser Amplifiers 21.2 Spontaneous Emission Noise in Laser Oscillators 21.3 Some Mathematical Background 21.4 The Laser Equations 21.5 The Laser Spectra 21.6 The Laser Spectra Experiments 21.7 The a Parameter 21.8 The Measurement of (Av)jaser CHAPTER 22 Guided Wave Optics—Propagation in Optical Fibers 22.0 Introduction 22.1 The Waveguide Modes 22.2 Mode Characteristics of the Planar Waveguide 22.3 Coupling Between Guided Modes 22.4 The Periodic Waveguide—Distributed Feedback Lasers 22.5 The Coupled-Mode Solutions 22.6 The Distributed Feedback Laser 22.7 Electrooptic Modulation and Mode Coupling in Dielectric Waveguides 22.8 Directional Coupling—Supermodes , 22.9 The Eigenmodes of a Coupled Waveguide System (“Supermodes’’} 22.10 Propagation in Optical Fibers APPENDIX 1 The Kramers—Kronig Relations APPENDIX 2 Solid Angle Associated with a Blackbody Mode APPENDIX 3 The Spontaneous Emission Lifetime for a Vibrational-Rotational Transition in a Linear Molecule APPENDIX 4 Quantum Mechanical Derivation of Nonlinear Optical Constants xix 570 570 570 577 582 584 586 592 594 396 600 600 600 603 606 608 6H 615 623 627 631 640 651 653 655 658 xX CONTENTS APPENDIX 5 The Interaction of An Electron and An Electromagnetic Field 663 APPENDEX 6 The Derivation of the Spontaneous Emission Langevin Fluctuation “Power” 666 669 Index cH APTER I Basic Theorems and Postulates of Quantum Mechanics 1.0 ENTRODUCTION In this chapter we shall consider some of the basic postulates and theorems of quantum mechanics. These are general and independent of the specific sys- tem studied. The application of these results to special problems will be the concern of Chapter 2 and, to a lesser extent, of the rest of the book. 1.1 THE SCHRODINGER WAVE EQUATION According to quantum mechanics, the behavior of a particle is described by the wavefunction (r, t), which is a solution of the Schrédinger wave equa- tion # - at |-Lw+ veo] a= at (ab Vir, t) is the potential energy function of the particle and & = /2m where h is Planck's universal constant. By associating the differential operator —i# V with particle linear mo- mentum p, that is, po -ii¥ (1.1+2) the operator on the left side of (1.1-1) can be associated with the sum of the kinetic and potential energy of the particle. po Eps ven = ae 113 Dra (= (11-3) Statistical Interpretation of the Wavefunction Consider a very large number of independent spaces with identical potential functions V(r, ). The motion of a particle in each one of these spaces is described by the same (r, ). The a-priori probability P(r, ¢) dv of finding any given particle inside a volume dv (centered about r) is taken as the fraction of the particles found, by measurernent, to be inside dv at time 1. According to quantum mechanics, the probability density Pir, f) is given by P(r. t) = We, Abt, (lea) 1 2 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS where the asterisk superscript stands for the complex conjugate of the quan- tity in question. The statistical interpretation of (Wp) is made plausible by showing, as willbe done in Section 1.2, that the average motion of a particle as deter- mined from the statistical point of view agrees with its classical counterpart. The final arbiter of the validity of this statement is, of course, the agreement of the results derived using it with experiment. The first condition resulting from the statistical interpretation is that the total probability of finding the particle somewhere in space is finite and is a constant, that is, Ponape POE 9 = Tanya Wr(e, Qin, t) dv = constant (1.15) or that a * Filategae prin, Quinn dv = 0 (1-6) The proof proceeds as follows a in _ ay. _ oe ob & [wre av = [5 ow ay = L(y + Su) a substituting for y/t and ay*/at from (11-1), the terms involving V(r, 1) drop out, and the result is BE egy — y Per So Ly WH Vb — WU) dv Use is now made of Green's theorem [UVa 9 VA) dv = [UV — 9 Wh nda (7) where fand g are two arbitrary scalar functions, A is the surface bounding V, and n is the unit outward normal vector. This leads to 2 yyy dv = 2 | ye vu we YH) nda (11-8) aly 2m ta If the volume V is that of ail space, the admissible solutions of (1.1-1) are indeed those in which the behavior of & as r—> * is such that the integration specified by (1.1-8) leads to a zero result. If the volume V is finite, the same result is obtained by choosing @ so that its value on the bounding surface A leads to a zero result in (1.1-8). This point is discussed further in Problem 1.1. Having proven (1.1-5), we are free to normalize ¢ so that the constant appearing in (1.1-5) is unity. This is consistent with the probabilistic interpre- tation, since the probability of finding the particle somewhere in all of space is unity, that is, Jeane pir.) dv = J) yor, gr, dv = 1 (11-9) 1.1 THE SCHRODINGER WAVE EQUATION 3. Particle Density Current We start by rewriting (1.1-8) d if SE yy dv = = [, wt Ve - ww on da dt iv 2m ta where V is any arbitrary volume. The use of Gauss’s theorem fev B) dv = [ Benda (11-10) for any arbitrary vector function B leads to 8 pynyy = -9- | 4 2ew = -v [Zw wil Q.L1) in direct analogy with the charge conservation condition in electricity . 4 : Veins a where p is the charge density and i, is the electric current density, we define i-% Vy - Ve 1.1-12) i= 3, th et ~ rv) (11-12) as the particle probability current density. Equation (1.1-11) is thus a state- ment of the conservation of particle probability. The quantum mechanical counterpart of the classical motion of a particle can be viewed as the specifica- tion of the particle probability current density as a function of space and time. Expectation Value ‘The expectation value of a physical observable is the ensemble average, in the sense described in the paragraph preceding (1.1-4), of the measurements of this observable, Alternatively, it can be defined as the mathematical expecta- tion for the result of a single measurement of this observable. The expectation yalue for the position radius vector r is a) =f rete. queen de = [pred ay (14-13) So far we have no reason to prefer one or the other of the two expressions for {r) given in (1.1-13). When the observable whose expectation value is to be evaluated is a function of the coordinates only, the two methods lead, obvi- ously, to the same result. This is not necessarily true when the observable is a function of momenta and/or of energy, This point can be illustrated if we write the expectation value for the momentum component P,. If we use (1.1-2), the contenders for (p,) become (pa) = ~it [ we) 4 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS and . oo infor Bar It is clear that these two procedures will lead, in general, 10 different results. The issue can be settled by making the reasonable requirement that when calculating the expectation value of the particle energy (£), we obtain the same result by using either of the two energy operators given in (1.1-3), that is, ( = V+ Vir, 4) = (it 4 (1-14) From Schrédinger’s equation (1.1-1), it follows that 2(-# + ob Ww (- w+ v) w= ay = so that the first procedure for finding the expectation value of an operator applies, since a a @-felEvrvd=a lobe (1-15) The second procedure, on the other hand, requires that a a (Eo ve + v) ray = 8 2 roy Carrying out the indicated operations of the last equation shows that it cannot be reconciled with the Schrédinger equation. The procedure for finding the expectation value of an operation A, which depends explicitly on the coordi- nates, momenta, energy, and time, is thus (At, p, E. )) = { yA (. —ih V, ih i. ‘ue ay (11-16) We next show that our choice of (1.1-16) for calculation (A) leads to expecta- tion values of physical observables that agree with their classical counterparts. The proofs, which tend to be rather tedious, are greatly simplified here be- cause all the operators considered are Hermitian. Hermitian Operators. The Hermitian adjoint At of an operator A is defined as the operator satisfying the relation’ [raga = [ catpty ay aay where f and g are two arbitrary scalar functions. If At = A, that is, when ' See, for example, N. Dunford and J. T. Schwartz, Linear Operators (New York: Interscience, 1958), Part I, p. 350. 1.1 THE SCHRODINGER WAVE EQUATION 5 [run av = | (aprgav (11-18) the operator A is called a Hermitian operator. It will be left for Chapter 3 to prove that operators representing physical observables are indeed Hermitian. In the meantime, we shall use this fact in some of the following proofs. The Time Rate of Change of the Expectation Value In addition to the procedure for obtaining the expectation value of an opera- tor (4), we can derive a very general, and most useful, result involving the time rate of change of the expectation value. di d aad 2 | geay av = [2 gran av =[ (days S ao) av ‘GA a abe = fle aG) + Fran] Defining the Hamiltonian operator H by ye-Ewey 139) =o (1.139) we have, according 10 (1.1-1), By = ih sd ms which, substituted for é%p/at above, results in 4A _ [yn 24g —Lyrany + Z creas] = &) + if (WtHAy ~ YAH) dv (11-20) = Gi) + pea where, owing to the Hermiticity of H, we put { (aby A dv = f ytHAa av and the commutator of H and A is defined by [H, A] = HA — AH ‘This definition will be used for any pair of operators. 6 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS Ehrenfest’s Theorem The theorem shows that the classical equations of motion for a single particle . dr m 2 _ gy (11-21) ae ae are obeyed, according to quantum mechanics, when all the vectors appearing in (1.1-21) are replaced by the expectation values of the corresponding quan- tum mechanical operators. Part 1: J wx av = | (ux + + a) av {th i th cae Lage = | [vee GR wow — 508) +a (gps = gw) | if = Bly vy - waren) dv If we use the Hermiticity of V?, the last expression becomes & [ure Wy — wt vox | av but Veep) = x VA + 2 so that 40) Rf yy BH gy = (1.1-22) Part 2: We use Eq. (1.1-20). Kpx) =f levee) x ~ f (ue %y) we (14-23) --@ Use has been made of the fact that the operators V? and d/dx commute. 1.2. THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 7 The Momentum Wavefunction. Consider the Fourier transform of the wavefunction (xr, #) defined by? LVR pee Op. = (34) {C eke Trib, f) dv (11-24) so that vr, = (aa) [E ewntap. 9 &p where d'p = dp, dp, dp,. Using the last equation in Schrédinger’s equation (1.1-1), we find—the actual derivation being assigned as a problem—that ®(p, t) satisfies F [z + Vir i Vp. De alowp, ha hi ip. #) m or (11-25) where r — id V, signifies that wherever x;(i = 1, 2, 3) appears in V, itis to be replaced by ih d/dp:. Equation (1.1-25} is the wave equation in p space, the expression inside the square brackets being in energy operator. Since p?/2m is the kinetic energy term, we identify p in (1.1-24) as the momentum vector. The statistical interpretation of ®(p, #} is similar to that of w(r, #), and (p, 1) dp, dp, dp; is the probability that the particle momentum at time 1 is within the differential volume dp, dp, dp, centered on p in momentum space. The expectation value of any operator A is calculated accordingly by 7 a (a) = f Op. nd (pi Vp. HS. 1) tp. dre. doy doe (11-26) The normalization of O(p, 1), J &&* @p = 1, follows from Parseval’s theorem, according 1o which JU rage ak = J" [fen dv (1.1-27) where F{k) and f(r) are a Fourier transform pair. [1 will be left as a (nontri- vial) exercise (Problems 1.4 and 1.5) to show that (A) calculated by (1.1-26) is identical to that obtained from (1.1-16). 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION Starting with the time-dependent wave equation B 3 -Evy+ went (12-1) 2 Notice that we choose the coordinate of the conjugate space as pif rather than the conven tional k. 8 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS let Yr, ) = u(r)g(t). Equation (1.2-1) becomes l(# o, _ th dg . L(-E vas wu) -24 (1.2-2} For a time-independent potential function V(r), we can separate {1.2-2) by means of a separation constant E R 7 Vue) + Vira = Eu (1.2-3) and wih dat) _ wo a =F (1.2-4) The solution of (1.2-4) is gt) = glOpeem Equation (1.2-3) is referred to as the time-independent Schrédinger equation, and its solutions z,(r) as the time-independent energy wavefunctions. The set {discrete or continuous) of allowed values of £ are the eigenvalues and are determined (as will be shown in Chapter 2) by the boundary conditions of u(r) or by the necessary behavior at infinity. We will assume that the set we(r) generated by (1.2-3) is a complete set and will proceed to derive some general Tesults conceming these wavefunctions. The Orthonormality of the Wavefunctions We will show first that any two members of the set 2(r) with different F’s are orthogonal in the sense J eb truetn) dv = The starting point is the time-independent wave equation (1.2-3) BE BR ~ Fg Vee + Vite = Ble where we dropped the functional notation. We rewrite the equation for uz, multiply the first equation by 2 the second by zz, and subtract one equation from the other. The result is # — on (ue Vue — ue Vuk) + V(ugui — ued) = (E — Buen Integrating the last equation over a volume V and invoking Green’s theorem (11-7) give B - oh (uh Vug — ug Yul) mda = (E ~ E’) fe up dv 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 9 The integral on the lefi side vanishes, as discussed in connection with Equa- tion {.1-8, with the result that f up(Kyug(t) dv=0, BFE (1.2-5) We have used the fact that, since the E's are the eigenvalues of an Hermitian operator, they are real and (E')* = 2’. The normalization of f(r, t) is achieved by requiring that g(0)g*(0) = 1 and by the condition Jf tus? dal (1.2-6} so that #e(r)exp[—(iE/%)t] is a particular (normalized) solution of the time- dependent wave equation. The Significance of E The separation constant E has, up to this point, a mere mathematical signifi- cance as the eigenvalue of the time-independent wave equation (1.2-3). Con- sider the case when the wavefunction is given by the particular solution We, th = up(e)e CEA (1.2-7) The expectation value of the total energy operator, which from now on will be called the Hamiltonian operator, is Ae a w=(-Ewrv=(ad-£ (1.2-8) where use has been made of (1.2-3) and (1.2-6). E emerges as the expectation value of the total energy operator H for the tinae-independent potential func- tion Vir). Some Mathematical Properties of the Wavefunctions Since the wp{r) constitute a complete orthonormal set, they can be used for the expansion of any arbitrary function G(r) G{t} = = ake (t) (1.2-9) and using the orthonormality conditions {1.2-5, 6), we obtain ge = | Oe’yuk(r') av’ G(r} = 2 [| Hb (r')G(r') a] uz(t) (1.210) = | avew') = Ere (t) where the integration extends over all space. We can write in general Gir} = [ave é(r’ — r} 10 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS. from which we conclude that D ahr u(r) = Er" — 3) (L.2-11) t and that . =1 ifrisinv’ [2 ub(e’)ug(r) dv’ (1.2-12) 0 otherwise The Complex Vector Space The orthonormal and complete set of eigenfunctions generated by a linear operator can be viewed as a set of mutually orthogonal unit vectors in an abstract (Hilbert) vector space. The advantage of this point of view is that many of the mathematical properties and manipulations involving these functions become analogous to familiar properties and operations involving vectors in ordinary space. By an extension of this idea we may regard an arbitrary function G(r) as a vector in the complex vector space. To illustrate this point of view, consider the energy eigenfunctions uw. Defining the operation (uz, us) by (ue, te) = f ubug dv (1.2-13} and calling it the scalar (dot) product of the ‘vectors’ up and up, we can write the orthonormality condition (1.2-5) as (ug, Ue) = E The expansion of an arbitrary function G{r) in terms of the u,’s, Eq. (1.2-10), takes the form G= > tue, Gag (1.2-14) = and is interpreted as expressing the vector, that is, function, G in terms of the unit vectors uy, so that (ug, G) is the projection of G along ur. The effect of linear operators in this space is to alter, in general, the directions and magnitudes of the vectors on which they operate. The eigenvectors of a given operator, say A, ate those vectors whose direction in space is not altered when operated on by A. Fundamental Postulates of Quantum Mechanics According to the abstract formulation of quantum mechanics, a physical state of a system corresponds to a complex state vector. The set u,, for example, constitutes all the possible energy states of the system. A physical measure- ment on the system constitutes a disturbance and, in general, alters the state of the system, Quantum mechanically, the process of measuring a physical quan- tity (energy, momentum, etc.) corresponds to operation on the state vector by the operator corresponding to the physically observable. 