327

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 2: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Wendell T. Hill and Chi H. Lee

Light-Matter Interaction

Page 3: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Each generation has its unique needs and aspirations. When Charles Wiley firstopened his small printing shop in lower Manhattan in 1807, it was a generationof boundless potential searching for an identity. And we were there, helping todefine a new American literary tradition. Over half a century later, in the midstof the Second Industrial Revolution, it was a generation focused on buildingthe future. Once again, we were there, supplying the critical scientific, technical,and engineering knowledge that helped frame the world. Throughout the 20thCentury, and into the new millennium, nations began to reach out beyond theirown borders and a new international community was born. Wiley was there, ex-panding its operations around the world to enable a global exchange of ideas,opinions, and know-how.

For 200 years, Wiley has been an integral part of each generation’s journey,enabling the flow of information and understanding necessary to meet theirneeds and fulfill their aspirations. Today, bold new technologies are changingthe way we live and learn. Wiley will be there, providing you the must-haveknowledge you need to imagine new worlds, new possibilities, and new oppor-tunities.

Generations come and go, but you can always count on Wiley to provide youthe knowledge you need, when and where you need it!

William J. Pesce Peter Booth WileyPresident and Chief Executive Officer Chairman of the Board

1807–2007 Knowledge for Generations

Page 4: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Wendell T. Hill and Chi H. Lee

Light-Matter Interaction

Atoms and Molecules in External Fields

and Nonlinear Optics

WILEY-VCH Verlag GmbH & Co. KGaA

Page 5: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

The Authors

Prof. Wendell T. Hill

University of Maryland

Institute for Physical Science and Technology

USA - College Park MD 20742

[email protected]

Prof. Chi H. Lee

University of Maryland

Electrical and Computer Engineering

USA - College Park Md 20742

[email protected]

Cover

False color representation of the probability

density (square of the wavefunction) for an

electron in the 03lmd state of the hydrogen atom.

All books published by Wiley-VCH are carefully produced. Nevertheless, authors, editors, and publisher do not warrant the information contained in these books, including this book, to be free of errors. Readers are advised to keep in mind that statements, data, illustrations, procedural details or other items may inadvertently be inaccurate.

Library of Congress Card No.:applied for

British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library.

Bibliographic information published by the Deutsche Nationalbibliothek Die Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available in the Internet at <http://dnb.d-nb.de>.

2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

All rights reserved (including those of translation into other languages). No part of this book may be reproduced in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into a machine language without written permission from the publishers. Registered names, trademarks, etc. used in this book, even when not specifically marked as such, are not to be considered unprotected by law.

Typesetting Uwe Krieg, Berlin

Printing Strauss GmbH, Mörlenbach

Binding Litges & Dopf GmbH, Heppenheim

Printed in the Federal Republic of Germany

Printed on acid-free paper

ISBN: 978-3-527-40661-6

Page 6: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

V

Contents

Preface XIII

Part 1 Light-Matter Interaction: Atoms, Molecules and External Fields 1

1 Hydrogen-Like Ion: An Atom (Ion) With One Electron 31.1 Bohr Model of the Atom 41.2 Hydrogen-Like Ions, Quantum Approach: Bound States 71.2.1 Angular Wavefunctions 71.2.2 Radial Wavefunction and Energy States 101.2.3 Exact Radial Solution, Hydrogen-Like Ions 131.2.4 Energy Units and Atomic States 171.3 Classification of Nonrelativistic States 191.3.1 Parity 201.3.2 Degeneracy 201.4 Corrections to the Energy Levels 211.4.1 Relativistic Motion 211.4.1.1 Electron Spin and the Dirac Equation 221.4.1.2 Classification of Relativistic Hydrogen States 261.4.1.3 Hydrogen-Like Ion Wavefunction Including Spin 271.4.2 Fine Structure and Spin–Orbit Interaction 281.4.3 Rydberg Series 291.5 Continuum States 30

Further Reading 31Problems 31

2 The Structure of the Multielectron Atom 332.1 Overview 332.2 Angular Momentum Coupling Schemes 402.2.1 LS or Russell–Saunders Coupling 402.2.2 jj Coupling 45

Page 7: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

VI Contents

2.2.3 Intermediate or Pair Coupling 452.2.4 Recoupling Between Coupling Schemes 482.3 Fine Structure 49

Further Reading 51Problems 51

3 Atoms in Static Fields 533.1 External Electric and Magnetic Fields 533.1.1 Stark Effect 533.1.1.1 Linear Stark Effect 573.1.1.2 Quadratic Stark Effect 593.1.2 Zeeman Effect 603.2 Hyperfine Structure 613.2.1 Magnetic Interaction 613.2.2 Explicit Expression for Al 623.2.3 Hyperfine Zeeman Effect 653.2.4 Electric Quadrupole Correction 65

Further Reading 66Problems 66

4 Atoms in AC Fields 694.1 Applied EM Fields 694.1.1 Radiation Hamiltonian 694.1.2 Coulomb or Radiation Gauge 714.2 Free-Electron Wavefunction 714.3 Radiative Transitions 724.3.1 One-Photon Transitions 744.3.2 Two-Photon Transitions 754.3.3 Transition Rate: Fermi’s Golden Rule 764.3.3.1 Degeneracy 784.3.3.2 Narrow and Broad Sources 794.3.4 Transition Strength: Absorption 794.3.4.1 Line Strength 794.3.4.2 Cross Section 804.3.4.3 Oscillator Strength 804.3.5 Transition Strength: Emission 814.4 Selection Rules for Atomic Transitions 824.4.1 Electric Dipole (E1) Transitions 824.4.2 Magnetic Dipole (M1) Transitions 844.4.3 Electric Quadrupole (E2) Transitions 844.5 Atomic Spectra 854.5.1 Rydberg Series 85

Page 8: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Contents VII

4.5.2 Autoionization 894.5.3 Photoionization with Intense Lasers 92

Further Reading 94Problems 94

5 Diatomic Molecules 975.1 The Hamiltonian 985.2 Born–Oppenheimer Approximation 995.3 Nuclear Equation 1015.3.1 Harmonic Approximation of U(R) 1025.3.2 Beyond the Harmonic Approximation of U(R) 1045.3.3 Vibrating Rotator 1055.3.4 Analytic Expression for U(R) 1075.3.5 More Accurate Techniques 1075.4 Electronic States 1095.4.1 Angular Momenta in Cylindrically Symmetric Fields 1105.4.1.1 Orbital Angular Momentum 1115.4.1.2 Spin Angular Momentum 1115.4.1.3 Multiplet Splitting 1125.4.1.4 Total Angular Momentum 1125.4.1.5 Labeling Nomenclature 1135.4.2 Angular Momenta Coupling: Hund’s Cases 1145.4.2.1 Hund’s Case (a) 1145.4.2.2 Hund’s Case (b) 1165.4.2.3 Hund’s Case (c) 1175.4.2.4 Hund’s Case (d) 1185.4.3 Molecular Symmetries: Electronic Motion 1185.4.3.1 Inversion Symmetry 1195.4.3.2 Reflection Symmetry and Σ States (Λ = 0) 1215.4.3.3 Reflection Symmetry and Λ = 0 States 1225.4.3.4 Exchange of Nuclei 1225.4.4 Molecular Symmetries: Nuclear Motion 1235.4.5 Molecular Symmetries: Herzberg Bookkeeping Diagram 123

5.4.5.1 The caseJ = N (S = 0) 124

5.4.5.2 The caseJ = N + 1 (S = 1) 1245.4.6 Molecular Symmetries: Nuclear Spin 1265.4.6.1 Example 1275.4.7 Molecular State Labeling Convention 1285.4.7.1 Rule 1: Ground State 1285.4.7.2 Rule 2: Excited States with Ground-State Multiplicity 1285.4.7.3 Rule 3: Excited States with Different Multiplicity 1285.4.8 Molecular Orbital Theory 128

Page 9: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

VIII Contents

5.4.8.1 United Atom Construction 1315.4.8.2 Separated Atom Construction 132

Further Reading 135Problems 135

6 Molecules in External Fields 1376.1 Introduction 1376.2 Electronic Transitions 1376.2.1 General Selection Rules 1396.2.2 Case-Specific Selection Rules 1396.2.2.1 Hund’s case (a) 1406.2.2.2 Hund’s case (b) 1406.2.2.3 Hund’s case (c) 1406.2.2.4 Hund’s case (d) 1416.2.3 Examples 1416.3 AC Tunneling Ionization 142

Further Reading 146Problems 146

Part 2 Light-Matter Interaction: Nonlinear Optics 149

7 Nonlinear Optics 1517.1 Introduction 1517.2 Phenomenological Description 1527.2.1 Second-Harmonic Generation 154

7.2.2 Electrooptic Effect, χ(2)(ω; 0, ω) 156

7.2.3 Optical Rectification χ(2)(0; ω,−ω) 157

7.2.4 Parametric Generation χ(2) (ωs; ωp,−ωi

)158

7.2.5 Third-Order Nonlinear Effect 1587.2.6 Nonlinear d coefficient 159

Further Reading 165Problems 165

8 Wave Propagation 1678.1 Nonlinear Wave Equation 1678.2 Phase Matching in SHG 1708.2.1 Phase Matching of SHG in KDP 1718.2.2 Noncollinear Momentum Matching 1768.2.3 Experimental Arrangement 1788.3 Parametric Interaction 1798.3.1 Coupled Equations 1798.3.2 Parametric Amplification 184

Page 10: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Contents IX

8.4 Parametric Oscillation 1888.4.1 Tuning of OPO 1908.5 The Manley–Rowe Relations 1928.6 Parametric Upconversion 194

Problems 196

9 Quantum Theory 1999.1 Introduction 1999.2 Density Matrix Formalism 1999.3 Perturbation Method 2019.4 Transition Probability 2039.5 Two-Photon Absorption 2089.6 Scattering Cross Section 2119.7 Three-Photon Absorption 2129.8 Doppler-Free Two-Photon Absorption 2139.9 Calculation of Susceptibility 2159.10 Third-Order Nonlinear Susceptibility 218

Problems 221

10 Applications 22310.1 Optical Harmonic Generation 22310.1.1 Theory of Light Wave at Boundary 22410.1.2 Criteria of Null Transmitted SHG 22810.1.3 Experimental Observations of the Second-Harmonic Generation at

Oblique Incidence 23110.2 SHG due to Reflection from Media 23710.3 Nonlinear Electroreflectance 24210.4 Near-Field Second-Harmonic Microscopy 24410.5 Terahertz Pulse Generation 250

Problems 257

Appendices

A Atomic Physics Definitions 259A.1 Air and Vacuum Wavelengths 260A.2 Wavenumber 260A.3 Fine-Structure Constant 260A.4 Atomic Energy Unit (Hartree) 261A.5 Rydberg Energy Unit 261A.6 eV Energy Unit 261A.7 Mass 262

Page 11: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

X Contents

A.8 Length 262A.9 Atomic Velocity and Momentum 263A.10 Atomic Time Scale 263A.11 Atomic Field Strength 264A.12 Atomic Unit of Dipole Moment 264A.13 Magnetic Moments 264A.13.1 Electron Magnetic Moment 264A.13.2 Proton Magnetic Moment 266A.13.3 Neutron Magnetic Moment 267A.13.4 Magnetic Moment of the Nucleus 267A.14 Quadrupole Moment of the Nucleus 267A.15 Frequently Used AMO Quantities 268

B Mathematics Related to AMO Calculations 269B.1 Kronecker Delta, δij 269B.2 Dirac Delta Function, δ (x − xo) 269B.3 Hypergeometric Series 270B.4 Confluent Hypergeometric Series 271B.5 Associated Laguerre Polynomials 271B.6 Legendre and Associated Legendre Functions 271B.7 Spherical Harmonics 273B.8 Mathematical Formalism of Quantum Mechanics 274B.8.1 Summary 279B.9 Schrödinger’s Equation in Parabolic Coordinates 280B.10 Voigt Line Profile 283

Further Reading 284

C Atomic and Molecular Data 285C.1 NIST Online Data 285C.1.1 NIST Online Atomic Data 285C.1.2 NSIT Online Molecular Data 285C.2 Molecular Constants 285C.3 Filling Subshells 286C.4 Electronic Configurations 287

D Coupling Angular Momenta 289D.1 Two Angular Momenta and 3-j Symbols 289D.2 Properties of 3-j Symbols 290D.3 Three Angular Momenta and 6-j Symbols 291D.4 Four Angular Momenta and 9-j Symbols 292

Page 12: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Contents XI

E Tensor Algebra 293E.1 Spherical Tensors 293E.2 Commutation Relations 294E.3 Reduced Matrix Elements 294E.4 Matrix Elements of Products of Operators 296

Further Reading 300

References 301

Index 305

Page 13: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 14: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

XIII

Preface

With several Nobel Prizes to its credit over the past decade, research in atomic,molecular and optical (AMO) science and engineering finds itself at the apexof investigations ranging from fundamental studies in quantum mechanicsto the development of new nano-technologies. Whereas in years past, mostresearchers working in AMO fields had strong backgrounds in physics, op-tics and electromagnetic theory, this is not always the case today as the fieldand its participants become more and more diverse. Consequently, there is aneed for a text designed to provide an overview of the field to get researchersnew to AMO areas up and running as quickly as possible. This two-volumeseries entitled Light-Matter Interactions is designed to meet this need and tobe user-friendly to investigators from physics, chemistry and the engineeringdisciplines. The purpose of Volume II is to provide a firm understanding ofatoms and molecules, their interaction with external fields and how to exploitnonlinear optical processes to manipulate light. The text is geared towardthe advanced undergraduate and the first- or second-year graduate student.While some familiarity with quantum mechanics and electromagnetic theoryis assumed, readers with limited or no knowledge of atoms, molecules andnonlinear optics will find this text a good introduction. The more advancedreaders should appreciate having available in one reference a vast array ofideas.

The number of subjects included in this volume is far too great for all tobe covered in detail. As the details of each subject can be found readily inthe literature, much of which we cite, it is not necessary or even prudent towrite such a text. Our approach has been, rather, to give key results, for themost part. The derivations we have included were chosen to help illuminatethe physics. To encourage the reader to delve deeper, a list of references andsuggested extra reading is given at the end of each chapter. Again, our goalis to present enough information for readers from a variety of backgrounds tounderstand the important concepts.

The book is organized into three parts. Atoms and molecules are the subjectof the first six chapters. The focus will be on the valence electrons as they

Page 15: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

XIV Preface

are primarily responsible for how the system interacts with its environment– chemical reactions, cooling and trapping, etc. Nonlinear optics will occupythe balance of the main portion of the text. Finally, to aid the reader’s effortto become familiar with the numerous energy units, atomic parameters andspecial functions used in AMO research, we have compiled a set of appendicessummarizing those most frequently used.

Our discussion of atoms and molecules will be confined to three specifictypes of systems – one-electron atoms, multi-electron atoms and diatomicmolecules. In Chapter 1 we begin with hydrogen-like ions and look in somedetail at their structure both nonrelativistically and relativistically; we intro-duce spin, spin-orbit interaction and the radiative hamiltonian all via the Diracequation. We present the one-electron wavefunction, parity, the Rydberg con-stant and series as well as bound and continuum states in this chapter. Chap-ter 2 is devoted to multi-electron atoms where we classify states, discuss an-gular momentum coupling schemes and fine structure. Static field effects,Stark and Zeeman, are discussed in Chapter 3. Hyperfine structure as wellas field ionization are also subjects covered. Radiative transitions in the pres-ence of electromagnetic fields dominate the discussion in Chapter 4. Oscilla-tor strengths, cross sections, line strengths, multiphoton transitions, selectionrules (for electric and magnetic dipole along with electric quadrupole transi-tions) and strong-field ionization processes round out our discussion in thischapter.

We turn our attention to diatomic molecules in Chapters 5 and 6. Chapter 5begins by presenting solutions to the separable nuclear and electronic equa-tions. The consequences of this separation are discussed as well. Electronicstates along with Hund’s four primary cases are discussed in this chapter. Inaddition, molecular symmetries, molecular labeling conventions and molec-ular orbital theory are reviewed. Radiative transitions in diatomic systems isthe subject of discussion in Chapter 6, which includes a summary of selectionrules for electric dipole transitions for each Hund’s case.

The second part of this volume deals with nonlinear optics. Ultrashort op-tical pulses and extremely high field strengths (intensities) are now routinelygenerated in contemporary laboratories. When exposed to such fields, con-densed and gas phase systems respond nonlinearly. In particular, these fieldsinduce a nonlinear polarization leading to a variety of phenomena, the mostbasic of which is optical second harmonic generation. From its inception in theearly 1960s, nonlinear optics has blossomed into a research field of its own. Asis the case with atoms and molecules, there are many excellent texts on nonlin-ear optics. We have made no attempt to duplicate these presentations. Rather,we seek to provide the reader with a straightforward and practical view ofnonlinear optics. Although nonlinear interactions include higher order effects,involving interaction with lattice vibration, etc., the basic principle can be un-

Page 16: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Preface XV

derstood in terms of optical second harmonic generation. Consequently, themain focus of this section will be harmonic generation and parametric inter-action.

Chapter 7 provides a phenomenological description of nonlinear optics.Nonlinear phenomena, including the electro-optic effect, sum- and difference-frequency generation, optical rectification, terahertz generation and paramet-ric generation, are described in terms of the nonlinear susceptibility, with var-ious combinations of frequencies. More complicated nonlinear effects can bedescribed by a nonlinear polarization expressed as a product of higher ordernonlinear susceptibilities and higher order products of optical fields. Thesegive rise to the phenomena of four wave mixing, intensity dependent indexchanges. Temporal effects give rise to self-phase modulation, while the spatialeffects cause self-focusing or spatial solitons. All of this is discussed in Chap-ter 7. Space requirements dictate omitting some subjects, however. We havechosen not to include important phenomena requiring more of a backgroundin condensed matter than this volume provides.

Once the nonlinear polarization is obtained, it can be regarded as the driv-ing source term for the nonlinear wave equation. This leads us to a gen-eral, albeit somewhat detailed, treatment of wave propagation in a nonlinearmedium, the subject of Chapter 8. Since the source wave and the generatedharmonic wave propagate at different speeds through the medium, conceptsand practical methods for achieving phase-matching are discussed. To in-troduce parametric processes (oscillation and amplification), we extend ourdiscussion of degenerate 3-wave interaction to three waves with different fre-quencies.

For those interested in learning how to calculate the nonlinear susceptibili-ties from a quantum picture, Chapter 9 provides an overview of the quantumtheory of nonlinear processes. The density matrix formulism is introducedand combined with the perturbation method. This is followed by a treatmentof multiphoton absorption. Calculations of linear and nonlinear susceptibil-ities, up to the third order, are then presented. Doppler free two-photon ab-sorption is presented in this chapter as an example of using nonlinear effects torealize high resolution measurement in the presence of Doppler broadening.

Finally, a few additional applications of nonlinear optical effects are givenin Chapter 10. Since many well-known applications are covered extensivelyin standard texts on nonlinear optics, we decided to cite a few applicationsthat have received less attention. These include optical harmonics generationat oblique incidence angles where a nonlinear Brewster’s angle is introduced,harmonic generation from surfaces, nonlinear electro-reflectance, near fieldsecond harmonic microscopy and terahertz pulse generation by optical recti-fication.

Page 17: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

XVI Preface

Five appendices appear at the end of the book covering atomic physicsdefinitions, mathematics related to AMO calculations, atomic and moleculardata, coupling angular momenta and tensor algebra. These appendices are in-tended as a reference source and include links to several web-based databaseswhere the latest atomic and molecular information can be downloaded.

We wish to thank Professor Marshall Ginter for valuable discussions andhelpful advice and Professor William DeGraffenreid for generating many ofthe figures. We acknowledge collaborations with our graduate students whohelped generate some of the data presented in this text and the technical as-sistance of Mrs. Patsy Keehn and Mr. Alen Shu-Zee Lo who typed portionsof the manuscript. Much of the material came from teaching graduate coursesat the University of Maryland over the years. We thank the many graduatestudents for feedback that helped focus our thoughts. We trust the readerswill find this volume useful and we welcome suggestions for improvement.

College Park, Maryland, summer 2006 Wendell T. Hill, III and Chi Lee

Page 18: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

Part 1Light-Matter Interaction: Atoms, Molecules and ExternalFields

Page 19: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 20: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

3

1Hydrogen-Like Ion:An Atom (Ion) With One Electron

The atom can be viewed in a variety of ways. From a macroscopic point ofview, the atom is the smallest entity of significance and the building blocks ofmatter. From a microscopic point of view, the atom is a perplexing compositeof more basic particles of which only the electron appears to be fundamentalor elementary; the proton and neutron are composed of more bizarre con-stituents. Our focus will not be at the subatomic level where the strong andweak forces must be considered along with the electromagnetic force nor themacroscopic and mesoscopic levels involving clusters of particles. Rather, wewill concentrate on the atomic level where the Coulomb interaction betweenthe electrons and the nucleus as well as between electrons (and between dif-ferent nuclear centers in molecules) is the dominant interparticle force of con-cern, at least in the absence of an applied strong field. We will assume thenucleus (composed of protons and neutrons) and the electron to be the basicconstituents of concern. Our primary goal is to model atoms and moleculesboth in the absence and presence of weak and strong external electromagneticfields. We will treat the motion of the electrons and nuclei quantum mechan-ically, for the most part, and the external field classically. Consequently, wewill seek solutions to Schrödinger’s equation,

ih∂

∂tΨ (r, t) = HΨ (r, t) , (1.1)

for specific Hamiltonians, H.We begin our discussion by considering the hydrogen-like, one-electron

atom, in the absence of an external field. This atom will be described by afamily of stationary states of well-defined energies, that can be ascertainedfrom the time-independent Schrödinger equation

Hψ (r) = Eψ (r) . (1.2)

Equation (1.2) is obtained from Eq. (1.1) by letting Ψ (r, t) → e−iEt/hψ (r).The Hamiltonian is taken as the quantum mechanical analog of the classicalHamiltonian given by T + V, where T = p2/2me and the potential, V (r),is just the Coulomb potential, −Ze2/4πεor, with Z being the charge on the

Page 21: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

nucleus (Z = 1 for neutral atoms). Since p → −ih∇ (see Appendix B.8), thequantum mechanical Hamiltonian for this single-particle nonrelativistic casetakes the simple form

H = − h2

2me∇2 − Ze2

4πεor, (1.3)

where me is the mass of the electron, e is the elementary charge (which wetake to be positive in this book), h is Planck’s constant divided by 2π and εo

is the vacuum permittivity. We will refer to this as our one-electron, field-freeHamiltonian.

1.1Bohr Model of the Atom

Before solving Eq. (1.2) with the Hamiltonian given in Eq. (1.3), it is interest-ing to see that it is possible to determine the energy spectrum classically byassuming the existence of stationary states – Bohr’s hypothesis that the elec-tron does not radiate when circling the nucleus. For stationary states to exist,the Coulomb force (−Ze2/4πεor2) must equal the centripetal force (−mev2/r,where v is the tangential velocity). This leads to a kinetic energy of

T =12

mev2 =12

(Ze2

4πεor

)(1.4a)

and a total energy of

E = T + V = −12

(Ze2

4πεor

). (1.4b)

Here, we adopt the usual convention that negative energy means that the sys-tem is bound. Prior to the development of quantum mechanics, it was knownthat atoms emitted light at specific wavelengths. In the late 19th century itwas determined that the energy of the light, in wavenumbers (see Eq. (A.4) ofAppendix A), associated with a transition between energy levels n and n′ wasfound to obey1

νn′n = C(

1n2 − 1

n′2

), (1.5)

where C is a constant that is related to the Rydberg constant. If one furtherpostulates that the energy levels are quantized such that r → n2r1, where r1 is

1) This general formula was proposed by Rydberg in 1889. Prior tothis date, Balmer used the empirical formula (4/B)

(1/22 − 1/n2)

to fit the visible hydrogen spectral lines he observed. With n =3, 4, 5, 6, . . . , he extracted a value of 3645.6 Å for B.

Page 22: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.1 Bohr Model of the Atom 5

the minimum stable radius, Eq. (1.4b) becomes

En = −12

(Ze2

4πεon2r1

). (1.6)

It follows that

hcνn′n = En′ − En =12

(Ze2

4πεor1

)(1n2 − 1

n′2

). (1.7)

If we identify the constant as

C =1

2hc

(Ze2

4πεor1

), (1.8)

the energy spectrum for hydrogen and hydrogen-like ions 2 takes this simpleform:

En = −hcC 1n2 . (1.9)

Since the potential is not harmonic, the energy levels are not equally spaced.Figure 1.1 shows that the second level, the first excited state, is three-quartersof the way to the ionization limit while subsequent levels become more closelyspaced.

It is possible to determine C by fitting the spectrum to Eq. (1.5), from whichwe can estimate r1. Furthermore, it is possible to predict C and r1 theoreticallyfrom the correspondence principle.3 When the radius of the electron orbitbecomes large, n becomes large and the circulating electron will radiate, asany charged particle must classically, with a frequency given by

ν =v

2πn2r1. (1.10)

When a classical particle radiates, it loses energy. Quantum mechanically, ra-diation is accompanied by transitions between quantum levels, as implied byEq. (1.5). To remain classical, radiation must be between levels with large n.Consider a transition between two adjacent levels where n′ = n + 1. Equa-tion (1.5) then becomes

νn′n = C(

1n2 − 1

(n + 1)2

) 2C

n3 . (1.11)

2) Hydrogen-like ions are ions with a single electron plus the nucleus.These include H, He+, as well as all multiply-charged ions that haveall but one electron removed; Ca19+ would be an example.

3) The correspondence principle states that in the limit of large sizesthe quantum description must approach the classical description.

Page 23: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

6 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

Fig. 1.1 Bound energy levels of the traditional Bohr hydrogen atom.

Multiplying Eq. (1.11) by c, the speed of light, and equating it with Eq. (1.10),using Eq. (1.4a) to express v and Eq. (1.8) to express C , we can show that

r1 (

4πεo

e2

)h2

Zme, (1.12)

from which we identify the minimum radius for Z = 1,

ao ≡(

4πεo

e2

)h2

me, (1.13)

called the Bohr radius.4 The Rydberg constant, for an infinitely massive nu-cleus,5 is then defined as

R∞ ≡ 14π

(e2

4πεo

)1

hcao=

CZ2 . (1.14)

The energy spectrum can then be expressed as

En = −12

(e2

4πεo

)Z2

n2ao= −hcR∞

Z2

n2 = −12

(e2

4πεo

)2 Z2me

n2 h2 , (1.15)

which depends only on the integers, Z and n, and fundamental constants.

4) We note that r1 = ao/Z.5) An atom with an infinite nuclear mass is sometimes called a fixed

nucleus or stationary nucleus atom.

Page 24: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 7

For an infinite nuclear mass, we identify n2ao as the radius of the orbit of thenth state for hydrogen. This is not exact because the nucleus has finite massmN. To account for this, we can let me → µ, the reduced mass. The nth radiusthen becomes

rn =me

µ

n2ao

Z n2ao

Z, (1.16)

assuming that the nucleus has charge Z, where the reduced mass is given by

µ ≡(

1me

+1

mN

)−1

. (1.17)

If we let me → µ in Eq. (1.15), however, we shift the energy levels, requiringthe finite-mass Rydberg constant to be defined as

RM = R∞

(1 +

me

MN

)−1

. (1.18)

1.2Hydrogen-Like Ions, Quantum Approach:Bound States

Although we were able to determine the spectrum for the simple hydrogen-like ions from our classical approach, to study any of the details of thestructure (energy levels) requires the time-independent Schrödinger equa-tion (Eq. (1.2)) to be solved. In this section we will look at exact solutions inthe nonrelativistic approximation for hydrogen-like ions.

1.2.1Angular Wavefunctions

When we generalize the field-free Hamiltonian (Eq. (1.3)) by letting me → µ,the reduced mass (Eq. (1.17)), the Schrödinger equation takes the form

∇2ψ (r) +2µ

h2

(E +

Ze2

4πεor

)ψ (r) = 0. (1.19)

Again, Z is the residual charge on the nucleus. That is, Z = 1 for H, Z = 2for He+, etc. The Coulomb potential only depends on

∣∣−→r ∣∣ and, thus, fallsinto the class of centrally symmetric potentials. It is natural to seek solutionsto Eq. (1.19) in spherical coordinates in this case. If we write the Laplacian inspherical coordinates,

∇2 → 1r2

∂r

(r2 ∂

∂r

)+

1r2 sin ϑ

∂ϑ

(sin ϑ

∂ϑ

)+

1r2 sin2 ϑ

∂2

∂ϕ2 , (1.20)

Page 25: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

the radial and angular variables separate and the solutions to Eq. (1.19) takethe form

ψ (r) = Rnl(r)Ylm (ϑ, ϕ) , (1.21)

where Rnl(r) is known as the radial wavefunction and Ylm (ϑ, ϕ) the sphericalharmonics (see Appendix B.7). Such wavefunctions describe stationary stateswith definite energy E and angular momentum l. For centrally symmetricpotentials, the angular momentum is conserved.6 The Ylm (ϑ, ϕ)s describe theangular part of the electron distribution and are given by

Ylm (ϑ, ϕ) = (−1)m

√(2l + 1)

(l − m)!(l + m)!

Pml (cos ϑ) eimϕ, (1.22)

with Pml (cos θ) being the associated Legendre polynomial (see Appendix B.6)

given by

Pml (x) =

12l l!

(1 − x2

)m/2 dl+m

dxl+m

(x2 − 1

)l, (1.23)

where m = −l,− (l − 1) , . . . , 0, . . . , l − 1, l.7 The spherical harmonics are or-thogonal, normalized and obey∫ 2π

0dϕ

∫ π

0Y∗

l ′m′ (ϑ, ϕ) Ylm (ϑ, ϕ) sin ϑdϑ = δl ′lδm′m, (1.24)

where δij is the Kronecker delta (see Appendix B.1). The first six sphericalharmonics are

Y00 (ϑ, ϕ) =√

1/4π

Y10 (ϑ, ϕ) =√

3/4π cos ϑ

Y1±1 (ϑ, ϕ) = ∓√3/8π sin ϑe±iϕ

Y20 (ϑ, ϕ) =√

5/4π(

32 cos2 ϑ − 1

2

)Y2±1 (ϑ, ϕ) = ∓√

15/8π sin ϑ cos ϑe±iϕ

Y2±2 (ϑ, ϕ) =√

15/32π sin2 ϑe±2iϕ

. (1.25)

Figure 1.2 shows spherical polar representations of the first few Ylms.

6) We will find that when the spherical symmetry is broken, there willbe a preferred direction in space and l will not be conserved in gen-

Page 26: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 9

Fig. 1.2 Spherical polar plots of the spherical harmonics, |Ylm(ϑ, ϕ)|2,for l = 0, . . . , 3 and m = −l, . . . , l. The z-axis is vertical in these 3Dimages. Note that only Y00 has a nonzero contribution at the origin.The Yl0 and Yll harmonics are oriented along and perpendicular to thez-axis, respectively.

The radial wave function is a solution to the equationd2

dr2 +2r

ddr

+2µ

h2

[E − Ve f f (r; l)

]Rnl (r) = 0, (1.26)

eral. This occurs in diatomic molecules, for example, or when anatom is placed in an external field. However, in both cases the pro-jection of l along the axis of symmetry often will remain conserved.

7) The phases for the spherical harmonics are consistent with Mathe-matical Methods for Physicists by Arfken and Weber [1], The Theory ofAtomic Structure and Spectra by Cowan [2] and MATHEMATICA®.

Page 27: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

Fig. 1.3 Effective single-electron potential for l = 0, 1, 2, 3.

where

Ve f f (r; l) = − Ze2

4πεor+

l (l + 1)r2

h2

2µ(1.27)

is the effective potential. Figure 1.3 shows that the potentials for l = 0 becomerepulsive at the origin. As we will see below, this causes the wavefunction forstates with l = 0 to vanish at the origin. For l = 0 the wavefunctions do nothave to vanish at the origin.

1.2.2Radial Wavefunction and Energy States

It is straightforward to show that the radial wavefunction leads to the sameenergy spectrum predicted by the Bohr model.8 To see this, consider the as-ymptotic behavior of Eq. (1.26). As r → ∞, Ve f f → 0 and Eq. (1.26) reducesto [

d2

dr2 +2µ

h2 E]

R (r) = 0 (1.28)

and has solutions of the form

R (r) = e±ir√

2µE/h2. (1.29)

For positive energy, E > 0, the exponent is imaginary and the electron is de-scribed by unbounded plane waves. These correspond to continuum states.

8) Our treatment follows that found in Bethe and Salpeter, QuantumMechanics of One- and Two-Electron Atoms [3] and Messiah, QuantumMechanics [4].