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 11 Assume, for example, that the energy state of the system is wz, that is, the system is known, with certainty, to have a total energy £. Ifa second mea- surement of a physical observable A is undertaken, this corresponds to oper- ating on the state vector #; with the linear operator A. If zg is not an eigenvec- tor of A, the operation alters, by definition, the state uz so that one is no longer certain about the energy of the system. If, on the other hand, uz is also an eigenvector of A, the energy state is unaltered and we can state precisely both the energy and the value of the physical quantity that corresponds to the operator A. The necessary and sufficient condition that any two linear operators, in this case H and A, have the same eigenvectors is that they commute. To prove the sufficiency part, assume HA = AH. It follows that AHug = HAug = AEug = EAug. The vector Aug is thus an cigenvector of H with an eigenvalue E. If there is only one eigenvector associated with E, it follows that Aus = Cue, where C is any constant. If there is more than one eigenvector ue with an eigenvalue £ (ie., the set E is degenerate), it is possible to take linear combi- nations of the we’s of this subspace that are orthonormal and also eigenvectors of A. The proof of the necessity condition is left as an exercise. Let us summarize: The necessary and sufficient condition that any two physical observables be simultaneously and precisely measurable is that their respective opera- tors commute. When the operators A and H do not commute, the interpretation of the measurement process, due to Born, is the following. The wavefunction w(r, f) is expanded in terms of the eigenfunctions (here we revert to talking about functions rather than vectors) v, of the operator A, that is, wer, ) = & enktvatr) where Avy it} = / ft) The eigenvalues a, constitute the set of all possible measurements of the observable A. The probability that any given value a, is found by the measure- ment at time t is |c,(f)|?. According to this interpretation, (Ace) =D ate len? (12-15) using on = J va (Ryo, t) dv leads to Cate) = Za f wee ovale) foc, avgiry ae = | avure nat [ arvee. 9 vecepente) = J wr jaye, dv 12 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS where A‘ operates on functions of r’ and use has been made of the closure property, 2, vf(rjv.(r’) = 5(r — r'). The value (A(1)) obtained according to (1.2-15) is thus consistent with the previous procedure, Eq. (1.1-16). The Uncertainty Principle Having established that members of certain pairs of physical observables can- not be measured simultaneously with certainty, it is of interest to discover the magnitude of the uncertainties involved. One such pair of variables is the position x and momentum p, of a particle, because the operator (4/i}(8/8x) corresponding to p, does not com- mute with x. We have already shown that the wavefunctions @(p/h, i) and w(x, ) form a Fourier transform pair. It follows directly from the mathematical properties of such pairs that reasonably defined (this will be done in the proof that follows) widths of these functions denoted as A(p,/i) and dx will satisfy the relation A(p./#) Ax ~ 1. It should be possible, however, to extract this information from the wave equation since the behavior of the particle is completely specified by it. The uncertainties in coordinate and in momentum are characterized by the respective mean-square deviations from the average _ hey = fate - Go) dv (1.2-16) (set) = [v(t 2 - twa) ae If we take (x) = (p,) = 0 [this will be true if p(x) is an even function of x], the product of uncertainties becomes 2 = orf yt (aphex) = ~# fu SE av [ yeey dv From the Hermiticity of i a/ax (or from simple integration by parts), it follows that 2 a aut ay Jugea=- fo and in consequence caniy(are) = 4 [ 2H dy [ goat dv 247) By the Schwarz inequality? [mau f org do [Mf ye av + fap av) {1.2-18) 3 The Schwarz inequality states that the surn of the lengths of two sides of a triangle exceeds or equals the third. That is, |At |B| > A+ B where A and B are any two vectors. When exterdded to complex vector space. it becomes (E fig. g) = GLE g) + (& DJF for two arbitrary functions {vectors) f and g. This form is identical to (1.2-18). 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 13. if we let f = ab/ax, g = 2p gives A (op at 2 #8 aye aye (ApiXax) > ( ae ht ay + Jw tv) Ee a PLB Fl qwa) =F The term within the last brackets is equal to —1, as can be shown by integra- tion by parts (or the Hermiticity of i 4/ax). Defining Ax as (A2)*? and similarly for Ap,, we write Ap, Ax > 8 (1.2-19) Minimum Uncertainty Wavepacket It is of considerable interest 10 discover when the equality sign of {1.2-19} applies, that is, under what conditions is the uncertainty product Ap, Ax a minimum. It is clear from the geometrical interpretation that the equality in Eq. (1.2-18) applies when f and g are equal within a multiplicative constant. If we use again the substitutions f = dy/ax, g = xp, the equality sign is satisfied when a = Newer Be ay, y= nee where a is a constant. The normalization constant N is determined by requiring that f wy dx = 1, The result is W = (a/m)"*. In a similar fashion, we can identify the constant a by ar) = [Sera =(8)" [Peet aed al bow 2a The normalized expression for the minimum uncertainty }{x) becomes Wx, 0) = (4) errata) (1.2-20) . 2r{Ary where the time coordinate, chosen arbitrarily at { = 0, was restored. An important result is the fact that the minimum uncertainty wavepacket has a Gaussian form. Since the Fourier transform of one Gaussian is another one, it follows that ®(p, 0), the corresponding particle momentum wavepacket, is also a Gaussian. Since, according to {1.1-27), ®(p, 0) is normalized, we can write it directly as 1 \4 = (4+) war - (p, 0) uae) € {1.2-21) Tc will be left as an exercise to show that (Ap’) in (1.2-21) satisfies the uncer- tainty principle (1.2-19) with an equality sign. 14 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS Other pairs of conjugate observables obeying an uncertainty relation are tk AEA? & Ag, AJ 5 (1.2-22) Anabel where E and tare the energy and time of a system, @, and J, are the azimuthal angle and the angular momentum of a system about an arbitrary (z) axis, and nand © refer to the number of quanta and oscillation phase of an harmonic oscillator.* The Uncertainty Principle in Communication In this section we shall illustrate how the uncertainty principle can be used to give, in a quick fashion, certain answers concerning communication prob- lems. The Spread of an Antenna Beam. Consider first the problem of designing a large microwave antenna for communication purposes. One of the most im- portant problems in antenna design is that of concentrating the radiated power into the smallest possible divergence angle @. The beam spread can be considered to be due to the uncertainty Ap, in the transverse momentum p, of photons that are emitted by the antenna of diameter 2R. Taking the uncer- tainty Ax = & leads to 0 ~ = Sak (1.2-23) where the relation p, = 4/A was used. This is, of course, a very old problem and the result (1.2-23) is well known. Approached as a problem in electromagnetic theory, Eq. (1.2-21) results from the Fourier transform relationship that exists between the field intensity distribution in the antenna plane and the far-field angular distribu- tion of the radiated field. This Fourier transform relationship is also the basis for the uncertainty relations in quantum mechanics, as shown in the discus- sion surrounding Eq. (1.2-20). Information Capacity of a Communication Link. Having just become ex- perts in antenna design, we may contemplate, for a moment, the information capacity of the beam radiated by the antenna. It is known from the work of Shanon that the capacity of a communication channel in bits per second is «R, Serber and C. H. Townes, Symposium on Quantum Electronics (New York: Columbia University Press, 1960}, p. 233. 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 15 Electric field ry FIGURE 1.) A sequence of radiation pulses with the information coded into the amplitudes. given by” c= [loz (i + jy B (3.2-24) where S and N are the average signal and noise powers, respectively, and B is the transmission bandwidth in cycles per second. There are numerous sources of noise in communication links and their study is of no concern here. Even if all other noise sources are “cleaned up,” there remains, however, an effective noise source that we may consider as being of quantum mechanical origin. To be specific, consider the information carried by a sequence of pulses as shown in Figure 1.1. The average power is taken as p, so that the average energy per pulse is At. The transmitted information is coded into the amplitude of the pulses so that each predeter- mined value of pulse energy is associated with a message. According to infor- mation theory,‘ the information contents, in bits, of each pulse is given by the logarithm (to the base 2) of the number of distinguishable values (in this case energies) it may possess. The average information contents per pulse is thus Tyutse = loge (a4 (1.2-25) where the energy resolution AE is the smallest energy difference between pulses that can be measured. Since the time available for measuring the energy of a given pulse is 2 At, the energy resolution AE is given according to the uncertainty principle (1.2-22) by AF ~ #/2 At. Substitution in (1.2-25) gives Tyutse ~ 1OR2 [22a | The average information transmission rate is equal to the average information comtents per pulse multiplied by the number of pulses per second 1/(2 An). 5. B, Shanon and W. Weaver, The Mathematical Theory of Communication (Urbana, Ill.: University of Mllinois Press, 1949), p. 67. “Ibid. 16 BASIC THEOREMS AND POSTULATES OF QUANTUM MECHANICS C~ log: (eer) (1.2-26) Itis clear that C increases as Af is made smaller. The limit for this procedure is when At becomes comparable to the oscillation period f-'.? At this point, (1.2-26) becomes ¢ ~ log, [Zl B (12-27) where the transmission bandwidth is taken as B = 1/At. Thus, according to (1.2-24}, we may associate an average noise power N ~ fwB with what is classically a noise-free channel. The origin of this noise is, as shown above, quantum mechanical, and it is indeed the limiting factor on channel capacity when the condition Aw > kT is satisfied. A rigorous treatment of this problem has been given by Gordon.* We shall return to this subject in Chapter 13, where the noise properties of some specific systems are described. Supplementary References 1. Leighton, R. B., Principles of Modern Physics (New York: McGraw-Hill, 1959}. 2. Messiah, A. Quantum Mechanics (New York: Interscience, 1961). 3. Schill, L. 1., Quantum Mechanics (New York: McGraw-Hill, 1959) Problems 1.1 Ifthe behavior of w(r, f) as r— & is dominated by x", what values can n assume if the integral (1.1-8) J, 0 We — 6 904) on da taken over the surface at infinity is to vanish. 1.2. Show that (a(p,)}/at = —(av/ax) without using Hermiticity. 1.3 What is the value of [p,. x]? Using this result and {1.1-20), derive de) _ (pe) dt m 1.4 Show that if wir, ) defined by wr, = (Ay” QC ePrhb(p, 1) dp is to satisfy Schrédinger’s equation, d{p, #} satisfies the equation ? aip, & > = @ COP [z + Vir ik V,. p. 4] D(p, t) = a where r —> i V, means that x; is to be teplaced by if a/ap; 7 For shorter Af the concept of frequency becomes meaningless. * J.P. Gordon, in Advances in Quantum Electronics, 3. R. Singer, ed. (New York: Columbia Univer- sity Press, 1961), p. $09. Also in Proc. IEEE 30, 1 (1962). 1.6 17 18 1, © 1.10 1.2 THE TIME-INDEPENDENT SCHRODINGER WAVE EQUATION 17 Hint: Show that 2 rk ap, = — Ec pe += ab [2 OP, A for b(—%) = D(+) = 0. Show that the expectation value (A) is the same whether calculated by (1.1-26) or by (11-16). Prove that operators A and B commute if they have the same eigenfunctions Prove that au; on dv = (E> Bi) [E wpan, dv ao where Hu; = Bye; Hint: Calculate first the commutator {H, x]. Show that if the uncertainty product Ap, Av of a free particle obeys & Ap, Ax 2 5 the uncertainty product for the time ¢ and energy E = pi/2m obeys AE At > h/2. Show that [x, pi] = inp"! Hint: What is (a°/ax") (xu)? Using (1.1-17), show that the eigenfunctions of a Hermitian operator are or- thogonal. CHAPTER 2 Some Solutions of the Time-Independent Schrédinger Equation 2.0 INTRODUCTION In this chapter we shall solve two eigenvalue problems. This will provide concrete demonstrations for the abstract material in Chapter 1, as well as working formulas for subsequent chapters. Specifically, we shall find the energy eigenvalues and corresponding eigenfunctions for the harmonic oscil- lator and for angular momentum operators. 2.1 PARITY Consider the wave equation Le vi = om Vede) + Virja(r) = Bu(r) replacing r by —r gives ® v2 = ~ Fy VMN) + Vine )e(—4) = Bu(—r) 11 follows that if V(—r) = Vir), u(—r) is also a solution of the wave equation with an eigenvalue E. We can now construct two new solutions u(r) = u(r) + u(-3} Q.1-1) u(t) = utr) — u(—r) where 4, is an even function and x, is odd. The functions u, and 4, are also solutions of Schrédinger’s equation with the same eigenvalue E. If the set Fis nondegenerate, all four functions must be multiples of the same function, Two cases are possible: (1) «(r) is a multiple of u,(r) and u,(r) is zero, (2) u(r) is a multiple of u,(r) and u,(r) is zero. The eigenfunction u(r) is thus seen to Possess even or odd parity, that is, u(r) = u(r) If E belongs to a degenerate set containing n eigenfunctions, it is possible to construct » linearly independent superpositions of these functions that have 18 2.2 THE HARMONIC OSCILLATOR 19 definite parity. It should be noted that the eigenfunctions possess a definite parity only when Vir) = V{—r), that is, when the potential field has inversion symmetry, There are cases, such as the potential in noncentrosymmeuic crys- tals, when this condition is not satisfied so that the functions are neither odd nor even. 2.2 THE HARMONIC OSCILLATOR The harmonic oscillator consists of a mass m acted on by a restoring force that is proportional to its displacement from some point (which is taken as the origin). The solution of this problem provides an illustration of the manner in which the necessary behavior of the wavefunction at infinity is used to deter- mine the eigenvalues. Other situations in quantum mechanics, such as elec- tromagnetic radiation in an enclosure and the propagation of sound waves in solids and liquids, can be shown, as will be done in Chapter 4, to be formally equivalent to that of the harmonic oscillator and their treatment in the follow- ing chapters will rely on the material developed here. The wave equation of the harmonic oscillator in one dimension becomes ge _ ae | & + tKe?)e = 2m dei + 1KxPu = Ew {2,2-1) using the following substitutions: sit (a- n= 0 (2.2-2) for £? > d the behavior of u is dominated by the e~"? term, so that it is natural to assume a solution of the form u(E} = HE Oe (2.2-3) where H(€) is a polynomial of a finite order. Substituting 4(€) from (2.2-3) in (2.2-2) leads to the equation 2H EL _ ae Get AH = 0 {2.2-4) » In fine with the comment following (2.2-3), we assume a power series expan- sion HUE) = § (dy + ag + ag? +--+) (2.2-5) where, for definiteness, ay # 0. Substituting (2.2-5) in (2.2-4) and equating separately the coefficients of 20 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION the various powers of £ 10 zero yield 5(s — lap = 0 . (5 + Hsa, = 0 (s+ 2){s + 1)a, — (25 + 1 — Ajay = 0 (2.2-6) istu+ 2st yt laa — (s+ vt 1- da, = Since aq # 0, it follows from the first of Eqs. (2.2-6) thats = O ors = 1, from the second that s = 0 or a, = 0 or both. The last equation shows how the general coefficient a,+2 can be determined from ay. The first case to be considered is that of s = 0. Since a) # 0, the only way to terminate the sequence of a, with v even is to have =lvt1 (22-7) for some v. Since v is even, A can take on the values 1, 5,9,.. . . This choice of ® will not terminate, as is made clear from the last of (2.2-6), the odd a, sequence. The only way to guarantee a finite number of terms in H(é) is to put a; = 0. This prevents the odd a, series from ever ‘getting off the ground.” The same argument is repeated with s = 1; again, only even v terms are allowed and a, = 0, \ takes on the sequence of values A= Qvt3=3,7, 11 {2.2-8} and the resultant polynomial H,(€) is odd (since the even polynomial is now multiplied by €). If we combine (2.2-7) and (2.2-8), the allowed values for A become RK=2n+1 n=31,2,3,... (2.2-8a) which when using \ = 25/iw gives B, = fain +4) (2.2-9) for the energy of the nth eigenstate. According to {2.2-9), the harmonic oscillator, even in its lowest energy state, 2 = 0, has a finite amount of energy, iw. The lowest energy of a classical harmonic oscillator is zero. This essential difference is a manifesta- tion of the uncertainty principle. For the classical harmonic oscillator to have zero energy, both its momentum p, and position x must be simultaneously zero. This, according to the uncertainty principle, is impossible. The division of uncertainty between p, and x, which minimizes the total energy Ey while satisfying the uncertainty principle, gives Fy ~ fw. To show this, we take the total energy as ea( rian + a) 2.2 THE HARMONIC OSCILLATOR = 21 letting Ap, Sx = A/2 2 2 eee which when minimized with respect to Ap, gives Emin = Hie This result suggests that the lowest energy wavefunction (x) is a minimum uncertainty wavepacket, that is, a Gaussian. This indeed is the case since, according to (2.2-3), uo(x) * Holaxje "2" where Ho(ax) = av. Hermite Polynomials The solutions of Eq. (2.2-4} that correspond to the different values of A = 2n + L are seen to be polynomials of order 7. The polynomials are even when n is an even integer or odd when n is odd. If we put A = 2 + 1, the differential equation for the polynomials H,(é), which are known as the Hermite polynomials, becomes ?H, dH, eee ag tH = 0 {2.2-10) These polynomials are conveniently derived by means of the power scries expansion of the function e~**? according to Hel) on nt GE, 8) = eB UTEP cs gtd = > =o (2.2-11) Itis a simple matter, left as a problem, to show that H,{) defined by (2.2-11) satisfies the differential equation (2.2-10). To generate the Hermite polynomials, we notice that, according to (2.2-11), anes = {Seles (2.2-12) = ef cused] = ey—aye fhe = eff wig = e(-1) ee $ Applying (2.2-[2) lo generate, as an example, the first three H,’s gives HolE)= 1, (= 2, HY) = 4€?7— 2 (2.2-13} This particular sequence of H,(£) corresponds to choosing a = 1 in Eq. (2.2- 5) and is common practice. The Harmonic Oscillator—Creation and Annihilation Operators The harmonic oscillator wavefunction is, according to (2.2-3), HylX) = Nye O28" Hex) (2.2-14) 22 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION The normalization constant N,, is determined by requiring that f*2 uxu, a=. +0 rs 2 ae : [02 thay dx = NBL PP ax) dx = “ [Seung ag = The evaluation of N, is most conveniently performed with the aid of the generating function. Consider the integral [Reeaeence “4-353 sei Hyl€)Hy(G)e C8 dé (2.2-15) AO =o Replacing the definite integral on the lefi side of (2.2-£5) with its value w!?¢2" gives mite = Hayle) Hint Be ae Equating the coefficients of equal powers of sf” on both sides results in 2 # [i mtevmnterer® de = Cragin MEM (22-16) Substitution in (2.2-14) identifies N, as a N= (te (2.2-17) The normalized wavefunction is thus ue 4,40) = — Hydaxye OI? (2.2-18) A plot of some of the low-order u,(x) is given in Figure 2.1. We will now use the generating function to evaluate the integral ty flu Ga Consider first the integral [2 ce, new & pote. ne 21 de (2.2.19) which with (2.2-11) becomes te a : Hse nste (gn Op etnys2g-r {° € e age 14 “355 EPS yey 2 aatgyeON a (2.2-20) = QC eras 2 P(E + 28 ak — 4) ao - [us]? —— Ph ) —— FIGURE 2.1 Harmonic-oscillator wavefunctions, The solid curves represent the functions a” "u,(ax) with ax = € form = 0, 1,2, 3, and 10. The dotted curves represent a” ',4, forthe same values of n. The dashed curves represent the probability distri- bution for a classical oscillator having the same energy as the corresponding quantum-mechanical oscillator. The vertical lines define the limits of the classical motion. Source: R. B Leighton, Principles of Madern Physics (New York, McGraw-Hill, 1959), 23 24 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION —— mle) woo feof? —— Pulft -5 co) FIGURE 2.1 (Continued) This last integral is next replaced by the value of the two definite integrals (in the same order) that compose it A mIDs + tet + 221% = BE — sper nent gntiyn nt Equating equal coefficients of s*¢” in the last expression and in (2.2-20), and using (2,2-18) give n+ ye af z +8 din ve fC Mn Gy OK ~a(%) ma=n-1 (2.2-21) a otherwise We can use (2.2-21) to obtain another useful integral involving 1, and ty. 2.2 THE HARMONIC OSCILLATOR = 25 We make use of the following relation that holds (see Problem 1.8} for any pair of one-particle wavefunctions: > Rope du, [72 wfteyeniey av = mE! af Sade (2.2-22) ae SE) = A direct substitution of (2.2-21) gives V(t? _ ff ua - i > } = Vig ty m=ntl = we I Madr )xum(x) dx = VL (AVN? I Bae man (2.223) a \2 Ime 0 otherwise where the relation E,+1 — E, = feo was used. Another important integral, whose proof is assigned as Problem 2.3, is an +1 2a? os ; —— fc UglX)X7 eX) dx at fe oe (22-24) 2a 0 n#tm#tn2£2 Consider next the operators 4 and @* defined by « i « la a= Six4— p= %,+ 44 VI Vata Va" * Via and toy! pi mw, 1 io 22-25 OME Viti V2" ~ Vin x ‘ ) With these definitions and with the aid of Eqs. (2.2-21) and (2.2-24), it follows directly that fc dee [van meant) jn Matin OX = g otherwise (2.2-26) 4 _ iva m=n-1 Joo gat = { 0 otherwise To lea more about the operators at and a, we notice that Eqs. (2.2-26) are consistent with the relations atu, = Vn + 1 unas att, = VA By To show that Eqs. (2,2-27) are indeed correct, we need only prove that the functions atv, and ax, satisfy the Schrddinger equation with the eigenvalues B,+, and E,.r, respectively. This is left as an exercise. An alternative approach uses the recursion formulas for H,({€) as discussed in Problem 2.8. (2.2-27) 26 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION The operators at and a usually are referred to as the creation and annihi- lation operators, respectively. These names are used because, according to (2.2-27), the operation with a* (or 4) on the wavefunction x,, corresponding loa state with # quanta of energy fim, leads to a new wavefunction with 1 + | (or n — 1) quanta, thus “creating” (or “‘annihilating’’) one unit (quantum) fw of excitation. The commutator of a and at is my =[(2x+—bep) (4 ,-- fa, at] (Sx Taz Ge Fae! =~ sg lord + og real which using the commutator relation [pee x} = th becomes Iaat}=1 (2.2-28) Using (2.2-25), we can express x and 7, in terms of a and at as x= Tm (at + ay (2.2-29) iho = EX Cat = al P= 7k ) Substituting (2.2-29) in the Hamiltonian -#y) Hatt 3k and using the relations a* = mK/#? and w = VKim yield = 8 (aat + atay 2 Using {a, at} = 1, we obtain H = heo(ata + 4) (2.2-30) This is a most useful form of the harmonic oscillator Hamiltonian and it will be encountered in several subsequent developments. The operator ata commutes with H and has the number of quanta x for its eigenvalue. To show this, we use Eq. (2.2-27) atau = atVn tly) = Hy so that Hu, = fee(n + $)u, in agreement with (2.2-9). A far simpler procedure for evaluating integrals involving w,, such as those of Eqs. (2.2-21), and (2.2-23), is by use of the recursion formulas for the Hermite polynomials. This is discussed in Problem 2.8. 2.3 THE SCHRODINGER EQUATION 27 2.3 THE SCHRODINGER EQUATION IN SPHERICALLY SYMMETRIC POTENTIAL FIELDS A spherically symmetric potential is one in which the potential depends only on the distance from the origin, that is, Vir) = V(r). In classical dynamics, a particle moving under the influence of such a force field has a constant angular momentum L, where L = r X p. at _ ae ap. en aX TTX GHIXF where F is the force acting on the particle F = —VVir). For V(r) = V(r), we have where a, is a unit vector in the r direction. Classically, both £ and L are conserved in a spherically symmetric potential field. The corresponding state- ment in quantum mechanics is that the three operators H, L?, and L, (where z is any arbitrary direction) commute, so that the energy, the magnitude of the angular momentum, and its projection along any arbitrary axis (z) can be measured simultaneously with precision. Schrédinger Equation and Its Solutions in a Spherically Symmetric Potential Field ‘The Schrédinger equation in spherical coordinates is a [52 22 or aaele z * For a spherically syrametric potential, 4 can be separated into a product of functions 1 e 3 af. a = 75 (sno 3) + zak + Virju = Eu (23-1) wir) = Rr} YG, b) Substitution in (2.3-1) and introducing the separation constant A yield ld ,dR\ | 2mr? (e®) + [E- viz Rar w and (2.3-2} lef. ia 1 @) - [ag (snes) + sroap') very The radial differential equation can be rewritten as 44(P8) + fe [2 - voy - Al =0 (2.3-3) 28 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION whereas the angular equation is a 1 PY sin 6 30 wisn 8 4) + Gare age t Y= 0 (2.3-4) This last equation can be separated further into the product ¥ = 0(6)@() by the introduction of the separation constant #?.!' The two equations that result are 126, earn” 23-5) and 1 dj. de we _ sin a (sin 0%) * ( ~ Sn? 5/8 =0 (2.3-6) The solution for @ becomes = Aem + Be“™ sm #0 b=A + BG m=O Requiring that w(r, 0, 6 + 2m) = u(r, @, 6), that is, that u be a single-valued function of the real space coordinates, limits us to integral values of m and makes B’ zero. Both of these conditions are satisfied by writing m= 0,41, £2, 43 {2.3-7) where the (27)? factor was introduced for normalization purposes so that BBY Ob db = 1. Introducing the variable @ = cos 6 and taking @{8) = P(w), we obtain de la on S] + (- Using the series substitution method in a manner similar to that em- ployed in the harmonic oscillator, we find there is a physically acceptable solution for P(e) that remains finite for # = +1 (8 = 0 and a). This occurs only when A = i(/+ 1) and bn| 7-@ goeta@ Since P!,(a), where w = cos @, is equal to (1 — w)!"!? times a polynomial {in @) of order f — |m|, we have Yitar — 8, + aw) = {- 1) IPL (aemteime = Pi(dem(—1)! = ¥n(8. 6(— 1)! 30 SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION so that the parity is even when / is even and odd when / is odd. Otherwise stated, ¥}, has the parity (—1)'. 2.4 THE ANGULAR MOMENTUM OPERATORS AND THEIR EIGENFUNCTIONS The classical angular momentum L of a particle is L = r x p where L is measured with respect to the origin. Using the operator equivalent p> —iAV, we can write the operators L,, Ly, and L, as Ly = YP, ~ Py = ih (v2-72) Ly = 2Py — XP, = —ih (2-x3) (2.4-1) Ly = By — Px = ~it (x 2-3) In spherical coordinates, these operators become . a a L, = ih {sin o 3 + 019 cos 6 $5) Ly = th (-cos ¢ 3 + cap sin gd a {2.4-2) weak By using (2.4-2), the operator L? = L - L can be writien as la a loa = -# no 30 (sin 8 3) + wal (2.4-3} This operator is equal 10 A? times the operator appearing on the left side in the last of Eqs. (2.3-2). This operator was found to have as its (normalized) eigenfunctions the set Y},(8, ¢), as given by (2.3-12), with eigenvalues A = I + 1). Consequently, we can write directly the eigenfunctioris and eigen- values of L? LPYZ(8, 6) = WUE + 1)¥n(8, 6} (2.4-4) The possible values of the magnitude of the angular momentum are thus AVI? + 1). Since L? commutes with the Hamiltonian, being proportional to the angular part of the later, the magnitude of the angular momentum can be measured simultaneously with the total energy. We consider next the operator L, = —if 8/a@. L, also commutes with the Hamiltonian operator (2.3-1) and consequently has the same eigenfunctions. From (2.3-12), it follows that LY n(8, 8) = ALY (8, 6) (24-5) 2.4 THE ANGULAR MOMENTUM OPERATORS AND THEIR EIGENFUNCTIONS 31 so that the result of measuring the projection of L along zis one of the values mh. Tt can be verified straightforwardly that both £, and L,, as well as L;,, commute with L? (and H) so that the component of L along any arbitrary direction commutes with L*. We consider next the possibility of simultaneous measurement of any two components, say Z, and Ly, of L. We must consider the commutator [Le Ey] [Le Ey] = [ype — 2Py)e (ex — XP2)T = [yee Pel + Uepy Ape] = [Pe Pe — Pa YPx + WyXP2 — AP Py| Using [z, p.} = ih, so that zp, = pz + ih, gives [Le, Ly] = iilxpy — yp) = til, The other two commutator relations are obtained by cyclic permutation. The result is [Le, Ly] = AL, [Ly, Le = diy (2.4-6) Lez, La] = iby Consequently, no two components of L can be measured simultaneously. Supplementary References 1. Leighton, R. B., Principles of Modern Physics (New York: McGraw-Hill, 1959). 2. Messiah, A.. Quantum Mechanics (New York: Interscience, 1961). 3. Schiff, L. 1. Quantum Mechanics (New York: McGraw-Hill, 1959). Problems 2.1 Find the eigenvalues and eigenfunctions of the three-dimensional harmonic oscillator where V(r) = L(x? + + 2). 2.2 (a) Derive the second of Eqs. (1.2-22) using the relation Ap, Ax > A/2. (b) Use the reference mentioned in connection with the last of Eqs. (1.2-22) and outline the proof for AN A > 1. 2.3 Prove that dnt) 2a? [oo Bone dx = [Pena atte de = oa Vint lyr + 2) where 1, is the harmonic oscillator wavefunction. 32 2. - 2. we 2.6 27 28 2.9 2.10 24k SOME SOLUTIONS OF THE TIME-INDEPENDENT SCHRODINGER EQUATION Calculate the expectation value of the kinetic energy (p?/2m) of a harmonic oscillator in the state uy. Hint: Can you make use of Problem 2.3? Calculate Ap, Ax for the state 4, of the harmonic oscillator. Answer: Ap, Ax = (n + 4h Prove that atu, = Van #1 tte Atty = VA yy There is no loss of generality in taking the operators as an-e-g anbt where = ax. Starting with the generating function G(s, €) = e~"#? = Drao (H,(£)/n!|s", show that #,(£} obeys the following recursion formulas: dH, ag $y = Mine + Hy) = 2H; With the aid of the recursion formulas of Problem 2.7, show that for wt = a= fn nt 28mm FEY ZT Smet ty = VU ua) atu, = Va ¥ 1 tye where the u,'s are the harmonic oscillator wavefunctions. UpPatty dv = and that Express the total energy of a three-dimensional harmonic oscillator in terms of annihilation and creation operators. Show that £, and L, commute with E - L and H in a spherically symmetric V(r}. Show that if fap] = ih 2.4 THE ANGULAR MOMENTUM OPERATORS AND THEIR EIGENFUNCTIONS — 33 then (a. p'] = hp! Q) [p. @] = ~ alg’) 2) Hint: Use an induction proof. 2.12 Show, using (1) and (2) in Problem 2.11, that aa la, FB. @)] = th ap; = —i9 2F [pi FQN = ih Bai Cc HAPTER 3 Matrix Formulation of Quantum Mechanics 3.0 INTRODUCTION In Chapters ] and 2 the solutions and eigenvalues of the Schrédinger wave equation were obtained from a conventional solution of the differential equa- tion. Eigenfunctions and eigenvalues of Hermitian operators can also be ob- tained by matrix methods. This formulation is equivalent, formaily, to the differential equation approach, and this equivalence will be brought out in this chapter. 3.1 SOME BASIC MATRIX PROPERTIES It will be assumed that the student is familiar with the basic definitions and operations involving matrices. Consequently, the review of the necessary background of matrix algebra is very sketchy. The & element of the product of the matrices 4 and B is (AB)er = 2 AtmBrt and by extension (ABC) = DD) Ate BrnnCar Bi) The unit matrix J is defined as the matrix that, when multiplying a matrix B, leaves the latter unchanged. BL=B (3.1-2) IB=B It follows that fy: = 34. The inverse matrix 87! of B is the matrix satisfying the conditions B'B=1 (3.1-3} BE =1 The inverse of a product of matrices is equal 10 the product of the inverse matrices taken in a reverse order. (AB) = BAT! (3.1-4) 34 3.2 TRANSFORMATION OF A SQUARE MATRIX 35 The elernents of A=! are related to those of A by the relation cofactor of Ay -) = : Ait > determinant of A (3-1-5) Notice the reverse order of k and / on both sides of the equality sign. The Hermitian adjoint of a matrix A, denoted as At, is defined by Al, = Ah (3.1-6) A matrix is Hermitian when it is its own Hermitian adjoint, that is, when At=A G.1-7) Equation (3.1-7) can be written, if we use (3.1-6), as Al, = Aw = Ak (3.1-8) so that in a Hermitian matrix the interchange of rows and columns is equiva- Jent to replacing each element by its complex conjugate. This applies, conse- quently, only to square matrices, It also follows that the matrix elements Ay along the main diagonal are real. A unitary matrix is defined as the matrix whose Hermitian adjoint is equal to its inverse At=a-t (3.1-9) From (3,1-9}, it follows that (4At) = I for a unitary matrix A that when written in a component fonn becomes (AA =D Aves =X AAs = 8 (3.1-10) where use was made of (3.1-6). 3.2 TRANSFORMATION OF A SQUARE MATRIX A matrix A’ derived from a square matrix A by the operation Al = SAS" GB.2-1) is called the transformation of A by (the square matrix) S. Matrix equations are left invariant under a transformation of the individual matrices. A typical equation containing matrix products and sums such as AB + CDE =F can be written as SAS~'SBS"! + SCS~'SDS™! SES~! = SFS~* since S'S = I. The last equality, according to (3.2-1), is AB’ + CDE’ =F (3.2-2) 36 9 MATRIX FORMULATION OF QUANTUM MECHANICS 3.3 MATRIX DIAGONALIZATION Of special interest in quantum mechanics is the transformation A (of a square matrix A’) that is diagonal. We will illustrate in Section 3.6 that finding the elements Aj and those of S takes the place, in the matrix formulation of quantum mechanics, of solving the wave equation. A is assumed to be diago- nal with (unknown) elements Ay = A, so that (SA'S ag = An Sig Postmultiplying the relation SA’S~' = A by S, we obtain SA‘ = AS and taking the &/ element yields D Sent = 3 AcoSnt = AeSet or D Sen(Ant — Ak Bua) = 0 B31) if S is an N-dimensional matrix, we obtain N equations by fixing < and writing (3.3-1) for each value of / between 1 and N. These N simultaneous and homogeneous equations for the N unknowns Sia... - + Sew have nontrivial solutions only when the determinant of the matrix made up of the coefficients vanishes det [Amt — At Bm} = (3.3-2) The N solutions 4;,.. . , Aw of Ae that result from solving (3.3-2) are the diagonal matrix elements sought. They are also referred to as the N eigenvalues of the matrix A’. The remainder of the problem is to find the elements Sg, of the transfor- mation matrix. A given A, is substituted in (3.3-1). The N homogeneous equations can be solved to yield Sa. Siz...» + Sin to within an arbitrary multiplying constant. This procedure is repeated with each of the N A;’s, thus generating the matrix Sj». The additional condition necessary to determine Sim uniquely is, as will be shown in Section 3.4, that S be unitary, so that, according to (3.1-10), D Sees = bu 3.4 REPRESENTATIONS OF OPERATORS AS MATRICES In the matrix formulation of quantum mechanics, an arbitrary operator A is represented by a matrix A. (We will depend on the context to avoid confusion between a matrix A and the operator A.) A particular matrix representation A 3.4. REPRESENTATIONS OF OPERATORS AS MATRICES 37 is derived through the relation Ain = [ E(t) At (F} dv B41) where un(r) is any arbitrary complete orthonormal set of functions. The represen- tation Ay, of an operator A is, consequently, not unique and depends on the {arbitrary} choice of the set %,(r). The operator A can have other representa- tions as well. Let one such representation, obtained in the space v,, be A’. It follows that Ain = f VE AV, dv (34-2) A Unitary Transformation Matrix We can expand an arbitrary member of the set v, in terms of the set v4 vale) =D Sted) (3.4-3} @ The reverse expansion becomes 4 =D Shab (3-4-4) It follows from (3.4-3) or (3.4-4) and the orthonormality of the sets u, and v, that Sin = J ubv, dv (3-4-5) The set of numbers Sj, can be regarded as a matrix. We will refer to it as the transformation matrix from the space v, to uz. The matrix S defined by (3.4-5) is a unitary matrix. As a proof, we may show that SSt = I where Tis the unity matrix (SSta = SinShe =D SeoSHe =D f ubteyvnte) dv f ware) dv” = f ainutery f ee’) S vetryeneey det = J nate J u(r") 8(r — 2°) dv" = | utente) dv = be An alternative proof, and one thal may shed some more physical insight into the nature of unitary transformations, is to show that a unitary transforma- tion is required so that the (squared) magnitude of an arbitrary vector f (in abstract vector space) (on =f pray remains the same when fis expanded in terms of ¥, or wy. For further discus- sion, see Problem 3.1. 