Page 28: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 11

For bound states we want R → 0 as r → ∞. This requires E < 0 and the minussign in the exponent. The solution to Eq. (1.26) must then take the form

R (r) = f (r) e−r√

−2µE/h2, (1.30)

where f (r) is a function that must be finite at the origin and blows up slower

than exp[

r√−2µE/h2

]as r → ∞. The minus sign was inserted under the

radical to take care of the fact that E < 0. Substituting this expression for R (r)into Eq. (1.26) produces

f ′′ (r) + 2

[1r−

√−2µ

h2 E

]f ′ (r)

+

[2r

(Ze2

4πεo

µ

h2 −√−2µ

h2 E

)− l (l + 1)

r2

]f (r) = 0. (1.31)

The general solution for f (r) will be polynomials,

f (r) = rβ∞

∑ν=0

bνrν. (1.32)

Equation (1.32) can be written in the form of Laplace’s equation with a regularsolution of confluent hypergeometric series that will be introduced in the nextsection. Substituting this back into Eq. (1.31) gives

∑ν=0

[(β + ν) (β + ν + 1)− l (l + 1)] rβ+ν−2

− 2

h2

[h√−2µE (β + ν + 1)− Ze2

4πεoµ

]rβ+ν−1

= 0. (1.33)

In order for Eq. (1.33) to hold for all r, the coefficient of each term must vanish.For the lowest-order term, the rβ−2 term, we must demand that

β (β + 1) = l (l + 1) → β =

l− (l + 1)

(1.34)

if b0 is nonzero. In order that R (r) be finite as r → 0, we must choose β = l.Setting the coefficient of the second term in Eq. (1.33), the rβ−1 term, to zerowith β = l yields

b1 = 2b0

√−2µE/h2 (l + 1)− (

Ze2/4πεo)

µ/h2

(l + 1) (l + 2)− l (l + 1),

Page 29: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

12 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

from which the recursion relation,

bν = 2bν−1

√−2µE/h2 (l + ν)− (

Ze2/4πεo)

µ/h2

(l + ν) (l + ν + 1)− l (l + 1), (1.35)

follows. It is convenient to write bν as

bν = bν−12ζ

ν

ν + l − λ

ν + 2l + 1, (1.36)

where

ζ =

√−2µE

h2 and

λ =Ze2/4πεo√−2µE/h2

µ

h2 .(1.37)

Using Eq. (1.36), we can write bν in terms of b0 as

bν =(

ν

ν + l − λ

ν + 2l + 1

)·(

bν−22ζ

ν − 1ν − 1 + l − λ

ν − 1 + 2l + 1

)(1.38)

=(

ν

ν + l − λ

ν + 2l + 1

)·(

ν − 1ν − 1 + l − λ

ν − 1 + 2l + 1

)· · ·

(b0

11 + l − λ

1 + 2l + 1

)(1.39)

= b0(2ζ)ν

ν!(ν + l − λ)!(2l + 1)!(l − λ)!(ν + 2l + 1)!

, (1.40)

and f (r) becomes

f (r) = rlb0

∑ν=0

(2rζ)ν

ν!(ν + l − λ)!(2l + 1)!(l − λ)!(ν + 2l + 1)!

. (1.41)

We need to consider again the behavior as r → ∞. This can be done by lookingat the asymptotic form of f (r) as ν becomes large. From Eq. (1.41) we write

f (r) ν→∞→ rlb0

∑ν=large

(2rζ)ν

ν!= b0rle2r

√−2µE/h2

,

from which it follows that

R (r) = e−r√

−2µE/h2f (r) → rler

√−2µE/h2

.

This exponential divergence at large r can be avoided if we truncate the powerseries with a judicious choice for the value of λ, namely an integer. The integer

Page 30: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 13

must be chosen to make bν=n−l = 0. Thus, λ = l + 1 + k where k = 0, 1, 2, 3, 4and we find that

Elk = −12

(e2

4πεo

)2 Z2µ

(l + 1 + k)2h2 , (1.42)

from Eq. (1.35), which we see is the energy spectrum for bound states. Thisenergy spectrum is identical to Eq. (1.15) when we recognize n ≡ λ = l + 1 + k,the principal quantum number, and we let me → µ.

1.2.3Exact Radial Solution, Hydrogen-Like Ions

The exact solutions to the radial equation, Eq. (1.26), can be determined fromEq. (1.41) by recognizing that

∑ν=0

(2rζ)ν

ν!(ν + l − λ)!(2l + 1)!(l − λ)!(ν + 2l + 1)!

= F(l + 1 − λ; 2l + 2; 2rζ), (1.43)

where F(l + 1 − λ; 2l + 2; 2rζ) is the confluent hypergeometric function (seeAppendix B.4). Truncating the series at n − l − 1 as discussed in the previoussection leads to the normalized wavefunction expressed as

Rnl (r) =1

(2l + 1)!

√(n + l)!

2n (n − l − 1)!

(2Znao

)3/2

e−Zr

nao

(2Zrnao

)l

× F(

l + 1 − n, 2l + 2;2Zrnao

), (1.44a)

which can also be expressed as an associated Laguerre polynomial (see Ap-pendix B.5)

Rnl (r) = −√

(n − l − 1)!

2n [(n + l)!]3

(2Znao

)3/2

e−Zr

nao

(2Zrnao

)l

× L2l+1n+l

(2Zrnao

). (1.44b)

Page 31: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

14 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

Fig. 1.4 First three radial wavefunctions, Rnl, for hydrogen-like ions.

Explicit expressions for the first four functions are

R10 (r) = 2(

Zao

)3/2e−Zr/ao

R20 (r) = 2(

Z2ao

)3/2 (1 − Zr

2ao

)e−Zr/2ao

R21(r) = 2√3

(Z

2ao

)3/2 (Zr2ao

)e−Zr/2ao

R30(r) = 2(

Z3ao

)3/2[

1 − 2(

Zr3ao

)+ 2

3

(Zr3ao

)2]

e−Zr/3ao

. (1.45)

Additional radial functions can be generated from one of the two expressionsusing a symbolic manipulation program.

The radial wavefunctions are normalized and orthogonal and thus satisfy∫Rnl (r) Rn′ l (r) r2dr = δn′n, (1.46)

where δn′n is the Kronecker delta (see Appendix B). The first three solutionsare shown graphically in Fig. 1.4. We note that (1) Rnl is finite at r = 0 forl = 0, while Rnl → 0 as r → 0 for l = 0, and (2) the number of times thewavefunction crosses the axis (number of zeros) is given by n − l − 1.

The probability density, r2R2nl ,

9 for finding the electron at radius r in an effec-tive potential (given by Eq. (1.27)) is displayed in Fig. 1.5 for a few wavefunc-tions. It is clear from this figure that we are more likely to find the electron

9) We will discuss this probability density at the end of this chapter.

Page 32: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 15

at larger r as the principal quantum number increases. Specifically, the meanvalue of r is given by 〈r〉 = (ao/2Z)

[3n2 − l (l + 1)

]. This and other moments

of r can be determined from

〈rγ〉 =∫

r2+γR2nl (r) dr. (1.47)

A few common moments are10⟨r2⟩ = a2

on2

2Z2

[5n2 + 1 − 3l (l + 1)

]⟨r3⟩ = a3

on2

8Z3 [35n2 (n2 − 1)− 30n2 (l + 2) (l − 1)

+ 3 (l + 2) (l + 1) (l − 1)]

⟨r4⟩ = a4

on4

8Z4 [63n4 − 35n2 (2l2 + 2l − 3)+ 5l (l + 1)

× (3l2 + 3l − 10) + 12]⟨r−1⟩ = Z

aon2

⟨r−2⟩ = Z2

a2on3(l+ 1

2 )⟨r−3⟩ = Z3

a3on3(l+1)(l+ 1

2 )l

⟨r−4⟩ =

Z4[3n2−l(l+1)]a4

o2n5(l+3/2)(l+1)(l+ 12 )l(l− 1

2 )

. (1.48)

Combining our radial and angular solutions, the three-dimensional nonrel-ativistic wavefunction for a one-electron atom can be written as

Ψnlm(r, ϑ, ϕ) = Rnl(r)Ylm(ϑ, ϕ), (1.49)

which is the same as Eq. (1.21). With this definition Ψ is normalized accordingto11 ∫

|Ψnlm(r, ϑ, ϕ)|2r2 sin θdrdϑdϕ = 1. (1.50)

We must remember, however, that even though Ψ depends on n, l and m, theenergy,

En = −12

(e2

4πεo

)2 Z2µ

n2h2 = −hcZ2

n2 RM, (1.51)

10) These expressions are taken from Bockasten [5].11) Caution! Some authors define the radial wavefunction as rRnl(r),

leading to a different normalization from that in Eq. (1.50).

Page 33: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

16 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

Fig. 1.5 Radial probability densities, r2R2nl , for the first three radial

wavefunctions, n, l = 1, 0; 2, 0 and 2, 1, respectively, for hydrogen-likeions displayed in conjunction with the associated effective potentials ofFig. 1.3. The solid curve corresponds to l = 0 and the dashed curve tol = 1.

only depends on n, nonrelativistically. The reader will note that the differencebetween Eqs. (1.15) and (1.51) is that the former describes the energy spectrumfor an infinitely massive nucleus while the latter accounts for the finite massof a particular nucleus; RM and R∞ are given respectively by Eqs. (1.18) and(1.14). This one-electron wavefunction is often called an atomic orbital. We willuse the atomic orbitals to create states for more complex multielectron atomsand molecules in Chapters 2 and 5, respectively.

Probability Density. We are now in a position to put everything together andtalk about how to visualize the electron in an atom. To that end, we recallthat one of the main tenets of quantum mechanics, due to Max Born, is that|Ψnlm(r, ϑ, ϕ)|2 represents the probability per unit volume or the probabilitydensity for measuring the electron with a specific set of physical coordinates.One should not think of the electron, however, as a corpuscular particle actu-ally whizzing about the nucleus as planets do the Sun so that at time t1 it is atpositionr1 and at time t2 it is at positionr2 as one would classically. Rather,Ψ describes the electron as a wave – the electron’s wavefunction – that existsin an abstract space. We can picture this density as a cloud with the electroneverywhere within the cloud simultaneously. Figure 1.6 shows a few repre-sentations of the density for the first few states for the case of Z = 1. WhileΨ contains all the information we can know about the electron, according toquantum mechanics, what we measure in the laboratory is related to |Ψ|2. Forphysical coordinates where |Ψ|2 is large, there is a large probability for find-

Page 34: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.2 Hydrogen-Like Ions, Quantum Approach: Bound States 17

ing the electron. The probability density should be understood in a statisticalsense. When we make a measurement, we project the electron into a physicalspace with classical parameters, position and momentum, but subject to theuncertainty principle.12,13

Several comments can be made concerning Fig. 1.6, which can be deducedfrom Figs. 1.2 and 1.5 as well. In fact, the images in Fig. 1.6 were generated bymapping R2

nl onto the distributions in Fig. 1.2. The first observation we canmake is that the atom increases in size as n increases (〈r〉 scales as n2) and theelectron cloud becomes more diffuse. Only l = 0, s states that are sphericallysymmetric contribute to the electron density near the origin. Finally, we noticethat the contribution to the density from the np, m = ±1 state is about half thatof the np, m = 0 state. This is because each m = ±1 state is actually spreadbetween two orientations. This is more easily seen in rectangular coordinates.To show this, we first recognize that Ψn,l=1,m represents a spherical vector asdiscussed in Appendix E. From the general conversion between spherical andCartesian vectors, Eq. (E.8), it is straightforward to show that

ψnpz = Ψn10, (1.52)

ψnpx =1√2

(Ψn11 + Ψn1,−1) , (1.53)

ψnpy =1

i√

2(Ψn11 − Ψn1,−1) , (1.54)

where ψnpx,y,z are the Cartesian vectors – the wavefunctions in rectangular co-ordinates. Whereas ψnpz = Ψn10, ψnpx and ψnpy are reduced by 1/

√2 while

being linear combinations of Ψn11 and Ψn1,−1 and vice versa. In Cartesiancoordinates, we would view the electron distribution as being double-lobedclouds along the x-, y- and z-axes, each similar to that of the 2pm=0 distribu-tion in Fig. 1.6 (see also Y10 in Fig. 1.2). The Cartesian representation is the oneusually presented in chemistry text books.

1.2.4Energy Units and Atomic States

Generally, there are two conventions for energy levels and several energyunits (see Appendix A) in use. Theorists tend to prefer atomic units (Hartree,see Appendix A) and to set the zero of the energy at the ionization threshold.

12) The wave nature of quantum mechanics leads quite naturally to anuncertainty in our knowledge of the position, ∆x, and momentum,∆p, of the electron or any particle and a limit on the determinationof both simultaneously, ∆x∆p ≥ h/2.

13) It is interesting to note that Schrödinger actually misinterpreted Ψ asthe density distribution of matter, with some regions of space richerin matter than others.

Page 35: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

18 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

Fig. 1.6 The probability per unit volume,the probability density |Ψnlm(r, ϑ, ϕ)|2, forseveral states of the hydrogen atom. Theseare cross-sectional images of the densitieswith the symmetry axis – the z-axis – inthe plane of the paper as indicated by thearrows. The top row represents the prob-ability densities for the first three s states,the middle block all m values for the 2p and3p states and the bottom row all m valuesfor the 3d state. The spatial and intensityscales vary from image to image. Relativeto the 1s image, the ratios between the spa-

tial scales are 1:5:10:5:5:10:10:10:10:10,respectively, for 1s:2s:3s:2p, m=0:2p,m=±1:3p, m=0:3p, m=±1:3d, m=0:3d,m=±1:3d, m=±2. That is, the 1s distribu-tion extends over about 0.1 nm while the 2sdistribution extends over about 0.5 nm. Tomake it visible, each image is multiplied by1:200:2000:67:133:1000:2000:1000:1333:1000respectively from top left to bottom right. Wenote that to make the weaker componentsvisible, the stronger contributions are satu-rated in several of the images.

Bound states then will have negative values and continuum states will havepositive values. When energy levels are determined experimentally, they areoften expressed as wavenumbers,14 as in the energy-level tables provided by

14) It is interesting to note that wavenumbers originated at a time beforecomputers. While we think nothing of multiplying by hc today, itwas very time consuming in the past. Wavenumbers were inventedto aid calculations.

Page 36: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.3 Classification of Nonrelativistic States 19

the National Institute for Standards and Technology (NIST).15 Typically, theenergies of both bound and continuum states have positive values when ex-pressed as wavenumbers. This is due simply to shifting the zero of the scaledown from the ionization threshold to the ground state. To see this, we writethe energy of level n relative to level n′ as

En = En′ − hcZ2RM

(1n2 − 1

n′2

). (1.55)

If we let n′ → ∞, then we have

En = E∞ − hcZ2

n2 RM, (1.56)

the energy of level n relative to the ionization energy IP. That is, we let E∞ =IP instead of 0. Thus, in wavenumbers the energy of level n becomes

νn = IP − Z2

n2 RM. (1.57)

We point out that the fact that such a shift is possible illustrates the importantpoint that only the relative energy is important. Furthermore, it would be pos-sible to use atomic units with the zero set to the ground state or wavenumberswith the zero set at the ionization threshold.

The energies of the first three states of hydrogen and the ionization thresh-old in wavenumbers and atomic units are

n1 2 3 ∞

cm-1 0 82 259.07 97 492.23 109 678.8Energy

a.u. −0.50 −0.12 −0.06 0

1.3Classification of Nonrelativistic States

The full nonrelativistic hydrogen (one-electron atom or ion) wavefunction,ψnlm (r, ϑ, ϕ), has three quantum numbers:

n = the principal quantum number, integers ranging from 1 to ∞ in stepsof 1

l = the (orbital) angular momentum, integers ranging from 0 to n − 1 insteps of 1

15) Other possible units in use are eV (electron volt), Hz (s−1) andkcal/mol (kilocalorie per mole, a unit often used by chemists). Theconversion between these units is given in Appendix A.

Page 37: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

20 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

m = the magnetic quantum number, integers ranging from −l to l in stepsof 1 .

The states are named according to their angular momentum,

l = 0 1 2 3 4 5s p d f g h

The state with n = 3 and l = 2 is called the 3d state. An electron in the 3d stateis also said to be in the 3d subshell, where a shell is designated by nl.16

1.3.1Parity

The wavefunction is said to have positive parity if ψ(−−→r )

= ψ(−→r )

(i.e.,inversion x → −x, y → −y, z → −z) and negative parity if ψ

(−−→r )=

−ψ(−→r )

. In spherical coordinates this translates into r → r, ϑ → π − ϑ

and ϕ → ϕ + π. The parity is thus determined by the angular part of thewavefunction, Ylm (ϑ, ϕ) ∝ Pm

l (cos ϑ)eimϕ. Since

ϑ → π − ϑ → Pml (cos [π − ϑ]) = (−1)l−m Pm

l (cos ϑ)ϕ → ϕ + π → eim(ϕ+π) = (−1)m eimϕ

, (1.58)

Ylm (π − ϑ, ϕ + π) = (−1)l Ylm (ϑ, ϕ) .

Consequently, ψ(−−→r )

= (−1)l ψ(−→r )

implies that

s, d, . . . even parity,p, f , . . . odd parity.

Neither n nor m participate in the determination of parity.

1.3.2Degeneracy

Each energy state in Eq. (1.42) is n2 degenerate, with n2 unique combinationsof values for l and m. The energy, however, does not depend on l nor m.This is because the potential is spherically symmetric so that all the forcesare isotropic. As we mentioned before, if an external field were applied tothe atom the symmetry could be broken and part of the degeneracy would belifted. It should be understood that the l-independent Hamiltonian of Eq. (1.3)is a result of ignoring the spin of the electron. When spin is included, we willsee that part of the l degeneracy will be lifted.

16) Shell and subshell are often used interchangeably. We will reserveshell for states with the same n regardless of l.. As we will discuss inthe next chapter, the 3d designation also refers to an atomic orbital.

Page 38: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.4 Corrections to the Energy Levels 21

1.4Corrections to the Energy Levels

Thus far, we have examined idealized hydrogen-like ions in the absence of ex-ternal fields. These ions were composed of a nonrelativistic spinless electronand a nucleus that we have treated as a spinless point charge. In reality, ofcourse, the nucleus has a finite size (see Problem 1.3), the electron and nucleushave spin and the electron can move with relativistic velocities when near thenucleus. These lead to shifts of the energy levels (Eq. (1.42)) and symmetrybreaking. In this section we will examine three corrections to the energy lev-els for hydrogen-like ions: (1) relativistic motion, (2) the spin of the electronand (3) the spin of the nucleus. The first two will give rise to what is calledthe fine structure and partially splits the l degeneracy, while the latter givesrise to one form of the hyperfine structure. In addition, we will set up theframework for looking at atoms in applied electric and magnetic fields.

1.4.1Relativistic Motion

The relativistic energy of the electron is given by

ER = V +√

p2c2 + m2e c4

V + mec2 +[

p2

2me− p4

8m3e c2

+ · · ·]

︸ ︷︷ ︸kinetic energy

. (1.59)

The first term is the usual Coulomb potential, the second is the rest mass, thethird is the nonrelativistic kinetic energy while the fourth is the first-orderrelativistic correction term to the kinetic energy. If we define

Ho =p2

2me+ V (1.60)

ζR = − p4

8m3e c2

(1.61)

= − 12mec2

[Ho − V

]2, (1.62)

we can apply first-order perturbation theory to ζR (i.e., find the expectationvalue of ζR with the state vectors of Ho) to determine the energy shift, ∆ER.Doing this gives

∆ER =⟨

ψn

∣∣∣ζR

∣∣∣ ψn

⟩= − 1

2mec2

⟨ψn

∣∣∣H2o − 2HoV + V2

∣∣∣ ψn

⟩. (1.63a)

Page 39: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

22 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

But,17

Ho |ψn〉 = En |ψn〉 →⟨∣∣∣H2

o

∣∣∣⟩ = E2n, (1.63b)

while⟨∣∣∣HoV∣∣∣⟩ = −En

Ze2

4πεo

⟨1r

⟩= −

(Z2e2

4πεo

)En

n2ao(1.63c)

and ⟨∣∣∣V2∣∣∣⟩ =

(Ze2

4πεo

)2 ⟨1r2

⟩=

(Z2e2

4πεo

)2 1n3a2

o(l + 12 )

. (1.63d)

The expressions in Eq. (1.48) were used to evaluate the expectation values for1/r and 1/r2. Collecting the terms leads to

∆ER = − E2n

2mec2

1 + 2Z2e2

4πεon2ao

1En

+(

Z2e2

4πεon2ao

)2 n

E2n

(l + 1

2

) . (1.64)

Finally, with the help of Eq. (1.42) the shift can be written as

∆ER = − E2n

2mec2

(4n

l + 12

− 3

)(1.65a)

= −α2En

4

(4n

l + 12

− 3

)Z2

n2 (1.65b)

= −α2RM

4

(4n

l + 12

− 3

)Z4

n4 , (1.65c)

where α is the fine-structure constant (see Appendix A). We see that the shiftdepends on both n and l; thus, the degeneracy between states with the samen but different l is broken.

1.4.1.1 Electron Spin and the Dirac Equation

The electron, a fermion, has half-integer spin and thus obeys Fermi–Dirac sta-tistics. Spin, which plays a major role in atomic and molecular physics, is notincluded in the nonrelativistic Hamiltonian or its correction term (Eq. (1.61)).Instead of inserting a series of terms to account for spin and its interactionwith other angular momenta, we will endeavor to understand the origin of

17) Here, |ψn〉 is an eigenstate of Ho with eigenvalue En. See Appen-dix B.8 for further discussion of the mathematical formalism ofHo|ψn〉 = En|ψn〉.

Page 40: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.4 Corrections to the Energy Levels 23

the spin terms in this section. Spin, as it turns out, is relativistic in origin. Toderive the terms we need, we must start with an equation that is relativisti-cally invariant. We will use the free-particle Dirac equation,[

cα˜ · −→p + β˜mec2]

ψ = Eψ, (1.66)

where α˜ and β˜ are 4 × 4 matrices given by18

α˜ =(

0 σ˜σ˜ 0

)β˜ =

(I˜ 00 −I˜

), (1.67)

σ˜ are the 2 × 2 Pauli spin matrices and I˜ is the identity matrix.19

As before, −→p is our momentum operator (−ih−→∇ ) and the wavefunction, ψ,

is now a four-component column vector,

ψ =

ψ1ψ2ψ3ψ4

, (1.68)

where we can define ψu and ψv as

ψu =(

ψ1ψ2

), ψv =

(ψ3ψ4

), (1.69)

which correspond to the positive and negative energies, respectively. Insert-ing this wavefunction into the Dirac equation leads to two equations that mustbe solved simultaneously,

cσ˜ · −→p ψv +(mec2 − E

)ψu = 0,

cσ˜ · −→p ψu −(mec2 + E

)ψv = 0,

(1.70)

In the nonrelativistic limit, v c, these reduce to the Schrödinger equation.From the second equation in Eq. (1.70), we have

ψv =cσ˜ · −→p

mec2 + Eψu. (1.71)

18) We will use a tilde under a symbol to indicate a matrix.19) The Pauli spin matrices (see, for example, Eq. (4.73) of Ref. [6]) can

be represented as

σ˜(x) =(

0 11 0

), σ˜(y) =

(0 −ii 0

), σ˜(z) =

(1 00 −1

), I˜=

(1 00 1

).

Page 41: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

24 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

When the velocities are small∣∣−→p ∣∣ → mev and E → mc2, which implies

|ψv| ≈ v2c

|ψu| |ψu| . (1.72)

Thus, most of the physics will be represented by the positive-energy wave-functions at nonrelativistic velocities. Substituting Eq. (1.71) into the upperexpression in Eq. (1.70) to eliminate ψv gives

c2(

σ˜ · −→p)2

ψu =(

mec2 + E) (

E − mec2)

ψu (1.73a)

= E2 − m2e c4ψu (1.73b)

≈ 2mec2Enrψu, (1.73c)

where Enr = E − mec2. Using the spinor identity(σ˜ · A

) (σ˜ · B

)= A · B + iσ˜ ·

(A × B

), (1.74)

we can write

c2 p2ψu = 2mec2Enrψu, (1.75)

which is just the free-particle Schrödinger equation.To account for interactions with the Coulomb potential of the nucleus and

external magnetic and time-varying electric fields, we introduce two addi-tional terms into the free-particle equation (Eq. (1.70)). We incorporate theexternal fields with the vector potential, A, by letting

−→p → −→p − ec−→A ≡ −→π , (1.76)

where the electric and magnetic fields are given by

E = −−→∇φ − ∂A∂t

, (1.77a)

B =−→∇ × A. (1.77b)

A static potential, −eφ, was added to Eq. (1.77a) to account for the nuclearcharge. In Chapter 4, π will lead to the radiation Hamiltonian used in pho-toexcitation. The modified Dirac equation takes the form

cσ˜ ·(−→p − e

cA)

ψv =(E − mec2 − eφ

)ψu,

cσ˜ ·(−→p − e

cA)

ψu =(E + mec2 − eφ

)ψv.

(1.78)

Page 42: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.4 Corrections to the Energy Levels 25

Following steps similar to those taken above, the lower expression in Eq. (1.71)becomes

ψv =cσ˜ ·

(−→p − ecA)

E + mec2 − eφψu. (1.79)

Substituting this into the upper expression and expanding the denominator,

1E + mec2 − eφ

≈ 12mec2

[1 − Enr − eφ

2mec2 + · · ·]

(1.80)

(where terms of order v2/c2 have been kept), gives

12me

[(σ˜ · π)2 − (

σ˜ · π) Enr − eφ

2mec2

(σ˜ · π)]

ψu = (Enr − eφ) ψu. (1.81)

We will solve this equation by allowing ψ = ξψu, where

ξ = 1 +(σ˜ · π)2

8m2e c2 . (1.82)

To first order in v2/c2,

ξ−1 ≈ 1 −(σ˜ · π)2

8m2e c2 . (1.83)

Making this substitution for ψu and multiplying the resulting equation on theleft by ξ−1 gives 1

2me

(σ˜ · π

)2 −(

σ˜ · π)4

8m2e c2 +

(σ˜ · π

)2

8m2e c2 (Enr − eφ)

− 14m2

e c2

(σ˜ · π

)(Enr − eφ)

(σ˜ · π

)+ (Enr − eφ)

(σ˜ · π

)2

8m2e c2

ψ

= (Enr − eφ) ψ. (1.84)

Recognizing that

(σ˜ · π)2 =(−→p − e

cA)2 − eh

c σ˜ · ∇× A,

(σ˜ · π) (Enr − eφ) − (Enr − eφ) (σ˜ · π) = ieh(

σ˜ · ∇φ)

,

(σ˜ · π)(

σ˜ · ∇φ)

−(

σ˜ · ∇φ)

(σ˜ · π) = −ih∇ · ∇φ + 2iσ˜ · π × ∇φ,

(1.85)

Page 43: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

26 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

we can rewrite Eq. (1.84) as[1

2me

(−→p − ecA)2 − eh

2mecσ˜ · ∇× A −

−→p 4

8m3e c3

+eh2

8m2e c2

∇ · ∇φ − eh4m2

e c2 σ˜ · ∇φ ×−→p]

ψ = (Enr − eφ) ψ. (1.86)

Again, only terms of order v2/c2 have been kept. The quantity in parentheseson the left- and the right-hand sides comprises the nonrelativistic Schrödingerequation with an interaction term represented by −e

−→A /c. The rest of the

terms all have a relativistic origin and are responsible for spin-1/2 particleinteraction with magnetic fields as well as spin–orbit interaction. A brief de-scription of each term along with its approximate magnitude is given in Ta-ble 1.1.

The second perturbation in Table 1.1 allows us to define the Bohr magneton,µB = eh/2mec while the fifth, the spin–orbit interaction, further splits thedegeneracy between states of different l. Before we calculate the magnitudeof this splitting, we will say a few words about spin.

1.4.1.2 Classification of Relativistic Hydrogen States

The spin quantum number, s, assumes only one value, 1/2. Similar to theorbital angular momentum, the eigenvalues of −→s 2 will be s (s + 1) = 3/4.There is also a magnetic or z component of s, sz, that will have eigenvaluesms = ±1/2. We will define a new operator called the total angular momen-tum,

−→j =

−→l +−→s , (1.87)

with eigenvalues (of−→j 2) j (j + 1). The sum in Eq. (1.87) is a vector sum. The

values that−→j can assume are

j =

l ± 12 , l = 0,

12 , l = 0.

(1.88)

The magnetic quantum number or the eigenvalue of the z component of−→j is

mj, with 2j + 1 values, mj = −j,−j + 1, . . . , j.In Section 1.3 we introduced the concept of a shell, designated by nl. Each

l will have two values of j associated with it when l > 0. For example, anelectron in a state of n = 4 and l = 2 will have j = 3/2, 5/2. Thus, weintroduce the concept of a subshell designated by nlj. This electron is said tobe in the 4d shell and will be in either the 4d3/2 or 4d5/2 subshell. Note thatnow we have two angular momenta about which to be concerned, l and j. To

Page 44: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.4 Corrections to the Energy Levels 27

Tab. 1.1 Summary of relativistic and spin-dependent interaction terms.

Perturbation Description Magnitude

12me

(−→p − ecA)2

The nonrelativistic motion, p2/2me, and the inter-

action terms e2mec

A · ∇ + e2

2mec2A · A. The lat-

ter terms are responsible for absorption and emis-

sion; they can be written as −e−→r · −→E (see Section4.1.1).

105 cm−1

eh2mec

σ˜ · ∇× A The spin-1/2 (−→s = σ˜/2) interaction with a mag-

netic field (−→B ≡ −→∇ × −→

A ). This term gives thecorrect g-factor, g = 2, and magnetic moment ofthe electron, µ = eh

2mec g−→s .

≈ 1 cm−1

−→p 4

8m3e c3

Relativistic mass correction term. ≈ 0.1 cm−1

eh2

8m2e c2

∇ · ∇φ The “Darwin” term responsible for s-state shifts.It represents the relativistic nonlocalizability of theelectron and is related to both the negative-energysea and its rapid motion.

< 0.1 cm−1

eh4m2

e c2σ˜ · ∇φ ×−→p Spin–orbit interaction. As shown in Problem 1.4,

this term can be written as eh2

2m2e c2

1r

dφdr

(−→l · −→s

),

where−→l is the orbital angular momentum. In con-

trast to the previous term, this does not affect sstates.

10–103 cm−1

distinguish their magnetic quantum numbers, we will add subscripts l and jto the orbital, ml , and total, mj, quantum numbers.

1.4.1.3 Hydrogen-Like Ion Wavefunction Including Spin

We are now in a position to write down the three-dimensional wavefunctionfor a one-electron atom including the spin of the electron. This wavefunctiontakes the form

Ψnlmlms = Rnl(r)Ylml(ϑ, ϕ)φms(sz), (1.89)

where we have placed the subscripts l and s on the magnetic quantum num-bers to distinguish orbital from spin. Since this wavefunction describes thestate of a single electron, we only need to specify the direction of the spin.

Page 45: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

28 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

1.4.2Fine Structure and Spin–Orbit Interaction

In general, subshells have different energies. The splitting between the statesis determined in part by the spin–orbit interaction

Hso =eh2

2m2e c2

1r

dr

(−→l · −→s

). (1.90)

This interaction causes a level to shift by ∆Eso =⟨

Hso

⟩.20 To calculate

⟨Hso

⟩,

it is convenient to write−→l · −→s as21

−→l · −→s =

j2 −l2 −s2

2. (1.91)

The energy shift then becomes

∆Eso =eh2

2m2e c2

[j (j + 1)− l (l + 1)− s (s + 1)

2

]⟨1r

dr

⟩. (1.92)

But,

1r

dr=

Ze4πεor3 . (1.93)

Using the expressions in Eq. (1.48), it is possible to write

∆Eso =Z4e2 h2

4πεo

h2

2m2c2a3on3

j (j + 1)− l (l + 1)− s (s + 1)

2l(

l + 12

)(l + 1)

. (1.94)

Combining this with the relativistic energy correction (Eq. (1.65a)), we obtain

∆Enlj = ∆ER + ∆Eso = α2RM

(3

4n− 1

j + 12

)Z4

n3 . (1.95)

At the maximum value of j + 1/2 = n, ∆Enlj is always < 0, which meansthat the combination of the relativistic and spin–orbit interactions binds thesystem more tightly as shown for a few levels in Fig. 1.7. The figure shows aswell that the degeneracy has not been lifted completely. States of different l

20) We saw in Eq. (1.65a) that the relativistic correction shifted levels ofdifferent l by different amounts. Below, we will combine the spin–orbit with the relativistic correction to remove the degeneracy of ashell partially.

21) This can be shown using the algebra developed in Appendix E; seeEq. (E.29) and surrounding discussion.

Page 46: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

1.4 Corrections to the Energy Levels 29

2

2

3MZ R

4/3C

2/1,2/3

2/3,2/54/C

0l 1l 2l

2/1j 2/1j

2/3j 2/3j

12/C2/3j

Fig. 1.7 Fine structure of hydrogen-like ions in wavenumbers for n = 3where C = α2Z4RM/n3 and ∆j+1,j = ∆Enl j+1 − ∆Enl j = C/l (l + 1) .The levels are all shifted below –Z2RM/n2, the nonrelativistic energy.

but with the same j have the same energy. The splitting between j = l ± 1/2levels will be given by

∆j+1,j =α2Z4RM

n31

l(l + 1). (1.96)

Example. For j = 1/2 and 3/2 in hydrogen, Eq. (1.95) gives ∆j+1,j = 0.36,0.12, 0.044 cm−1, respectively, for n = 2, 3, 4.