38 MATRIX FORMULATION OF QUANTUM MECHANICS 3.5 TRANSFORMATION OF OPERATOR REPRESENTATIONS We have considered in Section 3.4 two arbitrary representations of an opera- Yor A. one in the space v,, the other in x,. Aug = | ub Are dv Au = f ve Ay, dv (3.5-1) w= 5 Shy We can derive the matrix A from A' and vice versa by the transformation A = SA'S" = SA’St (3.5-2) where S$ is the unitary tansformation matrix defined by the last of Eqs. ann proof of Eq. (3.5-2) consists of replacing #f and u, in the first Eqs. (3.5-1) by their expansion according to (3.4-4). Ay = { ufaw dv =f (2 Set) A D Stn de - DIS J Av, dust, = DD SiAimSte = (SA'S If the matrix A’ is Hermitian, so is the matrix A. In other words, the Hermiticity of a matrix is invariant under unitary transformations. The proof consists of showing that Aw = Ai where A is given by (3.5-2) Au = & StmAnnSts = 3, Steer) *Sb, An =D Sim ArnSin where the relations Am = (Anm)* were used. Taking the complex conjugate of the last equation and interchanging m and n lead to (Aa)* =X Sh (Ann) *Sum = Aur which completes the proof. A corollary of this result is the following: The eigenvalues of an Hermitian 3.6 DERIVING THE EIGENFUNCTIONS AND EIGENVALUES OF AN OPERATOR 39 matrix are real. This result follows from the fact that any unitary transforma- tion of an Hermitian matrix is Hermitian and therefore has real diagonal elements. This applies also to the diagonal transformation whose elements are the eigenvalues. 3.6 DERIVING THE EIGENFUNCTIONS AND EIGENVALUES OF AN OPERATOR BY THE MATRIX METHOD Finding the eigenfunctions and eigenvalues of an arbitrary Hermitian opera- tor A consists of solving the equation AUy = Antin 3.61) The set of functions x, is referred to as the eigenfunctions of A, and the (real) numbers A,, are its eigenvalues. As an example of an eigenvalue problem which we have already solved, we may take Eq. (2.2-1). The eigenfunction u, is given by (2.2-18), while the eigenvalues are, according to (2.2-9), Ey = ha(n + 9). A second example is provided by the operator L? in a spherically symmetric potential field. The eigenfunctions and eigenvalues are given by Eq. (2.4-4) as LY m(8, O) = E+ Vl @, An alternative approach to solving an eigenvalue problem, such as (3.6-1), is by use of the matrix methods developed above. The matrix repre- sentation of A in the Hilbert space , (where the u,’s are the eigenfunctions of A) gives a diagonal matrix with the eigenvalues A, as the elements Au = f utAuy dv = Bf utu dv = Eby It follows immediately that if the representation A’ of an operator A in any complete and orthonormal set v, is known, we can obtain directly its eigenvalues and eigenfunctions. The procedure consists of finding, by the method discussed in Section 3.3, the matrix S that diagonalizes A’ Ag = (SA'St)u = Ag Bir The diagonal elements A, are the eigenvalues of A. Since, according to Section 3.5, the same matrix Sis used in transforming from the v, to the %,, space, the eigenfunctions are given by . ude) =D Stavele) 6.62) if only the eigenvalues of A and not the eigenfunctions are desired, it is not necessary, according to Section 3.3, to obtain the transformation matrix $. The eigenvalues A; are the solutions of the determinantal equation det [Ant — Ay Sm] = 0 (3.6-3} To convince ourselves that the eigenvalues of A obtained by matrix diagonalization are the same as those obtained by solving (3.6-1}, we con- 40 MATRIX FORMULATION OF QUANTUM MECHANICS sider the matrix elements of the operator A in the function space u,. Using (3.6-1), we obtain Aum = | UpAtty d= Ay Ban so that the matrix A,» is diagonal with its elements being equa! to the eigen- values A,.. Since the two matrices A and A’ are derivable from one another by a unitary transformation, they possess the same eigenvalues A,. The procedure described above applies to the eigenvalues and eigenfunc- tions of any Hermitian operator. As an illustration, let us consider formally the Hamiltonian operator H. R H=- Im w+ Vir) whose eigenfunctions and eigenvalues are u, and &;, respectively, so that Huy = Eyily (3.6-4) Let the Hamiltonian matrix in the (arbitrary) space v, be denoted as 4’ Hay = J vAHY, dv G.6-5) The transformation coefficient S is defined by i = > Sh¥n We will show below that the transformed matrix A= SH'St is diagonal with its elements equal to the eigenvalues E,. (SH'St)y = > = Sim Enon Shy = DE J aberration av [ vatervtesry av’ f naetrger) de = feewtay f ave watercraft) Sve yeste de = | andes fa ace — rye [ ute") 80 - 2) a" = f vat ey { (er — r')Huair’) dv’ = J uf (1) Hug(r) dv = Ey 8a where use has been made of (1.2-11} and of (3.6-4), This completes the proof. We are now finally in a position to justify the assumption that operators that correspond to physical observables are Hermitian. The necessary and sufficient condition for an operator to have Hermitian matrix representations 3.7 THE HEISENBERG EQUATIONS OF MOTION 41 is that it be Hermitian (see Problem 3.5). Since an Hermitian matrix has real eigenvalues (see Problem 3.4) and these correspond to the possible results of physical measurements, it follows that the corresponding operator must be Hermitian. 3.7 THE HEISENBERG EQUATIONS OF MOTION In Section 1.1 we derived the equation of motion for an arbitrary operator B as d{B) _ (2) i “a Nal th (LH. By) GB.7-1) where (B) = J w*By dv and ¢ satisfies the time-dependent Schrédinger equa- tion i . Boge (3.7-2) From the last equation, it follows that! wir, t) = ehp(r, 0) so that (8) can be written as (B) = J [e-hye(r, 0)]*Be“ p(x, 0) dv = { Wir, Oe Be Mehr, 0) dv (3.7-3) = f wr, O)* By (b(r, 0} dv (3.7-4} Moving the operator ¢~#* to the right of #(r, 0} in (3.7-3} can be justified by expanding it in a power series and treating each power of H as an Hermitian operator. By(t) is defined by the last equation as Bult) = eH Bette (3.7-5} and is referred to as the Heisenberg form of B. According to (3.7-4), in the Heisenberg formulation the expectation val- ues of physical observables are evaluated using the wavefunctions at a fixed time (¢ = 0) so that the wavefunctions are stationary. The operators, however, evolve in time in accordance with (3.7-5}. To evaluate (B), we thus need to know B,(t). The differential equation describing the evolution of B,(¢) can be derived straightforwardly from (3.7-5) as aB;At) i to =F arene ii — cranny + =! uaa + (2) on aq | ale 1 The operator e”#** is equal to Zn-o (—iHE/hy"/n!. 42° MATRIX FORMULATION OF QUANTUM MECHANICS This formulation is especially useful in problems involving interactions with quantized boson fields. It will be used in Chapter 17 to describe energy exchange in nonlinear optical processes. 3.8 MATRIX ELEMENTS OF THE ANGULAR MOMENTUM OPERATORS As an illustration of the ideas developed in this chapter, we shall derive next some matrix elements involving the angular momentum operator. These rela- tionships will play a central role in later developments concerned with the addition of angular momenta and with the orbital and spin magnetic mo- ments, The raising and lowering operators L* and L~, respectively, are defined by Lie i,t il, (38-1) Lo = Ly ily These operators obey the commutation relations [E*, LJ = ERL* (2, 27] =0 (3.8-2} [L*, L-] = 2hL, These relations can be proved with the aid of Eqs. (2.4-6). As an illustration, consider the first of Eqs. (3.8-2) (LE, Le] = [(Le * iby), Ea] = Uae Le] * ily, Le] = ~ifily + AL, = ERL* The result of operating with L* on the eigenfunctions ¥;,(8, 6) of L? (here we replaced / by j) can be studied by the following development: LLY) = (Lt Ly + BL*) Yh, = (MALE = BLEVY), = (rm = LYA(L* Fa) where we used the relation 2,¥%, = mh Yi, and the first of Eqs. (3-8-2). The last equality states that £* Y}, is an eigenfunction of L, with eigenvalues (m + 1h. These have been found to be the functions ¥j21. We can thus write directly L*¥4,(8, 0) = BC Vine (8, b) (3.8-3) where the constants Cj remain to be evaluated. Consider next the matrix elements of L?, £,, and L* in the function space vi,(0, $). Using Eqs. (2.4-4} and (2.4-5) results in Osmo = J PT Le, GIMLE YS 8, 6) sit. 8-40 cep HG + YR? 8) Bnet Ladjmg me = Mh 357 Sim (3.8-4) (L°)jmrigm = Gah = Caf where Cy = Cy by definition CE Yamjmss = Cutite = (Cu) 4H = Cube 3.8 MATRIX ELEMENTS OF THE ANGULAR MOMENTUM OPERATORS 43, The last two equations involving (L*) were derived using Eq. (3.8-3). The proof of the relation Ga+1 = (Cn)* is left as an exercise (Problem 3.9). In order to derive the constant G,,, consider the mth diagonal element of the third of Eqs. (3.8-2) (LAL> = LOL Yano = 28 (Le nm = Drath? where the subscript j is omitted since matrix elements involving states with a different j have been shown, in (3.8-4), to be zero. All the relations developed below consequently will assume a constant j, Expanding the matrix elements gives DS Sim Lin me — La inn = Liem semmti ~ Laas eee ae = Dat? or Co 1 Cr where use has been made of the last two equations of (3.8-4). The last telation is rewritten as 1 GrCm = 2m \m—al? = [Cl = 2are GB.8-5) |, we use the relation W= 12+ KEL + LL) (3.8-6) To evaluate |C,| and take the m, m diagonal matrix elements of beth sides. The matrix elements of the first two terms are taken directly from (3.8-4), the result being ‘ HU + A = mH + ED asl + Lasts) = mh? + HC iCm + Cam) G.8-7) R met + EG? + al) Combining this result with (3.8-5) gives Cn = Vii + 1) — mom + 1) where the phase of the wavefunction is assumed to be such that C,, is real and positive. Substituting C,, in Eqs. (3.8-3) gives Lig = BVIG +L) — mera + 1) Yee (3.8-8) LoVe = AGG FY — me — VY) Yn or (LV jmjimer = (LY menue = AVIG +L) — mon + 1) 3.8-9) These are the desired results. Using Eqs. (3.8-4) and (3.8-9}, we can construct the matrices corre- 44 MATRIX FORMULATION OF QUANTUM MECHANICS sponding to the various angular momentum operators in the space ¥3,(0, 6). Since the matrix elements involving states with different j values are zero, we may limit ourselves to the submatrix within a constant j manifold. (The term * manifold will be applied often in this context to describe the subspace made up of the 2j + 1 eigenfunctions Y},(@, 6) with —j-< m = Carn ma| jijjmach, B, (3.10-2) where the (|) symbol represents the elements of the unitary uansformation matrix connecting the two sets of eigenfunctions. It is clear that only the product wave functions «82, where m, + m = m are to be included in the summation on the right side of (3.10-2). This can be verified by operating on both sides with L, = L,, + L2,. In consequence, we have oe =D Chia (m — mdMijrimers, a _, (3.10-3) Another property of the (|) coefficients is Ubiblili +b +h) =1 (3.10-4) so that wd = ahah G.10-5) This results from the fact that according to (3.10-3}, m = m, + m=), + jp and there is only one product function, given by (3.10-5), where m, + m, = du + jz. Since j = m, it follows that j = j, + jy. It follows from the above discussion that the maximum value of is equal to j, + j2. With each value of j, there are 2j + 1 wavefunctions with —j < mm © j, Since the total number of wavefunctions. yi, must be equal to that of the function space a/,p%,, namely (2/, + 1)(2j. + 1), it follows that the smallest value of j is |; — j2|. This can be verified from the summation ide S +1) = Bi + Ria + 1) bid The coefficients {|} are known as the Clebsch—Gordan coefficients. Condon and Shortley (Reference 1), as an example, tabulate these coefficients for a number of j; and jz. 3.11 TEME-INDEPENDENT PERTURBATION THEORY 47 For simple cases, it is quite easy to derive the eigenfunctions of L = L, + L, starting with (3.10-5). Consider, for example, “adding” the angular momenta of two particles with j, = j2 = 1. We have from (3.10-5) 3 aif! applying L~ = Ly + ZL; to both sides and using (3.8-8) result in Loh = yt = V2¢088i + a8) so that ut = Je taial + alah The next function yj can be written according to (3.10-3} as WU} = aeaBt + baiBo Requiring that (q}, 1) be unity gives (for real a and b) +h =1 whereas the condition (}, Wi) = 0 gives a+b=0 The last two equations are satisfied by that, if we choose arbitrarily the upper sign, gives L 1 tol Lah == - 3. 10-6} Oh = 5 (abt — api) (3.10-6) This procedure can be used to generate the remaining eigenfunctions. 3.12 TIME-ENDEPENDENT PERTURBATION THEORY In this section we pose the following problem: Given a Hamiltonian Hy and its spectrum of eigenfunctions u,, and eigenvalues £,, Holle = Eyl G.U-l) What are the new eigenfunctions and eigenvalues when the Hamiltonian is changed, adiabatically, from Hp to Hp + H'? One method of solution would be to diagonalize the matrix Hp + H" in the u, representation as discussed in Section 3.6. This method is often used in practice. If Hy > H’, we can employ perturbation theory and obtain analytic expressions for the perturbation of u, and £, to any desired ordér. This is the concern of this section. The Hamiltonian operator is taken as Hy + AH’ where 0 (3.11-3) and substituting in (3.11-2) give (Ho + AH’ Vly + Ady + AMa + +) = (Wo + AW, + Wee Vho + Aya + Me +--+) Equating the coefficients for 4°, \!, and 42 on both sides of the last equation gives Holho = Warbo Hots + Hho = Wot + Witt (311-4) Foz + Hy = Woe + Wid + Wao respectively. Comparing the first of Eqs. (3.12-4) with (3.11-1) identifies the zero-order solutions as bo = Ue Wo = Em where # and £,, are the eigenfunction and eigenvalue at the absence of perturbation. Next, we expand , in terms of u, as w= > aa, (3.11-5) and substitute it in the second of Eqs. (3.11-4}. The result is D AWB jtin + Hig = By > A! uy + Wit Multiplying by uf and integrating give Bal! + Hoy = Eqal!) + Widen (3.11-6) that for k # m yields Hine (1) oe ——E - oth km (3.11-7) Putting & = m in (3.11-6) gives Wi = Hoag (B.11-8) 3.11 TIME-INDEPENDENT PERTURBATION THEORY 49 We still need to evaluate al). This is done by requiring that = um + th be normalized to wnity [ [un Satan] [an +S a] av ans, = 1+ a + at +S altat) = 1 that, if we neglect the second-order term, gives a{!! = 0 for a choice of phases that renders aj? real. The eigenfunction and eigenvalue to first-order perturbation are thus given as = Hon Y= tn +2 Ent (3.11-10) W = Ey + Hn Second-Order Perturbation The starting point is the third of Eqs. (3.11-4). Expanding #. according to da = 2 ai lty and substituting in (3.11-4) give Da Eyl + HD) tee =D a Ege + With + Watt Substituting for its expansion according to (3.11-5), and then multiplying by uf and integrating result in ae. + 3) albag, = aPley + Wral + WoBmx Q.-11} Setting k = m gives We = al Hie, — Wral? =D abn + aH ~ Wiel! nem If we use (3.11-7) for af!’ and (3.1 1-8) for W;, the last ewo terms cancel each other with the result |F Ey — Ey Q.11-12) 50° 9 MATRIX FORMULATION OF QUANTUM MECHANICS: Going back to (3.11-11) for the case k # m, using (3.11-7), (3.11-8), and the result a!’ = 0 give . oy) Hin hin Hn Hi i 2 eee a) AP To find a!?!, we go back to the normalization integral (3.11-9). Adding the second-order correction to gives [nt Sates Saluda eS oes Baa] If we use the result af! = 0, the last equation yields op a = 13 fail = Heel (3.11-13} 7 154, (En ~ En} The eigenfunction and the energy, to second order, can be written as High owes SE Se 1 The = ExEw ~ Be) - “tate — [Hin ,, } 3.11-14) (By ~ 8)! 2B = BP Baris = ‘ [En |? - W= Ext Ham + 3. Ek (3.11-15) Notice that the second-order correction tends, according to (3.11-12), to increase the energy separation | — E,|. This fact is often expressed in the physics jargon as “Energy levels repel each other.” 