1.4.3Rydberg Series

Figure 1.1 shows a progression of energy levels to the ionization limit asn → ∞. To lowest order, the energy of each level in the series is given byEq. (1.56). For each level n, there are n nearly degenerate l sublevels. For allbut s (l = 0) states, there are two possible j (= l ± s) values for each l andthus two series for each l. Consequently, there are an infinite number of seriesof levels marching their way toward the ionization limit. Each series, nlj forfixed l and j, is called a Rydberg series. For hydrogen-like systems (ions withjust one electron) these series have a simple behavior. Systems with two ormore electrons behave very differently because the members of the series caninteract in very profound ways leading to a variety of interesting phenomenasuch as autoionization. We will look at Rydberg series for multielectron atomsin Chapter 2.

Page 47: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

30 1 Hydrogen-Like Ion: An Atom (Ion) With One Electron

1.5Continuum States

Thus far, we have focused on bound states, which correspond to solutions tothe radial equation (Eq. (1.26)) when E < 0. However, when E > 0 the elec-tron energy lies above the n = ∞ level in Fig. 1.1. The asymptotic (r → ∞)

solutions for the positive-energy states take the form R (r) = e±ir√

2µE/h2

(Eq. (1.29)), which oscillates sinusoidally and does not tend to zero. The radialsolution for these states can be obtained by following the steps we took for thenegative-energy states up to the recursion relation in Eq. (1.35), which becomecomplex when E > 0. Since the wavefunctions are finite at r = ∞ (Eq. (1.29)is bounded by 1), there is now no need to truncate the series. As a result, thestate of the electron can assume all positive energies; hence, the name contin-uum state. Since Eq. (1.35) is still valid, the wavefunctions can be written interms of hypergeometric functions; however, the principal quantum number,n, becomes imaginary in Eq. (1.37),

n ≡ λ = −iZe2/4πεo√

2µE/h2

µ

h2 . (1.97)

The solution, Eq. (4.23) of Bethe and Salpeter, can be written as

Rεl(r) = (−1)l+12√

Ze2

4πεo√1 − e−2πn′

×l

∏ν=1

√ν2 − n′2 (2kr)l

(2l + 1)!e−ikrF(in′ + l + 1, 2l + 2, 2ikr), (1.98)

where n′ = in is real, and k =√

2µE/h2 also real with dimensions of 1/r.In general, each nl Rydberg series in the bound region is associated with

a series of the same-parity energy states in the continuum region. The con-tinuum states are designated by the label εl. An nl Rydberg series plus its εlcontinuum is called a channel. The channel is said to be closed when E < 0 (theelectron is bound to the nucleus) and open when E > 0 (the electron is free).

Page 48: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Further Reading 31

Further Reading

1 H. A. Bethe and E. E. Salpeter, QuantumMechanics of One- and Two-Electron Atoms,Plenum, New York, NY, 1977.

2 A. Messiah, Quantum Mechanics, DoverPublications, 2000. ISBN: 0486409244.

3 R. D. Cowan, The Theory of Atomic Structureand Spectra, University of California Press,Los Angeles, CA, 1981.

4 I. I. Sobelman, Atomic Spectra and RadiativeTransitions, Springer Series in ChemicalPhysics, Springer, New York, NY, 1979.

5 J. Mathews and R. L. Walker, MathematicalMethods of Physics, Benjamin Cummings,San Francisco, CA, 1970. ISBN 0-8053-7002-1, Second edition.B

6 N. N. Lebedev, Special Functions and TheirApplications, translated and edited by R. A.Silverman, Dover, New York, NY, 1972.

7 J. J. Sakurai, Advanced Quantum Mechanics,Addison-Wesley, Menlo Park, CA, 1973.

8 L. I. Schiff, Quantum Mechanics, McGraw-Hill, San Francisco, CA, 1968.

9 F. Yang and J. H. Hamilton, Modern Atomicand Nuclear Physics, McGraw-Hill, SanFrancisco, CA, 1996.

10 J. D. Bjorken and S. D. Drell, RelativisticQuantum Mechanics, McGraw-Hill, SanFrancisco, CA, 1964.

11 For additional problems with worked so-lutions see D. Budker, D. F. Kimball and D.P. DeMille, Atomic Physics: An ExplorationThrough Problems and Solutions, Oxford Uni-versity Press, New York, NY, 2004.

Problems

1.1 Find the lowest-energy state of the one-dimensional hydrogen atomwhere the Schrödinger equation takes the form

− h2

d2

dx2 ψ(x)−(

14πεo

Ze2

|x| − E)

ψ(x) = 0 (−∞ < x < ∞). (1.99)

1.2 Calculate

〈rγ〉 =∫

r2+γR2nl (r) dr (1.100)

for γ = −3,−2,−1, 1, 2, 3.

1.3 Since the wavefunctions of the s levels are finite at the origin, the potentialfor s electrons changes when the electron penetrates the nucleus and gives riseto a shift in the energy levels relative to Eq. (1.42). If the radius of the nucleusis ro, the potential will be given by

Q|r| when r > ro,

∫N

ρ (r′)|r −r′|d

3r′ when r < ro,

(1.101)

Page 49: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

32 Further Reading

where ρ is the charge density of the nucleus. Use first-order perturbationtheory to show that the s states are shifted by

(∆Es)finite size =23

e2

4πεo

⟨r2⟩

N

a3o

, (1.102)

where⟨r2⟩

N=

∫N

r′2ρ(r′

)d3r′. (1.103)

If you take the size of the nucleus to be about 8 fm, estimate the magnitude ofthis shift for the ground state, i.e., calculate (∆Es)finite size /E1s.

1.4 Show that

eh4m2

e c2 σ · ∇φ ×−→p (1.104)

in Table 1.1 can be rewritten as

eh2

2m2e c2

1r

dr

(−→L · −→S

). (1.105)

1.5 It is possible to derive the spin–orbit interaction by going to the rest frameof the electron and allowing the nucleus to orbit the electron. The electronthen feels a magnetic field due to the circular current. Determine the expres-sion for the Hamiltonian, −−→µ · −→B , that describes the interaction.

1.6 Sketch the probability densities similar to Fig. 1.6 for the 1s through 3dstates of hydrogen in Cartesian coordinates. You might find it helpful firstto express the spherical harmonics in terms of Cartesian coordinates. That is,relate the spherical components (l, m) to combinations of x, y and z.

Page 50: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

33

2The Structure of the Multielectron Atom

Thus far, we have restricted our investigation to atomic systems with a singleelectron. We will now turn our attention to atoms with more than one electron.We will focus primarily on one-electron and two-electron atoms – atoms withone and two electrons outside closed shells, respectively.

2.1Overview

We begin by introducing the nonrelativistic multielectron Hamiltonian,

Ho = ∑i

[p2

i2me

− Ze2

4πεori

]+

12 ∑

i =j

Ze2

4πεorij+ ∑

iζ (ri)

(−→l i · −→s i

). (2.1)

The first sum is the single-particle Hamiltonian, the second is the mutual in-teraction between electrons, also known as the electrostatic interaction, andthe last is the spin–orbit interaction. It is the second sum that distinguishesthe Hamiltonians for multielectron systems from those of hydrogen-like ions.Since the Hamiltonian involves more than two particles, we will not be ableto generate an exact solution to Schrödinger’s equation but must resort to ap-proximations. The simplest approximation that captures much of the impor-tant physics is known as the central field approximation. We will first look at theproblem without the last term. At the end we will put it back in as a pertur-bation.

This single-particle approximation assumes that the interaction betweenelectrons can be represented as an average central potential that can be addedto the Coulomb potential of the nucleus to give an effective potential U (r).The Schrödinger equation,

∇2ψ (r) +2µ

h2 [E − U (r)] ψ (r) = 0, (2.2)

takes a form similar to Eq. (1.19) with −Ze2/4πεor replaced by U (r). Assum-ing that ψ (r) is again separable into radial and angular portions (Eq. (1.21)),

Page 51: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

34 2 The Structure of the Multielectron Atom

the radial wavefunction is then a solution to

1r2

ddr

(r2 d

drRnl (r)

)+ [E − Ul (r)] Rnl (r) = 0, (2.3)

where

Ul (r) = U (r) +h2

l (l + 1)r2 . (2.4)

The solutions Rnl (r), from which we can describe the structure and dynam-ics of multielectron systems depend on the exact expression for U (r). Beforelooking at details, we will state a few general features of multielectron sys-tems.

1. Limiting values for U (r) when E < 0:

U (r) → V (r) = −Ze2/4πεor as r → 0

U (r) → 0 as r → ∞

The first is true because as r → 0, all other electrons will have largerorbits and the electron will sense only the nucleus. The second must betrue to have bound states.

2. In general, the system will be uniquely determined by E, l, ml . Forhydrogen-like ions we saw in the nonrelativistic approximation thatthere will be 2l + 1 degenerate states. In multielectron systems partof this degeneracy is lifted. We will discuss this in more detail in (4)below.

3. The states of atomic systems consist of shells and subshells, which arecomposed of collections of atomic orbitals (one-electron wavefunctions,Eq. (1.49)).

• An atomic orbital is designated as nl, e.g., 1s, 2p, etc., where s and pindicate l = 1, 2, respectively.

• A subshell is very similar to an orbital, the difference being that onedesignates the number of electrons contained within. We wouldcall a 2s orbital with two electrons the 2s2 subshell. In general,a subshell with # electrons would be designated by nl#, wherethe maximum value # can assume is given by (2s + 1) (2l + 1) =2 (2l + 1). The maximum number of electrons per subshell for

Page 52: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.1 Overview 35

l = 0–3 is

l Designation Maximum number0 s 21 p 62 d 103 f 14

• A shell is composed of all the states for a given n. For example, thereare two subshells (2s and 2p) that form the n = 2 shell. One willoften find shells referred to by following X-ray absorption names:

K shell n = 1 l = 0

L shell n = 2 l = 0, 1

M shell n = 3 l = 0, 1, 2

etc.

• A shell or subshell is said to be closed when it contains its maximumnumber of electrons. A closed subshell bears the special property ofbeing spherically symmetric, i.e.,J = 0.1 It should be understoodthat when shell is used in the literature, it can be in reference towhat we have defined as a subshell or to what we have definedas a shell. Consequently, a closed shell can mean a shell with itsmaximum number of electrons, e.g., 2s22p6, or a subshell with itsmaximum number of electrons, e.g., 2s2.

• A configuration describes the states of all the electrons in an atom.For example, the ground-state configuration for carbon, which hassix electrons, is 1s22s22p2. In general, a configuration will lead toseveral different terms (which we define next). With the exceptionof a closed shell or subshell, however, one should view the place-ment of electrons into specific configurations as only an approxi-mation to the truth. Configurations should be used primarily toaid bookkeeping. Although there are cases where a pure configu-ration can be a reasonable approximation for defining the state ofan atom, in general, including ground states of atoms, admixtures

1) When an orbital contains its maximum number of electrons, thespherical symmetry is due to the fact thatJ = ∑i

ji = 0 orL = ∑i

li = 0 andS = ∑isi = 0 so thatL +S = J = 0, for ex-ample, depending on the appropriate coupling scheme (discussed inSection 2.2).

Page 53: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

36 2 The Structure of the Multielectron Atom

of several configurations are required to determine the energy of aparticular state of an atom adequately.

• A term designates a particular grouping or coupling of all the elec-trons in an atom. For example, the electrons in the lowest subshellsof carbon (1s22s22p2) couple together to give several LS terms,3P0,1,2, 1D2 and 1S0 (see Section 2.2 for these LS designations), eachhaving a different energy. A complete designation of the term re-quires the configuration (nl) to be specified as well. Since closedsubshells produce spherical charge distributions, it is customary togive only the highest or all open subshells. For our carbon examplethis would be 2p2 3P. In some cases, particularly for heavier atoms,other coupling schemes are more appropriate (see Section 2.2) lead-ing to multiple jj terms, for example. Ground-state configurationsare listed in Appendix C.4.

4. In many-electron systems, the electrons typically fall into one of twogroups – outer or valence electrons and inner electrons, which form whatwe refer to as the core electrons or the core.

• Valence electrons. The outer electrons determine the ground stateand the bound excited electronic states of an atom. They are re-sponsible for the chemical behavior of atoms in general. Valenceelectrons typically respond to low-frequency light, frequencies as-sociated with energies of about 1 to 10 eV (near-infrared to ultra-violet wavelengths). The wavefunction of the valence electrons ex-tends from the nucleus to the outer edges of the atom, in general.While outside the core, the valence electron in neutral atoms experi-ences a Coulomb attraction to the nucleus with an effective chargeof one electric charge. Inside the core, the nuclear charge is onlypartially screened and the effective nuclear charge increases alongwith stronger attraction to the nucleus. At the same time, the va-lence electrons interact with the core electrons more strongly wheninside the core. The net result is that the valence electrons are moretightly bound than an equivalent electron (one with the same n) inthe hydrogen atom. Consider Na, for example. The binding energyof a 3s electron in hydrogen, according to Eq. (1.15), is about 1.5 eV,whereas a bit more than 5.1 eV is required to remove the ground-state valence electron from Na. Figure 2.1 compares the hydrogenenergy levels with those of several alkali atoms.In Section 1.2.3 we learned that the wavefunctions for s states pen-etrate the nucleus while those for states with l = 0 are excluded,with those of higher l being pushed further away. This causes low-l states to spend more time near the nucleus than high-l states and,

Page 54: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.1 Overview 37

Fig. 2.1 A few observed energy levels of alkali atoms (Li–Cs) vs thoseof hydrogen.

thus, experience more interactions with the core electrons and astronger attraction to the nucleus. At the same time, the averageradius for the electron for hydrogen-like ions is given by

〈r〉 =32

n2

Zao. (2.5)

Thus, we expect the shift from the hydrogen position to decrease asn and l increase as a consequence of the electron spending less timein the core. The fact that the shift is larger for s states (see Fig. 2.1)is just a reflection of the simple fact that s states penetrate the coremost deeply.

• Effective Quantum Number. Even though many-electron atoms areinherently more complicated than hydrogen-like ions, it turns outthat we can specify the location of bound levels in Fig. 2.1, at least tolowest order, with a simple equation similar to Eq. (1.51) or (1.57).This is possible because we can account for the average effect ofthe core and other valence electrons with one parameter we call thequantum defect, µ. If we define an effective quantum number,2

n∗ = n − µ, (2.6)

2) The effective quantum number is sometimes given the label ν in-stead of n∗, which is not to be confused with the energy of the statein Eq. (2.7), νn, or the frequency of the light associated with transi-tions between levels.

Page 55: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

38 2 The Structure of the Multielectron Atom

2 4 6 8 10 12n

2

4

6

8

10

eulaV

2 3 4 5 6n

1

2

3

4

5eula

V Li NaK

Rb

Cs

Alkali Ground ns Levels

Na ns (solid) and np Levels

Fig. 2.2 Effective quantum numbers (squares) and the associatedquantum defects (diamonds) for neutral alkali atoms (upper) and alongthe ns (filled) and np Rydberg sequences for neutral Na. Data forthese plots were taken from the NIST Atomic Spectra Database [7,8].

the energy of a particular level can be determined from Eq. (1.57)when we replace the principal quantum number with n∗,

νn = IP − Z2

n∗2RM. (2.7)

Equation (2.7) leads to Rydberg series for multiple electrons as well.Figure 2.2 shows how n∗ and µ depend on n and l. In the upperplot, one sees that n∗ is nearly constant (changes by less than 20%)for the ground states of Li through Cs. At the same time, µ changesby one unit for each step down the column. This change, however,is due to the fact that n changes by one unit along the sequence. Inthe lower plot we see that n∗ increases with n for both s and p states,

Page 56: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.1 Overview 39

while µ is nearly constant in both cases. In addition, we see that µ

is different for s and p states, as we surmised above. These obser-vations are associated with the fact that the physics is embodied inthe noninteger portion of n∗ or µ.

5. What we learned in the preceding paragraphs under (4) can be usedto explain the electronic structure of the ground states of the atoms inthe periodic table. In general, the energy increases with the sum n + l. Astate of n, l + 2 can be above a state of n + 1, l, however. For example, the4p state sits at a higher energy than the 3d state, which sits higher thanthe 4s state. As a consequence, in going from K to Kr along the fourthrow of the periodic table, the 4s subshell fills first followed by the 3d andfinally the 4p subshell. A similar behavior is seen in the fifth, sixth andseventh rows of the periodic table as shown in tables in Appendix C.

6. The degeneracy of the energy within a shell is partially lifted not onlyby the spin–orbit interaction, the last sum in the many-electron Hamil-tonian, Eq. (2.1), but also by the electrostatic interaction, the second sumin Eq. (2.1). This interaction is also responsible for purely quantum ef-fects associated with correlated motion between the electrons as well asthe fact that the electrons obey Fermi–Dirac statistics (Pauli exclusionprinciple).

7. The electronic state of the atom is defined not in terms of the individualorbital and spin angular momenta but by what is called the total angu-lar momentum,

−→J . The total angular momentum defines the collective

angular momentum of the system and is the result of the coupling ofthe individual orbital and spin angular momenta in one of a variety ofpossible ways. At one extreme,

−→J is the vector sum of the total or-

bital and spin angular momenta of the system,−→L and

−→S , respectively,

where−→L = ∑i

−→l i and

−→S = ∑i

−→s i, the vector sums of individual or-bital and spin angular momenta. The individual total angular momenta−→j i (=

−→l i +

−→s i for the ith electron in the system) are not good quantumnumbers in this case. At the other extreme, the total angular momentumis given by

−→J = ∑i

−→j i. In this case,

−→L and

−→S are not good quantum

numbers. In these as well as other cases discussed in Section 2.2,−→J is a

good quantum number. In the absence of an applied external field, therewill be 2J + 1 degenerate sublevels (MJ) for each J.

8. When the spin of the nucleus, I, is included, the grand total angularmomentum for the atom will be F = I + J. The interaction between Iand J tends to be weak, so that J remains a good quantum number. Forthe hyperfine case, there are 2F + 1 degenerate sublevels (MF) for each F.

Page 57: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

40 2 The Structure of the Multielectron Atom

2.2Angular Momentum Coupling Schemes

The presence of more than one electron necessitates understanding how thevarious angular momenta (orbital and spin) couple to give a total angularmomentum for the atom. In general, the coupling scheme depends on severalfeatures. However, the relative magnitudes of the last two sums in the many-electron Hamiltonian in Eq. (2.1) set two important limits. These two termsdescribe two distinct interactions between the electrons: (i) electrostatic and(ii) spin–orbit. When either (i) or (ii) clearly dominates, the electrons coupleaccording to the LS-coupling scheme, (i) > (ii), or the jj-coupling scheme, (ii)> (i). The ground state of an atom is usually well represented by LS couplingwhile jj coupling describes systems where one electron is highly excited, forexample in a Rydberg state. When neither dominates, an intermediate cou-pling scheme must be employed, which can occur for low-lying excited statesof heavy systems.

2.2.1LS or Russell–Saunders Coupling

When the second sum in Eq. (2.1) (the electrostatic interaction) is larger thanthe last (the spin–orbit interaction of either electron), LS coupling is a goodapproximation. While the name might imply that there is a strong couplingbetween

−→l and −→s , this is not the case! Quite the contrary, in this scheme

the orbital angular momenta of the individual electrons couple to give a total−→L and the individual spins couple to give a total

−→S . The name originates

from the coupling of the total orbital angular momentum,−→L , to the total

spin,−→S , to give

−→J . The maximum and minimum values of J are

∣∣∣−→L ∣∣∣ +∣∣∣−→S ∣∣∣

and∣∣∣−→L −−→

S∣∣∣, respectively. The electrostatic interaction usually dominates in

low-lying states of light atoms where the electron(s) in the outer shell interactstrongly with electrons in the inner shells known as the core. This couplingscheme is often appropriate for the ground states of atoms as well.

As an example of LS coupling, consider a closed-shell system such as He,with a ground-state configuration of 1s2. The values of L, S and J are all zero.This is true for L because both electrons have l = 0. This is true for S becausethere are only two possible values for ms, ±1/2, commonly referred to as spinup (+1/2 or ↑) and spin down (−1/2 or ↓). Since no two fermions can be in thesame state at the same time (Pauli’s exclusion principle), one electron must bein the ms = +1/2 state while the other will be in the ms = −1/2 state. Theonly value MS can then have is zero, which means that S = 0. Since L and Sare zero J must be zero as well.

Page 58: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.2 Angular Momentum Coupling Schemes 41

To understand the general case, consider two electrons with quantum num-bers ζ1 and ζ2. We can form a joint state of the system by multiplying thewavefunctions of the two electrons:

Ψ = ψ1(ζ1)ψ(ζ2),

where 1 and 2 refer to the electrons 1 and 2. However, since the electrons areidentical particles, we have to take into account the fact that if we exchangethe two, we will not be able to tell. For example, the wavefunction will besymmetric upon exchange of electrons if both electrons have spin up (↑↑) orspin down (↓↓). However, the wavefunction will be antisymmetric upon ex-change of electrons if one electron has spin up while the other has spin down.Thus, we end up with two types of states:

Ψs = ψ1(ζ1)ψ(ζ2) + ψ1(ζ2)ψ(ζ1) (triplet state)Ψa = ψ1(ζ1)ψ(ζ2) − ψ1(ζ2)ψ(ζ1) (singlet state)

In general, these will be two different states. Typically, LS states with differentL, S and J will have different energies. We label the states in the LS-couplingscheme as

nl# 2S+1LJ or nl# 2S+1LoJ ,

where nl# is the last subshell to be filled and # is the number of electrons inthe subshell. If there are two open subshells, the second is often indicatedby adding n′l′#′ to the expression. In the labeling scheme 2S + 1 is called themultiplicity of the state (see Section 2.3) and the superscript o designates anodd-parity state (see Section 1.3.1) determined by

ψ(−−→r )

= (−1)∑i

∣∣∣−→l i

∣∣∣ψ

(−→r ), (2.8)

where the sum is over all the electrons in open subshells. In analogy to thenames of the subshells, a state of total L is indicated by

L = 0 1 2 3 4S P D F G

but with upper case letters. As mentioned earlier, these are usually referred toas LS terms.

Returning to our He example with a 1s2 configuration, since an s shell haseven parity, the LS-coupled state formed by this configuration also has evenparity. The term representing the state (the ground state in this case) is 1s21S0.The superscript 1 preceding S indicates that this is a singlet. In general, closedshells or subshells are spherically symmetric (S states), have even parity (an

Page 59: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

42 2 The Structure of the Multielectron Atom

even number of electrons) and zero spin. These states are then “singlet S zerostates.”

Within each term there will be one or more states of specific J and its pro-jection along the z-axis, MJ . We will represent these states by a state vector,3∣∣γ;

(limli

, . . .)

LML (simsi , . . .) SMS; JMJ⟩

,

where γ represents the other parameters necessary to define the state such asposition, n, etc. Within the state vector it is understood that the items insidethe parentheses are coupled first to give L and S before L and S are coupled togive J. There is a lot of freedom in writing this vector. It can be written morecompactly as∣∣γ; (li, . . .) L (si, . . .) S; JMJ

⟩or

∣∣γ; LML, SMS; JMJ⟩

. (2.9)

Often, the right-hand vector in Eq. (2.9) is written without ML and MS, whilethe left-hand vector is written without the subscript on MJ .

As another example, we will consider the open-shell system of sodium. Itsground-state configuration is 1s22s22p63s. The first 10 electrons are in closedshells and thus can be designated by 1S0. The last electron is in the 3s subshellwith s = 1/2 and l = 0. Thus, L = 0, S = 1/2 (there is only one electron) andJ = 1/2. We can designate the ground state of Na as

1s2(

1S0

)2s2

(1S0

)2p6

(1S0

)3s 2S1/2, (2.10)

or more compactly as 3s2 S1/2. This term is called a “doublet S one-half state.”The state vector takes the form |00, 1/2± 1/2; 1/2± 1/2〉. The 1S0 term in Eq.(2.10) contributes nothing to the final designation because S = J = 0, i.e., thetotal spin and total angular momentum are zero.

If the 3s electron were promoted to the 3p subshell, the total angular mo-mentum could have two values, L + S = 3/2 and L − S = 1/2. As a result,there are two final states with different J values, 3p2 P3/2 and 3p2 P1/2. Thestate vectors for these two cases could be written as |1ML, 1/2 ± 1/2; 3/2M〉and |1ML, 1/2 ± 1/2; 1/2 ± 1/2〉. These states are called “doublet P three-halves” and “doublet P one-half” states, respectively.

The meaning of the multiplicity is now clear. It is the number of differentJ states possible for a specific value of L and not the number of possible MJstates as Eq. (2.10) might imply. For S states (L = 0) there is only one J valuepossible regardless of the value of the total spin. Although misleading, themultiplicity, 2S + 1, is traditionally given for S states (L = 0) as well, as wasdone for the 3s configuration. To circumvent this anomaly with S terms, one

3) See Appendix B.8 for a discussion of the mathematical formalism ofstate vectors (the properties of bras, |〉 and kets, 〈|) and the complexalgebra they obey.

Page 60: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.2 Angular Momentum Coupling Schemes 43

must remember that the number of J levels is given by 2S + 1 (the multiplicity)when L S and 2L + 1 when L < S.

As a last example, we will look at the two-electron system of carbon witha ground-state configuration of 1s22s22p2. As in the previous examples, the1s22s2 subshells are closed so that we only need to look at the two 2p electrons.Each electron can be in one of six possible states.

s = 12 ms = − 1

212

l = 1 ml = −1 0 1

If we couple two orbital angular momenta with l = 1 indiscriminately, wewould find that there are three possible values for L: 0, 1 and 2. The possiblevalues for S are 0 and 1. Consequently, there would be 36 = 62 possible stateswith J ranging in value from 0 to 3. However, all of these states will not beallowed in every situation because of Pauli’s exclusion principle. Some com-binations are not allowed when so-called equivalent electrons, that is, thosethat have the same n and l, are coupled. The only allowed states will be thosewhere ml1 and ml2 are different and/or ms1 and ms2 are different. The numberof states meeting this criterion is given by the binomial coefficient (6

2) = 15.That is, of the 36 possible combinations, only 15 exist where their quantumnumbers are not identical; these will be the allowed states. To determinewhich 15 do not violate Pauli’s exclusion principle, we will use Table 2.1 tohelp us keep track of the quantum numbers for each electron.

In the first two columns, we give the magnetic quantum numbers for l ands of the first and second electrons as mms

l . In the third through fifth columnswe give ML = ml1 + ml2 , MS = ms1 + ms2 and MJ = ML + MS values, re-spectively. To determine the terms to which the entries in Table 2.1 are to beassociated, we will group the occurrences of ML and MS. We recall that themaximum values that L and S can assume are 2 and 1, respectively. WhenL = 0, ML takes on one value, 0, while when L = 1 and 2, ML = 0, ±1 and 0,±1 and ±2, respectively. We next must count the occurrences of the variousvalues for ML and MS, which we tabulate in Table 2.2.

We note that ML = MS = 0 occurs three times. One of these corresponds toL = S = 0, a singlet S state (2S + 1 = 1). We also notice that along the columnwhere MS = 0, ML takes on the values 0, ±1 and ±2. This corresponds to L =2 and S = 0 or a singlet D state. Finally, we are left with a single occurrenceof ML and MS running from −1 to 1. This corresponds to L = S = 1 and atriplet P state. Thus, we have 1S, 3P and 1D states. Now we need to assign Jvalues to these states. For the 1S and 1D states, the only choices are J = 0 and2, respectively. For the 3P state, we have J = 0, 1 and 2. To summarize, theallowed terms are 1S0,3 P0,1,2 and 1D2. It is straightforward to show that ourselections account for all the MJ values in Table 2.1 and that there are indeed15 states. Finally, we state a general property of LS coupling that allows one

Page 61: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

44 2 The Structure of the Multielectron Atom

Tab. 2.1 The 15 unique selections of quantum numbers for equivalent p2 electrons in LS cou-pling.

mms1l1

mms2l2

ML MS MJ

−1−1/2

−1+1/2

0−1/2

0+1/2

1−1/2

1+1/2

−2−1−100

0−10−10

−2−2−1−10

−1+1/2

0−1/2

0+1/2

1−1/2

1+1/2

−1−100

0101

−1001

0−1/20+1/2

1−1/2

1+1/2

011

0−10

001

0+1/2 1−1/2

1+1/211

01

12

1−1/2 1+1/2 2 0 2

Tab. 2.2 Occurrences of ML and MS values.

MS

ML–1 0 1

–2 1

–1 1 2 1

0 1 3 1

1 1 2 1

2 1

to quickly verify that the selections are correct when coupling two equivalentelectrons,

L + S must be an even number.

It is important to note that had we chosen oxygen to study instead of carbon,we would have had to consider a 2p4 configuration. However, this wouldcorrespond to two vacancies or holes in the 2p configuration and would leadto exactly the same LS terms. All the LS terms for coupling all combinationsof same-shell p, d and f electrons are given in Ref. [2].

Page 62: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.2 Angular Momentum Coupling Schemes 45

2.2.2jj Coupling

When the second sum in Eq. (2.1) is larger than the last, that is the spin–orbitinteraction of both electrons is stronger than the electrostatic interaction be-tween the electrons, jj coupling is most appropriate. In other words, the twoelectrons in question are sufficiently far from each other that the electrostaticinteraction between them is weak and thus only their total angular momenta,j, are constants of the motion. This is most often the case when the excitedelectron is in a high-n state and thus spends most of its time far from the oth-ers confined to the core near the nucleus. Consequently, for each electron (or ingeneral for the excited electron and the core electrons) we have

−→l i +

−→s i =−→j i

and−→j 1 +

−→j 2 =

−→J . The possible values for

−→J lie between

∣∣∣−→j 1

∣∣∣ +∣∣∣−→j 2

∣∣∣ and∣∣∣−→j 1 −−→j 2

∣∣∣. The usual designation for a jj-coupled term is

(j1, j2) J . (2.11)

Again, if the term has odd parity, a superscript is placed to the right of theright parenthesis. As in LS coupling, we can write a state vector,∣∣γ; (l1m1, s1m1) j1m1 (l2m2, s2m2) j2m2; JMJ

⟩. (2.12)

The application of Pauli’s exclusion principle in jj coupling leads to no twoelectrons having the same value for n, l, j, and mj simultaneously.

As an example, we will look at two equivalent p electrons again. Wewill keep track of the quantum numbers for each electron in Table 2.3. Thefollowing jj−coupled terms can be extracted from Table 2.3: (1/2, 1/2)0,(1/2, 3/2)1,2 and (3/2, 3/2)0,2. One should notice that there are again fivestates with two having J = 0 and 2, and one with J = 1. The reader is encour-aged to create a table similar to Table 2.2 to count the occurrences. The factthat we obtain the same J values as we did with LS coupling means that re-gardless of how we couple the angular momenta, J is a good quantum number,i.e., a constant of the motion.4

2.2.3Intermediate or Pair Coupling

The intermediate coupling scheme, which also involves excited configura-tions, occurs when neither of the last two terms in Eq. (2.1) clearly dominates.There are two limiting cases of this coupling scheme, jl coupling (also knownas jK coupling) and LK coupling (sometimes called Ls coupling, not to be con-fused with LS coupling discussed in Section 2.2.1). In either case, the spin of

4) The reader is referred to a standard quantum mechanics text for aproper discussion of quantities that are “constants of the motion.”

Page 63: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

46 2 The Structure of the Multielectron Atom

Tab. 2.3 The 15 unique selections of quantum numbers for equivalent p2 electronsin jj coupling.

j1 m1 j2 m2 MJ J

+ 12 − 1

2 + 12 + 1

2 0 0

+ 12 − 1

2 + 32

− 32

− 12

+ 12

+ 32

−2

−1

0

+1

2

2, 1

2, 1

2, 1

+ 12 + 1

2 + 32

− 32

− 12

+ 12

+ 32

−1

0

+1

+2

2, 1

2, 1

2, 1

2

+ 32 − 3

2 + 32

− 12

+ 12

+ 32

−2

−1

0

2

2

2, 0

+ 32 − 1

2 + 32

+ 12

+ 32

0

+1

0

2

+ 32 + 1

2 + 32 + 3

2 2 2

the excited electron is coupled last after all the other angular momenta arecoupled to give an intermediate angular momentum called K. The name paircoupling comes from the fact that coupling s (= 1/2) to K always produces apair of levels with total J differing by 1; see Fig. 2.3. In the jl-coupling limit,the order of strength of the interactions is (1) the spin–orbit interaction of theinner or more tightly bound electron, (2) the electrostatic interaction betweenthe two electrons and (3) the spin–orbit interaction of the outer or less tightlybound electron. In this case, j of the inner electron (which we will call j1) cou-ples to the orbital angular momentum of the outer (excited) electron to give K.Finally, s of the outer electron is coupled to give the total J. The state vectorfor this limit is given by∣∣γ; [(l1, s1) j1, l2] Ks2; JMJ

⟩. (2.13)

In the other limit, the relative strengths of (1) and (2) are reversed. In thiscase, the two orbital angular momenta couple to give L, which couples to s ofthe inner electron to give K. The state vector for this limit is given by∣∣γ; [(l1, l2) L, s1] Ks2; JMJ

⟩. (2.14)

Page 64: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.2 Angular Momentum Coupling Schemes 47

Fig. 2.3 This is an example of KL coupling involving a 2S+1LJ coreand an nl electron. Coupling the spin to k, splitting the k degeneracy toproduce two different values of J, gives this scheme its alternate nameof spin doubling.