3.12 TIME-DEPENDENT PERTURBATION THEORY—RELATION TO LINE BROADENING Another class of problems that is often treated with the aid of perturbation theory is one in which the Hamiltonian operator varies with time. This hap- pens, for example, when the system interacts with a monochromatic radia- tion field, so that the Hamiltonian is perturbed harmonically, or when the Hamiltonian is changed by an abrupt step from one value to another. These two situations will be treated in this section and it will be shown that in both cases the perturbation induces transitions between the stationary states of the nonperturbed Hamiltonian. The perturbation Hamiltonian is taken as H'(f) so that the total Hamilto- 3.12 TIME-INDEPENDENT PERTURBATION THEORY 51 nian operator can be written as H= Hot HY) (3.12-1) where Hp is the nonperturbed Hamiltonian obeying Hotty = Enn (3.12-2) The solution s(t) of the Schrédinger equation wot he (3.12-3) can be expanded, at any time ¢, in terms of the complete orthonormal set “4, obeying (3.12-2}. Wt) =D) aatpene Eh (3.12-4) The coefficients of the expansion are chosen as a,(t}e~*“* and have the con- venient property that if H'(#) is identically zero, the a,(t) become constants. Substituting (3.12-4) in (3.12-3), we obtain D un [a (- Sr) eiah 4 ag] =- i D 4n( Hy +H’ juego oh which after multiplying by uj and integrating becomes a= — >) AqHjneiate! (3.12-5} where e, is defined by Ex— En oy = AE Up to this point, the analysis is exact and solving Eqs. (3.12-5) is fully equiva- lent to a sohution of Schrédinger equation. In a manner similar to that used in Section 3.11, we introduce the “turning on” parameter A by taking the per- turbation as AH’, so that the Hamiltonian becomes H= Hy) +hH The power series expansion for a, is written as ay = ah + ay + Nal) + --- which when substituted in (3.12-5) becomes a 4 Aap + tae ee ; D fal? + ratll “+ a2a + Jatt! 52 MATRIX FORMULATION OF QUANTUM MECHANICS Equating the same powers of A results in the set of relations a =0 a = — ED alate af =~ i> ay, (theent (3.12-6) ap = — 3D a aire The solution of the zero-order equation is af! = constant. The af?’ are thus the initial values for the problem. These are chosen as ay = a=0 ntm so that at = 0 the system is known with certainty to be at the state m. For this special case, the second of Eqs. (3.12-6) becomes af! = — 5 Hine'omt (12-7) Since at ¢ = 0 the system is at the state m, |a(f)?? is the probability that between t = 0 and ¢ the system made a transition to the state k. Harmonic Perturbation As a special case, we consider a perturbation that is modulated harmonically according to H(t) = He ™ + (A )te t>o H'() = 0 Ey, and E; — Bm ~ fw, whereas the second term dominates when E, < £, and Ey — FE, ~ hw. The harmonic perturbation can thus cause both upward and downward transitions from state # to states k, separated, in energy. by ~fiw. To be specific, fet us calculate the transition probability from m to a group of states clustered about state k where Ey > E,,. Let the density of these final states per unit Of @um be p{@m). Using only the first term of (3.12-8), we obtain D de = 5p BS Uti jp Slee — ol Plan) dédin (312-10) If |Hin/? is not a function of the final state, we may take it outside the integral sign. In addition, we limit the discussion to times ¢ large enough so that the width A(dx_ — w) ~ 2a/t of the function sin?[}(@p2— ©)t}/ (cox — @)? is much smaller than the width of p(@p,). Under this condition, we may take p(e@m = «) outside the integral sign, and using the definite integral = sin? (xct! ce sme) 2) abe g (12-11) obtain . Peale lal?}? = Feb = w)t (3.12-12) The transition probability rate from m to the continuum of states near k is thus 2a) Worst = SF |Himl? P(E = En + fio) {3.12-13) where p(E) dE = p(w) dw is the number of states in the energy range dE centered on E. This result, often referred to as the “golden rule,” is of central importance in the study of atomic transitions. The condition that 21/f be small compared to the width of p(w) is equivalent to treating the function sin?[3(@im — ©)f]/(@n — @)? as a narrow sampling function. We can thus substitute sin?{Morm — o)t] _ at (ain aye 2 On 54 MATRIX FORMULATION OF QUANTUM MECHANICS in Eq, (3.12-9) and obtain 2 lal"? = Fo [Hinl* 8 (eq — ©) = 2% 1Hin?t 8(84 — Em — Bea) or Wiese = AE Hig? 8(B, ~ Ein ~ He) (3.12-14) for the transition rate from m to a single state k. Equation (3.12-14) is espe- cially useful as a starting point in analyses where the specific details of the density of final states must be taken into account. This is done by multiplying (3.12-14) by the density of states function and integrating over alt energies. For the general case where the density of states function in energy (E) space is taken as p(E), the integration yields Eq. (3.12-13). . ‘The consideration of downward (i.e., Ex < E») in addition to upward transitions involves using both terms of (3.12-8) rather than the first one alone, with the result Worst = 2 |g BE ~ Bi = fn) + Eig? BUF ~ Ea + Reo) B.12-15) for the case where both upward and downward transitions are considered. Asan example, consider the transition rate W,,.; to a single state k where, to be specific, E; > En. Let the position of the energy £; be so smeared that we may describe it only by the probability function f(E,) 4B; of finding it be- tween E, and E, + dE, so that [2 fe) de = 1 The transition rate is then 2 az Waren =F [Hbnl? [7 8 Be — Ein — Ho) fe) aE (3.12-16) = 7 (High? f(Ex = Em + fo) If we replace f(£) by the corresponding function g(v) in the frequency do- main, we have g(v%) = 27 f(Ex) so that [72 ate) de = 1 and g(v) is normalized. Equation (3.12-16) takes the form 1 Worse = 5 |Hhnl? gle = Yn +») (3.12-17) 1 = 9p Hil? 9) 3.12 TIME-INDEPENDENT PERTURBATION THEORY 35 where g(v), the (normalized) natural lineshape function, also describes the transition strength (i.e., rate) as a function of the applied frequency v. The “smearing out” of the wansition m— & and the (resultant) need to describe it probabilistically by a lineshape function g(v) may be due te a variety of causes. Two typical causes are (1) a transition originating in an atom (or ion) inside a crystal depends on the local surrounding crystalline electric field. This field may shift both £,, and E;, so that E, — &,, varies from one atomic site to another due to, say, strain induced fluctuations of the local field. (2) A transi- tion involving an energy difference ty = (£; ~ Em) undergoes a Doppler shift Av = (v/c)v in a moving atom where v is the component of velocity along the direction of observation. This causes the radiation emitted or absorbed by an ensemble of atoms (or ions) in a gas (or plasma) to have a linewidth of Av ~ viceVkT/m where T is the temperature, c the velocity of light, and m the mass of the atom. If the velocity distribution is Maxwellian, the resultant lineshape function g(v) is Gaussian. These two forms of broadening belong to a class referred to as inhomogeneous broadening. A second class of broadening is due to the finite lifetimes of states m and & involved in the transition. This gives rise, according to the uncertainty principle, to an uncertainty in the energy sepata- tion A( Ey, — Ex) ~ f(z,’ + 7%!) where 7» and 7, are the lifetimes of levels m and k. This form of broadening is called homogeneous broadening. The basic difference between homogeneous and inhomogeneous broadening is that in the former the concept of a broadened transition applies to a single atom, whereas inhomogeneous broadening results from a consideration of a large number of atoms, each atom with its own different resonance frequency. This point will be discussed in detail in Chapter 8, Limits of Validity of the Golden Rule. Two conditions were used in deriving Eqs. (3.12-13) and (3.12-15). The first was that 27/t be small compared to the width of Av of p(@:). The second condition results from our use of first-order perturbation theory and requires that |a{(4)? < 1; otherwise, higher-order terms must be considered. This second condition can be stated, using (3.12- 9), as [Hie 1 a ch Aven (3.14-4) Now suppose that the precise state of the system is unknown. This lack of knowledge is reflected in an uncertainty in the values of the ¢,’s in the expan- sion of w(r, t}. Assume, however, that we have enough information to calcu- late an ensemble average for c4¢,. The average will be denoted by a bar over the quantity in question. Thus, one can compute an average value of the expectation value of A according to (4) = FG Amn (314-5) It is convenient to define _ Pam = Caen {3.14-6) The matrix formed by the values of pi is known as the density matrix. Using (3.1-1), we obtain. (4) = > (Adon (314-7) This computation is indicated by the trace, abbreviated “tu.” Thus, {4) = wip) (3.14-8) and is (see Problem 3.2) independent of the choice of the basis functions u,{r). It follows from (3.14-6) that Pm = pi, $0 that p is an Hermitian matrix. Another important result is that up = 2» km = 1. This follows directly from the normalization condition of y(r, t) in Eq. (3.14-1). 3.15 THE ENSEMBLE AVERAGE The type of averaging indicated above by a bar is what is known. as the ensemble average. This ensemble averaging process can be interpreted physi- cally in the following manner. One prepares an ensemble of N systems (N large) so that the systems are as nearly identical as allowed by one’s incom- plete information. The systems are then allowed to evolve in time. Each system is thus characterized by a state function Hrs) = DP (eat) (3.05-1) 58 MATRIX FORMULATION OF QUANTUM MECHANICS fors = 1,2,. . . , N. The ensemble average of c4<, is computed according to the following formula: y : Ponti) = ERE = nD reno (348-2) Then, the ensemble average is an average over all N systems. In this physical interpretation, the density matrix represents certain prob- abilistic aspects of the ensemble. The diagonal term fn» is the probability of finding one of the systems in the ensemble in the state u,(r). The off-diagonal terM Pm(f} is equal to the ensemble average of c(f)cn(f) that will be shown in Chapter 8 to be related to the radiating dipole of the ensemble. 3.16 TIME EVOLUTION OF THE DENSITY MATRIX Since the wavefunction of each system in the ensemble satisfies Schrodinger’ s equation, then Abt, t) = ne 5 (3.161) Substituting (3.14-1) for p(x. t) yields . i & A astry = S ttt (316-2) Taking the inner product of (3.16-2) with u(r) and using the orthonormality of the u,’s, one obtains it SeglO= Del 00Hae (3.163) But from (3.14-6) 5P rm oe sy a tar Oe B.16-4} By making the use of (3.16-2) and the Hemmiticity of H, (3.16-4) reduces to eat = ip, HI (3.16-5) A more detailed discussion of the properties of the density matrix and meth- ods for evaluating it for physical systems can be found in Reference 2-4. It will be applied in Chapters 8 and 15 to describe the interaction of a two-level atomic system with a radiation field. In Appendix 4, we will use it to describe nonlinear optical processes. 3.17 THE TIME EVOLUTION OPERATOR—FEYNMAN DIAGRAMS Another very important formalism in our tool box of perturbation methods is that of the time evolution operator. It is closely related to the topic of the 3.17 THE TLME EVOLUTION OPERATOR—FEYNMAN DIAGRAMS 59 Feynman diagrams. We shall make extensive use of it in describing multipho- ton (2nd, 3rd, etc.) optical processes and, specifically, to obtain an expression in this chapter for the two photon absorption coefficient and in Chapter 16 for obtaining expressions for nonlinear optical coefficients. The eigenfunction V(t) of an atom subjected to an electromagnetic field can be obtained formally by solving the Schrédinger equation =» ab Hy = h> (3.E7-1) or, equivalently, by operating on the eigenfunction at time f, with the evolu- tion operator u(t,, f,) according to Wl) = Kt, ta) b(t) (3.17-2} where u(t,, tz) satisfies Ou{ fy, fa) ip eet) saat, G.17-3) aly If H does not depend on time, then it follows from (3.17-3) and the Hermitic- ity of A that u(fy, te) = exp [ 317-4) = J bm) dal expl—ien(ts ~ t2)] where @» = E,,/h, |m) is the eigenfunction of H with energy Ey, that is, H|m) = he,|m) and where we made use of the identity >, fm) Gn] = 1. In the cases of interest to us, we take the Hamiltonian as Hit) = Ho + Vid) (3.17-5) where Hp is time-dependent and V{z) represents the time-dependent intereac- tion of the atom with the optical fields. It follows directly from the definition (3.17-2) that w is a unitary operator. To prove the fast statement, we start with iG) (4) = 1 normalization condition (velop) = Qplutulp) = (ly) = 1 where ut is the Hermitian adjoint of v > utu = I(=identity operator) (3.17-6) and x is a unitary operator. Another property of is Ulty, fa) bas te) = Ute te) (3.17-7) From the last two numbered equations, it follows that H(las te) = (ty. ta) (3.17-8) 60 MATRIX FORMULATION OF QUANTUM MECHANICS Jt can be shown by simple substitution that the formal solution of (3.17-3) when H(0) is given by (3.17-5) is Hlty. fa) = Wty ta) f WOM, QVM, ta) at B.17-9) with iH uO (ty, t) = exp [- = (e- «| (3.17-9) is the propagator operator at the absence of perturbation. We can use the last result as a basis for a perturbation expansion. We substitute for u(t,, fa) Ute, ta) = WMD, 2) — aft HONE, 1) HEE ef fa) PY? pte pr + (- i) ae, ae te WUE eH tad Vlada fe) (B-AT=10} hou >h>te Note again the sole appearance of » rather than # at the very end of the integral. Repeating the process, we obtain wt fa) = HM te) — Ff wt WIM CWM, td + (= 5) ef ae dear, viene, MIM, ted BATA) 12 pw pr pe (HF) PR PE ae ats denen, vey Ut, AVEC IW la IVE IU te) We can thus write formally Ueto, ta) = HONE fa) + whe fa) + HOM, fe) (3.17-12) pe Mtge fe) Boe where 1) (ty, ta) = EXP [-i % (hy = «)| wlth = (= 2) fee everett) a w= (- iy Jo FF ats dente, VUE) W(t), fy} V(E2)UM E2, fa) (3.17-13) whe (- i ER i ty dty dtu (Ly, ) Vit HOt, Via) X al) (tp, ty) Vda )eeM ty, ted 3.17 THE TIME EVOLUEION OPERATOR—FEYNMAN DIAGRAMS 61. where >> hr hr >t To obtain (7) to, as an example, third order in ¥(¢}, we write directly, putting hat ant) = [oft ta) + WO ta) + HOHE a) + HONE Ue) 1714) YC) + YO + WL) + BHD) (3.17-15) A short inspection of (3.17-13) should enable us to write u!(t, £.) by inspec- tion. We note that it is composed of a sequence of unperturbed propagators WOE, Y= WOH 1 a) WOM ne fa) punctuated by the action of the perturbation V({) at times t, ~~ t,,. That the “flow’' of time is from right to left is indicated at the end of (3.17-13). To calculate the expectation value of physical observables as in (1.1-16}, we need to obtain an expression for the wavefunction to any desired order of perturbation. The optical field at the atom site is taken as E(t) = Be" + dpe! + cc. (3.17-16) and the interaction Hamiltonian as E E Vit) = we & et z ett ce). (3.17-17) where p is the negative of the dipole moment operator. The perturbation is assumed tumed on at tg = ~~, at which time the atom is in its ground state n. The eigenfunction at a subsequent time ¢ is Y= POY BEE EBM) He where yen (ty = EE, top|) Using (3-17-13), we write He pg = exp {i to | fet) = exp[—iea(t ~ fo)] |) igt fF wg = Ef ep[- Be e- ay] ren in -exp[- zn wo) sb which when we use (3.17-4) becomes 90) = — ED fm) Ol expl—ion(t — 111M) (3.17-18) + exp{—ite(t ~ fo)] fn) dt, Since V(t) is, according to (3.17-17), a sum of four terms, the integrand in (3.17-18) is made up of four terms. One such term, for example, resulting 62 MATRIX FORMULATION OF QUANTUM MECHANICS from the part of V(i) involving Ef exp(—iw,) will yield het 2) = 3D [an ta ET XP Hlren = 2) (3.17-19) + exp(—ieat) exP(iwyto) dt, |n7) where jz, is the projection of p along Ey and ayn = @m — @1 = (Em — En)/h. Carrying out the integration leads to giom-aiit — gi wan—oiKa 1 SO = 35 Dy (oi dmEt exp(— Herm) lem) (nn ~ @1) where the factor exp(iw,o) was dropped since it cancels out (through multi- plication by its complex conjugate) in the catculation of any physical observ- ables. The factor €xp[i(@m — @:)fo] becomes zero for fo > —® if we allow mq > We — EY (y > 0) This not only eliminates any initial “memory” effects that, having taken place at f = —» should not affect the present, but also introduces in a correct fashion the finite natural linewidth y of the wansition. The result is expli(= (rm — 1 WL) = og dn ET In a similar manner, we can calculate the contributions to ys‘) (t} due to the terms E, exp(iat), Ep exp(iwt), and E3 exp(—iant) in V(t) obtaining, respec- tively, ya = 980 + 00@ + WM + ve ai { Ef expli(an — @n)t)} bm) Re (er anne a $ (atadan ELeABile = ooh |r) Be Cn Ft — FY (3.17-20) + (uslee Ex expli(—@, — @)f] |) Opn ~ 02 — ty E; exp[i(o2 + (Hada Using (3.17-13), we write the second-order wave functions p(t) as ‘2 pt pa prea = wee tory =(~ 5) fh fl ae ae -exp{ k=] vexo| iM] im = (- BSL Viti) exp [-i Bolt — Vite) Y to dt 3.17 THE TIME EVOLUTION OPERATOR—FEYNMAN DIAGRAMS 63 + exp[ feast — 41) 1 Is) Os]Vitu) {3.17-21) + exp[—ieom(ty — £2)] bem) dnl Vie) + exp[—ftay (ta — fo)] |) dt dtr Since V(t) contains four frequency terms and appears twice in (3.17-21}, the full integration in (3.17-21) will yield 16 terms. A typical term, as an example involving the use of a BY exp(—iw;tp) at f and EF exp(—ieozt,) at ty, is ¥ \s) (3.17-23) daca = SY (fH eolentond np on = a= at {pn — Wy — Fy} — wo — 2 — HY) A convenient way to represent (3.17-23) is through the use of a Feyn- man diagram, as shown in Figure 3.1. Time increases from the bottom to the top. Each solid line segment represents an eigenstate. The atom starts at é in state n and “scatters” at t, into state m by absorbing a photon at ,. This scattering is accounted for by the factor (#i)mn Ebe "Henn — @1 — iy) in (3.17-22). The next scattering is at t, and involves an absorption of a photon al w,. A negative frequency denotes absorption and is represented by an arrow terminating at a vertex, whereas an arrow starting at a vertex denotes the emission of a photon. A case where the transition from state 1 to s involves the absorption of a photon at «, and the emission of a photon at w2, as an example, is shown in Figure 3.12. The corresponding contribution to p(t) can be written by inspection ce = > (Gy 2 EYE, (tha amr todo XP (Gn ~ oh + © (am — 0) — Mw — 0 + =F oy G72) We note that each scattering, that is, each vertex in the diagram, contributes one factor to the denominator, and the factor is equal to the energy in units of & of the atom and field after the scattering minus the initial (f = f) energy. The second factor in the denominator in (3.17-24), as an example, is obtained () (eh tod FIGURE 3.1 The Feynman diagrams used to obtain (a) W2 _.(#). (2) FS) se,(t} (FL end) 64 MATRIX FORMULATION OF QUANTUM MECHANICS from w, + @, — (w) + w,}. The remaining 14 wavefunctions are obtained by taking all the possible permutations of #,, ..,,i/ = 1, 2 where our conven- , ton is such that oi ~u,, for example, corresponds to emitting an «; photon at t, and absorbing an ; photon at 4; (f, > f2). We note that ¥"}, is not equal to ye Using the diagram technique, we can write the wavefunction to any order of perturbation. As an example, consider the process in which an atom makes a transition from state n to state s while absorbing one photon at ), one photon at w2, and emitting a photon at w;. The diagram describing this process 4, -o,.0,- is shown in Figure 3.Ic. We obtain by inspection _ ~BIEIE, (mn — 1) (Win — G1 — W2) (Hn — @, — 2 + Gs) The total number of 2), -.,2 Combinations is 6* = 216 inside the triple summation. (There are now six terms in V(t), two for each frequency.) The tremendous advantage of the diagrammatic representation is that it enables us to single out and write down very simply the terms that dominate in any given physical situation. This will become clearer in the examples that follow. Two-Photon Absorption Here, we will apply our formalism to derive the tansition rate for absorption of two photons—one at @, and a second at w,—by an atom. By allowing 2, = @, we will obtain the familiar expression for a two-photon absorption coefficient. We start by writing the second-order wavefunction ¢@ (t) corresponding to an atom that at / = 0 is at state # and that interacts with the radiation field (3.17-17), consisting of fields at m; and #2, We will assume that the largest contribution to #” comes from #@),,_., and not from W-,., -«, because of the existence of some level in such that (4, — #1) is very small. We thus write Halt = DD Ge - _ Lesbian oe) te [EXP Hoe — @) — w2)f — 1 (i — ey ~ 0, — 2} 6) G.a7-25) The —1 term is due to the fact that here the problem starts at f = 0 and not, as above, at —%, at which time the atom fs in state |n). The integration in (3.17-21) is thus from 0 to £. Since wy, = @;, the term J dominates and we rewrite (3.17-25) as By Can { Ha) me expLi(@s, — @ — @2) — 1 (mx — ©} — @ — @2} Pennell) = 8) (3.17-26) 3.17 THE TIME EVOLUTION OPERATOR—FEYNMAN DIAGRAMS 65 The probability of finding the atom in some state é at lime t is ARIE (O)P so that the transition probability due to two-photon absorption is Pe = Kelp (YP (3.17-27) 1 [EAIP|B 2 (#1 Yn #2 Dw si? iB (Wm = 1 ~ ant| — 16Kt 1 - (3.17-28) (mm ~ on)? [F fy ~ 07 ~ 0) We note that transitions occur to the state k that conserves energy, that is, to that state where £, — E, = fi(w; + 2). If the normalized lineshape function for the transition 1 — & is g(wn), then the average value of P; is obtained by multiplying (3.17-28) by g(@.) and integrating from ~~ to ». Using t sin lime Taye Dat BUN) we obtain for the transition rate a Wr = > eo Gm (3.17-29) — TUESPLEaP (141 )nn| 12nd (oir = 1 + 022) BA (gn = 01)? In the special case of one-frequency two-photon absorption, that is, #; = @2, we obtain the absorption coefficient by equating the change in the inten- sity f,, of Z, to the number of transitions per unit time per unit volume. dl, Fe = he Wns (Me — Nad (3.17-30) lk> Nex wy Aeon '=0 (a) 16) FIGURE 3.2 (a) The dominant Feynman diagram for the two- photon absorption process for the case x * an. (#) The corre- sponding energy level diagram. 66 MATRIX FORMULATION OF QUANTUM MECHANICS where /, is the intensity at w, and N; and N, ate the population densities of level k and n, respectively. Using f, = (ec/n) [Ei /?/2. we obtain areal Ny — Ne){ tyro Ba Yenttig (in = 20)y FB (mn — w)?e?c* @yy0 photon = Since @mn # 1, there is no actual transition 10 and dwelling in the intermediate level |mt). To obtain a “real” transition, 2 > ”, Gm — @, must, in practice, be smaller than a few times the transition linewidth. We refer to the transition 2 m indicated by Figure 3.2 as “virtual”, Also noteworthy is the fact that a transition 7! — k takes place even when jx = 0 ic. when the direct one-photon transition may be forbidden. References 1. E. U. Condon and G. H. Shortley, The Theory of Atomic Spectra (New York: Cam- bridge University Press, 1959), pp. 73-76. 2. C. Kittel, Elementary Statistical Mechanics (New York: Wiley, 1958). RC. Tolman, Principles of Statistical Mechanics (London, Oxford University Press, 1938). 4. U. Fano, Rev. Mod. Phys. 29, 74 (1957). Ao Supplementary References 1. Leighton, R. B., Principles of Modern Physics (New York: McGraw-Hill, 1959). 2. Messiah, A., Quantum Mechanics (New York: Interscience, 1961). 3. Schiff, L. L, Quantum Mechanics (New York: McGraw-Hill, 1959). Problems 3.1 A function f may be expanded in terms of two arbitrary complete orthonormal sets v, and u, in ihe form fe 3 fate =D fit The set v, can be expanded as = D Siite (0) t Show that the unitarity of the matrix S can be derived by requiring that [/*fdv be independent of the set {vq or %,) in which it is expanded. 3.2. Show that the trace of a square matrix A wA => Aw i is invariant under matrix transformations, that is, that tA = w(SAS*!) where S is unitary. 3.3 Prove that (AB)t = BAt. 34 35 3.6 3.7 38 3.9 3.10 3b 3.12 3.13 314 3.17 THE TIME EVOLUTION GPERATOR—FEYNMAN DIAGRAMS 67 Prove that a Hermitian matrix remains Hermitian under a unitary transforma- tion, Show that as a consequence, a Hermitian matrix must have real eigen- values. Show that if Ay = J vf Avidv = Ai where y, is any orthonormal function set, the operator A is Hermitian. Show that the matrix representation of a product of operators is equal to the product of the individual matrices. Use the result of Problem 3.6 and the known matrix elements Xpx for the har- monic oscillator to evaluate 7) mn = { UX? AY Check the result against the solution as given in Problem 2.3. Show that the necessary and sufficient condition that two matrices commute is that the same transformation diagonalize each one of therm. Prove that cn+) = (c4)* where ca) and ci, are defined by Eq. (3.8-4). Prove that y} = 1/V2 (ab! — at Ba). Eg. (3-10-6}, is an eigenfunction of L* where L = L, + Lz. What is the eigenvalue? Generate the set of eigenfunctions i, resulting from the addition of two angular momenta jy = 4 and j, = 1. . The potential energy of a harmonic oscillator is perturbed by adding a term bx. (a) What is the correction, to second order, in the energies? (b) Obtain an exact expression for the energies. Write by inspection, if you can, the function #2 anonltd costesponding (o absorption of a photon at w, and a simultaneous emission of photons at «, and a. What is the corresponding Feynman diagram? Assume an electric field 1 Ei) => (Bueiee + Byeiws + Eyeiont + ce) {a) Calculate the gain coefficient experienced by a field 1 By (t) = Exelon + 6. in the presence of an intense field 1 AM =5 Exgiast + CC. due to a process in which atoms initially in the ground state in |) absorb a photon at @, and emit simultaneously a photon at « ({;). The atom ends up ina state |&) so that By — E, = &(0; ~ 2). (Note the similarity to the two- photon absorption example weated in this chapter.) (b) Assuming wansition linewidths of 10? Hz and matrix elements tm ~ 107°? MKS, calculate the gain coefficient experienced by the field Z as a function of the intensity 1). CHAPTER 4 Lattice Vibrations and Their Quantization 4.0 INTRODUCTION The topic of lattice vibrations is treated in this chapter for a number of rea- sons: (1) The spectrum of lattice vibrations and the formalism for treating it are necessary to explain certain relaxation mechanisms such as the atom- lattice relaxation. {2) The formalism for quantizing the lattice vibration field serves as a model for the quantization of other boson fields such as that of electromagnetic radiation. (3) The formalism will be used in Chapter 18 for treating Brillouin scattering of light by sound and stimulated Brillouin scat- tering. 4.1 MOTION OF HOMOGENEOUS LINE Consider the one-dimensional problern of a uniform line of mass density p (kg-m~). Let the displacement of a point x from its equilibrium value be u(x, t). The force at x is taken as a constant times the strain du/ax ou Fee (41-1) which is equivalent to using Hooke’s law for the restoring force of a spring. If we consider an element Ax of the line, the net force acting on it is F(x + Ax) — F(x) = c(a?u/dx7)Ax. The equation of motion becomes eu eu ey aE (4.£-2) with a solution, if we assume et) dependence, 4 = xgeltoezen) %,= na (4.13) According to (4.1-3), the line can support waves with a phase velocity v, = Vip. The dispersion relation is @ = kv;. Since vy, is a constant, the line is dispersionless. The general solution of (4.1-2) is a linear superposition of individual 68 4.2 WAVE MOTION OF A LINE OF SIMILAR ATOMS = 69 solutions that can be written as ue, 1) = Cc ule)elat-"n dy (41-4) 4.2 WAVE MOTION OF A LINE OF SIMILAR ATOMS As the next improvement in the realism of the model, consider a line of particles of individual masses M separated from each other by a and con- nected by massless springs with a spring constant § as shown in Figure 4.1. If we let the deviation of the nth atom from its equilibrium position be q,, the equation of motion for the mth atom becomes Min = B(Qns1 — Gn) — Blan ~ Gna) (4.2-1) = Bas + Ant — 24x) In direct analogy with the continuous line, we assume the basic harmonic solution to consist of a wave with an effective propagation constant k corre- sponding to a phase shift (ka) between the sinusoidal motion of two neigh- boring particles fn = Eyer (4.2.2) With this substitution, (4.2-1) becomes math = ple 4 ee — 2) = plete? — e-Barep Since the phase velocity of the wave (4.2-2) is equal to w/k, we can get both positively and negatively traveling waves by keeping w positive and letting k assume both positive and negative values. With this convention, the disper- sion relation becomes = we sin 4] = (2) “21 — cos(ka)] (4.2-3) A plot of Eq, (4.2-3) is shown in Figure 4.2. Consider next the solution 4,» for a single wave when k is increased by i(2a/a) where / is an integer. From Eq. (4.2-3), we obtain w(k + [2n/a) = (k). Using this result in (4.2-2) gives Gcrinnsayn = Excited gent = gyeiiartnad = gg so that all the possible waves can be generated by allowing k to roam over a 2n/a interval. The conventionally chosen interval is — (a/a) < k < w/a and is shown in Figure 4.2. Iv is called the first Brillouin zone. In a cubic crystal with FIGURE 4.1 A line of similar atoms. 70 LATTICE VIBRATIONS AND THEIR QUANTIZATION o i ol 1 a {First + Brillouin | zone 1 1 ®, -= 9 RK an 7 a aoa FIGURE 4.2 The dispersion diayram for the discrete line of atoms. a = b =c, for example, the first Brillouin zone will correspond to the region —(wla) < ky, ky, k, S w/a in k space. The phase velocity, for ka < 1, is given by = ve a (4.2.4) which is the same as that of the homogeneous line, Section 4.1, if we use the appropriate tansformation p— M/a and c—> Ba. The condition ka < 1 means that the acoustic wavelength A( = 22r/k) is much larger than the atomic sepa- ration a and, consequently, as far as the wave motion is concerned, the line may be considered homogeneous. The expression for the group velocity is ya tit YB aco!) (423) so that v, = 0 at the edges of the first Brillouin zone (ka = +7). Itis a property of all lossless periodic structures that the frequency cutoff and zero group velocity condition obtain when the phase shift per unit of periodicity {ka in our case) is equal to a radians. This point is discussed extensively by Bril- louin.' The most profound difference between the homogeneous and discrete lines is that in the latter there exists an upper limit on the wave frequency given, according to (4.2-3), by la This cutoff frequency is equal, according to (4.2-4), to 2v,/a that, if we use the typical values v, = 3 X 10° cm/s (in solids), a = 3 x 10-* cm, gives Ja = SE 3X 10"? cps This corresponds to an electromagnetic {free space) wavelength of 0.1 mm and shows that in typical acoustic experiments, which employ frequencies ‘L. Brillouin, Wave Propagation in Period Structeres (New York: Dover, 1953}. 4.3 A LINE WITH TWO DIFFERENT ATOMS = 72 through the microwave region, one is still well within the dispersionless, ka < 1, region. Mode Enumeration The individual solution gin = Ei,né""**" of (4.2-2) is called an acoustic mode of propagation. The number of independent modes of the discrete line must be equal to the number of particles (atoms) in the line, This implies that only certain values of k are allowed. These can be found by consideration of the boundary conditions. One such condition is that, if the line consists of NV atoms, the first and last atom are “clamped” and cannot move. An alternative condition is that the total motion of any atom is reproduced dfter every N atoms. These two conditions give the same results, so that, for the sake of definiteness, we will use the periodic boundary condition. The total motion of the nth atom can be described, according to (4.2-2), as nll} = z Gen = > gxctert on (4.2-6) The periodic boundary condition ¢,++ = x is satisfied if X is restricted to the values __ amin ke Na (42-7) where m is an integer. This can be verified by substitution in (4.2-6). The mode spacing in k space is thus uniform and is given by Ak = 217/Na. Since kis restricted to the first Brillouin zone —(a/a) = & < /a, the total number of modes is 24/(a Ak) = N as required. 4.3 A LINE WITH TWO DIFFERENT ATOMS Most crystals are made up of a lattice containing a number of different atoms. An ionic crystal like NaCl, for example, can be viewed as a cubic array in which neighboring lattice points are occupied alternately by Na* and Cl- ions. As an approximation to such crystals, we consider a model made of a line in which each atom of mass M has as its neighbors atoms of mass m. Let the M atoms occupy odd lattice points and those of mass m even ones. Using the definitions of Section 4.2, we can write the equations of motion for two neighboring atoms as Mbrn = Bldanei + Gani ~ 242n) (4.3-1) Mans = Bldant2 + dan — 24mai) The basic solutions are taken as waves of the form Gane = Exeter 7m) (4.3-2) Gamera = gel ans Mal 72 LATTICE VIBRATIONS AND THEIR QUANTIZATION which substituted into (4.3-1) leads to —wime, = Bnele™ + eo) ~ 2BEx (43-3) ~e'Mny = Bele + eo) — ne The determinantat equation guaranteeing nontrivial solutions for ¢ and y is (28 — me}(2B — Mw?) ~ 48? cos*(ka) = 0 (4.3-4) a a An inspection of (4.3-5) reveals that for a given k there are two frequencies. For ka < 1 we have, to first order in (xa), o = al + a) sae ml - mai a from which we obtain so that 0 Vee ka) (43-6) 2p Mim wz When ka = 77/2, the two frequencies become The upper branch of the dispersion curve is called the optical branch. Its highest frequency is equal to (28(1/m + 1/M)]! and occurs at ka = 0, whereas its lowest value is (28/m)'2 when m < M, or (28/M}}2 when M < im and occurs at ka = 17/2. The low-frequency branch corresponds to the propa- gation in a line of similar atoms studied in Section 4.2. It is referred to as the acoustic branch, At low frequencies (ka < 1), it is dispersionless with a sound velocity, according to (4.3-6), of [28/(M + m)]!?a. The maximum frequency of the acoustic branch occurs at ka = 7/2 and is the lower of (28/m)}' and (2p/M)"2, The acoustic and optical branches are plotted in Figure 4.3 for the case of mM. The nature of these modes can be appreciated by considering a number of cases. According to (4.3-2), the relative motion of two neighboring atoms is given by (43-7) Gants _ gika ME Gn & 4.3 A LINE WITH TWO DIFFERENT ATOMS = 73 : i _-Slope = velocity of ee light in crystal Optical branch 1 28m Acoustic branch | 17H i I ' It ' t k aida FIGURE 4.3 The dispersion diagram fora one-dimensional crystal contain- ing two different atoms with masses wt and Mim Proc. IEEE (Oct. 1965), special issue on ultrasonics; especially articles by M. H. Seavey and by N. F, Foster regarding surface excitation of acoustic phonons. 44 LATTICE SUMS = 75 Refractive index oN bo @ A A 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 4 (in microns) FIGURE 4.5 The refractive index of quartz for the ordinary ray (E to the optic axis) as obtained from the dispersion analysis of the reflectivity. Source: W. G. Spitzer and D. A. Kleinman, Phys, Rev. 121 1324 (1961). Rewriting the summation for g, in terms of a normalized variable Q;, we have 1 Galt) = pia D Qulthe* 44-1 = The series e#* is a complete orthonormal series for expanding any arbitrary function of the N lattice points. We may expect a closure property similar to {1.2-11)} to exist. The corresponding relation is 3S ete = NB, (44-2) rs Proof: For s = x, the result follows immediately. For s # #, we define s — n =] Na Selitemarsat Wz 2 we = et gaemin “Na+1 mo In the first series change variables to m' = m + N = Not wot Seem = pitetiw nw = gitminav “NRL weNAtL won adding the two series and replacing m' by m give tnt gitmmin = © 1 e This completes the proof. The inverse Fourier transformation is x 1 %& = HD ae (44-3) a 76 LATTICE VIBRATIONS AND THEIR QUANTIZATION It can be verified by substituting (4.4-3) for Q, in (4.4-1) and using the closure property (4.4-2). Another lattice sum that will prove useful in the next section is x D eit = Ne (44-4) = For k = k’, the result follows directly, For & # k', we have k — k’ = 2a/Na (nm — m') = (2a/Na)l where | = m' — m is an integer. The summation is identical to that of the last step in the proof of Eq. (4.4-2), thus yielding the desired result. 