As an example, we will look at Xe, which has 54 electrons in closed shellsand whose ground-state configuration is 5p6 1S0. The low-lying excited levelsof Xe obey jl coupling. If a 5p electron is promoted to the 6s subshell (seebelow), four levels will arise, one each with J = 2, 0 and two with J = 1. In thelast three columns of the table below, the values for j, jl and J are underlined.

Configurationj: Couple l & s

for inner electrons

jl

: Couple

j to outer l

J: Couple jlto outer spin

5p5 6s 5p5 (2P3/2) [3/2, 0] 3/2 [3/2, 1/2] 25p5 6s 5p5 (2P3/2) [3/2, 0] 3/2 [3/2, 1/2] 15p5 6s 5p5 (2P1/2) [1/2, 0] 1/2 [1/2, 1/2] 15p5 6s 5p5 (2P1/2) [1/2, 0] 1/2 [1/2, 1/2] 0

Figure 2.4 shows the transition from LS to jj coupling.

Fig. 2.4 An example of the transition from LS coupling to jj couplinginvolving two electrons with orbital angular momenta of l1 = 1 andl2 = 0.

Page 65: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

48 2 The Structure of the Multielectron Atom

2.2.4Recoupling Between Coupling Schemes

One finds in practice that there is a need to express the wavefunction in onecoupling scheme in terms of another coupling scheme. This would be the case,for example, for an electron in a Rydberg state where the electron interactsstrongly with the other electrons in the system when it is near the nucleuswhere LS coupling might be appropriate, but only weakly when it is far fromthe nucleus where jj coupling is appropriate.

Consider optical transitions in a two-electron system. If we prepare thesystem in a well-defined state of J and M far from the nucleus, the state vectortakes the form

|[(l1s1)j1(l2s2)j2]JM〉.

The transitions occur near the nucleus so that we need state vectors of theform

|[(l1l2)L(s1s2)S]JM〉.

Since both coupling schemes form a complete set, we can express the former interms of a series of the latter. The numerical coefficients, the Clebsch–Gordoncoefficients (see Appendix D), can be determined from

〈[(l1l2)L(s1s2)S]JM|[(l1s1)j1(l2s2)j2]JM〉.

As a specific example, consider two nonequivalent p electrons (pp′) with totalangular momentum J = 1; LS coupling leads to the following possible states:

3S1, 1P1, 3P1 and 3D1

while jj coupling leads to(12

,12

)1

,(

12

,32

)1

,(

32

,12

)1

,(

32

,32

)1

.

We can build a transformation matrix by calculating the Clebsch–Gordon co-efficients for each case:

Page 66: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

2.3 Fine Structure 49( 32 , 3

2

)1

(32 , 1

2

)1

(12 , 3

2

)1

(12 , 1

2

)1

3D1 – 13

√23 – 1

3

√56

13

√56

23

√53

3P1 0 1√2

1√2

0

1P113

√5 1

3 – 13

13

√2

3S113

√103 – 2

3

√23

23

√23 – 1

3√

3

From the table it is clear that if the system were initially prepared in the jj-coupled state (1/2, 1/2)1 then the LS-coupled wavefunction would have thecharacter

|ψLS〉 =13

(2

√53|3D1〉+

√2|1P1〉 − |3S1〉√

3

).

In this case, although the dominant LS-coupled term is the 3D1 term, the 1P1and 3S1 terms are non negligible.

2.3Fine Structure

The energy shifts associated with the electrostatic interaction and the spin–orbit interaction have opposite signs. The electrostatic interaction leads to apositive shift in the energy levels, i.e., the levels are less bound. The magni-tude of the shift depends on both the total spin and the total orbital angularmomentum. In particular, it has been found empirically that for the groundconfiguration and for configurations with equivalent electrons that the lowest-energy state will be associated with the term of highest multiplicity (greatest S). Theterm of lowest energy of a given multiplicity will be that with the greatest L. Thisbehavior is often referred to as Hund’s rule. For both our carbon, 2p2, andoxygen, 2p4, examples, Hund’s rule implies that the lowest state will be the3P state and that the highest state will be the 1S state. This is exactly what onefinds.5

We saw in Section 1.4.2 that the spin–orbit interaction led to a splittingwithin a shell; not all levels with the same n quantum number have the same

5) See Atomic Energy Levels: Vol. I, NSRDS-NBS 35, which can be foundon the NIST web site [9].

Page 67: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

50 2 The Structure of the Multielectron Atom

energy. It is generally true that levels within nonsinglet LS terms, known asmultiplets, are split by the spin–orbit interaction. This fine-structure splittingis proportional to J, the Landé interval rule. This can be shown by calculating∆EγLS;J, which is given by the matrix element of the various multiplets over

the spin–orbit interaction, ζ (r)−→L · −→S , from Eq. (1.90):

∆EγLS;J =⟨γ; LML, SMS; JMJ

∣∣ ζ (r)−→L · S ∣∣γ; LML, SMS; JMJ

⟩. (2.15)

As was done in the fine-structure calculation for hydrogen, we express L · Sas J2 − L2 − S2 and write

∆EγLS;J = ξ (γLS) [J (J + 1)− L (L + 1) − S (S + 1)] . (2.16)

The spacing between levels J and J − 1 will be

∆EγLS;J − ∆EγLS;J−1 = ξ (γLS) J. (2.17)

The quantity ξ (γLS) is given by

ξ (γLS) =h2

2m2e c2

⟨1r

dU(r)dr

⟩, (2.18)

where U(r) is the effective potential in which the electrons move (see Eq. (2.2)).When ξ (γLS) > 0 the splitting of the multiplet is referred to as normal, wherethe level with lowest J will have the lowest energy. When ξ (γLS) < 0 thesplitting of the multiplet is referred to as inverted, where the level with highestJ will have the lowest energy. A configuration that has a subshell that is lessthan 50% full will have a normal multiplet while one that is more than 50%full will have an inverted multiplet. Configurations that are 50% full followno general tend. Carbon is an example of a normal multiplet while oxygen isan example of an inverted multiplet.

There are some important properties of ∆EγLS;J that should be realized.

1. ∆EγLS;J has no M dependence, which implies that there is no preferreddirection in space – the fine structure is spherically symmetric.

2. Sum rule:

L+S

∑J=|L−S|

(2J + 1) ∆EγLS;J = 0. (2.19)

This implies that the center of gravity of the multiplet,

EγLS;J =∑ J (2J + 1) ∆EγLS;J

∑ J (2J + 1)= 0, (2.20)

coincides with the unsplit term.

Page 68: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Further Reading 51

Example. Find the multiplet splitting for L = 0, S = 1/2. Since J = L ± 1/2,the energy for the L + 1/2 level will be given by

∆EγLS;L+1/2 = ξ (γLS) · L, (2.21)

while that for the L − 1/2 level will be given by

∆EγLS;L−1/2 = −ξ (γLS) · (L + 1) . (2.22)

For a normal splitting, ξ (γLS) > 0.

Further Reading

1 W. C. Martin and W. L. Wiese, Atomic Spec-troscopy: A Compendium of Basic Ideas, Nota-tion, Data and Formulas, Physical ReferenceData, National Institute of Standards andTechnology [http://physics.nist.gov/Pubs/AtSpec/].

2 R. D. Cowan, The Theory of Atomic Structureand Spectra, University of California Press,Los Angeles, CA, 1981.

3 I. I. Sobelman, Atomic Spectra and RadiativeTransitions, Springer Series in ChemicalPhysics, Springer, New York, NY, 1979.

4 F. Yang and J. H. Hamilton, Modern Atomicand Nuclear Physics, McGraw-Hill, San Fran-cisco, CA, 1996.

5 For additional problems with worked so-lutions see D. Budker, D. F. Kimball and D.P. DeMille, Atomic Physics: An ExplorationThrough Problems and Solutions, Oxford Uni-versity Press, New York, NY, 2004.

Problems

2.1 From the energy levels compiled by NIST,6 estimate the quantum defect,µ, for the s, p and d levels for Li through Ne. How does the trend change forthe first ion?

2.2 Using the same data as in the previous problem, estimate ξ (γLS) for Lithrough Ne. How does the trend change for the first ion?

6) At the time of writing this book, the link washttp://physics.nist.gov/PhysRefData/Handbook/periodictable.htm.If it is no longer active, try to search for “Physical Reference Data” athttp://www.nist.gov.

Page 69: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 70: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

53

3Atoms in Static Fields

3.1External Electric and Magnetic Fields

Thus far, we have looked at the atom in the absence of external fields. Externalfields will perturb the structure outlined in the previous sections. Weak fieldswill remove some of the degeneracies while strong fields can cause the atomto fall apart. In this chapter we will study the atom’s response to static electric(Stark effect) and magnetic (Zeeman effect) fields. We will treat the responseto AC fields, which are responsible for transitions between different states ofa system, in Chapter 4.

3.1.1Stark Effect

When an electric field, F, is applied to an atom in the z direction, a new term,eFz, must be added to the effective potential of Eq. (2.4). Before we look at howthe energy levels are perturbed, it is instructive to look at the gross classicalbehavior associated with this term. This is possible by letting U(r) → V(z),where

V(z) = − 14πεo

Ze2

|z| . (3.1a)

To ease numerical calculations, it is often convenient to use a smoothed ver-sion of this one-dimensional potential,

V(z) = − 14πεo

Ze2√

z2 + a2, (3.1b)

Page 71: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

54 3 Atoms in Static Fields

Fig. 3.1 Composite potential, smoothed Coulomb potential ofEq. (3.1b) with a = 0.01 plus the potential due to a static electric field,−eFz, with F = 3.2 × 108 V/cm. The horizontal line inside the potentialindicates the level of the ground state of H, which is absolutely boundfor all negative z but is unbound, as are all higher levels, for positive z.

first introduced by Eberly’s group [10].1 The parameter a ensures that thepotential is finite everywhere and can be used to scale the system’s responseto that of the actual three-dimensional atom. An examination of the propertiesof the atom associated with this potential can be found in Ref. [11]. Figure 3.1shows the composite potential,

V′(z) = V(z) − eFz, (3.2)

for positive F and small a. Electrons are bound by a potential that extends to ∞for negative z, but those in states that lie above the barrier at zc =

√Ze/4πεoF,

determined by setting ∂V′(z)/∂z|a → 0 = 0, can escape to the right, a processthat goes by the name field ionization.2 The barrier height, as a → 0, is givenby

Vc = V(zc)− eFzc = −2e

√eFZ4πεo

. (3.3)

The field, Fc, required to lower this barrier to the level of the ground state canbe determined by setting V′(zc) equal to the ionization potential. For one-electron, hydrogen-like ions the energy for a particular level n is given by

1) It is interesting to note that in honor of Professor Joe Eberly, theDivision of Atomic, Molecular and Optical Physics of the AmericanPhysical Society jokingly proclaimed on July 28, 2003 that a virtualatom (element) described by this potential would bear the nameEberlonium.

2) For the three-dimensional case, zc locates a saddle point in theCoulomb potential over which the electron escapes.

Page 72: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.1 External Electric and Magnetic Fields 55

Eq. (1.15). Thus, the field required for such an over-the-barrier ionization of anatom is the simple expression

Fc =e

4πεoa2o

Z3

16n4 volts (3.4a)

=Z3

16n4 a.u. (3.4b)

An electron in n = 50 would only require a field of about 130 kV/cm, i.e.,13 kV across a 1-mm gap. The ground state of hydrogen requires an elec-tric field (1/16)th the strength of the atomic field strength, ∼ 3.2 × 108 V/cm,to ionize via field ionization. Average field strengths of this magnitude caneasily be created in the focus of a pulsed laser where the intensity can exceed1.4× 1014 W/cm2. For a general atom, with ionization potential, Ip, the thresh-old intensity for over-the-barrier field ionization, often called the appearanceintensity, is given by [12]

Iapp =c

128π

(4πεo

e2

)3 I4p

Z2 , (3.5)

where c is the speed of light. With Ip in eV, Iapp will be in W/m2. Figure 3.2compares the threshold for field ionization determined by measurement tothat predicted by Eq. (3.5). The simple classical theory shows excellent agree-ment over a wide range of intensities and systems.

Quantum mechanically, when the energy level in Fig. 3.1 is below the bar-rier, the electron can tunnel through if the laser frequency is not too high. Thecondition for tunnel ionization is set by the so-called Keldysh parameter [13],

γ =ωL

ωt=

ωL√

2meIp

eF, (3.6)

where ωL is the laser frequency (typically of the order of 1014 Hz), 1/ωt is thetunneling time and Ip is the ionization potential of the atom. Tunneling occurswhen the tunneling time is less than the laser period. That is, the electronmust be able to make it through (over) the barrier within one optical cycle.The tunneling time is determined by requiring an electron to acquire enoughkinetic energy from the field in a time τ to pass over (through) an energybarrier of height Io. Thus, mv2/2 = Io, where v = aτ, with the accelerationgiven by a = eF. This leads to the Keldysh expression τ = 1/ωt

√2mIo/eF.

Clearly, tunneling occurs when γ 1. Consequently, field ionization inducedwith laser fields is often called AC tunneling ionization. In the other extreme,γ 1, multiphoton ionization dominates. Multiphoton ionization will bediscussed in Chapters 4 and 9.

Perelomov, Popov and Terent’ev [14] developed an expression for the tunnelionization rate for hydrogen. Ammosov, Delone and Krainov [15] extended

Page 73: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

56 3 Atoms in Static Fields

Fig. 3.2 A comparison between experimentally determined and theo-retically predicted appearance intensities associated with (Eq. (3.5))for increasing stages of ionization from the lower left to the upperright. (Reprinted figure with permission from S. Augst, D. Strickland,D. D. Meyerhofer, S. L. Chin and J. H. Eberly, Phys. Rev. Lett. 63, 2212(1989). Copyright (1989) by the American Physical Society.)

this to complex atoms. Commonly referred to as the ADK theory, for n∗ lthis rate in Hz can be written as 3

w =1

τao

(3eπ

)3/2 Z2

3n∗32l + 1

2n∗ − 1

[4eZ3

(2n∗ − 1)n∗3F

]2n∗−3/2

× exp[− 2Z3

3n∗3F

], (3.7)

where τao is the atomic time (see Appendix A), n∗ is the effective quantumnumber (Eq. (2.6), which in atomic units can be expressed as Z/

√2Ip where

Ip of the system is also in atomic units), e = 2.71828 . . ., l is the orbital angularmomentum of the state and F is the field strength in atomic units. The ADKexpression fits experimental data better at higher intensities where γ is smaller[19]. Modifications to Eq. (3.7) for molecular systems are discussed in Ref. [20].

For sufficiently weak fields, the low-lying states in an atom will not be ion-ized but perturbed, where eFz is the perturbation Hamiltonian. Since z → −z

3) This form of the expression is the same as that given in Eq. (2) ofRef. [16] but multiplied by a factor of the atomic frequency, 1/τao , togive the expression dimensions of Hz. Note that there is an error inEq. (3) of Ref. [16]; Cn,l should be squared (see Refs. [17, 18]).

Page 74: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.1 External Electric and Magnetic Fields 57

under inversion, eFz must couple states of opposite parity. Since the l statesare degenerate in the nonrelativistic treatment of hydrogen and one-electronions for the same principal quantum number (i.e., n > 1), they will exhibit alinear Stark shift. By linear, we mean that the shifts are calculable in first-orderperturbation theory,

∆E(1) = 〈ψ|eFz|ψ〉. (3.8)

Since the l degeneracy is lifted by electrostatic interaction in multielectron sys-tems, second-order perturbation theory must be applied,

∆E(2) =∫∑ψ′

i

〈ψ|eFz|ψ′〉〈ψ′|eFz|ψ〉Eψ − Eψ′

. (3.9)

The second-order process is usually referred to as the quadratic Stark effect.We will look at both in this section.

3.1.1.1 Linear Stark Effect

Schrödinger [21], Epstein [22] and Walker [23] showed that the linear Starkproblem is nicely solved in parabolic coordinates,4 where the Hamiltonian iswritten as

H′ =e2(ξ − η)F. (3.10)

In parabolic coordinates the perturbation energy is given by

∆E(1) =e2

F〈ψn′1,n′

2,m′ |(ξ − η)|ψn1,n2,m〉

=32

n(n1 − n2)eFao

Z, (3.11a)

where

n = n1 + n2 + |m|+ 1. (3.11b)

Figure 3.3 shows an example for n = 2. We notice that when n1 = n2 = 0and m = ±(n − 1) = ±1, ∆E(1) = 0, reflecting the fact that Eq. (3.11a) isindependent of m. The two extreme components corresponding to m = 0(n1 = n − 1, n2 = 0 and n1 = 0, n2 = n − 1) shift in opposite directions,with a separation of these two extremes 3n(n − 1)eaoF/Z. For n1 = n2, thedegeneracy between n1 and n2 is removed. Table 3.1 summarizes the allowedvalues for n1 − n2 for various values of m.

4) See Appendix B.9 for a discussion of parabolic coordinates and theSchrödinger equation.

Page 75: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

58 3 Atoms in Static Fields

Fig. 3.3 Linear Stark splitting for n = 2. The shift for a field of magni-tude 5.0 × 105 V/cm is ∼ 0.016 eV.

It is possible to express parabolic wavefunctions, un1n2m, in terms of a linearcombination of spherical wavefunctions, ψnlm, and vice versa. As an example,consider n = 1. This value of the parabolic n corresponds to the sphericalprincipal quantum number 2, leading to

u100 = Ce−ar [1 + ar(1 + cos θ)] ,

u010 = Ce−ar [1 + ar(1 − cos θ)] ,

u00±1 = Ce−arr sin θe±imφ.

After some manipulation, this leads to

ψ2p0 = C(u010 − u100),

ψ2s0 = C(u010 + u100).

To see what this means, let us prepare the system in the ψ2s0 state at t = 0,

Ψ(t) = C[

u010e−i(En=2−∆)t/h + u100e−i(En=2+∆)t/h]

. (3.12)

This can be written as

Ψ(t) = [C(u010 + u100) cos(∆t/h) + iC(u010 − u100) sin(∆t/h)] e−iEn=2t/h

=[ψ2s0 cos(∆t/h) + iψ2p0 sin(∆t/h)

]e−iEn=2t/h.

The probability density for the system is

|Ψ(t)|2 = |ψ2s0|2 cos2(∆t/h) + |ψ2p0|2 sin2(∆t/h), (3.13)

Tab. 3.1 Allowed values for n1 − n2 for selected values of m.

m n1 − n2

0 n − 1, n − 3, . . . , −n + 11 n − 2, n − 4, . . . , −n + 23 n − 3, n − 5, . . . , −n + 3...

...n − 1 0

Page 76: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.1 External Electric and Magnetic Fields 59

and the system oscillates between the 2s and 2p states with a period τ ∼2πh/∆. For a field of ∼ 300 V/cm, τ ∼ 4.3 × 10−10 s. Note that this valueis of the order of the radiative lifetime of atomic states, which means that thecharacter of a state will change during a transition. This oscillation causesthe 2s state to pick up some 2p character, allowing forbidden electric dipoletransitions (see Section 4.4) to occur.

For Rydberg states, the energy levels are closely spaced. For sufficientlylarge fields, the manifolds begin to overlap. This so-called crossing field (Fc),where the upper extreme from the lower level meets the lower extreme fromthe upper level, is given by the Inglis–Teller limit Fc = (3n5)−1 [24].

The orbital angular momentum, L, in the presence of an external electricfield is not a constant of the motion. However, m is still a good quantum num-ber. This gives some insight into why L is not a good quantum number in adiatomic molecule. The presence of the electric field from the other nucleusdistorts the spherical symmetry. While L is not conserved, Λ (≡ |m|) is a goodquantum number. The linear case, however, only applies to hydrogen, whichhas energy-degenerate states of different parity. For other systems, includingmolecules, we must go to the next order to determine the Stark shifts. As wediscuss in the next section, the degeneracy between states of different m is par-tially lifted, so states of different |m| (i.e., Λ in molecules) will have differentenergies in general. We will discuss diatomic molecules in Chapter 5.

3.1.1.2 Quadratic Stark Effect

The linear Stark shift is absent in the low-lying states of atoms other thanhydrogen where the l degeneracy is broken (by the electrostatic interaction,for example). In this case, the second-order term is the source of the lowest-order correction. The quadratic Stark shift was first calculated in paraboliccoordinates by Wentzel [25], Walker [23] and Epstein [22], and is given by

∆E(2) = −n4

16

[17n2 − 3(n1 − n2)2 − 9m2 + 19

] F2a3o

Z4 . (3.14)

We note that the degeneracy is now broken between states with different |m|values but a ±m degeneracy still persists.

In spherical coordinates, Khadjavi et al. [26] showed that the shift takes theform

∆E(2) = −12

[α0 + α2

3M2J − J(J + 1)J(2J − 1)

]F2, (3.15)

where α0 and α2 are the so-called scalar and tensor polarizabilities, respec-tively, which take the form

α0 = −23 ∑

J′

e2|〈J||r||J′〉|2(2J + 1)[E(J)− E(J′)] (3.16a)

Page 77: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

60 3 Atoms in Static Fields

and

α2 = 2[

10J(2J − 1)3(2J + 3)(J + 1)(2J + 1)

]1/2

∑J′

e2|〈J||r||J′〉|2[E(J)− E(J′)]

× (−1) J+J′+1

J J′ 11 2 J

. (3.16b)

These polarizabilities can also be measured experimentally.

3.1.2Zeeman Effect

When an atom is placed in a magnetic field, B, similar to those used in neutralatom magnetooptical traps or magnetic traps, the degeneracy associated witheach J level, i.e., the magnetic sublevels labeled by the magnetic quantumnumber, mJ , is removed. In this so-called Zeeman effect, each J levels splitsinto 2J + 1 components, which follows directly from the expectation values ofthe Hamiltonian for the magnetic interaction,

Hmag = −µ ·B, (3.17)

whereµ (= µj) is the effective magnetic moment of the electron in thej direc-tion as discussed in Appendix A.13.1.

The energy associated with each level is found by taking matrix elements ofthe Hamiltonian with states of the form |γLSJMJ〉, which leads to

∆E = 〈−µ ·B〉 = gJµBB〈Jz〉 = gJµBBMJ, (3.18)

where gJ, the Lande g factor, 5 is given by6

gJ = 1 +J(J + 1)− L(L + 1) + S(S + 1)

2J(J + 1), (3.19)

in the LS coupling scheme. Figure 3.4 shows an example of the Zeeman effectfor an alkali atom. Since each fine-structure level splits into 2j + 1 levels witha shift from the zero magnetic field level proportional to gjmj, we label eachlevel by mj and gjmj. In a field of 10 G, for example, the splitting betweenthe magnetic sublevels is of the order of 14 MHz ×gJ . The energy spacing

5) This quantity also goes by the name g factor. It is the ratio betweenthe quantum and classical gyromagnetic ratios, the ratio betweenthe magnetic moment and the angular momentum. There is alsoconfusion in the literature as some authors call the g factor the gyro-magnetic ratio.

6) It is helpful to have an understanding of magnetic moments andthe origin of Eq. (3.19). In Appendix A.13 we review the quantummechanical treatment of the magnetic moments for the electron andnucleons, and derive Eq. (3.19).

Page 78: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.2 Hyperfine Structure 61

2S1/2

2P3/2

gJmJ

2

2/3

-2/3

-2

1

-1

mJ

3/2

1/2

-1/2

-3/2

1/2

-1/2

Fig. 3.4 Zeeman effect for the fine structure of alkali atoms.

determined in this section only applies for weak fields. When the appliedfield is strong, that is when the shifts become comparable to or larger than thespin–orbit splitting, L and S decouple and B will act on the orbital and spincontributions to the magnetic moment separately. The splittings for this casehave been worked out in several places. See Sobelman [27], for example.

3.2Hyperfine Structure

The Zeeman effect discussed in the previous section is appropriate when thespin of the nucleus, I, is zero. For nonzero I, we find substructure within thefine structure known as hyperfine structure. This splitting of the J levels evenin the absence of an applied external field can be traced to the dipole momentand quadrupole tensor of the nucleus and their interaction with the magneticand inhomogeneous electric fields produced by the electrons and the nucleus.In this section we will look at both.

3.2.1Magnetic Interaction

The Hamiltonian for the hyperfine interaction has a structure similar to thatused to calculate the splittings for the Zeeman Hamiltonian,

Hh f = −µI ·Bel , (3.20)

Page 79: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

62 3 Atoms in Static Fields

where µI = gIµNI is the magnetic moment of the nucleus. There are severaldirections we need to consider. The nuclear magnetic moment points in thedirection of the nuclear spin,I. Since the electrons are in a state of total an-gular momentumJ, the magnetic field they produce will be directed alongJ.Consequently, Eq. (3.20) can be represented by

Hh f = AI · J, (3.21)

where A depends on the details of the structure and interaction as we will seebelow. The position of a level one would observe in the laboratory is relatedto the average value of this Hamiltonian because neitherI norJ are constantsof the motion. These two vectors couple to give a new vectorF =I +J that isa constant of the motion. The appropriate state in which the Hamiltonian is tobe averaged will be |γ(LS)J IFMF〉, leading to an expression similar to whatwe had for the fine structure,

∆EµF = 〈Hh f 〉 =

A2

[F(F + 1) − J(J + 1) − I(I + 1)]. (3.22)

The expression for A depends on the value of L. When L = 0, for example,A ∝ 〈1/r3〉. To indicate this l dependence, we will add the subscript l to A, Al .When L = 0, and the electrons are in s states that can penetrate the nucleus,As ∝ |ψs(0)|2, where ψs(0) is the amplitude of the electronic wavefunction atthe nucleus (proportional to the Fermi contact potential). When the hyperfineinteraction is taken into account, the fine structure J levels split into 2J + 1sublevels when I ≥ J or 2I + 1 sublevels when J > I. As an example, weshow the hyperfine structure of two fine-structure levels for a one-electronalkali system in Fig. 3.5. In particular, we show the splitting of the ground2S1/2 and the excited 2P3/2 states in 85Rb, which has a nuclear spin I = 5/2.Each hyperfine level is shifted away from the J level by the right-hand side ofEq. (3.22) and the spacing between adjacent levels is

∆EF,F−1 = ∆EF − ∆EF−1 = AF. (3.23)

3.2.2Explicit Expression for Al

To write an explicit expression for Al , we must look more closely at the mag-netic field generated by the electrons. This field has two contributions, onefrom the orbital motion and the other from the magnetic moment (spin) of theelectron. As was done to generate an expression for the electron in the field ofthe nucleus, we start by writing classical expressions and then converting theappropriate quantities to operators. From the Biot–Savart law, we know that

Page 80: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.2 Hyperfine Structure 63

2S1/2

2P3/2

gF

1/2

7/18

1/9

-1

1/3

-1/3

F

4

3

2

1

3

2

85Rb (I=5/2)

3AS

2AP

3AP

4AP

Fig. 3.5 Magnetic hyperfine structure of the cooling states of 85Rb.Each level is labeled by its F quantum number as well as its nuclear gfactor, gF.

the magnetic field of a charge in motion is given by

Bel,L =µo

ev ×rr3 , (3.24)

where µo is the permeability of the vacuum andr is the position of the nucleusrelative to the electron. In terms of the angular momentum, which of courseis an operator in quantum mechanics, we have

Bel,L = − µo

2πµB

L/hr3 , (3.25)

where µB is the Bohr magneton. The field from a classical magnetic dipole is

Bel,S =µo

4πµB

1r3

[−µ +

3(µ ·r)rr2

]. (3.26)

Page 81: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

64 3 Atoms in Static Fields

Substituting −2µBS/h for the magnetic moment of the electron (see AppendixA.13.1), we can rewrite this quantum mechanically as

Bel,S =µo

2πµB

1r3

[S − 3(S ·r)r

r2

]1h

. (3.27)

Combining the two fields gives

Bel = − µo

2πµB

1r3

[L − S +

3(S ·r)rr2

]1h

. (3.28)

The Hamiltonian in Eq. (3.22) can be written as

Hh f = − µo

2πgIµ

2B

me

mp

1r3

I ·(L − S

)+ I ·r

3 (I ·r)(S ·r

)r2

1h

, (3.29)

where me/mp is the electron to proton mass ratio. If we let

al =µo

2πgIµ

2B

me

mp

1r3 , (3.30)

the energy shifts for fine-structure levels where L = 0 will be given by

∆EF = 〈Hh f 〉 = 〈al〉(−1) J+I+F〈I‖‖I〉

J I FI J 1

×

[〈SLJ‖L‖SLJ〉 −√

8π〈SLJ‖ (XK) ‖SLJ〉]

, (3.31)

where 〈SLJ‖ (XK) ‖SLJ〉 = 〈SLJ‖ (Y2 × S1) ‖SLJ〉, the reduced matrix ele-ment of one of the spherical harmonics (Yl=2,m) and the electronic spin, respec-tively. Equation (E.36) can be used to obtain an expression for this reducedmatrix element. It can be shown that

Al = 〈al〉 l (l + 1)j (j + 1)

=µo

2πgIµ

2B

me

mp

⟨1r3

⟩l (l + 1)j (j + 1)

.(3.32)

For s states, the major contribution to As comes from the electrons penetratingthe nucleus, giving

As =µo

2πgIµ

2B

me

mp|ψs(0)|2. (3.33)

For one-electron systems, one can write an explicit expression for Al and As

using the results in Chapter 1. For many-electron systems, only an approx-imate expression can be written so that numerical calculations are required.Alternatively, Al can be extracted from experiments.

Page 82: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

3.2 Hyperfine Structure 65

3.2.3Hyperfine Zeeman Effect

In the presence of a weak external magnetic field, each F level splits into 2F + 1sublevels. The shift from the zero-field level is given by gF MFµBB. Usingarguments similar to those employed in deriving gj in Appendix A.13.1, it isstraightforward to show that gF is approximately given by

gF gJF(F + 1) + J(J + 1)− I(I + 1)

2F(F + 1). (3.34)

Figure 3.6 is an example in the case of 85Rb.

F=4

F=3

mF

43210-1-2-3-4

3210-1-2-3

B

gFmF

23/211/20-1/2-1-3/2-2

12/31/30-1/3-2/3-1

Fig. 3.6 Zeeman effect for the hyperfine structure of 85Rb with nuclearspin 5/2.

3.2.4Electric Quadrupole Correction

The nucleus, not being a perfect sphere, will also have an electric quadrupolemoment, Q. In general, Q will be a tensor (see Appendix A.14). As we men-tioned above, the interaction between Q and the inhomogeneous electric field

Page 83: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

66 Further Reading

produced by the electrons leads to hyperfine splitting of the fine-structurecomponents as well, albeit smaller than that due to the magnetic dipole. Anexpression for the energy shift due to the electric quadrupole interaction canbe found if we express ∆Eµ

F as AC/2 in Eq. (3.22). This shift can be shown tobe linear in C(C + 1) with the total shift given by [27–29]

∆Eµ + ∆EQ =A2

C + B[C(C + 1) − 34

J(J + 1)I(I + 1)]. (3.35)

The coefficients, A and B, can be determined from experimental measurementof the hyperfine splitting. While outdated, tabulated values can be found inRef. [30].

Further Reading

1 H. A. Bethe and E. E. Salpeter, QuantumMechanics of One- and Two-Electron Atoms,Plenum, New York, NY, 1977.

2 R. D. Cowan, The Theory of Atomic Structureand Spectra, University of California Press,Los Angeles, CA, 1981.

3 I. I. Sobelman, Atomic Spectra and RadiativeTransitions, Springer Series in ChemicalPhysics, Springer, New York, NY, 1979.

4 T. F. Gallagher, Rydberg Atoms, CambridgeMonographs on Atomic, Molecular andChemical Physics 3, Cambridge UniversityPress, New York, NY, 1994.

5 R. F. Stebbings and F. B. Dunning, RydbergStates of Atoms and Molecules, CambridgeUniversity Press, New York, NY, 1993.

6 W. C. Martin and W. L. Wiese, AtomicSpectroscopy: A Compendium of Ba-sic Ideas, Notation, Data and Formulas,Physical Reference Data, National In-stitute of Standards and Technology[http://physics.nist.gov/Pubs/AtSpec/].

7 For additional problems with worked so-lutions see D. Budker, D. F. Kimball and D.P. DeMille, Atomic Physics: An ExplorationThrough Problems and Solutions, Oxford Uni-versity Press, New York, NY, 2004.

8 L. Armstrong, Jr, Theory of the HyperfineStructure of Free Atoms, Wiley-Interscience,New York, NY, 1971.

Problems

3.1 The fine-structure splitting can also be viewed as a magnetic interactionsince in the electron’s rest frame, the nucleus is seen to be a circulating cur-rent. Estimate the order of magnitude of the magnetic field produced by thiscirculating current for the 3p and 4p 2P1/2,3/2 terms in Na.

3.2 Determine the possible LS and jj terms for a d2 configuration.

3.3 Follow the steps outlined in Appendix A.13.1 to combine µJ and µI , dueto the nuclear spin, to show that

〈µF〉 = µBgF

√F(F + 1) (3.36)

Page 84: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Problems 67

and that gF is given by Eq. (3.34). Describe why Eq. (3.34) is a good approxi-mation.

3.4 Use the relationships in Appendix E to show explicitly that⟨γLSJMJ

∣∣ J2 |γLSJM〉 = J(J + 1). (3.37)

3.5 Carry out the steps to show Eq. (3.32) and show that it agrees withEq. (3.22).

3.6 Calculate the energy-level diagram for the strong magnetic field (so-calledPaschen–Back) effect. Draw a diagram of the energy levels for 85Rb as a func-tion of magnetic field.

Page 85: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 86: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

69

4Atoms in AC Fields

4.1Applied EM Fields

In the last chapter we saw that when an atom is placed in a static electric ormagnetic field, its structure changes. Under severe conditions, the distortionscan be so great that electrons are no longer bound. In this chapter, we willexamine the details of atoms in electromagnetic fields – light. Here, we willfocus primarily on light-induced transitions where electrons jump from onestate to another. While we will write down equations that govern multipho-ton transitions in the perturbative regime, we will not discuss nonlinear phe-nomena, such as harmonic generation, four-wave mixing, etc., in this chapter.The general features of the nonlinear response of media will be the subjectsof Part II. Much of what we will discuss in this chapter will also apply tomolecules as well. Thus, the first part of this presentation will be generic for-malism for transitions. We will take a semiclassical approach. That is, we willtreat the radiation field classically and the atom (and molecule) quantum me-chanically. At the end of the chapter, we will look at a few specific atomicexamples. Molecular examples will be presented in Chapter 6.

4.1.1Radiation Hamiltonian

To begin, we write out the radiation Hamiltonian that alters the motion of theelectrons in electromagnetic fields,1

H =1

(p − eA

)2 − eφ + V′, (4.1)

1) The quantity inside the parentheses is what we added to the Diracequation in Section 1.4.1.1 that led to spin–orbit interaction.

Page 87: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

70 4 Atoms in AC Fields

in which

µ = the reduced mass,

e = electric charge (taken to be a positive quantity),

p =hi−→∇ = the electron’s momentum,

A = the vector potential,

φ = the scalar potential,

V′ = the static electric potential.

The E and B fields can be derived from A using Eq. (1.77). To make the prob-lem more tractable, we let H = Ho + Hrad, where Ho describes the motionof the electron in the absence of the radiation field and Hrad, the radiationHamiltonian, describes the interaction between the field and an isolated elec-tron in the absence of any other interaction (i.e., the Coulomb interaction ofthe atomic potential). Expanding the right-hand side of Eq. (4.1) with the aidof the definition of p gives

Ho = − h2

2µ∇2 + V, (4.2a)

where we have let V = −eφ + V′ and

Hrad =e

[2ihA · −→∇ + ih

(−→∇ · A)

+ eA · A]

. (4.2b)

The−→∇ inside the parentheses means that it operates only on A. We will pri-

marily be interested in electronic transitions that electrons make in the pres-ence of this field. There are two instructive limits to consider first – the weak-and strong-field limits. In the weak-field limit, |p|

∣∣∣eA∣∣∣, in which case onlythe first term need be considered in Eq. (4.2b) and we will seek solutions to theSchrödinger equation that are perturbations to wavefunctions that are eigen-states of Ho that we discussed in Chapter 1. At higher intensities both the firstand third terms are important and must be kept. As we will show in the nextsection, the second term in Eq. (4.2b) can be set to zero through a gauge trans-formation. This is because

−→∇ acts only on A, which is just a complex number.As the strength of Hrad increases, perturbation theory breaks down and newnonperturbative approaches must be employed. At extreme intensities, theSchrödinger equation will no longer be valid and one must use a relativis-tic equation, such as the Dirac equation (Eq. (1.66)), to describe the electron’s

motion. We point out that when |p| ∣∣∣eA∣∣∣, the tables are turned and Ho

becomes a perturbation of Hrad. In this chapter we will limit our discussion oftransitions and electron dynamics to the nonrelativistic regime.

Page 88: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.2 Free-Electron Wavefunction 71

4.1.2Coulomb or Radiation Gauge

Since the second term in Eq. (4.2b) is just a complex number, we can simplifyour calculations by choosing a gauge where this number is zero. This is pos-sible with a gauge transformation of the first kind:

A → A′ = A +−→∇χ, (4.3a)

φ → φ′ = φ − ∂χ/∂t. (4.3b)

We are free to choose χ(r, t) to be any value, so we will choose it such that−→∇ · A = 0. It is straightforward to show by direct substitution into Eq. (1.77)that the Coulomb gauge leaves the electric and magnetic fields unchanged.It is also straightforward to show that the Schrödinger equation is invariantunder this gauge transformation as well, since

ψ → eieχ/hψ.

Applying the gauge transformation to Eq. (4.2b) transforms the radiationHamiltonian into

Hrad =iheµ

A · −→∇ +e2

2µA · A. (4.4)

In the presence of the atomic nucleus and the radiation field, the time-dependent Schrödinger equation for the electron can be written as

ih∂

∂tΨ

(−→r , t)

=[

Ho + Hrad

(−→r , t)

. (4.5)

The remainder of this chapter is concerned with solutions to and conse-quences of this equation.

4.2Free-Electron Wavefunction

As we have seen in Chapter 1, exact solutions can be found to Eq. (4.5) whenHrad = 0. Exact solutions are also possible when Ho → −h2∇2/2µ. Thesesolutions take the form

ψ(r, t) = eik·r exp− i

2µh

∫t

[hk − eA(t′)

]2dt′

. (4.6)

We can evaluate the integral for various forms of light. For linearly polarizedlight, A can be written as

A(t) =εzA0(t) cos(ωt + ϕ), (4.7a)

Page 89: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

72 4 Atoms in AC Fields

while for circularly polarized light it takes the form

A(t) = A0(t)[εx cos(ωt + ϕ)−εy sin(ωt + ϕ)

], (4.7b)

where A0(t) is a time-dependent amplitude. With Eq. (4.7a) the solution takesthe form

ψ(r, t) = exp− i

h

[h2k2

2µ+

e2E20

4µω2 +e2E2

08hω3µ

sin 2(ωt + ϕ)

−(

eEµω2

)εz ·k sin(ωt + ϕ)

]+ ik ·r

.

(4.8)

In this expression, e2E20/4µω2 is the so-called ponderomotive energy, Up, and

eE/µω2 is the average vibration radius, α. In terms of the intensity of the lightin W/cm2 and the frequency in Hz, the approximate value for Up in eV is

Up = 5.42× 1016 Iω2 . (4.9)

The approximate value for α in nm is

α = 4.88× 1023√

Iω2 . (4.10)

4.3Radiative Transitions

To solve Eq. (4.5) when the atomic potential does not vanish, one must resortto a variety of approximation techniques appropriate to the relative sizes ofHo and Hrad. In this section we are interested in the nature of single and mul-tiphoton transitions in atomic and molecular systems. Our goal is to generateexpressions for transition rates and cross sections. We find that for many ap-plications making a Dyson expansion of the transition operator is convenientat intensities where the lowest-order term describes the physics well. Giventhe transition amplitude, the rates and cross sections we seek are just the am-plitude squared as we average over a transition cycle. This is perhaps moststraightforward when the wavefunction is expressed in the interaction picture(see, for example, Eq. (3.17) of Ref. [6] or a standard quantum mechanics text),

Ψ (t) = eiHot/hψ (t) , (4.11)

where ψ (t) is an eigenfunction of the unperturbed states, |i〉, with eigenvalueEi. Note that, since we are interested primarily in the time dependence ofthe process, we dropped the explicit reference to space in the wavefunction.

Page 90: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.3 Radiative Transitions 73

When we substitute this wavefunction into Schrödinger’s equation, Eq. (4.5),we find that ψ (t) must satisfy

ih∂

∂tψ (t) = eiHot/h Hrade−iHot/hψ (t) = V (t) ψ (t) . (4.12)

The matrix elements of this interaction will be

〈j| V (t) |i〉 = ei(ωj−ωi)t 〈j| Hrad (t) |i〉 , (4.13)

where we have made use of the fact that Ho |i〉 = hωi |i〉. We note that thematrix element of our new interaction differs from the original interaction bya phase. To solve Eq. (4.12) we will seek integral solutions of the form

∣∣ψj (t)⟩

= |j〉 +(− i

h

) ∫ t

−∞dt1V (t1)

∣∣ψj (t1)⟩

. (4.14)

One should view∣∣ψj (t)

⟩as a state of the system that evolves in time. In

general, we will be interested in transitions from an initial state |i〉 to a finalstate | f 〉. At this point, we have not specified if the states are bound or free.

The transition amplitude will be a projection of 〈 f | out of∣∣ψj (t)

⟩,

〈 f∣∣ψj (t)

⟩ ≡ 〈 f | T(t, t′

) |i〉 , (4.15)

where T (t, t1) is a unitary operator that takes the system from state |i〉 at timet1 to state |ψi (t)〉 at time t,

|ψi (t)〉 = T (t, t1) |i〉 . (4.16)

Now substituting this into Eq. (4.14) leads to

T (t, t1) = 1 − ih

∫ t

t1

dt′1V(t′1

)T

(t′1, t1

). (4.17)

In general, we can make the Dyson expansion

T (t, t1) = 1 +∞

∑n=1

(− i

h

)n ∫ t

t1

dt′1∫ t′1

t1

dt′2 · · ·∫ t′n−1

t1

dt′nV(t′1

) · · · V(t′n

).

(4.18)

We are interested in n-photon processes. Thus, we define T(n) (t, t1) such that

T(n) (t, t1) =(− i

h

)n ∫ t

t1

dt′1∫ t′1

t1

dt′2 · · ·∫ t′n−1

t1

dt′nV(t′1

) · · · V(t′n

). (4.19)

These are particularly useful for cases where the field is weak enough thatthe lowest-order term describes the physics well. When higher-order terms

Page 91: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

74 4 Atoms in AC Fields

are required, this approach quickly becomes inconvenient to use. The lowest-order n-photon transition amplitude, however, would be given by

T(n)i f (t, t1) = 〈 f | T(n) (t, t1) |i〉 . (4.20)

Typically, only certain terms within T(n)i f (t, t1) exist to describe the transition

from |i〉 to | f 〉. For example, if the two levels have opposite parity, only theodd terms will exist; the rest vanish. In order to perform a calculation to es-

timate T(n)i f (t, t1), we must write an explicit expression for V (t). The choices

include the length, velocity and acceleration forms. We will work with thelength form, which usually gives the best results with approximate wavefunc-tions. In the length form we have

V (t) = eiHot/h[d ·F (t)

]e−iHot/h, (4.21)

whereF (t) is the electric field and is expressed as

F (t) =12

[F∗eiωt +Fe−iωt

], (4.22)

whiled is the dipole moment and takes the form

d = −er. (4.23)

The dot product between d and F will select the component of d along thepolarization axis, ε, of the laser field. Thus, we will denote d ·F as dεF andwrite V (t) as

V (t) = eiHot/h[

dεF∗

2eiωt +

dεF2

e−iωt]

e−iHot/h. (4.24)

We are now in a position to look at n-photon transitions.

4.3.1One-Photon Transitions

The transition matrix element for a one-photon transition takes the form

T(1)i f (t, t1) = − i

h

∫ t

t1

dt′1

[ei(ω0+ω)t′1

F∗dεi f

2+ ei(ω0−ω)t′1

Fdεi f

2

], (4.25)

where ω0 =(

Ef − Ei

)/h and

dεi f = 〈 f | dε |i〉 . (4.26)

Page 92: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.3 Radiative Transitions 75

To perform the integration, we recognize that the interaction will take placeover the length of the pulse, which is typically very long compared with theatomic time scale (∼ 10−17 s, Eq. (A.22)). This allows us to let t → ∞ andt1 → −∞. Then, using the definition of the Dirac delta function, Eq. (B.7), wewrite

T(1)i f (∞,−∞) = −2πi

hdε

i f

[F∗

2δ (ω0 + ω) +

F2

δ (ω0 − ω)]

. (4.27)

The first term on the right contributes when ω = −(

Ef − Ei

)/h, which cor-

responds to the transition f → i, stimulated emission, while the second term

contributes when ω =(

Ef − Ei

)/h, which corresponds to the transition

i → f , stimulated absorption. We note that there is no spontaneous emis-sion. In this semiclassical treatment, spontaneous emission must be added ar-tificially and treated phenomenologically with the Einstein A coefficient (seeChapter 2 of Volume 1). Spontaneous emission emerges naturally when theradiation field is quantized as done in Chapter 5 of Ref. [6]. The photons thatpermeate the vacuum are said to stimulate spontaneous emission.

4.3.2Two-Photon Transitions

For n = 2, two-photon transitions, T(n)i f (t, t1) takes the form

T(2)i f (t, t1) =

(− i

h

)2

∑j

∫ t

t1

dt′1ei(ω f j−ω)t′1

(dε

j f F∗

2+

dεj f F

2

)

×∫ t′1

t1

dt′2ei(ωji−ω)t′2

(dε

ijF∗

2+

dεijF

2

), (4.28)

where ω f j = ω f − ωj and ωji = ωj − ωi. The summation is a result of insert-ing a complete set between V(t′1) and V(t′2). We solve the integrals from rightto left. We first let t1 → −∞, assume everything vanishes at −∞ and solve thet′2 integral. We then let t → ∞ and solve the t′1 integral. This gives

T(2)i f (∞,−∞) =

2πi

h2 ∑j

[δ(ωo + 2ω)

(F∗

2

)2 dεj f dε

ij

ωji + ω

+ δ(ωo − 2ω)(

F2

)2 dεj f dε

ij

ωji − ω

+ δ(ωo)∣∣∣∣ F

2

∣∣∣∣2(

dεj f dε

ij

ωji + ω+

dεj f dε

ij

ωji − ω

)],

(4.29)

Page 93: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

76 4 Atoms in AC Fields

where ωo = ω f − ωi. The first two terms correspond to two-photon emissionand absorption, respectively, while no real photons are absorbed nor emittedin the last two terms. These terms describe dynamic Stark shifts, if the initialand final states are the same or a mixing between degenerate states.

Following this same prescription, it is possible to write down an expressionfor an arbitrary number of photons, at least to lowest order. The reader isencouraged to compare the one- and two-photon expressions with those ofChapter 9.

4.3.3Transition Rate: Fermi’s Golden Rule

What one observes in the laboratory is not the transition amplitude but quan-tities related to the transition rate, which is proportional to the modulussquared of the amplitude. In lowest order, we define the transition rate for

an n-photon absorption as the probability, P(n)i f ,2 per unit time as we let time

approach infinity,

R(n)i f ≡ lim

t→∞

1t

∣∣∣T(n)i f ,a (t/2,−t/2)

∣∣∣2, (4.30)

where the subscript a stands for absorption and is associated with the sec-ond term in Eq. (4.27) or (4.29). An equivalent expression can be written forstimulated emission.

At first glance, the structure of the rate appears to be a bit odd, mathemati-cally, because it contains a delta function squared. We see this explicitly fromthe one-photon rate (n = 1),

R(1)i f = lim

t→∞

1t

∣∣∣∣∣(−2πi

h

) (dε

i f F

2

)δ (ω0 − ω)

∣∣∣∣∣2

. (4.31)

One of the delta functions, however, is associated with the line shape, as wediscuss below, while the other is proportional to the period and can be elimi-nated. To see this, we write Eq. (4.31) as

Ri f =2π

h2

∣∣∣∣∣ dεi f F

2

∣∣∣∣∣2

δ (ω0 − ω)(

limt→∞

1t

∫ t/2

−t/2dt′ei(ω0−ω)t′

),

2) We point out that one must be careful with probabilities,

P(n)i f =

∫R(n)

i f (t′)dt′,

because they grow with time and eventually tend towards 1, evenfor very weak transitions if one waits long enough. Thus, if one isnot careful, one will find oneself in a situation where the probabilityis greater than 1.

Page 94: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.3 Radiative Transitions 77

where we used Eq. (B.8) to express 2πδ(ωo − ω) as the integral and combinedthe two limits. Since the entire expression for Ri f vanishes when ω = ωo, theintegral reduces to t, which leads to the one-photon transition rate

Ri f =2π

h2

∣∣∣∣∣ dεi f F

2

∣∣∣∣∣2

δ (ω0 − ω) . (4.32)

Equation (4.32) is often referred to as Fermi’s Golden Rule. The steps for themore general n-photon rate can be obtained in a similar way.

The one-photon rate can be written in several useful forms. In terms of theresonant Rabi frequency, Ω0, for example, it takes the form

Ri f =π

2Ω2

0δ (ω0 − ω) , (4.33)

where

Ω0 =dε

i f F

h, (4.34)

which is the same as Eq. (2.9) of Ref. [6]. It is often convenient to expressthe rate in terms of the intensity of the incident field, Iω = hωηω, where ηω

is the photon flux at frequency ω. Typically, the intensity is expressed withdimensions of W/cm2 and the flux with dimensions of cm−2 s−1. The ratethen takes the form

Ri f =4π2α

hIω

∣∣∣rεi f

∣∣∣2δ (ω0 − ω) (4.35a)

= 4π2αηωω∣∣∣rε

i f

∣∣∣2δ (ω0 − ω) , (4.35b)

where |rεi f | = |dε

i f /e| and α is the fine-structure constant. Finally, if we let thematrix element have units of Bohr radii, ao, we obtain

Ri f = 4π2a2oαηωω

∣∣∣rεi f

∣∣∣2δ (ω0 − ω) . (4.35c)

We distinguish the matrix element in units of Bohr radii by the symbol∣∣∣rε

i f

∣∣∣.Since |rε

i f | is of order 1, we estimate the order of magnitude of the transitionrate in the case where neither the initial nor the final state is degenerate to be

∼ 8 × 10−18ηωδ(ω − ωo). (4.35d)

The rate given in Eq. (4.35c) is not well defined because of the presence of thedelta function. Literally, it says that the transition rate is zero for all ω exceptfor ω = ωo, where it is infinite! Dirac pointed out that since the state repre-sented by the delta function is infinitely narrow, the probability for detecting

Page 95: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

78 4 Atoms in AC Fields

such a line would be of the order zero [31, 32]. Thus, to measure the probabil-ity for a transition, one must integrate this rate over an energy or frequencyrange in the vicinity of the line. In addition, at least one of the states involvedin a real transition is usually not infinitely narrow, but has a definite shapeand width. Consequently, as mentioned above, the delta function should bereplaced by the line shape, L(ω − ωo).

The line shape can be due to one or more of a variety of effects such asthe natural lifetime of the state, collisions during the transition or perhapsthe translational energy of the system. We will consider, as an example, thenatural lifetime, which leads to a Lorentzian line profile that was defined inEq. (1.50) of Ref. [6] as

L(ω − ωo) =2π

γ 12

(ω − ωo)2 + γ212

, (4.36)

where in this case γ 12

is the half width at half maximum and is equal to half

the radiative lifetime between the two states.3 The line shape obeys∫ ∞

0L(ω − ωo)dω = 1.

Replacing 2πδ (ω0 − ω) in Eq. (4.35c) with L(ω − ωo), the rate is written as

Ri f = 2πa2oαηωω

∣∣∣rεi f

∣∣∣2L(ω − ωo), (4.37)

which corresponds to the transition rate for a monochromatic source at ω.While we have derived Eq. (4.37) assuming a Lorentzian profile, a simi-

lar expression holds for other profiles such as a Gaussian shape (Doppler-broadened line) or even for Voigt profiles (a convolution of a Gaussian with aLorentzian as discussed in Appendix B.10).

4.3.3.1 Degeneracy

The transition rate given in Eq. (4.37) corresponds to the case where there isno degeneracy. We know from our study of the structure of atoms that a stateof angular momentum J will be (2J + 1)-fold degenerate. The appropriate ratein the degenerate case for polarized light is obtained by averaging over theinitial states and summing over the final states to give

Ri f =2π

giαηω

∣∣∣rεi f

∣∣∣2L (ω0 − ω) , (4.38)

3) The radiative lifetime is given by the Einstein A coefficient couplingthe two states. Compare Eq. (4.36) with Eq. (2.33) of Ref. [6].

Page 96: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.3 Radiative Transitions 79

where gi = 2Ji + 1, the number of degenerate initial states of different Mi, and|rε

i f |2 now takes the form∣∣∣rεi f

∣∣∣2= ∑

Mi,M f

∣∣∣⟨γ f J f Mf |rε| γi Ji Mi

⟩∣∣∣2. (4.39)

An additional sum over the polarization components is required when thelight is unpolarized.

4.3.3.2 Narrow and Broad Sources

When the transition is induced with a laser that has a line width small com-pared with γ 1

2,

L(ω − ωo) 2πγ 1

2

,

on resonance, and the rate takes the form

Ri f (ωo) = 4a2oαη

ωo

giγ 12

∣∣∣rεi f

∣∣∣2. (4.40)

For a source broad compared with the width of the line, one needs to integrateover the line. Assuming that the intensity is flat over the line, one can write

Rbroadi f =

gia2

oαηωo

∣∣∣rεi f

∣∣∣2 ∫ ∞

0dωωL(ω − ωo). (4.41)

4.3.4Transition Strength: Absorption

Although the rate is often what is needed when deciding the feasibility of anexperiment, it depends on the strength of the light source. It is often useful tohave an expression for the strength of the transition that depends only on theatomic system. The common ways to express the strength of a transition arethe line strength, the cross section and the oscillator strength. We will look at eachin this section.

4.3.4.1 Line Strength

Exploiting the Wigner–Eckart theorem (Appendix E.3) we see that |rεi f |2, from

Eqs. (4.38) and (4.39), is related to the reduced matrix element for the transi-tion. Specifically, as shown in Chapter 9 of [27], a single polarization compo-nent, ε, can be written as

∣∣∣rεi f

∣∣∣2=

∣∣∣ri f

∣∣∣2

3=

∣∣∣⟨γ f J f ||r| |γi Ji

⟩∣∣∣2

3. (4.42)

Page 97: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

80 4 Atoms in AC Fields

The modulus squared of the reduced matrix element,⟨

γ f J f ||r| |γi Ji

⟩, defines

the line strength,

S ≡ e2a2o

∣∣∣⟨γ f J f ||r| |γi Ji

⟩∣∣∣2, (4.43)

which is a sum over all transitions and all three directions. The line strengthgives the probability for a transition from a specific initial level to all 2J f + 1levels of the final state. The line strength has units of C2 cm2.

4.3.4.2 Cross Section

The cross section is defined by dividing the transition rate of Eq. (4.38) by thephoton flux,4

σ(ω) =Ri f

ηω=

gia2

oαω∣∣∣rε

i f

∣∣∣2L(ω − ωo). (4.44)

The cross section as defined has units of area, typically cm2.

4.3.4.3 Oscillator Strength

The oscillator strength, introduced in Section 2.2.5 of Volume 1 [6], can bethought of as a fudge factor to bring the classical cross section into agreementwith the quantum mechanical cross section for absorption. Specifically, it isthe dimensionless number by which the classical cross section must be multi-plied to give the quantum cross section for a particular transition. Classically,the absorption cross section for a damped oscillating electron can be expressedas

σcl(ω) =1

4πεo

πe2

mecL(ω − ωo), (4.45)

where e, me, c and εo have their usual meanings. The absorption strength isgiven by the total or integrated cross section (sometimes called the spectralcross section),

σclT =

∫ ∞

0dωσ(ω)

=1

4πεo

πe2

mec. (4.46)

The oscillator strength is then defined as

σT = σclT f =

14πεo

πe2

mecf = 2.65 × 10−2 f cm2/s. (4.47)

4) One should understand that while we have been using L(ω − ωo) toindicate a Lorentzian, in Eq. (4.44) it is meant to convey the appro-priate line shape (Lorentzian, Gaussian, etc.).

Page 98: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.3 Radiative Transitions 81

Comparing Eqs. (4.43)–(4.45) and (4.47) along with the definition of the fine-structure constant (Appendix A.3), an explicit expression for the absorptionoscillator strength, fi f , takes the form

fi f =2mea2

gih

∣∣∣rεi f

∣∣∣2=

2meω

3he2Sgi

. (4.48)

Sometimes the strength of the transition is given in terms of g f , where forabsorption and emission we have respectively

g f = (2J f + 1) fi f , (4.49a)

= −(2Ji + 1) f f i. (4.49b)

The oscillator strength defined by Eq. (4.48) is related to the integrated crosssection (see Eq. (4.47)) and assumes that the transition strength is confined toa small region of frequency space. In essence, it is only appropriate for tran-sitions between bound states. For transitions to continuum states, it is moreappropriate to talk about the oscillator strength density, d f /dE. Expressionsfor d f /dE are discussed in Ref. [33].

4.3.5Transition Strength: Emission

There are corresponding expressions for the downward transition – emission– to quantities presented in the preceding section. We give the oscillatorstrength here, but the rest can be determined from Table 2.1 of Volume 1 [6].The oscillator strength for stimulated emission takes the form

f f i = −2meω

3he2Sg f

, (4.50)

where g f = 2J f + 1. Note that f f i is usually expressed as a negative value. Themeaning of Eq. (4.50) is the probability for emission from a specific upper levelto all the lower levels. The oscillator strength is also related to the lifetime5 ofthe state through

A f i = − e2ω2

2πεomc3 f f i (4.51a)

=gi

g f

e2ω2

2πεomc3 fi f . (4.51b)

5) The lifetime of the state refers to how long the electron will stay inan excited state before it spontaneously emits a photon and makesa transition to a lower state. Spontaneous emission occurs in theabsence of an applied external field and is due to photons in the vac-uum stimulating a transition. This interpretation follows naturallywhen the radiation field is quantized.

Page 99: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

82 4 Atoms in AC Fields

Combining Eqs. (4.50) and (4.51a) (or Eqs. (4.48) and (4.51b)) allows us relatethe lifetime to the line strength as well,

A f i =1

3πεo

ω3

hc3Sg f

. (4.51c)

4.4Selection Rules for Atomic Transitions

There are three types of transitions commonly observed in atomic systems:the electric dipole (so-called E1), the magnetic dipole (M1) and the electricquadrupole (E2) transitions. Of these, the electric dipole is by far the mostcommon and strongest. The other two are often referred to as electric-dipoleforbidden. Since the E1 transitions are the most common, we will work outthe selection rules for them in some detail and only give the final results forthe other two.

4.4.1Electric Dipole (E1) Transitions

From Eq. (4.39) we know that the transition rate for a dipole transition is pro-portional to a matrix element of the form < f |r|i >. It is not necessary tocalculate this matrix element for all possible states because most of them willbe zero. What we will do in this section is to determine which are nonzero.There are a few ways we can approach this problem. We will limit ourselvesto two. First, we will use our knowledge of the symmetry of the wavefunc-tion to work out much of the detail. We will then calculate the nonzero matrixelement more formally with the aid of the Wigner–Eckart theorem. (See Ap-pendix E.3 for a discussion of the Wigner–Eckart theorem.)

We know that in coordinate space, the integral represented by the matrixelement represents a volume integration over 4π. Thus, we know that if theintegrand is odd with respect to inversion, the matrix element will be identi-cally zero. We have three parts of the matrix element to consider – the wave-functions of the initial and final states and r. Looking first at r, we know thatunder inversion r → −r. Such a function has odd parity. The first requirementfor a nonzero matrix element is that < f | and |i > have opposite parity. Welearned in Section 1.3.1 that the parity of the state is given by l. Consequently,for a nonzero matrix element we must have

∆l = ±1. (4.52)

We can say a few more things about the selection rules from what is missingfrom Eq. (4.39). Specifically, the matrix element does not involve the principal

Page 100: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.4 Selection Rules for Atomic Transitions 83

quantum number, n, or the spin. Thus, there is no restriction on the principalquantum number nor on the spin quantum number.

The fact that there is a restriction on the change in value of l would implythat there should be a restriction on L and perhaps J as well. To explore thiswe need to turn to the Wigner–Eckart theorem and the properties of the n − jsymbols discussed in Appendix D. To that end, we expressr as a sphericalvector. In Cartesian coordinates the components ofr are

r1 = x,

r2 = y,

r3 = z,

while in spherical coordinates they are

r0 = z,

r+1 = − 1√2(x + iy),

r−1 = +1√2(x − iy).

Expressing r in spherical coordinates and using the Wigner–Eckart theorem,Eq. (E.14), the transition matrix is seen to be related to the line strength, a 3-jsymbol and a phase,6

〈γjm| rq∣∣γ′ j′m′⟩ = (−1)j−m

(j 1 j′

−m q m′) ⟨

γj ‖r‖ γ′ j′⟩

, (4.53)

where q = 0,±1 and the last factor is the reduced matrix element, which isindependent of the magnetic components (see Appendix E.3) and gives theprobability for a transition from a specific initial level to all 2j + 1 levels ofthe final state.7 From the properties of the 3-j symbol (see Appendix D.2), weknow that the matrix element will be zero unless j, 1 and j′ add as vectors and−m + q + m′ = 0. In other words, we have the two additional selection rules,

∆j = 0,±1, (4.54a)

∆m = 0,±1, (4.54b)

with two exceptions,

j′ = 0 j = 0, (4.54c)

m′ = 0 m = 0 when ∆j = 0. (4.54d)

6) The reader should note that in Appendix E.3, the rank of the tensoroperator is also included in the subscript in the general expression ofthe matrix element. We drop the rank on r to avoid confusion withthe +1 component.

7) We note that the reduced matrix element in the Wigner–Eckart theo-rem is not in atomic units, so S ≡ |e 〈γj ‖r‖ γ′ j′〉 |2.

Page 101: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

84 4 Atoms in AC Fields

In LS coupling, one can show that

∆L = 0,±1, (4.54e)

∆S = 0, (4.54f)

L′ = 0 L = 0. (4.54g)

4.4.2Magnetic Dipole (M1) Transitions

The magnetic dipole operator, which has a form similar to Eq. (3.17), couplesstates of the same parity. The Wigner–Eckart theorem (Eq. (E.14)) can be usedto determine the selection rules as we did for E1 transitions. The resultingselection rules are

∆l = 0, (4.55a)

∆j = 0,±1, (4.55b)

∆m = 0,±1, (4.55c)

again with two exceptions,

j′ = 0 j = 0, (4.55d)

m′ = 0 m = 0 when ∆j = 0. (4.55e)

In LS coupling,

∆L = 0, (4.55f)

∆S = 0, (4.55g)

∆J = ±1. (4.55h)

4.4.3Electric Quadrupole (E2) Transitions

The electric quadrupole operator is a second-rank tensor Q2,q, which is com-monly written in Cartesian coordinates as Qij = erirj and couples states of thesame parity. Using the Wigner–Eckart theorem (Eq. (E.14)), the selection rulesfor transitions are determined to be

∆l = 0,±2, (4.56a)

∆j = 0,±1,±2, (4.56b)

∆m = 0,±1,±2, (4.56c)

Page 102: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.5 Atomic Spectra 85

again with the exceptions,

j′ = 0 j = 0, (4.56d)

j′ = 1/2 j = 1/2, (4.56e)

j′ = 0 j = 1. (4.56f)

In LS coupling,

∆L = 0,±1,±2, (4.56g)

∆S = 0. (4.56h)

4.5Atomic Spectra

Before we conclude our discussion of atoms, it is instructive to look at the sim-plest, unique features in the spectra of one- and two-electron atoms. A morecomprehensive review is beyond the scope of this introductory text. Thereare a variety of sources to which the interested reader can turn for additionalexamples. For a few classic examples, see Photoabsorption, Photoionization, andPhotoelectron Spectroscopy by Berkowitz and Atomic and Molecular Spectroscopy:Basic Aspects and Practical Applications by Svanberg listed under further read-ing at the end of this chapter and references therein. Here, we will look atRydberg series, autoionization and above-threshold ionization.

4.5.1Rydberg Series

From the selection rules of the previous section, it is clear that from the groundstate, transitions to any state n are possible. However, the strength of the tran-sition will vary with n. In the absence of strong interactions between differentseries, the strength will decrease as 1/n∗3 for a so-called hydrogenic system.Figure 4.1 shows this reduction theoretically for H. This trend, however, is vi-olated in general due to a variety of interactions between series. We will lookat a classic example later in this chapter.