4.5 QUANTIZATION OF THE ACOUSTIC BRANCH OF LATTICE VIBRATIONS**¢ The purpose of this section is to show that the ensemble of N lattice vibrations in a line of (similar) N atoms is formally equivalent to N independent har- monic oscillators. The Hamiltonian of the (one-dimensional) crystal is the sum of the ki- netic and potential energies of the individual atoms lz + Bldrn ~ «| (4.5-1) m H le where f is the effective “spring constant.” Consider first the kinetic energy term, Expanding p, in a lattice Fourier expansion as in (4.4-1) 1 1 P= 3a = Pe ee (pt =yad prev) (4.5-2) gives z 2mN Spee tertire Es 1 =o LSE ectiee tie Tmin Dy PP De 1 = Tan Dy PPEN Bee 1 “Sa 2 PP_x “C. Kittel, Quantum Theory of Solids (New York: Wiley, 1963) 5G. Weinreich, Solids (New York: Wiley, 1965). *M. Born and K. Huang, Dynamical Theory of Crystal Lattices (Oxford, England: Clarendon Press, 1954). 4.5 QUANTIZATION OF THE ACOUSTIC BRANCH OF LATTICE VIBRATIONS 77 where, it should be recalled, P, is an operator, being the transform of the operator p,. Since p, and q, are Hermitian operators, it follows from (4.4-1) that Qn=Q@ and Py=Pt (4.5-3) This makes the sum of the k and —& terms in (4.4-1) an Hermitian operator. The potential encrgy is B 7 Ge 4 = gy (afer + 92 — 24re14) that, when using Eq. (4.4-1), becomes ELS Quay fettirrnagners ia — ties nagit'ra — gimp tra 4 lkragh na] 2N ek = A> [orate = ete = OE z ert va} If we replace the last summation over r by N & -y in accordance with (4.4-4), the last expression becomes. 8 2 QQ-x[1 — cos(ka)} (45-4) ‘The total Hamiltonian is thus 1 a= {5 PP + BOQaLl = costkayl} (4.5-5) Equations (4.4-1) and (4.5-2) can be used to evaluate the commutator of P, and Q, (45-6) ell Be = (ih) Say where use has been made of (4.4-4) and the commutator relation [p,, 4,] = —ih 6,, that applies to the momentum p and displacement q of a particle. Except for the mixing of the k and ~x terms, the Hamiltonian (4.5-5) is in the form of a sum of single harmonic oscillator Hamiltonians. To eliminate the mixing, we introduce the creation and annihilation operators that are 78 LATTICE VIBRATIONS AND THEIR QUANTIZATION defined in correspondence with Eq. (2.2-25) as eVig, tp . Vi Vig ta * (45-7) aan i = +e a= & Vig hat k where a is defined, as in Chapter 2, by and g, = V2[E — cos(ka)]. Equations (4.5-7} are consistent with (4.5-3). Solving for Px and Q, gives haVGn _ , fener Phy = = EM al — aay =i fe tat — a Va (4.5-8) hee Ph Py EE (ate ~ ae) l fe Oh = a= eee tal + a0) = Vimoy (al 8) tg = —t a 4.5-9 Oty = Qe = Fe ala + a) (4.5-9) a n= (2) 5 The commutator of a, and af is + a Vi i i ap] = + Pa), (— QQ - ——— Pe tab = [XB a+ he ea) ME on ei | =~ Sp (Qe Pel + og [Pe Qe] = Pee (49-10) where use has been made of Eq. (4.5-6). Substituting for P; and Q; in a single term of H in (4.5-5) results in 1 Hy = 5 PuP-« + BQQ-4[1 — cos(ka}] 2m Hae; =~ SEM ad aaytats — a) B 207g, (ak + a-ghaty + aL — cos(ka)] Multiplying out the last equation and using the relations a? = Vinh, gx = V2 — cos(ka)], lay, af] = dxe, and Eq, (4.2-3) ov (2)" var contean = (8)" m 4.5 QUANTIZATION OF THE ACOUSTIC BRANCH OF LATTICE VIBRATIONS 79 give, after some algebra, igh ‘ Me 2 onlay + aha, t 1) (4.5-11) The total Hamiltonian is a= Sma > [Malt ye ata + Hf 7 = (4.5-12) = = fax(ata, + 4 since the summation includes, symmetrically, both positive and negative val- ues of k. Comparing (4.5-12) to the Hamiltonian of the harmonic oscillator, Eq. (2.2-30), and using (4.5-10), we establish that the one-dimensional crystal is equivalent, quantum mechanically, to N independent harmonic oscillators, one harmonic oscillator for each mode x, Each harmonic oscillator is charac- terized by eigenfunctions w,, that are functions of Q; and that obey, in direct analogy with (2.2-27), tig, = Vite FT ings Bidkag, = Vg by (4.5-13) Magen = Thy where n,, the number of phonons in the k mode, can range from 0 to «. The wavefunction of the complete crystal is the product of the single-mode wave- functions x © = tgteg ny = TL tay (4.5-14) ket and, according to (4.5-12) and (4.5-13), corresponds to a total crystal energy (eigenvalue) B= hein + 1) ¥ We can obtain the explicit time dependence of the annihilation and creation operators in the Heisenberg representation from dai _i + 1 GRA] = fond ahay + 3), @ where we used (4.5-12}. The only nonvanishing term is k = j so that da; _ He 7 ieilalay. a] = —ieyay 80 LATTICE VIBRATIONS AND THEIR QUANTIZATION and a(t) = afOye™ (4,5-15) , aft) = af(oyen* The motion of the nth atom can be expressed in terms of the annihilation and creation of the individual modes as 1 A(t) = tyetna anl8h = Fyre B Quine ey sie itn aye & (ue + Quer") (4.5-16) 1 =— Ld (ayer + aterivny Tye & (etme + Qe 1 tale + aye*™) where use has been made of (4.5-3), (4.5-8), and (4.5-9). The summation includes both positive and negative values of k. If we substitute for af and a, from (4.5-15}, we get Ls - tog n(t) = a —— [ah (oy eroe 2) + ay(O)e Hoe 45-17 ate) HD yg HO} 10) J (45-17) In this form, the traveling-wave nature of the individual modes and their explicit time dependence are brought out. 4.6 AVERAGE THERMAL EXCITATION OF LATTICE MODES According to Section 4.5, a lattice vibrational mode at a (radian) frequency w has an energy E, = (n+ Siw If the lattice is in thermal equilibrium at a temperature T, the probability of finding a given made excited to the state 7 is given by the Boltzmann factor ett Pn) = Se ae (4.6-1) where k is the Boltzmann constant. The average excitation energy of the mode is given, consequently, by _ Sheott + hae MIT figs Expl) = Sse rae Me E= me where B = (kT). 4.6 AVERAGE THERMAL EXCITATION OF LATTICE MODES 81 The denominator of (4.6-2) forms a geometric progression whose sum is = 1 stop = —1__ 2 ® T= oh Taking the derivative of the last expression with respect to B gives fe tok D swe 08 = am 0 Substitution in (4.6-2) leads to (4.6-3) The average excitation of a mode is often characterized by the average num- ber of quanta # (4.6-4) When kT > hw, the average energy per mode is KT, whereas for kT < fiw, it is hw/2. The latter is, of course, just the zero-point energy of the harmonic oscillator. In Section 4.2 it was shown how the periodic boundary condition qu+w = 4n Fesulis in a discrete spectrum for the allowed & values with the difference between adjacent k values being Ak = 2m/L. Extending the periodicity re- quirement to three dimensions with the basic periodic cell taken as having edges A, B, and C gives 2a 2 _ 2a Fo bk a Ake & (4.6-5) Ak, = = for the minimum spacings of the three components of the propagation vector k. Accordingly, we can associate with each k vector a volume (2n)3 _ (20? 3s = - (Ak) "ABC ¥ (4.6-6) in k space, where V is the volume of the periodic cell in (real) space. The number N(k} of modes between 0 and < is given by the volume of a sphere of radius k divided by the volume per mode, that is, ankV NO ~ SeamP For the case of no acoustic dispersion, we take k = 2av/v, where y, is the sound velocity. This gives anv Ny = (46-7) for the number of modes with frequencies smaller than y. We have multiplied the result by 3 to account for the three independent acoustic polarizations. 82 LATTICE VIBRATIONS AND THEIR QUANTIZATION The number of modes per unit volume per unit frequency p(v) is 1 dN@v) _ 12av? PO Vay (4.6-8) Whereas the average vibrational energy at temperature T per unit volume per unit frequency is = lanv? fh hi piv) = pol = 25" (7 + sear) (4.6-9) where we used (4.6-3). In situations where one is concerned with excitation energy, that is, that energy that can be extracted from the oscillators, one neglects the term hy/2 in (4.6-9) since this zero-field energy cannot be removed from the oscillator. Problems 4.1 Describe the relative motion of adjacent m and M ions at ka = 7/2 for both the optical and acoustic modes. 4.2 Show that when m = M the dispersion (w versus k} diagram for the diatomic line is identical to that of the monatomic line. Hint: Investigate doo,/dk, and dan/dk, at k = 1/2a and extend the diagram to k = n/a. A statement was made in Section 4.2 that increasing X by 2spiva, where lis an integer, does not yield new modes (see Figure 4.2). How can one reconcile this statement with the fact that the phase velocity v, is equal to w/k and is, conse- quently, different for modes with different k values (but the same @)? 4. Hint: Consider the physical significance of the “different velocities” in view of the discrete nature of the structure, Can you conceive of an experiment that distin- guishes among such different velocities. 4.4 Show, using arguments similar to those used in Section 4.2, that the total number of modes in a “three-dimensiona!” crystal containing N atoms, two per unit cell, is 3N, Hint: Include both the acoustic and optical branches. 4, ‘Assuming that in a line of N similar atoms the initial conditions at ¢ = 0 are a0} = 920) = A, 4i(0) = 92(0} = 0, (a) derive the expression for the deviation q,(¢) of the nth atom. (b) Assuming w = kv,. show that the disturbance propagates asa pulse with a velocity ¥,. c HAPTER 5 Electromagnetic Fields and Their Quantization 5.0 INTRODUCTION This chapter provides the main background for the electromagnetic theory used in this book. The following topics are discussed. (1} Propagation of plane waves in homogeneous anisotropic media. This material is used in the treat- ment of the electrooptic, magnetooptic, and photoelastic manipulation of light. (2) Energy storage and power dissipation.-(3) Normal mode expansion in a generalized resonator. This material will be used to derive the laser oscillation condition in Chapter 9 as well as in the treatrnent of laser noise in Chapter 13. (4) The quantization of the electromagnetic field. This formalism is the background for describing spontaneous emission processes. 5.1 POWER TRANSPORT, STORAGE, AND DISSIPATION IN ELECTROMAGNETIC FIELDS in this section we derive the formal expressions for the power transport, power dissipation, and energy storage that accompany the propagation of electromagnetic radiation in material media. The starting point is Maxwell’s equations (in MKS units) (5.1-1) al VxES-— (5.1-2) and the constitutive equations relating the polarization of the medium to the displacement vectors D=eE+P (5.13) B = jao(H + M) {5.1-4) where i is the current density (amperes per square meter); E(r, f) and H(r, ) are the electric and magnetic field vectors, respectively; D(r, t) and Bir, t) are the electric and magnetic displacement vectors; P(r, f) and M(r, f) are the electric and magnetic polarizations (dipole moment per unit volume) of the medium: and é and jp are the electric and magnetic permeabilities of vac- 383 $4 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION uum, respectively. For a detailed discussion of Maxwell’s equations, the reader is referred to any standard text on electromagnetic theory such as, for example, Reference | at the end of the chapter. + Using (5.1-3) and (5.1-4) in (5.1-1) and (5.1-2) leads to é Vx Wait 5 (ook +P) (5.1-5} a Vx E =~ F Holl + M) (5.1-6) Taking the scalar (dot) product of (5.1-5) and E gives &q 3 E-VXH=E-it 55°) +E (5.1-7) where we used the relation lé oE tu (E-E) = E> ue Next, we take the scalar product of (5.1-6) and H nevx =~“ 2 gem - pot (5.1-8) 2 ae at Subtracting (5-1-8) from (5.1-7) and using the vector identity V(AXB)=B:VXA-A‘UXB (5.1-9} result in a ap aM “Vex H) B+ So (2 E-E+= yn) +E Se Sy + HoH Gp (51-10) We integrate the jast equation over an arbitrary volume V and use the Gauss theorem (Reference 1). [valde = [acne where A is any vector function, n is the unit vector normal to the surface enclosing V, and dv and da are the differential volume and surface elements, respectively, The result is ~ [+8 x Hav - [ex -nda (1-11) fle. i+2@ E- E) + 3 (Hen) BS + uo FA av According to the conventional interpretation of electromagnetic theory, the left side of (5.1-11), that is, 5.1 POWER TRANSPORT, STORAGE, IN ELECTROMAGNETIC FIELDS 85. ~ [0h x Hynde gives the total power flowing into the volume bounded by S. The first term ou the right side is the power expended by the field on the moving charges, the sum of the second and third terms corresponds to the rate of increase of the vacuum ¢lectromagnetic stored energy 8,a: where ace | (SE-B + Saw) éy (5.1-12) Of special interest in this book is the next-to-last term ap E- a that represents the power per unit volume expended by the field on the electric dipoles. This power goes into an increase in the potential energy stored by the dipoles as well as to supply the dissipation that may accompany the change in P. We will return to this subject again in Chapter 8, where we treat the interaction of radiation and atomic systems. Dipolar Dissipation in Harmonic Fields According to the discussion in the preceding paragraph, the average power per unit volume expended by the field on the medium electric polarization is Power _ — P Volume at (5.1-13} where the horizontal bar denotes time averaging. Let us assume for the sake of simplicity that E(t) and P(#) are parallel to each other and take their sinusoidally varying magnitudes as E(t) = Re(Ee) (5.1-14) P(t) = Re(Pe™) (5.2-15) where £ and P are the complex amplitudes. The electric susceptibility y- is defined by P= e0xeF (5.1-16) and is thus a complex number. Substituting Eqs. (5.1-14) and (5.1-15) in (5.1-13) and using (5.1-16) give Power = RaEe™)RetiwPe™) Volume = Rete )Re(iwPe) = 4 Re(imenx.FE*) (5.1-17} = F ealBlPRetixed 86 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION Since x. is complex, we can write it in terms of its real and imaginary parts as Xe = Xe — Exe (5.1-18) that, when used in (5.1-17), gives Power Volume = ont 1eP (5.1-19) which is the desired result. We leave it as an exercise (Problem 5.8) to show that in anisotropic media in which the complex field components are related by P= 60>) XuEi (5.1-20) the application of (5.1-13) yields Power o . Voume 7 7" >) Re(ixjE?E) (5.1-21} a The Vector Potential The electric field E(r, ‘} and magnetic field H(r, #) can be derived from a single vector A(r, f), the vector potential. The relation V - B = 0 is satisfied by choosing B=VxA (5.1-22) Since V + V x (any vector) = 0, the relation V x E = letting B/at is satisfied by B= -%4_ wy (5.1-23) since ¥V x VV = 0 where V(r} is any scalar function. Our basic defining relations (5.1-22,23) must also satisfy V+ E = p/e where pir) is the charge density. This leads to a ~ GVA) VW = ple =0 We need to specify V - A, since only V x A was specified so far. The choice V-A=0 (5.1-24) is called the Coulomb gauge. This ties down the scalar potential V(r} intro- duced by (5.1-23) that now must satisfy V-eWe vend (5.1-25) We still need to make sure that E and H derived from A satisfy the second Maxwell equation V x H = e dE/at. Using (5.1-22,23), we find that this is the case provided PA VA — pe = 0 5.1-26' Me ( y 5.2 PROPAGATION OF ELECTROMAGNETIC WAVES IN ANISOTROPIC CRYSTALS 87 5.2 PROPAGATION OF ELECTROMAGNETIC WAVES IN ANISOTROPIC CRYSTALS An understanding of wave propagation in anisotropic crystals is a prerequisite to the treatment of a number of important topics. Some of these that are treated in this book are (I) electrooptic, magnetooptic, and acoustooptic modulation and (2) phase matching in nonlinear optical interactions. Inan anisotropic crystal, the polarization induced by an electric field and the field itself are not necessarily parallel. The electric displacement vector D and the electric field E are consequently related by means of the dielectric tensor 2, defined by’ Dy = eyE; (5.2-1) where the subscripts refer to a Cartesian coordinate (Kk, / = x. y, z}), where x, y, 2 are fixed with respect to the crystal axes and the convention of sammation over repeated indices is observed. Taking the stored electric energy density, as in an isotropic medium, by @, = HED = 4ByeuE, (5.2-2) we obtain . Ed ye i = FF (BE + FB) {5.2-3) According to the derivation of Poynting theorem in Section 5.1, the net power flow into a unit volume is -V-(Ex H)=E-D+H'B that, if we use (5.2-1) for D, can be written as -V-(E XH) = Reuis + HB (5.2-4) If the Poynting vector is to correspond to the energy flux in anisotropic media, as it does in the isotropic ones, then the first term on the right side of (5.2-4) must be equal to @, and must consequently be the same as @, as given by {5.2-3). If we write EyeyE; as 0, = MeuEE; + ex BE) and compare to (5.2-3), it follows that 8a = x and the dielectric tensor eq has, in general, only six independent elements. The electric energy density w, can be written, if we use (5.2-1) and (5.2-2), as Dey = gEl + ByBl + egE? + 2eycByEy + 2egEpEe + 2eyEEy (5-2-5) 1 Equation (5.2-1) assumes that no dispersion exists so that the instantaneous values of D, and E, are related by a single constant, This neglect of dispersion is valid whenever the medium is optically lossless in the spectral region of interest. 