The oscillator strength defined in Section 4.3.4.3 strictly refers to bound–bound transitions, where fi f is linked to the integrated (over the absorptionline) cross section. For bound–continuum transitions, photoionization, thespectrum is no longer composed of discrete lines but, rather, a continuum. Todiscuss the strength, it is more appropriate to define a spectral density of theoscillator strength, d f /dE. Thus, in Fig. 4.1, d f /dE for bound–bound transi-tions is plotted as histograms, where the area corresponds to f . Such a figureallows us to compare the strengths of the discrete and continuum portions of

Page 103: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

86 4 Atoms in AC Fields

Fig. 4.1 The theoretical oscillator strengthdistribution, d f /dE, to within a proportionalityconstant for bound-bound (histogram) andbound-continuum transitions from the groundstate in H taken from Ref. [33]. The horizon-tal axis is in eV. The oscillator strengths forthe bound states correspond to the area of

the histogram. It is clear that the distributionin the discrete region continues smoothlyinto the continuum. (Reprinted figure withpermission from U. Fano and J. W. Cooper,Rev. Mod. Phys. 40, 441 (1968). Copyright(1968) by the American Physical Society.)

the spectrum on an equal footing. The key observations one should make fromthis figure are that f varies with n and d f /dE continues smoothly from the dis-crete into the continuum. The latter is true in general even when the intensitydoes not fall as 1/n∗3. The reason for this behavior is due the fact that theRydberg series and its adjoining continuum (a channel) are intimately related,as we pointed out in Section 1.5. Typically, there is more than one channel thatcan be accessed from a particular level. When the Rydberg series converge tothe same ionization limit, the members march in lock-step, without crossingeach other. In the absence of any other perturbation, their intensities follow asimple, smooth progression similar to that displayed in Fig. 4.1.8 This simplestructure is common in one-electron (e.g. alkali) systems.

The spectrum is richer in two-electron systems. Specifically, it is commonfor two or more Rydberg series to occur in the same energy region but con-verge to different ionization thresholds. In this case, the spacing between lev-els is not the same and members of the two series periodically cross. This inter-mingling of the levels leads to an interaction between the series, particularly

8) While the interaction between Rydberg series may be small, moreexotic interactions can change the simple behavior observed inFig. 4.1 dramatically even for one-electron systems. See Ref. [33]for examples.

Page 104: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.5 Atomic Spectra 87

if the energies of two levels are nearly degenerate, resulting in energy shifts ofthe bound states away from their proper locations as dictated by Eq. (2.7). Atthe same time, the strength of the lines can vary rather dramatically near thecrossing points.

The classic example is the Xe isoelectronic sequence (same electronic config-uration). Part of the Ba2+ spectrum near the lowest ionization limit is shownin Fig. 4.2. From the ground state, a 5p6 1S0 configuration, transitions to con-figurations of the form 5p5nl consist of five distinct channels,9 three of whichconverge to the first limit producing Ba3+ in its 2P3/2 state with the other twoconverging to the second limit producing Ba3+ in its 2P1/2 state. The spectrumin Fig. 4.2a is plotted as a function of the effective quantum (ν2, see Eq. (2.6))associated with the upper limit. On this scale, levels that are part of the Ryd-berg series converging to the upper limit are equally spaced. Consider the lev-els marked ns′ and nd′, which are equally spaced and do not cross each other.On the other hand, the spacing between levels marked ns and nd decreasesas the effective quantum number increases causing these unprimed levels tocross the primed levels periodically; crossing occurs near ν2 = 4.8, 5.8 and 6.8.One notices that the intensity of the lines increases significantly near thesecrossing points. We see this again in the bottom graph of log g f (see Eq. (4.49)and Fig. 4.2c), which compares the calculated values of the oscillator strengthfor this Ba2+ case (the lines) against the ideal values for the hydrogenic case(∝ 1/n∗3, indicated by the dots). The connection between the Rydberg seriesand its continuum is clearly seen as well, in that the same intensity anomaly inthe discrete region is repeated in the continuum – the physics associated withthe perturbation in the two cases is essentially the same.

Figure 4.2b contains a plot of the positions of the bound resonances relativeto the upper limit (ν2) vs their positions relative to the lower limit (ν1) mod 1,a so-called Lu–Fano plot [36–38]. In general, the unprimed levels lie on curvesthat are more or less horizontal while the primed levels lie on vertical curves.The interaction between the channels is strongest where the curves cross. Thestronger the interaction, the larger the avoided crossings between the nomi-nally vertical and horizontal curves. Resonances that fall close to where thecurves would cross in the absence of any interaction experience significant en-ergy shifts. ν1 for the 19d level, for example, is shifted to lower values relativeto other members of the sequence. At the same time, we notice that 20s is alsoshifted, but by much less than 19d, indicating that the interaction betweenthe nd and nd′ channels is much stronger than the interaction between thens and ns′ channels. The differences in the interaction are more acute abovethe threshold (near ν2 = 7.8) in the autoionization (see next section) region.

9) Only four of the Rydberg series are observed in Fig. 4.2 because twoof those converging to the lower limit are nearly degenerate andunresolved.

Page 105: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

88 4 Atoms in AC Fields

Fig. 4.2 An example of interacting Rydbergseries in a 2-electron system, Ba2+, takenfrom Ref. [34]: (a) photoabsorption spec-trum plotted as a function of ν2 (Eq. (2.6)),(b) Lu-Fano plot of position of the resonancesrelative to the first and second limits (i.e.,ν1( mod 1) vs. ν2) and (c) bar graph of cal-culated oscillator strengths (g f ) for the boundnd series with the length of the lines repre-

senting the theoretical value, including fullinteractions, and the dots representing the hy-drogenic values falling as 1/ν3

1 . (Reprintedfigure with permission from W. T. Hill, III,J. Sugar and T. B. Lucatorto and K. T. Cheng,Phys. Rev. A 36, 1200 (1987). Copyright(1987) by the American Physical Society.)

The stronger interaction between the nd and nd′ channels compared with theinteraction between the ns and ns′ channels makes the nd′ resonance widerthan the ns′ resonance. Figures 4.3 and 4.4 show that as electrons are removedalong the isoelectronic sequence, the interaction is reduced. The intimate con-nection between the interaction in the bound and continuum regions is againseen. The wider autoionization resonances in Fig. 4.3 lead to larger avoidedcrossings about the bound levels in the Lu–Fano plots in Fig. 4.4.

Page 106: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.5 Atomic Spectra 89

Fig. 4.3 Autoionization widths along the Xe-isoelectronic sequence taken from [35]. Thespectrum of Xe, Cs+ and Ba2+ are shownrespectively from top to bottom betweenν2 = 7.5 and 9.0. The change in the inter-actions is seen in three ways: a change in thewidths and shapes of the nd′ resonances anda shift in the position of the ns′ resonances.(Reprinted figure, Fig. 3b) on page 188, from

W. T. Hill, III and C. L. Cromer, “Laser-DrivenIonization and Photoabsorption Spectroscopyof Atomic Ions,” in Lasers, Spectroscopy andNew Ideas: A Tribute to Arthur L. Schawlow,Opt. Sci. Series Vol 54, 183 (1987), with kindpermission from Springer Science and Busi-ness Media. Copyright (1987) by Springer-Verlag.)

4.5.2Autoionization

Autoionization is an interaction between a bound state that belongs to a closedchannel (energy below the ionization threshold) and a continuum state (be-longing to an open channel) that is energy degenerate. Thus, the bound stateis only quasibound because the interaction leads to its decay into the contin-uum state and a free electron. The stronger the interaction the faster the decayand the wider, in energy space (∆E∆t ∼ h), the resonance. The nd′ resonancesare wider than the ns′ resonances in Fig. 4.2, indicating that they autoionizefaster.

The formal treatment of autoionization was done first by Fano [39], whosesteps we follow here. The analysis begins by assuming eigenstates of ourHamiltonian describing the state of the system in the absence of excitation.The eigenstates will include bound (|n〉), quasibound (|φ〉) and continuum(|E〉) states. In the autoionization region (above the first ionization threshold)

Page 107: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

90 4 Atoms in AC Fields

Fig. 4.4 Lu-Fano plots of the bound Ryd-berg resonances associated with the au-toionization resonances in Fig. 4.3 takenfrom [35]. As electrons are removed from thesystem, the interactions decrease (smalleravoided crossings) and the resonances shift.(Reprinted figure, Fig. 4 on page 190, fromW. T. Hill, III and C. L. Cromer, “Laser-Driven

Ionization and Photoabsorption Spectroscopyof Atomic Ions,” in Lasers, Spectroscopy andNew Ideas: A Tribute to Arthur L. Schawlow,Opt. Sci. Series Vol 54, 183 (1987), with kindpermission from Springer Science and Busi-ness Media. Copyright (1987) by Springer-Verlag.)

we have

〈φ|Ho|φ′〉 = Enδφφ′ , (4.57a)

for the quasibound state and

〈E|Ho|E′〉 = Eδ(E − E′), (4.57b)

for the continuum state. We define the matrix element of Ho between thequasibound and continuum states,

V = 〈E|Ho|φ〉, (4.57c)

Page 108: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.5 Atomic Spectra 91

Fig. 4.5 Various observed autoionization line profiles from [33]: (a) He(q = −2.8, ρ2 = 1) 2s2p1Po, (b) Ar (q = −0.22, ρ2 = 0.86) 3s3p6

4p 1Po, (c) Ne (q = −2.0, ρ2 = 0.17) 2p4 3s3p 1Po and (d) Xe (q ∼200, ρ2 ∼ 0.0003) 4d95s25p66p 1Po. Here, ρ = 〈E|φ〉 and what isplotted is ρ2[σ(ε) − 1] + 1. (Reprinted figure with permission fromU. Fano and J. W. Cooper, Rev. Mod. Phys. 40, 441 (1968). Copyright(1968) by the American Physical Society.)

which indicates the strength of the interaction between the two. We next de-fine a transition matrix element from the initial (bound) state to the quasi-bound and to the continuum states,

tb = 〈φ|T|i〉 (4.58a)

Page 109: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

92 4 Atoms in AC Fields

and

tc = 〈E|T|i〉. (4.58b)

The eigenstate of the system above threshold with eigenvalue E will be a su-perposition of the continuum and quasibound states,

|Ψ(E)〉 = a|φ(E)〉+∫

dE′b|E(E′)〉, (4.59)

where a (a function of E) and b (a function of E and E′) are coefficients. It canbe shown [39] that the transition matrix element and cross section will takethe form

t(ε) =q + ε√1 + ε2

tc, (4.60a)

σ(ε) =(q + ε)2

1 + ε2 σc, (4.60b)

where

σc ≡ |tc|2, (4.60c)

q ≡ tb

πV∗tc , (4.60d)

ε ≡ E − Er

π|V|2 . (4.60e)

In Eqs. (4.60), Er is the energy where the enhancement in the continuum occurs(the resonance) and q is real and measures the relative strength of transitionsto the quasibound and continuum states. The magnitude and sign of q tend tocontrol the shape of the resonance while |V|2 is related to its width. Clearly,these parameters are different for each member of the isoelectronic sequenceof Xe shown in Fig. 4.3. It is possible to fit the profiles to determine the pa-rameters in each case. Several additional shapes along with the parametersare displayed in Fig. 4.5.

4.5.3Photoionization with Intense Lasers

Thus far, we have only looked at weak-field transitions involving one-photontransitions between bound and continuum states and two-photon transitionsbetween bound states. At higher intensities, n-photon transitions, where ncan become quite large, are possible between bound states as well as to con-tinuum states. Transitions to continuum states is referred to as multiphotonionization (MPI). It is straightforward, albeit tedious, to calculate the n-photon

Page 110: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

4.5 Atomic Spectra 93

transition rate by following the procedure used to determine the one- and two-photon rates in Section 4.3. While the lowest order is usually sufficient, if thisperturbative approach is valid it is also possible to calculate higher orders.Diagrammatic approaches have been developed to aid bookkeeping [40]. Inlowest order, the n-photon rate can be written as [40, 41]

R(n)i f =

h

(2e2

εoc

)n

In

×∣∣∣∣∣ ∑i,j,...,k

〈E|r|k〉〈k|r|j〉 · · · 〈i|r|g〉(Ek − Eg − (n − 1)hω) · · · (Ei − Eg − hω)

∣∣∣∣∣× δ(E − Eg − nhω), (4.61)

where |g〉 is the initial state with the associated energy Eg and |E〉 is the con-tinuum state with the associated energy E. Energy conservation leads to pho-toelectron energies given by

E = nhω − Eg , (4.62)

where n is the minimum number of photon to exceed the ionization threshold.At sufficiently low intensities, Eqs. (4.61) and (4.62) do a reasonable

job describing the photoelectron energy spectrum. At intensities of 1012–1013 W/cm2, Eq. (4.62) begins to fail because a new phenomenon, called

Fig. 4.6 Above-threshold ionization: (a) energy level schematic takenfrom Ref. [42] and (b) photoelectron momentum spectrum showingthe characteristic energy spacing corresponding to hωl. Note, on amomentum scale, pe =

√2mehωl. (Reprinted figure, Fig. 4.6a, with

permission from P. Agostini, F. Fabre, G. Mainfray, and G. Petite, Phys.Rev. Lett. 42, 1127 (1979). Copyright (1979) by the American PhysicalSociety.)

Page 111: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

94 Further Reading

above-threshold ionization (ATI), becomes active. The central feature of ATIis that electrons do not cease to absorb energy from the laser field once nhω

exceeds the ionization threshold, as they do at lower intensities. Rather, asshown in Fig. 4.6, the electrons can leave the atom having absorbed an integralnumber of additional or excess photons [42].

In order for the ATI spectrum to conserve energy, one must modify Eq. (4.62)such that

E = (n + s)hω − Eg, (4.63)

where s is an integer. We can understand why the spectrum contains peaksspaced by the photon energy when we recognize that some of the intermedi-ate states in Eq. (4.61) will be continuum states. The summations, which re-sulted from inserting complete sets, involving the continuum states are reallyintegrals. Thus, the rates peak every hω because of the energy denominatorsinvolving those continuum states.

Further Reading

1 F. H. Faisal, Theory of Multiphoton Processes,Plenum, New York, NY, 1987.

2 I. I. Sobelman, Atomic Spectra and RadiativeTransitions, Springer Series in ChemicalPhysics, Springer, New York, NY, 1979.

3 H. A. Bethe and E. E. Salpeter, QuantumMechanics of One- and Two-Electron Atoms,Plenum, New York, NY, 1977.

4 R. D. Cowan, The Theory of Atomic Structureand Spectra, University of California Press,Los Angeles, CA, 1981.

5 J. Berkowitz, Photoabsorption, Photoioniza-tion and Photoelectron Spectroscopy, Aca-demic, New York, NY, 1979.

6 S. Svanberg, Atomic and Molecular Spec-troscopy: Basic Aspects and Practical Applica-tions, Springer, Berlin, 1991.

7 R. F. Stebbings and F. B. Dunning, RydbergStates of Atoms and Molecules, CambridgeUniversity Press, New York, NY, 1993.

8 T. F. Gallagher, Rydberg Atoms, CambridgeMonographs on Atomic, Molecular and

Chemical Physics 3, Cambridge UniversityPress, New York, NY, 1994.

9 J.-P. Connerade, Highly Excited Atoms, Cam-bridge Monographs on Atomic, Molecularand Chemical Physics 9, Cambridge Uni-versity Press, New York, NY, 1998.

10 G.W.F. Drake (ed.), Atomic, Molecular, & Op-tical Physics Handbook, American Instituteof Physics, Woodbury, NY, 1996.

11 http://physics.nist.gov/PhysRefData/contents.html. This web site, maintainedby the US National Institute of Standardsand Technology (NIST), contains an assort-ment of tables and links to the latest atomicspectroscopic information and fundamen-tal constants. The link was valid at the timeof publication.

12 For additional problems with worked so-lutions see D. Budker, D. F. Kimball andD. P. DeMille, Atomic Physics: An Explo-ration Through Problems and Solutions, Ox-ford University Press, New York, NY, 2004.

Page 112: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Problems 95

Problems

4.1 Show by direct substitution that E′ = E and B′ = B when the Coulombgauge is made.

4.2 Show by direct substitution that the Schrödinger equation is invariant un-der the Coulomb gauge transformation.

4.3 Show that the Lorentzian function (Eq. (4.36)) is properly normalized foroptical frequencies.

4.4 Estimate the relative magnitudes of Γo and ΓD (see Appendix B.10) thatdetermine when the Lorentzian and Doppler profiles dominate the line shape.Determine the peak of the absorption when Γo = ΓD.

4.5 Using the equations given in Appendix B.10, show that when ω = ωo theline shape takes the form

S(ωo) =(

4 ln 2π

)1/2 eb2

ΓDerfc(b), (4.64)

where the complementary error function is given by

erfc(x) =2√π

∫ ∞

xe−u2

du. (4.65)

4.6 Using the result of Problem 4.5, estimate the peak absorption magnitudesin the limit where the Lorentzian and Doppler widths dominate.

4.7 Carry out the steps to determine the transition amplitude for two-photon

absorption,⟨

f∣∣∣T(2)(∞, ∞)

∣∣∣⟩, and then write an expression for the transition

rate. Estimate the rate for two-photon absorption between the 2s 2S and 3d 2Dstates in Li. Assume that the dominant intermediate state is the 2p 2P state.You can use the values in Atomic Transition Probabilities: Hydrogen ThroughNeon, NSRDS-NBS 4, Vol. 1 (which can be found online at Ref. [9]) to do thisproblem.

4.8 Starting from the Hamiltonian

H =1

[p − eA(t)

]2 − eφ + V′, (4.66)

where p is the electron momentum, µ the reduced mass, e the elementarycharge, A the vector potential, φ the scalar potential and V (= −eφ + V′) thestatic potential in which the electron moves, complete the steps to show thatthe radiation Hamiltonian takes the form

Hrad =iheµ

A(t)2 · ∇ +e

2µA(t) · A(t).

Page 113: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

96 Further Reading

Make a gauge transformation of the second kind,

ψ (r, t) = ei ehr·A(t)Ψ (r, t) ,

and show that the radiation Hamiltonian can also take the form

Hrad = −er · E(t).

4.9 Carry out the steps to verify the M1 and E2 selection rules of Eqs. (4.55)and (4.56).

4.10 Use the Wigner–Eckart theorem to determine the selection rules for atwo-photon transition in one-electron atoms.

4.11 Derive the restrictions on ∆L, ∆J and ∆S in LS coupling for one-photonand two-photon transitions in the dipole approximation.

4.12 Show that the following pairs of equations lead to the same physics:

a) Eqs. (4.27) and (9.37) and

b) Eqs. (4.29) and (9.49).

Page 114: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

97

5Diatomic Molecules

A diatomic molecule is an atom withone atom too many.

– Arthur L. Schawlow

In this section we will look at the simplest case of two atoms working in con-cert – the diatomic molecule. The collective motion of the two atoms will leadto symmetries and electronic states related to but distinct from atomic states.Just as in atoms, we will be able to define electronic configurations for the elec-trons from which molecular states and potential curves can be derived. Oneof the major differences between an atom and a diatomic molecule is that thestates in which the electrons move will depend upon an additional parameter,the internuclear separation, R, between the two atomic nuclei. This parameterR will, hence, couple the electronic motion to the nuclear motion. The nucleican vibrate and rotate, introducing new structure. Electronic states will notbe the result of a single configuration, in general, which will again produceconfiguration interaction. While we will not explore them in this book, thisinteraction leads to interesting radiationless decay channels – autoionizationand predissociation.1

We will employ a simple model to classify and analyze these systems. Thismodel has proven itself over the years to be able to capture most of the essenceof the physics of the diatomic system. The cartoon in Fig. 5.1 helps us to pic-ture the treatment. Specifically, the atoms are treated as heavy masses in closeproximity (∼ 0.1 nm) attached to each other via massless springs. The energyof the system will have two parts, an electronic part and a nuclear part. Thenuclear part will be determined by a combination of vibrating and rotatingdumbbells. The fact that the atoms can rotate and vibrate at the same timemeans that the vibrations will not be exactly harmonic nor will the moment ofinertia remain fixed during rotation. The electrons, which are shared betweenthe nuclei, will move in a distorted, cylindrically symmetric potential estab-lished by the Coulomb fields of the two atomic nuclei and the other electrons.

1) Readers interested in these phenomena are encouraged to readRef. [43].

Page 115: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

98 5 Diatomic Molecules

Fig. 5.1 The classical model of a diatomic molecule, two atoms sepa-rated by a distance R, that we will employ in our discussion in this sec-tion. The two atoms are assumed to be attached as if by a masslessspring allowing them to vibrate. In addition, the system is free to rotateabout an axis perpendicular to R that contains the center of mass.

5.1The Hamiltonian

We begin by looking at the Hamiltonian for the general case, which takes theform

H = − h2

2

(∇2

1M1

+∇2

2M2

)︸ ︷︷ ︸

A

− h2

2me

N

∑i=1

∇2i︸ ︷︷ ︸

B

+Z1Z2e2

4πεoR+

e2

4πεo

N

∑i<j

1rij︸ ︷︷ ︸

C

− e2

4πεo

N

∑i=1

Z1∣∣∣−→r i −−→R 1

∣∣∣ +Z2∣∣∣−→r i −−→

R 2

∣∣∣

︸ ︷︷ ︸D

+ ζ, (5.1)

where A is the nuclear kinetic energy, B the electronic kinetic energy, Cthe electrostatic interaction, D the Coulomb potential and ζ is composed ofsmaller relativistic interaction terms, such as spin–orbit, spin–spin, etc., thatwe will not consider explicitly. For a neutral diatomic molecule, N = Z1 + Z2.In A, ∇2

1,2 acts only on the nuclear coordinates, R1 and R2, while in B, ∇2i

acts on the electronic coordinates, ri; R = |R1 − R2| is the internuclear sepa-ration and rij is the distance between electrons i and j. In the center-of-masscoordinates, expression A can be written as

h2

2 (M1 + M2)∇2

cm +h2

2µ∇2, (5.2)

where the first term corresponds to the translation of the entire molecule,which does not affect the internal energy, and the second describes the relativemotion of the two nuclei; ∇ only operates on R and the nuclear reduced massis represented by µ (= M1M2/ (M1 + M2)). In center-of-mass coordinates the

Page 116: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.2 Born–Oppenheimer Approximation 99

Hamiltonian takes the form

H = − h2

2µ∇2 − h2

2me

N

∑i=1

∇2i + V(ri; R1, R2), (5.3)

where to first order

V(ri; R1, R2) =Z1Z2e2

4πεoR+

e2

4πεo

N

∑i<j

1rij

− e2

4πεo

N

∑i=1

Z1∣∣∣−→r i − M2M1+M2

−→R∣∣∣ +

Z2∣∣∣−→r i + M1M1+M2

−→R∣∣∣ . (5.4)

In Eq. (5.4) we have assumed that M1 is to the left of the center of mass andM2 is to the right. Applying this Hamiltonian to Eq. (1.2) yields

h2

2µ∇2 |ψ〉 +

h2

2me

N

∑i=1

∇2i |ψ〉+ (E − V) |ψ〉 = 0. (5.5)

5.2Born–Oppenheimer Approximation

We will seek a separable solution (known as the Born–Oppenheimer separa-tion) of the form

|ψ〉 = |ψN〉 |ψel〉 , (5.6)

where ψN and ψel are the nuclear and electronic wavefunctions, respectively.These wavefunctions are solutions to

h2

2me

N

∑i=1

∇2i |ψel〉 + (Eel − Vel) |ψel〉 = 0 (5.7a)

and

h2

2µ∇2 |ψN〉 + (E − Eel − VN) |ψN〉 = 0. (5.7b)

Equation (5.7a) describes the motion of the electrons when the nuclei are fixedin space. In the center-of-mass coordinates,

Vel(ri; R1, R2) =e2

4πεo

N

∑i<j

1rij

− e2

4πεo

N

∑i=1

Z1∣∣∣−→r i − M2M1+M2

−→R∣∣∣ +

Z2∣∣∣−→r i + M1M1+M2

−→R∣∣∣, (5.8)

Page 117: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

100 5 Diatomic Molecules

which is a function of the positions of all the electrons, ri, as well as the in-ternuclear separation, R. This implies that |ψel〉 and Eel both depend on R.Equation (5.7b) describes the motion of the nuclei in a field produced bythe electrons, Eel, and the electrostatic interaction with each other, VN =Z1Z2e2/4πεoR. Thus, E is the total energy of the molecule because it dependson both nuclear and electronic motion. We define U(R) = Eel + VN , the aver-age electronic potential in which the nuclei move, as

U (R) =Z1Z2e2

4πεoR+ 〈ψel|

e2

4πεo

N

∑i<j

1rij

− e2

4πεo

N

∑i=1

Z1∣∣∣−→r i − M2M1+M2

−→R∣∣∣ +

Z2∣∣∣−→r i +M1

M1+M2

−→R∣∣∣ |ψel〉 . (5.9)

Putting all this together, Eq. (5.5) becomes

h2

(∇2 |ψN〉

)|ψel〉 + (E − Eel − VN) |ψN〉 |ψel〉︸ ︷︷ ︸

A

+h2

[2(∇ |ψN〉

)·(∇ |ψel〉

)+ |ψN〉∇2 |ψel〉

]︸ ︷︷ ︸

B

+h2

2me

N

∑i=1

|ψN〉 ∇2i |ψel〉+ (Eel − Vel) |ψN〉 |ψel〉︸ ︷︷ ︸

C

= 0, (5.10)

where, again, ∇ only operates on quantities inside the parentheses.Equation (5.10) is subject to the constraints of Eqs. (5.7a) and (5.7b), so A and

C are identically zero. It is straightforward to see that B is nearly zero as well.Consider h/2µ∇ |ψel〉, which acts on the nuclear coordinate, R. Specifically,it measures the variation of the electronic wavefunction as R changes. Since−ih∇ |ψel〉 acts over nearly the same dimensions as −ih∇i |ψel〉, it is ∼ pel.Thus, B is of order me/µEel , which is very small. To a good approximationwe can ignore B. Setting B to zero is called the Born–Oppenheimer approx-imation. Essentially, the approximation relies on the fact that in most cases,the electrons move much faster than the nuclei and thus adjust very rapidlyto any small changes in R – the nuclei are assumed to be fixed in space whenestimating the electron dynamics. This is generally true for low-lying boundstates.

Page 118: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.3 Nuclear Equation 101

5.3Nuclear Equation

We will now look at the nuclear equation (Eq. (5.7b)) in some detail. As men-tioned at the beginning, the nuclear dynamics consists of vibration and rota-tion. In general,

|ψN〉 = |ψvib〉 |ψrot〉 . (5.11)

Writing ∇2 in spherical coordinates and substituting Eq. (5.11), the nuclearequation becomes

− h2

[1R

∂r

(R2 ∂

∂r

)− κ2

R2

]|ψvib〉 |ψrot〉 = (E − U (R)) |ψvib〉 |ψrot〉 , (5.12)

where

κ2 =1

sin ϑ

∂ϑ

(sin ϑ

∂ϑ

)+

1sin2 ϑ

∂2

∂ϕ2 (5.13)

is the angular momentum operator. When R is constant (say at R = Re, theequilibrium separation) we have the so-called rigid rotator and the equationreduces to

h2

2µR2o

κ2 |ψrot〉 = (E − U (Re)) |ψrot〉 . (5.14)

The solutions are spherical harmonics, |ψrot〉 = YKm (ϑ, ϕ) (see Appendix B.7),and the rotational energy is2

Erot = (E − U (Re)) =K (K + 1) h2

2µR2e

. (5.15)

Usually, the rotational energy will be written in wavenumbers (see Appen-dix A.2), Erot/hc. Thus, we define the rotational term energy, F (K), which isgenerally taken to be positive, as

F (K) =Erot

hc= BeK (K + 1) . (5.16)

Here, Be is the rotational constant,

Be =h

8π2µcR2e

=h

8π2cIe, (5.17a)

2) We point out that in this section, we use Kto denote the quantum number for rota-tion, as was done in Ref. [44]. This is notthe agreed upon convention used by con-temporary books and papers, however.The international agreed upon symbol for

the rotational quantum number is R, as weshow in Table 5.2. We use K in this tutorialenvironment to avoid possible confusionwith the internuclear separation, which isalso denoted by R.

Page 119: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

102 5 Diatomic Molecules

Fig. 5.2 Rotational energy level structure.

where

Ie = µR2e =

M1M2

M1 + M2R2

e (5.17b)

is the moment of inertia. The spacing between adjacent levels,

∆F (K) = F (K + 1) − F (K) (5.18a)

= Be [(K + 1) (K + 2)− K (K + 1)] (5.18b)

= 2Be (K + 1) , (5.18c)

is not constant but increases with increasing K as shown in Fig. 5.2.

5.3.1Harmonic Approximation of U(R)

If we relax the restriction that R is fixed at Re then we must solve Eq. (5.12) todetermine how R deviates from Re. To that end, we make the substitution

|ψvib〉 =ξ (R)

R, (5.19a)

which leads to

|ψN〉 =ξ (R)

RYKm (ϑ, ϕ) . (5.19b)

Equation (5.12) then becomes

d2

dr2 ξ (R) +2µ

h2

[E − U (R) − K (K + 1) h2

2µR2

]ξ (R) = 0. (5.20)

Page 120: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.3 Nuclear Equation 103

To solve this equation we must know something about U (R), the averagepotential in which the nuclei move. In general there are no analytic solutionsfor U (R), so approximations must be made. If we assume that R does notdeviate too much from its equilibrium value, then we can expand U (R) intoa Taylor series about Re,

U (R) = U (Re) + (R − Re)dU (R)

dR

∣∣∣∣R=Re

+12

(R − Re)2 d2U (R)dR2

∣∣∣∣R=Re

+ · · · .(5.21)

For bound states, U (R) represents a potential well with its minimum locatedat U (Re), so the second term vanishes. Without loss of generality we canchoose an energy scale such that U (Re) = 0, so the first term also vanishes.As a result the potential will be harmonic, described by

U (R) =12

(R − Re)2 d2U (R)dR2

∣∣∣∣R=Re

=ks

2(R − Re)2 (5.22)

to lowest order, near Re. Our radial equation now becomes

d2

dr2 ξ (R) +2µ

h2

[E − ks

2(R − Re)2 − K (K + 1) h2

2µR2

]ξ (R) = 0. (5.23)

When K = 0, the equation reduces to

d2

dr2 ξ (R) +2µ

h2

[E − ks

2(R − Re)2

]ξ (R) = 0, (5.24)

which describes a harmonic oscillator with energy

Evib = h

√ks

µ

(v +

12

). (5.25)

The wavefunctions are given by

ξ (R) = Nve−αR2/2Hv(√

αR)

, (5.26)

where Hv are the Hermite polynomials and α =√

ksµ/h. We identify thequantized vibrational energy as

E(v) = Evib = hcωe

(v +

12

)(5.27a)

or, in wavenumbers,

G(v) =E(v)

hc= ωe

(v +

12

), (5.27b)

Page 121: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

104 5 Diatomic Molecules

where

ωe =1

2πc

√ks

µ=

αh2πc

. (5.27c)

We note that the levels are equally spaced. However, the lowest level, v = 0,does not sit at the bottom of the well, but has zero-point energy, ωe/2.

It is straightforward to see that the rotational energy is considerably smallerthan the vibrational energy. Consider H2 where Re = 0.074 nm and µc2 ≈ 469MeV. If we express hc as 197 eV nm then the rotational energy is

Erot =(hc)2

2µc2R2e

= 7.6 × 10−3 eV (5.28)

and

Erot

hc→ 60.9 cm−1. (5.29)

Using Eq. (5.57) (see Problem 5.1), we have for the spring constant

ks = 2(

e4πεoRe

)e

R2e

= 7.1 × 103 eV/nm2. (5.30)

For v = 0, the vibrational energy is

Evib =hc2

√ks

µc2 = 0.38eV (5.31)

and

Evib

hc→ 3100 cm-1, (5.32)

about 50 times larger.3

5.3.2Beyond the Harmonic Approximation of U(R)

Our harmonic oscillator potential, which is parabolic, is unrealistic becauseit implies that the system remains bound as R → ∞. We know that at somepoint the system separates into two individual atoms. We can improve our ap-proximation by carrying out our expansion of U(R) to higher orders – cubic,quartic, etc. For the first correction we have

E(v) = hc

[ωe

(v +

12

)− ωexe

(v +

12

)2]

, (5.33)

3) The actual vibrational energy is about 2200 cm−1.

Page 122: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.3 Nuclear Equation 105

Fig. 5.3 This figure shows the effect of adding a cubic term (dashedcurve) to the harmonic oscillator potential.

where typically ωe > ωexe > 0.4 It should be recognized that just adding a cu-bic term can lead to unphysical behavior as well (see Fig. 5.3). Consequently,depending upon how far out one wants to determine the curve, additionalterms must be included. Huber and Herzberg [45] provided comprehensivetables that include ωe and ωexe for numerous systems, while Herzberg [44]tabulated values for ωeye ( ωexe) in some cases. Constants can be foundonline as well [46]. Thus, we redefine G(v) as

G(v) = ωe

(v +

12

)− ωexe

(v +

12

)2

+ ωeye

(v +

12

)3

+ · · · . (5.34)

5.3.3Vibrating Rotator

Thus far, we have ignored the R dependence for rotation. That is, we havedecoupled rotation from vibration. This too is unphysical because, since weare modeling the molecule as two masses attached by a spring, the atoms willexperience a centrifugal force as they rotate, which will tend to increase R asthe energy increases. This must be the case to conserve angular momentum.This can be taken into account by adding terms that are nonlinear in K(K + 1)to Eq. (5.16),

F(K) = BvK(K + 1)− DvK2(K + 1)2 + · · · . (5.35)

An approximate expression for Bv can be written in the form of Eq. (5.17a)with 1/R2

e replaced by an average of 1/R2v,

Bv =h

8π2cµ

⟨1

R2v

⟩. (5.36)

4) The sign of the ωexe term is chosen to make ωexe positive definite.Thus, this anharmonicity causes the potential curve to flatten out asshown in Fig. 5.3 – the potential becomes softer.

Page 123: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

106 5 Diatomic Molecules

To account for its v dependence, we write

Bv = Be − αe

(v +

12

)+ γe

(v +

12

)2

+ · · · , (5.37)

where αe > 0 and γe > 0, so that as v → ∞, R → ∞ and the energy ofthe system decreases.5 The subscript e refers to the equilibrium internuclearseparation, Re. In a similar way we write

Dv = De − βe

(v +

12

)+ · · · . (5.38)

Values for Be, De, αe, βe and γe can all be determined spectroscopically.Putting all this together, the expression for the energy of a rovibrational

state is

ν(v, K) = νe + G(v) + F(K), (5.39)

where νe is the energy of the bottom of the electronic state (which is zero forthe ground state).

Before we end this section it is worthwhile pointing out that Dunham [47](see also Refs. [44,48]) introduced a different expression for the energy for thevibrating rotator,

T = ∑ij

Yij

(v +

12

)i

Kj(K + 1)j, (5.40)

which leads to expressions for the vibrational and rotational constants thatdeviate from our earlier expressions. These deviations are quite small andusually ignorable except for H2. Nevertheless, the vibrational–rotational levelwill be given by [48]

ν(v, K) = νe + Y00 + G(v) + Fv(K), (5.41)

where

Y00 =Be

4+

αeωe

12Be+

α2e ω2

e144B2

e− ωexe

4, (5.42)

νe = Te − G0(X). (5.43)

The last expression gives the zero-point-adjusted electronic term energy. ForH2, G0(X) = 2197.3 cm−1 [48].6

5) Here again, the sign of the αe term is chosen to make αe positive defi-nite. In so doing, Bv reduces as the molecule stretches – the potentialbecomes softer, again.

6) Note that the last term in Eq. (5.42) is a correction to the expressiongiven in Ref. [44]. See Ref. [48].

Page 124: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.3 Nuclear Equation 107

5.3.4Analytic Expression for U(R)

The first correction to the harmonic potential leads to an exact expression forthe potential curve called a Morse potential that can be written as

U(R) = Uel + Do[1 − e−a(R−Re)]2, (5.44)

where Uel is a constant that sets the overall energy of the bottom of the po-tential well relative to the zero of the energy scale. For the example shown inFig. 5.4, Uel = 0. For this potential, at Re we have

U(Re) = Uel,

dUdR

∣∣∣∣R=Re

= 0,

d2UdR2

∣∣∣∣R=Re

= 2a2Do ≡ ks.

Substituting the Morse function (Eq. (5.44)) into Eq. (5.20), the nuclear equa-tion takes the form

d2

dR2 ξ (R) +2µ

h

E − Uel − Do

[1 − e−a(R−Re)

]2

− K (K + 1) h2

2µR2

ξ (R) = 0. (5.45)

When K = 0, the solution takes the form [49]

G(v) =Uel

hc+ ωe

(v +

12

)− ωexe

(v +

12

)without any higher-order terms, where

ωe = a(

hDo

πcµ

) 12

,

ωexe =ω2

e4Do =

ha2

4πµc.

Measured or tabulated values of ωe and ωexe can be used to determine Do, thedissociation energy, or to draw the potential curve. One is cautioned, however,that a2 h/4πcµ is not equal to experimental values of ωexe in general.

5.3.5More Accurate Techniques

While the Morse approach is simple and quick to implement, it is not themost accurate means of estimating the potential. The RKR method (named for

Page 125: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

108 5 Diatomic Molecules

Fig. 5.4 The Morse potential, for the ground state of H2 where Uel, theenergy of the bottom of the potential well, is 0 and Do, the dissociationenergy – the height of the well, is 4.46 eV in this case.

its inventors Rydberg [50], Klein [51] and Rees [52]) is a powerful technique.This method relies on knowledge of the energy levels and constants obtainedfrom spectroscopy to relate the potential energy curve, U, to the internuclearseparation. Specifically, the classical inner and outer turning points for U aregiven analytically by

Rmax =(

fg

+ f 2) 1

2+ f , (5.46a)

Rmin =(

fg

+ f 2) 1

2 − f . (5.46b)

The functions f and g, which depend on the energy of the molecule, E =Gv + Bv J(J + 1), are defined in terms of partial derivatives of

S (E, κ) =1

π√

∫ I′

0

√E − U (I, κ)dI, (5.47)

where U(I, κ) is the sum of the vibrational and rotational energies of the mol-ecule,

ωe

hI − ωexe

h2 I2 +ωeye

h3 I3 +1

R2e

κ − De

B2e R4

eκ2 − αe

hωeBeR2e

Iκ + · · · ,

I = h(v + 1/2), κ = h2 J(J + 1)/8π2µ and I = I′ when E = U; f and g arethen

f =∂S∂E

, (5.48a)

Page 126: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 109

and

g = −∂S∂κ

. (5.48b)

We identify I above as the action when we recognize that the RKR method isthe inverse of the JWKB procedure. In the JWKB procedure Ue f f (R) is knownand E is sought. This involves quantizing the action,

I =∮

pdq = h(v + 1/2)

=∮ √

2µ[

E − Ue f f (R)]dR, (5.49)

where the integration is over a complete cycle – Rmin to Rmax and back.One advantage of the RKR approach is that it is possible to generate poten-

tial energy curves without a good value of the dissociation limit. At the sametime, one of the drawbacks of this method is that it does not give an estimateof the dissociation limit. Traditionally, a Birge–Sponer extrapolation [53] givessuch an estimate. The technique uses the spacing between vibrational levels,which is given by

∆Gv+1/2 ≡ G(v + 3/2) − G(v + 1/2) = ωe − 2(v + 1)ωexe, (5.50)

up to the first anharmonic term (the cubic term), to determine how many vi-brational levels a potential well can hold. Since the spacing decreases linearlywith v, it is possible to plot a few values of ∆G(v) and extrapolate the curve to∆Gv+1/2 = 0 to determine vmax and hence the onset of dissociation (i.e., D0). Itis well known that near the dissociation limit the plot of the spacing deviatesfrom the extrapolation because of the influence of long-range interatomic po-tentials. LeRoy and Bernstein [54] and Stwalley [55] have developed methodsto correct for these long-range forces, giving more accurate results.

5.4Electronic States

In this section, we will look at the electronic potentials determined byEq. (5.34) in more detail. Our primary goals are to determine: (1) the rele-vant quantum numbers (conserved quantities) for bound electronic motion,(2) the symmetries that distinguish various states and (3) the selection rulesfor transitions (which we will investigate in the next chapter). The first twowill allow electronic configurations similar to what we have for atoms to beconstructed while the third will enable molecular spectroscopy.

Before we get to the details, however, we can make a few general commentsthat will allow us to use what we already know to reduce our workload. First,

Page 127: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

110 5 Diatomic Molecules

since the diatomic molecule is composed of two nuclei in close proximity, thespherical symmetry we have for atoms will be broken. As a consequence, itis necessary to revisit angular momentum and angular momenta coupling incylindrical symmetry. Second, the energies of the electronic states depend onthe internuclear separation as shown in Fig. 5.4. We will find that some statessupport binding, a lowering of the energy as R (the internuclear separation)decreases, while others will not. Equation (5.7a) must be solved to determinethe electronic energy, Eel. Third, the spacings between the electronic stateswill be on par with those of the corresponding atomic states (see below) of theconstituent atoms and, in general, much larger than the spacing between ro-tational and vibrational levels. Finally, each electronic state will be associatedwith a distinct pairing of atomic states. We will exploit this to build molecularconfigurations (molecular orbitals) from linear combinations of atomic config-urations (atomic orbitals), which will provide a rudimentary understanding ofmolecular structure.

5.4.1Angular Momenta in Cylindrically Symmetric Fields

Electrons bound to molecules have orbital and spin angular momenta, as theydo in atoms, which couple to the molecular rotation and spin to produce a to-tal angular momentum for the system. This coupling can occur in a variety ofdifferent ways. Frequently, the molecular analogue to LS coupling is appro-priate where all the electronic orbital angular momenta couple to produce Las do the electronic spins to produce S. This occurs, as we found with atoms,when the spin–orbit interaction is small compared with the electrostatic in-teraction. Electronic orbital angular momentum, however, is a good quantity(constant of the motion) in spherically symmetric fields. In our study of theStark effect (see Sections 3.1.1.1 and 3.1.1.2), we found that L is not constantwhen an electric field is applied, which breaks the spherical symmetry. Specif-ically, we found that

• the applied field reduces the spherical symmetry to cylindrical symme-try;

• L is not conserved;

• ML, the projection of L along the electric field direction, is a good quan-tum number and

• part of the 2L + 1 degeneracy for a given n shell is removed, but ±Mstates are still degenerate – all states with L ≥ 1 are doubly degenerate.7

7) This remaining degeneracy can be seen by reversing either the senseof the orbital motion of the electrons or the direction of the electricfield; ML → −ML, but the energy would be unchanged.

Page 128: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 111

LS

Fig. 5.5 The projections of the orbital and spin angular momenta ofthe electrons of a diatomic molecule onto the internuclear axis, defin-ing Λ, Σ and Ω.

The electric field produced by the two atoms also has cylindrical symmetry.Thus, the results from the Stark effect provide a framework to begin our anal-ysis.

5.4.1.1 Orbital Angular Momentum

Since L is not conserved, it would be better to classify the states according toML, which is conserved. However, since there is a degeneracy between thepositive and negative values of ML (see Eqs. (3.14) and (3.15)), it makes moresense to classify the states in terms of |ML|.8 Thus, we define

Λ ≡ |ML|, (5.51)

the projection of the orbital angular momentum of the electrons along the in-ternuclear axis as shown in Fig. 5.5. Each Λ corresponds to a distinct electronicstate. For each L there will be L + 1 distinct states (e.g., for L = 2 the possibleprojections along the internuclear axis are Λ = 0, 1 and 2.) Table 5.1 gives thedesignations of the first few molecular states.

Tab. 5.1 Molecular state designation

Λ0 1 2 3 . . .

Molecular designation Σ Π ∆ Φ . . .Atomic analogy S P D F . . .

5.4.1.2 Spin Angular Momentum

The aggregate electronic spin, S, unlike L, remains a good quantum opera-tor in the cylindrical field because the Hamiltonian, due to the fact that theelectric field of the nuclei does not involve the electron’s spin (see for exam-ple Eq. (3.10)). In addition, its projection onto the internuclear axis, MS ≡ Σ

8) We point out that building up the molecule at large internuclearseparation leads to the same conclusion.

Page 129: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

112 5 Diatomic Molecules

(see Fig. 5.5), is also a good quantum number.9 Unlike Ω, however, Σ hasboth positive and negative values. States of different spin are degenerate in Σstates (Λ = 0 states), as they are for S states in atoms. However, when Λ = 0,the motion of the electrons about the nuclei leads to 2S + 1 values for Σ (LScoupling) since it can be positive or negative (i.e., the projection of S can beparallel or antiparallel to the magnetic field produced by the electrons).

5.4.1.3 Multiplet Splitting

The multiple splitting for Λ = 0 is proportional to the interaction of the elec-tronic spin and the magnetic field, which has the form µ · Helec (see, for ex-ample, Eqs. (3.17) and (3.20)). This leads to a multiplet splitting that will besimilar to the Zeeman effect, which will be proportional to σ. As a first ap-proximation, the term values will obey

Te = T0 + AΛΣ, (5.52)

where T0 is the energy of the term in the absence of spin and A the propor-tionality constant independent of Λ and Σ. It is clear that unlike the multipletsplitting in atoms, this splitting is equally spaced.

5.4.1.4 Total Angular Momentum

Tab. 5.2 Table of standard symbols for angular momentum operators and quantum numbers,total and projections along the internuclear axis, used to describe the structure and dynamicsof diatomic molecules. This notation is consistent with that used in Refs. [43, 56]. We point outthat R that appears in this table to denote nuclear rotation is identical to K used in Ref. [44]and in earlier sections of this chapter (see footnote on page 101). Entries with “· · · ” indicateno standard notation.

Operator Description Total Projection

quantum number quantum number

R Nuclear rotation R ZeroL Electronic orbital L ΛS Electronic spin S ΣI Nuclear spin I · · ·J R +L +S J Ω = |Λ + Σ|F Total (J +I) F · · ·Ja

L +S Ja · · ·N R +L =J −S N ΛO J −L = R +S O · · ·

9) This choice for the name of the projection is unfortunate and mustnot be confused with a Σ (Λ = 0) electronic state.

Page 130: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 113

Tab. 5.3 Nomenclature example for Λ = 1, Π state.

S = 32 → Quartet state

Λ + Σ = − 12 , 1

2 , 32 , 5

2

Four states → 4Π− 12

, 4Π 12

, 4Π 32

, 4Π 52

By definition, the total angular momentum, J, in the absence of externalfields is always conserved. Regardless of how the electronic angular momentacouple to the internuclear axis,J is a result of coupling of the electronic andthe rotational angular momenta. According to Eq. (5.15), rotation is quantizedwith integer quantum numbers. We used the vector K in Eq. (5.15) to desig-nate pure rotation. The more conventional notation of Table 5.2 for pure rota-tion is R. For the rest of this chapter, we will assume that K ≡ R. In Table 5.2,J is due to coupling of R, L and S, where R is the total angular momentumignoring spin.

5.4.1.5 Labeling Nomenclature

We will define a total electronic angular momentum along the internuclearaxis as

Ω = Λ + Σ, (5.53)

which ranges from Λ − |Σ| to Λ + |Σ|.10 For light systems L and S coupleseparately to the internuclear axis as shown in Fig. 5.5. Following a labelingsystem similar to that used to define LS-coupled terms in atoms, the nomen-clature for electronic states is

2S+1ΛΩ.

Table 5.2 is an example for a Λ = 1 state. The independent coupling of Land S to the internuclear axis is just one example of how they can couple. Insome systems,L couples to R, the rotation vector, to form a resultant that thencouples toS to form the total angular momentumJ.

10) This convention agrees with contemporarypapers (see Refs. [56] and [43] for example)but differs from that of Ref. [44] where Ωis defined as |Λ + Σ|. The advantage ofthe latter definition is that Ω is positive

definite. The disadvantage is that for somecases, distinct states with the same valuesfor Λ and S can have the same value for Ω.This will not be the case with the definitionwe will use.

Page 131: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

114 5 Diatomic Molecules

5.4.2Angular Momenta Coupling: Hund’s Cases

There are several possible ways in which the various angular momenta cancouple to produceJ. Typically, these are organized into five major idealizedcategories called Hund’s cases (a)–(e). We point out that while real moleculescan fall between cases or change from one case to another as the nuclear ro-tation increases or the system becomes more excited, Hund’s cases providea good approximation for grouping or a starting point for analysis. We willreview four of them: cases (a)–(d). Table 5.4 summarizes the different ways inwhich the various angular momenta can couple in these four cases. A morethorough discussion of all five cases can be found in Herzberg’s book on di-atomic molecules [44].

Tab. 5.4 Summary of Hund’s cases (a)–(d).

Case Precoupling to axis Order of coupling(a) L → Λ & S → Σ Λ,Σ → Ω, R → J(b) L → Λ Λ, R → N,S → J(c) L, S → JA → Ω Ω, R → J(d) Nothing L, R → N,S → J

5.4.2.1 Hund’s Case (a)

In case (a) the interaction between the nuclear rotation and the electronic or-bital and spin angular momenta is weak. Thus, L and S couple to the inter-nuclear axis separately to give Λ and Σ, respectively, which then combine to

Fig. 5.6 Hund’s case (a).

Page 132: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 115

Fig. 5.7 Examples of the rotational structure for 2Π (left) and 2Σstates in cases (a) and (b), respectively. In case (a), S = 1/2, Σ = 1/2and Λ = 1, leading to two states, 2Π1/2 and 2Π3/2. Note that theJ = 1/2 level is missing for the 2Π3/2 state. In case (b), S = 1/2as well, but there is only one state. However, each R level (rotationallevel), except the lowest level, splits into two J levels.

give Ω. This was the situation in Fig. 5.5 for diatomics composed of lightatoms. Figure 5.6 shows that the resulting Ω couples to R to giveJ. For case(a), the rotational energy levels are given by

Fv(J) = Bv[J(J + 1) − Ω2] + . . . ,

while J takes on values

J = Ω, Ω + 1, Ω + 2, . . . ,

Page 133: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

116 5 Diatomic Molecules

and cannot be less than Ω. Since J must be positive, J ≥ Ω demands thatsome of the values of J may be missing from the spectrum. An example of therotational structure for a 2Π state in this case is shown in Fig. 5.7.

5.4.2.2 Hund’s Case (b)

When S = 0 but couples very weakly, or not at all, to the internuclear axis,case (b) applies. Weak coupling is easy to understand when Λ = 0 (Σ states),especially for the the case of zero nuclear spin. The internal magnetic fieldwith which the electronic spin can interact will be very weak. Consequently,in case (b),S tends to couple to the rotation vector instead of the internuclearaxis. Case (b) can also apply in situations where Λ = 0; L couples to theinternuclear axis but S does not. As sketched in Fig. 5.8, Λ and R couple togive N, which takes on integer values Λ + R,11

N = Λ, Λ + 1, Λ + 2, . . . .

Finally, N couples to S to produceJ, which takes on values

J = |N − S|, |N − S + 1|, . . . , (N + S − 1), (N + S).

When N ≥ S, there will be 2S + 1 J components for each N. The rotationallevels for 2Σ and 3Σ states are given in Table 5.5 and displayed graphically inFig. 5.7 for a 2Σ state.

Fig. 5.8 Hund’s case (b).

Page 134: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 117

Tab. 5.5 Rotational energy levels in Hund’s case (b) for S = 1/2 and S = 1. The splittingconstants γ (S = 1/2, S = 1) and λ (S = 1) are usually very small compared with Bv.

S = 12

FN+ 1

2(N) = BvN(N + 1) + 1

2 γN

FN− 12(N) = BvN(N + 1) − 1

2 γ(N + 1)

S = 1

FN+1(N) = Bv[N(N + 1) + (2N + 3)]− λ + γ(N + 1)−√(2N + 3)2B2

v + λ2 − 2λBv

FN(N) = BvN(N + 1)FN−1(N) = Bv[N(N + 1) − (2N − 1)]− λ − γN

+√

(2N − 1)2B2v + λ2 − 2λBv

5.4.2.3 Hund’s Case (c)

For heavy atoms, the spin–orbit interaction becomes larger than the electro-static interaction and L and S are expected to couple first to give Ja, beforecoupling to the internuclear axis. Its projection now defines Ω, which couplesto the rotation to give the total angular momentumJ. This constitutes case (c)and is illustrated in Fig. 5.9. In this case, however, the states are not classifiedin terms of Λ and Σ any longer because they are no longer conserved. Rather,they are classified by Ω, which is conserved. The energies of the rotational

11) Clearly, R ≡ N when Λ = 0.

Fig. 5.9 Hund’s case (c).

Page 135: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

118 5 Diatomic Molecules

Fig. 5.10 Hund’s case (d).

levels and the values over which J ranges are the same as in case (a). We pointout that some investigators doubt that this case exists [57].

5.4.2.4 Hund’s Case (d)

The final case we will mention is that whereL couples to the rotation vector, R,instead of the internuclear axis, to give N. In this case there are 2L + 1 valuesfor N, which range over

N = (R + L), (R + L − 1), . . . , |R − L|.The rotational energies are given by

Fv(R) = BvR(R + 1),

with each R level split into 2L + 1 components. While the totalJ results whenN is coupled toS, this coupling is usually very small and neglected, as we doin Fig. 5.10.

5.4.3Molecular Symmetries: Electronic Motion

We have already seen that the electrons in diatomic molecules, and all linearmolecules for that matter, move in a cylindrically symmetric field. We nowwant to look at what impact this symmetry has on the electronic states.12 In

12) We acknowledge that some of the symmetry arguments used in thissection were inspired by those presented by G. Herzberg in Spectra ofDiatomic Molecules [44].

Page 136: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 119

particular, we will determine how the wavefunctions associated with the var-ious states transform under inversion, reflection and exchange of nuclei.

5.4.3.1 Inversion Symmetry

In a rectangular system, inversion means that r → −r or x → −x, y → −yand z → −z. What does it mean in cylindrical coordinates? We can determinethis with the help of Fig. 5.11, where the two positive charges are found alongthe z-axis, the internuclear axis. If we start at (a), inversions will take us to(c). We can make the trek via two paths. Path 1: rotate the system by π aboutthe y-axis to reach (b) and reflect through the xz-plane. Path 2: reflect throughthe xz-plane to reach (d) and rotate by π about the y-axis. The rotation takesz → −z (and x → −x) while the reflection takes φ → −φ (and y → −y). It isclear that applying either sequence of operations a second time brings us backto (a), which means that at best

ψel ⇒ ±ψel

under inversion.13 The possibility of a phase change to the wavefunction putsrestrictions on the values that Λ can assume. To see what these restrictionsare, consider two functions in cylindrical coordinates,

χ+(zi, ρi, ∆φi) and χ−(zi, ρi, ∆φi),

where i labels the electrons and the difference between χ± is the sign in frontof ∆φi. We define

∆φi = φi − φ1,

where φ1 is the angle for the first electron and ∆φi is the relative azimuthal an-gle between the first electron and all the others.14 The electronic wavefunctioncan be defined as

ψel = χ+eiΛφ1

= χ−e−iΛφ1 .(5.54)

Both representations of ψel are eigenfunctions of

Lz =hi

∂φ1,

13) Since the act of rotation cannot change the physics, we could havereflected the system through the yz-plane instead of the xz-planeand then rotated the system about the z-axis by π to get to (b), with-out introducing any new physics. In fact, we can reflect throughany plane containing the z-axis followed by an appropriate rotationabout the z-axis and the physics would remain the same.

14) The physics must also not depend on the initial coordinate systemso that the value of φ1 is irrelevant.

Page 137: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

120 5 Diatomic Molecules

Fig. 5.11 Inversion via rotation and reflection. Path 1: starting at (a),rotate by π about the y-axis to reach (b) followed by reflection throughthe xz-plane to reach (c). Path 2: starting at (a), reflect through thexz-plane to reach (d) followed by a π rotation about the y-axis toreach (c).

the angular part of Schrödinger’s equation. Thus, these functions describerotations in the opposite sense and Λ represents the projection of the orbitalangular momentum onto the Z-axis. To see that Λ must be an integer, assume

Page 138: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 121

that it is a half-integer. If we were to rotate the system through 2π, then weshould be back to where we started. However, for half-integer Λ,

ψel(φ1 = 2π) → ψel(φ1 = 0)e±iπ = ψel(φ1 = 0).

Now consider a reflection through the xz-plane. This takes φ → −φ, whichimplies that

χ+eiΛφ1 χ−e−iΛφ1 .

Since the general solution to Schrödinger’s equation takes the form

ψel = aχ+eiΛφ1 + bχ−e−iΛφ1 , (5.55)

the electronic wavefunctions fall into two categories – Λ = 0 and Λ = 0. Wewill look at the symmetry properties of each category separately.

5.4.3.2 Reflection Symmetry and Σ States (Λ = 0)

In this case, upon reflection

ψel → ±ψel,

which implies that

ψ±el = χ+ ± χ−,

to an arbitrary phase. The Σ+ states correspond to

ψ+el = χ+ + χ−,

and do not change sign upon reflection. The Σ− states correspond to

ψ−el = χ+ − χ−,

and do change sign upon reflection. In general, the two states are not energydegenerate and must be treated as independent electronic states. This can beshown as follows. Assume that

H|ψ±el >= E±.

Then

E+ =< ψ+|H|ψ+ >

=∫

χ∗+Hχ+d3r +

∫χ∗

+Hχ−d3r

+∫

χ∗−Hχ+d3r +∫

χ∗−Hχ−d3r,

Page 139: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

122 5 Diatomic Molecules

while

E− =< ψ−|H|ψ− >

=∫

χ∗+Hχ+d3r +

∫χ∗−Hχ−d3r

−∫

χ∗+Hχ−d3r −

∫χ∗−Hχ+d3r,

so that E+ = E−.

5.4.3.3 Reflection Symmetry and Λ = 0 States

When Λ = 0, the ensuing positive- and negative-symmetry states (such asΠ±, ∆±, etc.) will be energy degenerate. Following the same approach asabove, we have

E± =< ψ±|H|ψ± >

=∫

χ∗+Hχ+d3r +

∫χ∗−Hχ−d3r

±∫

χ∗+Hχ+e−2iΛφ1d3r,

±∫

χ∗−Hχ+e2iΛφ1d3r.

Since χ± does not depend on φ1 explicitly, we can break up the integrationinto cylindrical components, z, ρ and φ and carry out the φ integral first. Thelast two terms above become∫

zdz∫

ρχ∗±Hχ±ρdρ

∫ 2π

0e∓2iΛφ1dφ1.

Since Λ is an integer,∫ 2π

0e∓2iΛφ1dφ1 = 0,

which means that E+ = E−. As a consequence of this, two-fold degeneracy ofthe idealized molecule allow us to drop the ± designation on Λ = 0 electronicstates.

5.4.3.4 Exchange of Nuclei

As we mentioned above, when dealing with homonuclear diatomics we needto consider the wavefunction upon the exchange of the identical nuclei. Aswith reflection, the wavefunction can either remain the same or change sign.When the sign remains the same the electronic state is known as a gerade state(g) after the German word for even. When the state changes sign with the

Page 140: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 123

Tab. 5.6 Summary of Hund’s cases (a)–(d).

Λ 0 1 2State Σ±

g,u Πg,u ∆g,u HomonuclearΣ± Π ∆ Heteronuclear

exchange of identical nuclei the state is called an ungerade state (u, German forodd). We summarize the symmetry properties of electronic states in Table 5.6.

5.4.4Molecular Symmetries: Nuclear Motion

Now that we have a basic understanding of the symmetry of the electronicpart of the wavefunction we need to examine the nuclear part. Here, we areconcerned with vibration and rotation of the molecule. Recall that the nuclearpart of the wavefunction is given by ψN = ψvib · ψrot, where

ψvib =1

Rnζ(Rn).

But, since ψvib only depends on | Rn| (the internuclear separation)15 it alwayshas positive symmetry with respect to reflection and inversion. The rotationalwavefunction is proportional to the spherical harmonics, YR,m(Ω), and thushas the following symmetry under inversion:

even R ψrot → +ψrot,odd R ψrot → −ψrot,

5.4.5Molecular Symmetries: Herzberg Bookkeeping Diagram

We are going to want to examine the selection rules for transitions betweenmolecular states. For the most part, we will focus on dipole transitions wherethe parities of the two states involved must be opposite. For high-resolutionstudies, we will be interested in transitions between individual rotational lev-els. Consequently, we need to consider the parity of the rotational levels ofelectronic states.

We will first consider a heteronuclear system with Λ = 0. In this case,

Σ+ ⇒ even rotational levels have positive symmetryΣ− ⇒ odd rotational levels have positive symmetry

15) We have added the subscript n to help distinguish the internuclearseparation from the rotational quantum number.

Page 141: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

124 5 Diatomic Molecules

To specify the symmetry of the rotational levels it is convenient to use abookkeeping method introduced by Herzberg, which we will refer to as theHerzberg diagram. To demonstrate these diagrams we will consider twoHund’s case (b) states where L couples to the internuclear axis but S doesnot. In case (b) the good quantum numbers will be

Λ, R, N and J,

where

N = Λ + R,

J = N +S.

The Herzberg diagram for the structure of the levels has three lines and canbe written as follows for a singlet (S = 0) state:

(top) N = R = 0 1 2 3

(middle) 1Σ+ ⊕ ⊕ (bottom) J = 0 1 2 3

5.4.5.1 The caseJ = N (S = 0)

The top line of the diagram gives the quantum number for each rotationallevel, R (or N in this case since Λ = 0). The middle line gives the symmetry ofthe rotational levels, positive = ⊕ or negative = . The bottom line gives thetotal electronic quantum number, J, for each rotational level. When S = 1 thediagram takes the form

N = R = 0 1 2 33Σ− ⊕⊕⊕ ⊕⊕⊕

J = 1 0 1 2 1 2 3 2 3 4

5.4.5.2 The caseJ = N + 1 (S = 1)

For the triplet states, there are three different values for J associated with eachrotational level all of which have the same symmetry. There is only one valuefor the lowest level since N = 0.

Tab. 5.7 The symmetry of rotational levels of electronic states for homonuclear molecules.

g: ψel → ψel upon exchange of nucleiu: ψel → −ψel upon exchange of nucleis: after inversion and exchange of nuclei, ψ → ψ

a: after inversion and exchange of nuclei, ψ → −ψ

Page 142: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 125

Now what about homonuclear systems? In this case we have an additionalsymmetry we need to worry about, the exchange of the nuclei leading to ge-rade (g) and ungerade (u) states. Each level will now carry an additional label– symmetric (s) or antisymmetric (a). Levels labeled by s will have overallpositive symmetry. We summarize the symmetries in Table 5.7 and 5.8. As anexample, consider a 3Σ−

g state, again in case (b):

N = R = 0 1 2 33Σ−

g ⊕⊕⊕︸ ︷︷ ︸ ︸ ︷︷ ︸ ⊕⊕⊕︸ ︷︷ ︸a s a b

J = 1 0 1 2 1 2 3 2 3 4

Tab. 5.8 The symmetry of rotational levels of Σ states for homonuclear molecules.

Σ+g : ⊕ levels are symmetric

Σ+u : levels are symmetric

Σ−g : ⊕ levels are symmetric

Σ−u : levels are symmetric

Now we will turn our attention to Λ = 0. We will look at two examples:one in case (b), a singlet (S = 0) state, and one in case (c), a triplet (S = 1) state.Since each electronic state with Λ = 0 is doubly degenerate, each rotationallevel will have both ⊕ and symmetries. Among other things, the two-folddegeneracy means that we no longer have to carry the g and u specificationfor homonuclear molecules. The Herzberg diagram for the 1Π state takes theform

N = 1 2 3 41Π1 ⊕ ⊕ ⊕ ⊕

J = 1 2 3 4

Now in case (c)Ja = L + S (Ω is the projection of Ja onto the internuclear axis)and J = Ω + R. Consider a 3Π state. We will have three different situations,one for each Ω value, 3Π0,1,2. Specifically, for Ω = 0 (J = R) we have

R = 0 1 2 33Π0 ⊕ ⊕ ⊕ ⊕

J = 0 1 2 3

For Ω = 1 (J = R + 1), we have

R = 0 1 2 33Π1 ⊕ ⊕ ⊕ ⊕

J = 1 2 3 4

Page 143: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

126 5 Diatomic Molecules

Finally, for Ω = 2 (J = R + 2) we have

R = 0 1 2 33Π2 ⊕ ⊕ ⊕ ⊕

J = 2 3 4 5

5.4.6Molecular Symmetries: Nuclear Spin

Thus far we have not mentioned the nuclear spin and how it affects the wave-function. For a diatomic system, each nucleus has spin I that combine to givea total nuclear spin T, which takes on values

T = Ia + Ib, . . . , |Ia − Ib|.

There are a total of (2Ia + 1)(2Ib + 1) possible states. For example, for Ia = 1and Ib = 1/2 there will be 3 × 2 = 6 possible states and

T = 3/2︸︷︷︸, 1/2︸︷︷︸4 + 2 = 6 states

For the homonuclear molecule there are (2I + 1)2 states and

T = 2I, . . . , 0.

However, we also need to consider what happens when the two nuclei areidentical. Depending on whether I is an integer or half-integer, the nucleibehave as fermions or bosons.

Consider H2 where I = 1/2. The overall wavefunction will take the form

Ψ = ψelψvibψrotΦspin,

and will transform like a fermion under the exchange of nuclei,

RexchΨ = −Ψ.

That is, Ψ must be antisymmetric. In this particular case,

Ψ = ψel,rot,vib

Σ+

g⊕s

Φas.

To make the overall wavefunction antisymmetric, the even rotational levelsmust combine with the antisymmetric spin states and the odd rotational levelsmust combine with the symmetric spin states.

Page 144: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 127

5.4.6.1 Example

Question: How many of the possible (2I + 1)2 states for H2 are symmetricand how many are antisymmetric?

Solution: To solve this problem we recognize that since H2 is made of up oftwo protons, each with I = 1/2, Ψtot must be antisymmetric with respect toexchange of nuclei. We have two possible orientations for T,

T = 0, 1No. of states = 1, 3 ⇒ Four states all together

We will write them out in the form φ1(m1)φ2(m2), where m1,2 is mI for eachnucleus. The four possible combinations are then

1φ1(+ 1

2 )φ2(+ 12 )

2φ1(− 1

2 )φ2(− 12 )

3φ1(+ 1

2 )φ2(− 12 )

4φ1(− 1

2 )φ2(+ 12 )

The first two are obviously symmetric with respect to the exchange of nuclei.This leads to mT = m1 + m2 = ±1, which is associated with T = 1. Evidently,T = 1, then, must correspond to symmetric levels. The third and fourth com-binations are degenerate pairs that can be combined into a symmetric and anantisymmetric grouping. Specifically,

φ1(+ 12 )φ2(− 1

2 ) + φ1(− 12 )φ2(+ 1

2 ) (symmetric)

φ1(+ 12 )φ2(− 1

2 ) − φ1(− 12 )φ2(+ 1

2 ) (antisymmetric)

Both of these produce MT = 0 with one being associated with the singletcombination, T = 0, and the other with the triplet, T = 1. Clearly, T = 0 givesrise to the antisymmetric terms. In summary, there will be three symmetricterms and one antisymmetric term. Table 5.9 summarizes the general case.Special note: when I = 0, there are no antisymmetric terms and every otherrotational level will be missing!