88 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION A principal axis transformation can be used to diagonalize (5.2-5). In the new coordinate syste, «@, becomes Qe, = 6£2 + 2B + © (5.2-6) where the x, y, z symbols now refer to the new axes. The new coordinate axes are called the principal dielectric axes. Since all of our analysis will be carried out in this system, a confusion involving the original coordinates cannot take place and we will retain the x, y, z labeling. In the principal dielectric coordinate system, the tensor e;, is diagonal and is given by Dy sy 0 0 |e Dl =|0 « 0 | 15, (5.2-7) Dz o 0 & |l& Using (5.2-7) in (5.2-6), we obtain DD DF 2w, = + 4 (5.2-8) fe & OB so that the constant energy (w,) surfaces in the space D,. D,, D, are ellipsoids. Assume next a monochromatic plane wave of radian frequency w propa- gating in the crystal with a phase factor exp {iw le = : esl} (5.2-9) This corresponds to a wave vector k = (w#/c)s where s is a unit vector normal to the wavefront (plane of constant phase). The phase velocity is then wris (5.2-10) so that ¢ is the velocity of light in vacuum and n is the index of refraction. For a monochromatic plane wave, we can formally replace the operator V by —{ien/c)s so that the Maxwell equations V x H = (dD/at) and ¥V x E = —{éB/dt) become, respectively, p=-"sxH (5.2-11) H=6xE (5.2-12) pe According to (5.2-11), D is perpendicular to H and both D and H are perpen- dicular to s, The direction of energy flow as given by the Poynting vector EX H, consequently, is not collinear with the direction of phase propagation s. The vectors E, D, H, s, and B x H are shown in Figure 5. 5.2 PROPAGATION OF ELECTROMAGNETIC WAVES IN ANISOTROPIC CRYSTALS 89 (direction of propagation) H FIGURE 5.1 The relative orientation of E, D, H, s and the Poynting vector, E x H, in an anisotropic crystal. The vectors D, E, s, and E x H lie in a single plane. By using (5.2-11), (3.2-12), and the vector identity A x (BX C) = B(A+ C) — C(A - B), we obtain the following expression: w nw ai esxsx B= JE - 55-8] = aat™ aa! s(5-E)] > 5.21 ou Byanwene 3) and since s+ D = 0 and 7°/c je = we, nw DP=—-E+D=reE-D en By using Dy = e,F; and ex = e,/e0 in Eq. (5.2-13) and solving for Fy, we obtain “E ee ee (52-14) we and after multiplying by s, and summing over k = x, y, z S'E=5'E > kang which can also be written as i=5 (5.2-15) Equation (5.2-15), named after Fresnel, is quadratic in 2? (see Problem 5.9). The two solutions +, and +7, (the + signs correspond to a mere reversal in the sign of the phase velocity) are the indices of the two indepen- dent plane wave propagations that the crystal can support. To complete the solution of the problem, we use the values of n?, one at a time, in Eqs. (5.2-14), The three equations (resulting from setting k equal to x, y, z) can then be solved for the relative magnitudes of E,, Ey, and £,, which are then used with the aid of Eqs. (5.2-11) and (5.2-12} to solve for the allowed directions of H and D. 90 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION From (5.2-14), it follows that &) _ Sel (2)'? = 64] By?! slvr yl? — e's where the superscripts (1, 2) correspond to the two independent solutions. For real values of (n?)}?, as in nonabsorbing or nonamplifying media, the ratio Ej?/E}? is real, so that both solutions are linearly polarized. To summarize, along an arbitrary direction of propagation s, there can exist two independent plane wave, linearly polarized, propagation modes. These propagate with phase velocities +(c/m,) and +(¢/n) where nj and 13 are the two solutions of Fresnel equation (5.2-15). In practice, the indices of refraction (#7)!? and the direction of D, H, and E are found, most often, not by the procedure outlined above but by the formally equivalent method of the index ellipsoid. This method is discussed in the following section. 5.3 THE INDEX ELLIPSOID The constant energy density («,} surfaces in D space given by (5.2-8) can be written as De Dy De S++ S = we rer where s;, gy, and e; are the (relative) principal dielectric constants. If we replace D/V2c,£q by r and define the principal indices of refraction n,, ny, and 1, by i = e; (k = x, y, 2), the last equation can be written as 2 x42 Eel (5.3-1} mon ont This is the equation of a general ellipsoid with major axes parallel to the x, y, and z directions whose respective lengths are 2n,, 2”,, 2n,. The ellipsoid is known as the index ellipsoid or, sometimes, as the optical indicatrix. The index ellipsoid is used mainly to find the two indices of refraction and the two corresponding directions of D associated with the two independent plane waves that can propagate along an arbitrary direction s in a crystal. This is done by means of the following prescription: Find the intersection ellipse between a plane through the origin that is normal to the direction of propaga- tion s and the index ellipsoid (5.3-1). The two axes of the intersection ellipse are equal in length to 27 and 2%, m, and n) being the two indices of refraction, that is, the solutions of (5.2-15). These axes are parallel, respec- tively, to the directions of the D,,2 vectors of the two allowed solutions. To show that this procedure is formally equivalent to the methed of the last section, we follow the weatment of Born and Wolf (Reference 2). The procedure consists of solving for (P)!*—the two extrema of the intersection ellipse—and showing that they are equal 10 (#?)!? of (5.2-15). The proof is completed by showing that the radius vectors to these two extrema are paral- lel to D, 2. 5.3 THE INDEX ELLIPSOID 9E The ellipse is specified by the following two surfaces: 1. The index ellipsoid, S4+54+521 (53-2) 2. The normal to s, Ts = x5 + ys, + 2, = 0 (5.3-3) The principal semiaxes of the ellipse are given by the extrema of Pax tyre (5.3-4} subject to conditions (5.3-2) and (5.3-3). The problem of finding extrema subject to auxiliary conditions is handled by means of the Lagrange method of multipliers. We set up a function F(x, y, Z, Ai, A2) xt Fae ty ret Os tos ta tae + -1) (5.3-5) € x , e where A, and 4) are undetermined coefficients, We then solve for the quanti- ties x, y, and z at the extrema of r? and the ratio 4,/A2 from the equations oF oF _aF mow tae? (63-6) oF FF Ly 5.3-7) Oy OA, (. From (5.3-6), we obtain x ut Ma + hag kaxyz (5.3-8) i Equations (5.3-7) merely reproduce the auxiliary conditions (5.3-2) and (5.3-3). Multiplying Eqs. (5.3-8) by x;, adding them, and using (5.3-2) and (5.3-3) yield P+h=0 (5.3-9) Multiplying (5.3-8) by s,, summing over X, using (5.3-2) and (5.3-3), and recalling that s? = I lead to vy Se Bin (E+ B+ a0 (5.3-10) 2 ee ee Using (5.3-9) and (5.3-10) to eliminate A, and A, from (5.3-8) results in the relation ee (1-2) +8 E+ 3) 20 (53-11) 4 and two similar equations in which x is replaced, respectively, by y and z. 92 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION If, as in (5.3-1), we perform the substitution (5.3-12) to (5.3-11), we have r? > D?/E + Deo = x? and x/e, > E/ VE * Deg. With these substitutions, (5.3-11} becomes @,Ey = egt? [Ey — 5,(8 * E)] (5.3-13) which is identical to (5.2-13). Since r? — 7°, it follows that the semiaxes (the extrema of the intersection ellipse) are equal to the values of 2, that is, to the indices of refraction for the two propagation modes. Second, since the radius vectors r to the extreme points of the ellipse satisfy (5.3-12), they are parallel to the two allowed D vectors. This establishes the forrnal equivalence between the index ellipsoid method and the Maxwell equations solution of electromagnetic propagation in anisotropic crystals. 5.4 PROPAGATION IN UNIAXIAL CRYSTALS In uniaxial crystals, that is, crystals in which the highest degree of rotational symmetry applies to no more than a single axis, the equation of the index ellipsoid (5.3-1) simplifies to (4-1) where the axis of symmetry was chosen, following convention, as the z axis. it is also referred to as the optic axis, #, is called the ordinary index of refraction, whereas n, is the extraordinary one. If n, < ”,, we have a negative (optically) uniaxial crystal, whereas in a positive crystal, 1, > 1o- Figure 5.2 shows the index ellipsoid for a positive uniaxial crystal. The direction of propagation is along s. Since the ellipsoid in this case is invariant toa rotation about the z axis, the projection of the s vector on the x-y plane is chosen without loss of generality to coincide with the y axis. According to the “prescription” given in Section 5.3, we first find the intersection of the plane through the origin that is normal to s with the index ellipsoid. The intersection is an ellipse whose plane is cross-hatched in the figure. The length of the semimajor axis OA is equal to the index of refraction n, (6) of the “extraordinary” ray whose electric displacement vector D,(8) is parallel to OA. The “ordinary” ray is polarized (i.e., has its D vector) along OB and its index of refraction is equal to 7). It is clear from Figure 5.2 that as the angle between the optic axis and the direction of propagation s is changed, the direction of polarization of the ordinary ray remains fixed (along the x axis in the figure) and its index of refraction is always equal to ,. The direction of D,, on the other hand, depends, as shown, on 6. The index of refraction varies from #,(@) = 7, for 6 = 0° ton,(0) = 7, for 8 = 90°. The index of refraction 1,(6) of the extraordi- 3.4 PROPAGATION IN UNIAXIAL CRYSTALS 93 2 (optic) axis FIGURE 5.2 The construction for finding the indices of refraction and the allowed polarization directions for a given direction of propagation s. The figure shown is for a uniaxial crystal with i My = My, te = M% nary ray is equal to 04, which according to Figure 5.2, is given by 2 in? @ 1 cos’ | sin (5.4-2) ne) ne The three-dimensional surfaces giving the indices n.(@, 6) and 7,(0, @) as functions of the wave-norma! direction (6, @) are called the normal surfaces. Such surfaces can be constructed from the index ellipsoid by the methods given above. For a uniaxial crystal, the azimuthal angle @ is redundant, and the normal surface becomes an ellipsoid of revolution about the z (optic) axis. FIGURE 5.3 The intersection of the s — z plane with the normal surfaces for a positive uniaxial crystal (1, > 7.) 94 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION The intersection curves of these ellipsoids with the s-z plane are shown in Figure 5.3 for a positive uniaxial crystal (#, > #,). The exterior curve n,(@) isa plot of Eq. (5.4-2). Plots such as Figure 5.3 are very useful since they convey “at a glance the two indices. The orientation of the displacement vectors is also shown. D,{8) is in the s-z plane, whereas that of the ordinary ray is at right angles to this plane. 5.5 NORMAL MODE EXPANSION OF THE ELECTROMAGNETIC FIELD IN A RESONATOR Maxwell's equations and the current continuity equation in the MKS system of units are VxE= = aD ar V-D=p (5.5-1} V-B=0 vui--% at We will limit ourselves, for the moment, to charge-free, isotropic, and homo- geneous media so that $#=0 B=pH V-D=0 D=cE (5.5-2) where « is the dielectric constant, Consider the electric field E(x, #) and magnetic field H(r, 1) inside a volume V bounded by a surface S$ of perfect conductivity. The tangential component of E, —n X n x E, and the normal component of H, n - H, must both be zero on S (nn is the unit vector normal to 5). We will expand B and H in terms of two orthogonal sets of vector fields E, and H,, respectively. These sets, which were introduced originally by Slater (Reference 3) obey the rela- tions KE, = VX He (5.5-3) KH, = Vx E, (5.5-4} where k, is to be considered, for the moment, a constant. The tangential component of BE, on S is zero. nxE,=0 on $ (5.5-5) If we take the curl of both sides of (5.5-3, 4) and use the identity VxVKA=ViV-A)- VA 5.5 NORMAL MODE EXPANSION/ELECTROMAGNETIC FIELD RESONATOR = 95 they become WE, + KE, =0 (5.5-6) VM, + GH, = 0 that is, the familiar wave equation. It follows from (5.5-3), (5.5-4), and (5.5-5) that the normal component of H,, n+ H,is zero on S. To prove this statemeht, consider an arbitrary closed contour € on S surrounding a surface 5’. fBerdl= 4 (onxmx Bj + fake d= 0 (55-7) where E, is expressed as the vector sum of its tangential (—m x n x E,} and normal (n - E,)n components. The first term on the right side of (5.5-7) is zero because of (5.5-5} whereas the second one is zero since n is perpendicu- Jar to dl. Using Stokes’ theorem on the eft side of (5.5-7) gives $B dl = [07 x By emda = ky f (Bes) da = 0 and since C is arbitrary, it follows that H,-n=0 on s {5.5-8) We will next prove that the functions E, and H, are orthogonal in the sense fea Beav=0 ath (55-9) [HoH av =0 a#p To prove the first of Eqs. (5.5-9}, apply the vector identity V+ (AX B) = B-VxA-—A-V~* Bto (E, x V x E,), then to (E, x ¥ x E,) and subtract. The result is V-(E, x VX E,) — V- (BE, x VX Es) =VXE,-VXE,~ By Vx VXE,- VX Ey VXE, +B, VX VX Ey From (5.5-4), we have V x E, = k,H, and V x V x B, = KE,, which substituted in the last equation gives Ka¥ + (By X Hy) ~ ky W> (By X By) = 8 — KEE» Bs that, after applying Gauss’s theorem (see Section 5.1), becomes fg Wen (Be X HL) — kam (Ee Hb)} da = (8 ~ Ki) [Be Bay The left side of the last equality can be shown, with the aid of the identity A+B x C=C-A x Band (5.5-5), to be zero, so that for &, # k,, the first of Eqs. (5.5-9) is proved. If k, = &,, that is, when E, and B, are members ofa degenerate set, it is possible to construct linear superpositions of the degener- ate functions so that orthogonality is preserved. The proof of the orthogonal- ity of the H, functions follows along identical lines. We are free to choose the magnitude of H, and E, so that they are 96 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION normalized according to [te By av = Bue (5.5-10) | Ba Ee dv = Bee This choice will be used throughout this text. The total resonator fields of E(r, f) and H(r, #) can be expanded as Bj =-> PeAtE a(t) (5.5-11} Hin = % coagal)He(0) where w, = k,/Ve. Substituting (5.5-11) in the first of Maxwell equations (5.5-1) and using (5.5-3) result in Po = 4a (53.5-12) and in a similar fashion, from the second of (5.5-1}, we have wg, = —Ba (5.5-13} Eliminating 42 gives Ba + wir, = 0 (5.5-14) This identifies k.(we)"'2 = aw, as the radian oscillation frequency of the ath mode. 5.6 THE QUANTIZATION OF THE RADIATION FIELD Im this section we will show that the electromagnetic field inside a resonator can be considered, formally, as an ensemble of independent harmonic oscilla- tors. The formalism is related closely to that used in Chapter 4, to quantize the spectrum of lattice vibrations. To bring out the similarity (dot) multiply the first of Eqs. (5.5-11) by E, and integrate over the resonator volume. The result, after using (5.5-10), is pelt) = —Ve f Bee, 9 Bat) dv (5.6-1) This equation is analogous to the second of Eqs. (4.5-2). It follows from (5.6- 1) and from the relation Ve ant) = Ef He 1) te dy (562) that the state of the classical electromagnetic field can be specified by H(r, t) and E(r, t) or, alternatively, by the dynamical variables p.{t} and q,(t). The total energy (Hamiltonian) is #24) uM + BB) av (5.6-3) 5.6 THE QUANTIZATION OF THE RADIATION FIELD 97 Substituting for Hand E their expansion (5.5-11) gives H = > Ups + wig) (5.6-4) which has the basic form of a sum of harmonic oscillator Hamiltonians as in (2.2+1}. The dynamical variables p, and 4, constitute canonically conjugate variables. This can be seen by considering Hamilton's equations of motion relating #, to q, and gq. tO pa- . ax 8‘ Po = — 5 = — Wale 94 (5.6-5) 5, = 2h. 4s = 5, Pe These are identical with Eqs. (5.5-12} and (5.5-13) obtained from Maxwell's equations. The quantization of the electromagnetic radiation is achieved by consid- ering p, and 4, as formally equivalent to the momentum and coordinate opera- tors of a quantum mechanical harmonic oscillator, thus taking the commuta- tor relations connecting the dynamical variables as {Pa, Pot = [42-4] = 0 [4a Pol = th Bae In a manner analogous to that used in Chapter 2 [see (2.2-25)}, we define the creation operator a and the annihilation operator a; by 5.6-6) 1 aly = (Ge) toate ~ ioe (5.6-7) . ( 1 y . ait) = (oa) lomle) + deo The commutator relations are found directly from (5.6-6) to be [4), &e] = [al abet = 0 (56-8) (ar, ah] = Bim Solving (5.6-7) for p; and q: gives , (fata pdt) =i) la ~ 24)] ( 2 ) , , (5.6-9} a ati) = Ga)” lah) + este] We can express the Hamiltonian in terms of the operators a} and a; by substituting (5.6-9) in (5.6-4) and replacing aaf by aja; + 1 [according to (5.6-8)}. The result is # = 3 teala, + 4 (5.6-10) 7 98 ELECTROMAGNETIC FIELDS AND THEIR QUANTIZATION which is the same as Eq. (4.5-12) obtained for the case of the lattice vibra- tions. The formal analogy between the operators af, a; and their counterparts in the case of the harmonic oscillators shows that, quantum mechanically, a Stationary state of the total radiation field can be characterized by an eigen- function ®, which is a product of the eigenfunctions of the individual Hamil- tonians fn(alay + 4) = (5.6-11) where Bly, = Vea ET tet hn, = Vt Uy (5.612) Af aithy, = Mite The expectation value of the operator aha; is (| ala |) = (ou| ala fr) = me (5.6-13) and is equal to the number of quanta 7; in the 4th mode of the resonator. Plane-Wave Quantization The discussion just concluded uses a generalized resonator of unspecified shape. We will find it useful to consider the form of the field operators in the case of a plane-wave resonator. Although such a resonator, which requires an infinite cross-sectional area does not exist. most optical resonators that em- ploy curved mirrors as reflectors involve nearly plane-wave propagation. To be specific, consider the /th mode of a resonator of length £ along the axis and mode volume V. Let the electric and magnetic field vectors point along the y and x directions, respectively. Equations (5.5-3, 4, 9) are satisfied by _ fhw\u2 4 ; Efe, = — i&(-— ft) = alt] ky de = = CY ale — ate) sin ke 56-14 he sce = 9 G2)" tale + a] cos ke where Vis the mode volume. In the case of plane wave modes, the total field can be written as a summation over all modes k fi B= 3-48 Va (also — ae) ak fh i H= 3 ig X Gay Tav (a em + ase”) ras (5.6-15)

You might also like