Tab. 5.9 Statistical weights for rotational levels.

I = half-integer T = odd, associated with symmetric levelsI = integer T = even, associated with symmetric levels(2I + 1)(I + 1) Symmetric statistical weight(2I + 1)(I + 1 Antisymmetric statistical weight

Page 145: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

128 5 Diatomic Molecules

5.4.7Molecular State Labeling Convention

5.4.7.1 Rule 1: Ground State

The ground state is traditionally given the label X. Thus, the ground state forH2 and all the alkalis is

X1Σ+g .

5.4.7.2 Rule 2: Excited States with Ground-State Multiplicity

The first excited state of the same multiplicity as the ground state carries thelabel A. As an example, for Na2 we have

A1Σ+u .

The next state is labeled with B. This would be

B1Πu,

for Na2.

5.4.7.3 Rule 3: Excited States with Different Multiplicity

The excited states with a multiplicity different from the ground state are la-beled with lower case letters. For example, the first triplet state might be

a3Πu.

In general, the energy increases with the alphabet so that

Energy ⇒A < B < C < . . . Ground-state multiplicitya < b < c < . . . First with different multiplicity

As with most rules, there are a few notable exceptions, which stem pri-marily from history. This energy ordering is not followed in H2 and He2,for example. The upper case/lower case convention is not followed in N2where singlet states are labeled a, b, . . . and the triplets A, B, . . ., even thoughthe ground state is a singlet.

5.4.8Molecular Orbital Theory

To complete this introduction to electronic states, we will look at how electronscongregate into configurations. While pure configurations rarely describe the

Page 146: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 129

electrons’ motion exactly, they do provide a framework with which to orga-nize our thoughts, characterize symmetries and determine which states con-tribute to bonding. The relationship between these configurations (labeledwith lower case Greek letters) and the electronic states (upper case Greek let-ters) will be very similar to what we have for atoms. A single electron σ or π,for example, produces a 2Σ or a 2Π potential, respectively. In general, as withatoms, a molecular configuration will be composed of several states. Sincethe molecular symmetry is cylindrical (as opposed to spherical), the electronicstate (potential) composed of more than one electron is a bit more complicatedto construct than for atoms. It is the goal of this section to look at how to buildthe molecular state. In Section 1.2.3 we introduced the atomic orbital (AO)to represent the one-electron wavefunction. The corresponding one-electronwavefunction in the molecular case is called a molecular orbital (MO). In atoms,see Chapter 2, a state was represented by a collection of atomic orbitals. In asimilar fashion, we will represent a molecular state by a collection of molecu-lar orbitals.

The electrons, and hence the MOs, are characterized by one good quantumnumber, λ, the projection of the orbital angular momentum onto the inter-nuclear axis. This quantum number assumes integer values with a namingconvention similar to that for atomic states,

λ = 0 1 2 3σ π δ φ

Molecular orbitals sometimes carry two additional labels, the atomic nand l, which are not constants of the motion in general. These labels serveprimarily as indicators of the atomic limit from which the MO was derived.There are two possible limits from which to build the MO known as the unitedatom and separated atom limits. To appreciate these limits, let us build someMOs for H2.

In the united atom limit, we would start with an idealized He atom (onewith just two protons and no neutrons) in a well-defined nl configuration (i.e.,1s2, 1s2s, 1s2p, etc.) and imagine separating the nuclei until the equilibriuminternuclear separation, Re, for H2 is reached. If this separation is done adi-abatically, the electrons will constantly adjust themselves until finally, at Re,they will settle into a perturbed, cylindrically symmetric potential created bytwo charge centers.16 While there are two distinct charge centers, there is onlyone well at the equilibrium separation. However, the perturbation is just theStark effect. Recognizing this allows us to apply what we learned from theStark effect (see Section 3.1.1.1) to determine how the degeneracy among thenl components of the H atom is removed. The 2p atomic state, for example,

16) Of course for H2 the electrostatic interaction between the electronscontributes to the potential as well.

Page 147: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

130 5 Diatomic Molecules

splits into 2pm=0 (≡ 2pσ) and 2pm=±1 (≡ 2pπ) components. Thus, we des-ignate the new state or states by nlλ in this limit, where λ = |m|. Since wesplit He into two identical H atoms, we must account for the the g/u symme-try of the wavefunction upon inversion or exchange of nuclei. To determinehow the symmetry of the atomic state maps to the molecular state, we usethe results reported long ago by Wigner and Witmer [58], who demonstratedthat the molecular orbital has the same symmetry as the atomic configurationfrom which it was derived. Specifically, even-parity (odd) states have g (u)symmetry [58] (see also Ref. [44]). Written in descending order, the first fourconfigurations in the united atom limit with at most one excited electron are

1s2p →

(1sσg)(2pπu)(1sσg)(2pσu)

1s2s → (1sσg)(2sσg)1s2 → (1sσg)2

The g/u symmetry does not exist for heteronuclear systems and, thus, theselabels are omitted from the MOs that describe them. Again, since the electronis a spin-1/2 fermion, no two can have identical quantum numbers. Thus, σ

orbitals can hold up to two electrons (two different orientations of the spin)and π (and higher) orbitals can hold up to four electrons. The λ > 0 orbitalshold four electrons because λ ≡ |m| and m can be positive or negative. Con-sequently, there are two spin states for +m and two spin states for −m values.

In the opposite extreme, the separated atom limit, two H atoms are broughttogether adiabatically from infinity until their nl configurations, well definedat infinite separation, are again perturbed at Re. The molecular states gen-erated in this approach are distinguished from the united atom approach byplacing the nl label after λ, λnl. Since the two H atoms are again identical,when brought together to form molecular states, a pair of states will alwaysresult, one being symmetric (g) and the other antisymmetric (u) depending onhow the AOs are combined. The first few molecular orbitals in descendingorder with at most one excited electron are

1s + 2p →

(σg1s)(σu2p)(σg1s)(πg2p)(σg1s)(πu2p)(σg1s)(σg2p)

1s + 2s →

(σg1s)(σu2s)(σg1s)(σg2s)

1s + 1s →

(σg1s)(σu1s)(σg1s)2.

Again, the g/u labels are dropped for the heteronuclear MOs and replacedwith labels on the nl to indicate which atom belongs to which atomic state;

Page 148: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 131

(σ1sA)(π2pB), for example. The ordering of MOs above is not sacred, butdepends on the system.

Figure 5.12 compares the MOs constructed in the two limits. It is clear thatthe orderings on the left and right do not match up. Consider, for example,the first few states on the left – Σg (= σgσg), Σg (= σgσg), Σu (= σgσu), etc.However, on the right we have Σg (= σgσg), Σu (= σgσu) and Σg (= σgσg), etc.While the lowest level for both limits is a σg orbital occupied by two electrons(i.e., σ2

g ), the similarity stops there. Since symmetry and λ are sacred, a specificnlλg (nlλu) state of the united atom construction must connect adiabaticallyto a specific λgnl (λunl) state of the separated atom construction. Assumingthat the states are not degenerate, connecting states of the same symmetry andλ implies that the energy of a level changes as the internuclear axis changes.To connect the two limits adiabatically, we forbid two curves representing or-bitals of the same symmetry from crossing. The correlation between the twoconstructions is indicated by dashed lines in Fig. 5.12.

To see that these correlations makes sense, we need to say a bit more abouthow an orbital is constructed mathematically. This will allow us to comparethe charge distributions for specific orbitals generated from the two construc-tions.

5.4.8.1 United Atom Construction

To lowest order, an nlλg orbital for a molecule X2 is given by separating aconfiguration of an atom Y, where Y contains twice as many protons and elec-trons as atom X. For our H2 example above (X = H and Y = He, albeit withjust two protons), we generate the 1sσg orbital by starting with He in the 1s2

configuration. Separating the two protons leads to the 1sσg orbital occupiedby two electrons. The charge distribution for the orbital occupied by one elec-tron (i.e., the square of the wavefunction describing the electron in this orbital)is shown in the bottom row of Fig. 5.13, second from the left. The orbital hastwo distinct regions of charge centered on the protons, with charge spillingbetween them. The negative charge between the two positive centers is theglue holding the system together. Thus, this orbital contributes to bonding ofthe molecule and, hence, is called a bonding orbital. It is interesting to note thatif one views the charge distribution along the internuclear axis (the z-axis) thedistribution looks like a 1s state.

The first and second rows from the top in Fig. 5.13 show the constructionof the 3dπg and 2pπu orbitals respectively from atomic 3dm=±1 and 2pm=±1states. The charge distributions show that the 3dπg orbital does not contributeto bonding (an antibonding orbital) while the 2pπu orbital does (a bondingorbital). When either of these orbitals is viewed along the internuclear axis,they resemble 2p states.

Page 149: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

132 5 Diatomic Molecules

Fig. 5.12 Correlation between molecular orbitals derived from unitedand separated atom limits with one excited electron. Only the excitedelectron is indicated; each level has an additional 1s electron such thatthe ground-state limit is 1s2 and 1s + 1s for united and separated atomconstructions. The dashed lines are to guide the eye only and do notrepresent the energy of the orbitals in going from one limit to the other.

5.4.8.2 Separated Atom Construction

The σg1s orbital is constructed, to lowest order, by bringing two H atoms intheir 1s state together. Since the two H atoms are identical, the two degen-erate states will perturb each other leading to two orbitals that will be linearcombinations of the atomic states. One orbital will be symmetric with respect

Page 150: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

5.4 Electronic States 133

Fig. 5.13 Electron charge distributions (not toscale) for a few molecular orbitals along withthe associated united atom (left-hand column)and separated atom (two right-hand columns)limits. The z-axis in these images is horizon-tal and in the plane of the page. The unitedand separated atom designations appearrespectively in the upper left and lower righthand corners while the generic molecular la-bel is given in the upper right hand corner.

The molecular orbitals were constructed bysquaring symmetric (+) and antisymmetric (–)sums of the atomic amplitudes used to createthe separated atom distributions on the rightbut displaced from each other by Re. The +/–signs between the two separated atom dis-tributions indicate how the amplitudes wereadded. Each orbital, second column, is la-beled by the three labeling schemes, where ∗indicates an antibonding orbital.

Page 151: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

134 5 Diatomic Molecules

to exchange of protons (positive inversion symmetry) while the other will beantisymmetric (negative inversion symmetry). The symmetric combination,indicated by the + sign between the two 1s atomic charge distributions in thebottom row of Fig. 5.13, leads to a σg1s orbital occupied by two electrons. Tolowest order, the resulting charge distribution for one electron in this orbital isthe same as that generated with the united atom construction (see the chargedistribution shown second from the left in the bottom row in Fig. 5.13).

The antisymmetric combination of the two 1s states, indicated by the “−”sign in the second row from the bottom in Fig. 5.13, produces the σu1s or-bital. Since the wavefunction describing an electron in this orbital changessign upon inversion, there is a conspicuous absence of charge over a wide re-gion between the two protons since the charge distribution must vanish at theorigin. This exclusion of charge leads to a stronger Coulomb repulsion be-tween the protons. As a consequence, this orbital does not contribute to bond-ing and is called an antibonding orbital. The united atom construction leadingto nearly the same charge distribution originates from a 2pm=0 atomic state asshown in the second from the bottom row in Fig. 5.13. Thus, the 2pσu and theσu1s orbitals are correlated. Again, looking along the z-axis, the charge distri-bution resembles a 1s state. Figure 5.13 shows MO constructions originatingfrom 2s and 2pm=±1 states as well, along with how they correlate to the unitedconstructions.

We have just seen that to first order, the MOs can be constructed from ei-ther limit. The united atom label, however, is most appropriate for describingRydberg states, where one electron spends most of its time far from the twonuclear centers and the nucleus looks a lot like an atom. The separated atomlabel, on the other hand, is more appropriate near dissociation limits, wherethe two nuclei are far apart and leading to two distinct positive charge centers.It turns out that the separated atom construction forms the basis for the con-struction method called the linear combination of atomic orbitals (LCAO). Con-sequently, one often finds the MOs labeled according to the separated atomscheme. Typically, however, the orbitals are labeled by a more neutral scheme,mλg/u, where m is a positive integer that simply indicates which state of aparticular λ with a specific g/u symmetry is being discussed. While m mayappear to be a molecular principal quantum number, it must be rememberedthat it is only a label. The configuration for O2, which follows the ordering inFig. 5.12, in this neutral scheme is

(1σg)2(1σ∗u)2(2σg)2(1σ∗

u)2(3σg)2(1πu)4(1π∗g)

2. (5.56)

In Eq. (5.56) we added a ∗ to the antibonding orbitals to distinguish themfrom the bonding orbitals and indicate their antibonding nature. Figure 5.14is an example of the two lowest potential curves for H2 generated by twohydrogen atoms in their 1s state. This example shows the general property

Page 152: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Further Reading 135

Fig. 5.14 The two molecular potentials derived from two H (1s) atoms.The lower, bound state is a singlet where the spins of the two electronsare antiparallel while the upper, unbound state is a triplet where theelectron spins are aligned. These calculated potentials are taken fromRef. [48].

of molecular states; there are the same number of molecular states as thereare pairs of atomic states. In this particular case, there are only two distinctcombinations possible – one with the spins parallel (triplet state) and the otherwith them antiparallel (singlet state). The bonding and antibonding nature isalso shown as well.

Further Reading

1 G. Herzberg, Molecular Spectra and Molecu-lar Structure I: Spectra of Diatomic Molecules,Van Nostrand Reinhold, New York, NY,1950.

2 C. S. Johnson, Jr and L. G. Pedersen, Prob-lems and Solutions in Quantum Chemistry andPhysics, Dover, New York, NY, 1986.

3 J. T. Hougen, The Calculation of RotationalEnergy Levels and Rotational Line Intensitiesin Diatomic Molecules, NBS Monograph 115[http://physics.nist.gov/Pubs/Mono115/contents.html].

4 I. Kovács, Rotational Structure in the Spectraof Diatomic Molecules, American Elsevier,New York, 1969.

5 H. Lefebvre-Brion and R. Field, The Spectraand Dynamics of Diatomic Molecules, Aca-demic, New York, NY, 2004.

6 For a few additional problems withworked solutions see D. Budker, D. F. Kim-ball and D. P. DeMille, Atomic Physics: AnExploration Through Problems and Solutions,Oxford University Press, New York, NY,2004.

Page 153: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

136 Further Reading

Problems

5.1 Show that the spring constant for vibration is approximately given by

ks =e2

4πεo

2R3

e. (5.57)

5.2 Draw the Morse potential for the ground state of H2 using molecular con-stants from Herzberg or Herzberg and Huber.

5.3 Show explicitly that for N2, with nuclear spin 1, T = 2, 0 correspond tosymmetric terms and occupy six of the possible terms. How many antisym-metric terms are there? What value(s) of T are relevant?

5.4 Write the configuration for the ground state of N2, assuming that the elec-trons fill the orbitals in order according to the separated atom sequence of Fig.5.12. To what united atom state do σg1sσg2p correspond?

5.5 Use Fig. 5.13 to justify the correlation between the 2sσg − σg2s, 2pπu −πu2p and 3dπg − πg2p orbitals.

5.6 Use your favorite program to generate lowest-order charge distributionsfor the four MOs in the separated atom limit starting from 2p states of H. Youranswers should look something like the images in Fig. 5.15. Label the chargedistributions using the united and neutral schemes. Indicate which orbitalsare bonding and which are antibonding.

g2p u2p

g2p u2p

Fig. 5.15 Electron charge distributions for the four MOs generated inthe separated atom construction from 2p atomic states. Each labelcorresponds to the separated atom labels.

Page 154: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

137

6Molecules in External Fields

6.1Introduction

Much of what we know about diatomic molecules comes from spectroscopy,exposing systems to external fields to induce transitions or dynamics. As withatoms, the changes the system undergoes depend on the Hamiltonian per-turbing the system. We will consider only two types of perturbations – thoseleading to radiative transitions and to field ionization. Most of the work re-quired was done earlier for atoms so, here, we will be able to quote most ofwhat we need.

This chapter is divided into two parts. In the first part, we will look atthe conditions and selection rules for radiative transitions. We will then doa simple example to introduce Herzberg’s method for bookkeeping. In thesecond part, we will look at field ionization of a diatomic molecule inducedby an intense laser field.

6.2Electronic Transitions

Spectroscopy involves transitions between levels. With sufficient resolution,one is able to identify individual rotational and vibrational levels. The wave-length of the transition between states will be given by ∆ν = ν′ − ν′′, wherethe prime indicates the upper level and the double prime indicates the lowerlevel.1 With the aid of Eq. (5.39) this can be written as

∆ν = (ν′e − ν′′e ) + (G′(v′) − G′′(v′′) + (F′v(J′)− F′′

v (J′′)), (6.1)

where J is the total angular momentum apart from nuclear spin. As withatoms, not all values of J will be allowed and only a subset of the values ofv will contribute. To determine which values are relevant, we must examine

1) The use of primes in this way corresponds to the convention em-ployed by molecular spectroscopists.

Page 155: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

138 6 Molecules in External Fields

Abs

orpt

ion

Em

issi

on

Fig. 6.1 Franck–Condon principle. Transitions are strongest for ab-sorption and emission between the solid and dashed arrows, respec-tively.

the transition matrix element. As diatomic molecules have more degrees offreedom than atoms, it is not surprising that this exercise is more involvedthan it was with atoms. In this chapter, we will focus our attention on electricdipole transitions. In this case, the general matrix element takes the form⟨

ψ′eψ′(v′)ψ′(J′)

∣∣ r∣∣ψ′′

e ψ′′(v′′)ψ′′(J′′)⟩

, (6.2a)

which can be written as⟨ψ′

eψ′(J′)∣∣ r

∣∣ψ′′e ψ′′(J′′)

⟩ ⟨ψ′(v′)

∣∣ψ′′(v′′)⟩

. (6.2b)

The second factor in Eq. (6.2b) will only be appreciable when there is sig-nificant overlap between the vibrational wavefunctions – the Franck–Condonprinciple. This leads to a so-called vertical transition as shown in Fig. 6.1. Forthe example shown, absorption would begin near the bottom of the lower welland terminate at the inner turning point of the upper well. Emission, on theother hand, would begin near the bottom of the upper well and terminate nearthe outer turning point of the lower well. The transitions originate where theelectron probability density is highest.

The allowed transitions that render the first factor in Eq. (6.2b) nonzero de-pend on certain selection rules. Selection rules can be determined as was donefor atoms by applying the Wigner–Eckart theorem. The rules fall into two cate-gories – general rules that apply in all cases and case-specific rules. In the nextthree sections we will discuss these rules in the electric dipole approximationfor the first three Hund’s cases.

Page 156: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

6.2 Electronic Transitions 139

6.2.1General Selection Rules

We begin with the general rules for electric dipole transitions. Since J is agood quantum number, its selection rules are the same as for atoms given inEq. (4.54):

∆J = 0,±1, (6.3a)

with the restriction

J′ = 0 J′′ = 0. (6.3b)

Note that ∆MJ is not included here. It will come in, masquerading as dif-ferent parameters under the specific selection rules. In Section 5.4.5 we dis-cussed the symmetry of individual rotational levels. For both heteronuclearand homonuclear systems, each rotational level has either positive (⊕) or neg-ative () symmetry. These levels obey the selection rules

⊕ ↔ , ⊕ ⊕, . (6.4)

Homonuclear systems carry the additional g/u symmetry, which when com-bined with the ⊕/ symmetry leads to levels being either symmetric (s) orantisymmetric (a). The selection rules that these obey are

s → s, a → a, s a. (6.5)

When identical nuclei have equal charge,

g ↔ u, g g, u u. (6.6)

Given these few rules, we can organize spectral lines that link certain rota-tional levels into groups. The restriction on ∆J limits the allowed transitionsto three types, which go by the names P, Q and R branches. By convention,

Branch J′′ J′

P J J − 1Q J JR J J + 1

It should be understood that in absorption, the P branch means that J de-creases by 1 while in emission, the P branch means that J increases by 1. Thatis, J′′ > J′ for the P branch, J′′ < J′ for the R branch and J′′ = J′ for the Qbranch.

6.2.2Case-Specific Selection Rules

In this section we will give the rules for the first four cases. It is importantto understand that the rules apply rigorously only when both the initial and

Page 157: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

140 6 Molecules in External Fields

final electronic cases are the same. When two different cases apply, only thecommon rules apply. With regard to quantum numbers and operators, in thissection we are using the convention of Table 5.2.

6.2.2.1 Hund’s case (a)

∆Λ = 0,±1, (6.7a)

with the restriction

Σ+ Σ−, (6.7b)

∆S = 0, (6.7c)

∆Σ = 0, (6.7d)

∆Ω = 0,±1, (6.7e)

with the restriction

∆J = 0 when Ω = 0 → Ω = 0. (6.7f)

The reader should note that this last restriction is the same restriction we haveon ∆J for atoms in Eq. (4.54).

6.2.2.2 Hund’s case (b)

∆Λ = 0,±1, (6.8a)

with the restriction

Σ+ Σ−, (6.8b)

∆S = 0, (6.8c)

∆N = 0,±1, (6.8d)

with the restriction

∆N = 0 when Σ = 0 → Σ = 0. (6.8e)

6.2.2.3 Hund’s case (c)

∆Ω = 0,±1, (6.9a)

Page 158: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

6.2 Electronic Transitions 141

Fig. 6.2 Ab initio calculation (values taken from Ref. [59]) of three Na2potential curves.

with the restriction

Ω = 0+ Ω = 0−. (6.9b)

6.2.2.4 Hund’s case (d)

∆N = 0,±1, (6.10a)

∆L = 0,±1, (6.10b)

∆R = 0. (6.10c)

6.2.3Examples

We are now ready to look at a couple of examples in a real system. Considerthe potential curves for Na2 shown in Fig. 6.2. The sodium dimer is well de-scribed by Hund’s case (a) for transitions from the ground state to low-lyingexcited states. According to the selection rules above, transitions are possibleto both the A and B states. We will first consider transitions to the A state. In

Page 159: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

142 6 Molecules in External Fields

this case we have a 1Σ+u ←1 Σ+

g transition.2 This is a homonuclear system, sothe g ↔ u restriction is obeyed as is Σ+ ↔ Σ+. Since ∆Ω = 0, there will beno ∆J = 0 transitions. This means that only the P and R branches will exist;the Q branch is not allowed to this state. Consequently, the only transitionsallowed will be ∆J = ±1.

While this is a rather simple case, it is instructive to use Herzberg’s book-keeping method to keep track of the allowed transitions, which is very helpfulin more complicated cases. We partially introduced the method back in Sec-tion 5.4.5. Now we must simply put two diagrams together and draw linesbetween the states that are allowed as we do in Fig. 6.3a. Note that both the⊕ and the s a restrictions are obeyed.

Now we will consider transitions to the B state. The Herzberg diagram forHund’s case (a) is shown in Fig. 6.3b. One should note two aspects of thisdiagram. First, each J level for the upper state has both a ⊕ (a) and a (s)state due to Λ doubling. Second, all three branches are active.

6.3AC Tunneling Ionization

In Section 3.1.1, we discussed AC tunneling ionization in atoms induced bystrong laser fields. Tunneling, which is related to static-field ionization, occurswhen the Keldysh parameter (γ, Eq. (3.6)) is small. A similar ionization pro-cess occurs in molecules as well. However, things are a bit more interestingwith molecules for several reasons. For example, the electronic and nuclearmotions are both important – we cannot assume that the nuclei are static. Atthe same time, after a sufficient number of electrons are removed from theparent molecule, the highly charged nuclei tend to dissociate, and rather vi-olently at that. This strong-field-induced dissociative ionization often goesby the name of Coulomb explosion or laser-induced Coulomb explosion. Whilethe genesis of Coulomb-explosion experiments can be traced to beam foil ex-periments [60],3 laser-based experiments provide increased flexibility and ameans for preparing the initial state and/or inducing exotic dynamics priorto the explosion. We will end our discussion of molecules with a brief discus-sion of this laser-induced Coulomb explosion.

2) The backwards arrow is the custom when describing absorption inmolecular physics. The lower state is always on the right. Emission,for example, would flow from left to right.

3) Beam foil experiments involve stripping several electrons from ahigh-energy, low-charge molecular beam traversing a thin foil. Af-ter the foil, the highly stripped molecules undergo a Coulomb ex-plosion. The atomic ions are then detected with position-sensitivedetectors.

Page 160: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

6.3 AC Tunneling Ionization 143

Fig. 6.3 Herzberg’s bookkeeping diagram for (a) 1Σ+u ↔1 Σ+

g tran-

sitions and (b) 1Πu ↔1 Σ+g transitions. The allowed transitions are

indicated by the notation ∆J(J′′). The subscript of the upper state in(b) is Ω = Λ + Σ. Note that J = N = R + 1 in these diagramsbecause S = 0.

Page 161: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

144 6 Molecules in External Fields

We begin by modifying the classical equations used to study field ionizationof atoms (Eqs. (3.1b) and (3.2)) with a smoothed, Eberlonium potential. In thiscase, the potential energy resulting from the combined Coulomb and laserfields takes the form

V(z) = − e2

4πεo

(Z1√

(z − R/2)2 + a2+

Z2√(z + R/2)2 + a2

)− eFz, (6.11)

where Z1,2 is the nuclear charge and R is the internuclear separation. Thevalence electron, apart from its dissociation energy, is bound by

E0 = −12

[(Ip1 +

14πεo

Z1e2

R

)+

(Ip2 +

14πεo

Z2e2

R

)], (6.12)

where Ip1,2 is the ionization potential for each atom. At the equilibrium sep-

aration, Re, the threshold intensity, Ith, required for over-the-barrier ioniza-tion can be determined as was done in Section 3.1.1.4 For a model system,Ip1 = Ip

2 = 1 a.u., R = Re = 1 a.u. and Z1 = Z2 = 1,5 Fig. 6.4a showsthat Ith ∼ 1.4 × 1015 W/cm2. Ionization at Re, in the absence of any resid-ual bonding, would lead to an Re explosion energy, potential energy turnedinto kinetic energy, of 14.4 eV, if the nuclei have the charge and mass of pro-tons.6 For sufficiently massive nuclei that do not move during the pulse, theexplosion energy observed is nearly that predicted by Re. More commonly,however, explosion energies significantly less than this are observed. Theyare typically consistent with separations 2–3 × Re.

Explosion energy deficits can be understood in terms of a simple modelbased on peak tunneling rates occurring at extended separations [61–63]. Theessence of the model, often called enhanced ionization at Rc, is captured inFig. 6.4 for a model system. In the presence of the laser field, electrons are in-duced to move back and forth from one side of the molecule to the other. Therates for such transitions are sustained by field-enhanced charge-resonancestates [64] that shift with the strength of the laser field and internuclear sepa-ration according to [65]

E± = Ip± ± FR

2. (6.13)

Above, the valence level is shifted upward (+) in the uphill potential well anddownward (−) in the downhill potential well. It is important to recognize

4) Tunneling will occur at a slightly lower intensity, but Ith is moreconvenient to visualize. In this molecular case, a molecular ADKrate can be used to estimate the ionization rate [20].

5) Note that our model system does not correspond to the parametersof either H2 or H+

2 .6) The explosion energy, or so-called kinetic energy released, is the

sum of the final energies of the two nuclei, protons in this model, atinfinite separation subsequent to the explosion.

Page 162: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

6.3 AC Tunneling Ionization 145

Fig. 6.4 Model calculation of the potential distortion allowing ACtunneling ionization for a homonuclear diatomic system: (a) Ith =1.4 × 1015 W/cm2 at Re; (b) Ith = 0.47 × 1015 W/cm2 at Rc; (c) and (d)electrons are trapped with I = 0.47 × 1015 W/cm2 at Re and at 6Rc.The intensity required for ionization is a minimum at Rc. For this modelcalculation, Z1,2 = Ip

1,2 = Re = 1 a.u. and Rc 3.8 a.u.

that the shift increases with R. Eventually the uphill level will shift over thebarrier allowing the electron to escape. Thus, it is easier to ionize the systemat a larger R than at Re. One must not lose sight of the fact that as the nucleiseparate, an inner barrier develops that can trap the electron, which was free

Page 163: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

146 Further Reading

to move between the two nuclei, in one well or the other.7 Once this occurs,it is possible to define Ith and Rc as the point where E+ equals the heights ofboth inner and outer barriers. In the model system, for example Fig. 6.4b, Ithat Rc (> Re) is less than half that required at Re. Note as well that though theelectron is in the downhill well it is still bound. Theoretically [62, 65],

Rc 4Ip , (6.14)

where atomic units are used. As R continues to increase, Fig. 6.4 shows thatIth also increases – Rc is unique in that Ith increases when R > Rc or R < Rc.Even though we have presented these ideas in classical terms, they have beensubstantiated quantum mechanically [62, 65].

Further Reading

1 G. Herzberg, Molecular Spectra and MolecularStructure I: Spectra of Diatomic Molecules, VanNostrand Reinhold, New York, NY, 1950.

2 J. Posthumus (ed.), Molecules and Clusters inIntense Laser Fields, Cambridge UniversityPress, New York, NY, 2001.

3 J. I. Steinfeld, Molecules and Radiation: An In-troduction to Modern Molecular Spectroscopy,The MIT Press, Cambridge, MA, 1978.

Problems

6.1 Describe the spectrum for excitation between the ground state X 1Σ1g and

A 1Σ−u of Na2.

a) Write out the Herzberg diagrams to show the allowed transitions be-tween the two states.

b) Estimate the wavelength for the lowest P, Q and R branch transitionsin absorption to first order (i.e., only keep the lowest relevant constants)for the first v = 0 to v = 0 transition. Specify the J values for the lowerand upper states. Also, if one of the branches does not exist, so state.

6.2 The selection rules for a magnetic dipole are

⊕ , (6.15a)

∆J = 0,±1, (6.15b)

7) Clearly this must be the case, since at infinite separation an electronattached to either nucleus will simply behave like an atom or atomicion having a single potential well.

Page 164: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Problems 147

with the restriction

J′ = 0 J′′ = 0, (6.15c)

∆Λ = 0,±1. (6.15d)

The restriction on the latter is that when Λ = 0 only intersystem combinations(single ↔ triplet, etc.) are allowed.

Justify these selection rules using the Wigner–Eckhart theorem. Describethe spectrum for excitation between the ground state, X3Σ−

g , and the excitedstate, b1Σ+

g , of O2, which approximately obeys Hund’s case (b). Estimate thewavelengths of the lowest P, Q and R branch transitions to first order (i.e.,only keep the lowest relevant constants). Specify the J values for the lowerand upper states. If one or more of the branches does not exist, so state. Forthis problem, you will need to define these branches for ∆K and ∆J transitionsseparately.

6.3 Derive Eq. (6.14). Hint, see Ref. [65].

6.4 Verify Eq. (6.14) for O+2 numerically by determining Ith(Rc). Plot Ith(R)

for R = Re to 10Re.

Page 165: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 166: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

Part 2Light-Matter Interaction: Nonlinear Optics

Page 167: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 168: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

151

7Nonlinear Optics

7.1Introduction

The linear relationship between the polarization and electric field is strictlytrue only if the susceptibility is not a function of the field. With the advent ofthe laser one can easily focus the laser beam in a medium, creating a very highfield. In this case the susceptibility itself becomes dependent on the electricfield. The relation between P and E will be nonlinear:

P(t) = χ(1)E(t) + χ(2)E2 + χ(3)E3 + . . . (7.1)

= P(1)(t) + P(2)(t) + P(3)(t) + . . . , (7.2)

where the first term is the usual linear polarization and the others are referredto as the nonlinear polarization. The quantities χ(2) and χ(3) are the second-and third-order nonlinear optical susceptibilities, respectively. In general themagnitude of χ(n+1) is approximately equal to χ(n)/Eatm, where Eatm is theatomic electrical field. From Eatm ∼ e/a2

o , where −e is the charge of theelectron and ao is the Bohr radius of the hydrogen atom, one can calculateEatm ∼ 5 × 109 V/cm (Eq. (A.11)). Thus, the effect of χ(2) is a factor of E/Eatm

smaller than χ(1). This explains why the discovery of the optical nonlineareffect had to wait for the invention of the laser. Only by focusing a laser beamis one able to generate an E field with high enough strength such that it be-comes possible to observe the nonlinear optical effect. In an extreme case,such as a table-top high-power femtosecond laser system, the electric fieldstrength in the focus of the laser beam can exceed that of the atomic field, mak-ing higher-order nonlinear effects more probable than their lower-order coun-terparts and leading to the breakdown of perturbation theory, upon whichEq. (7.1) is based, altogether.

Historically, the first nonlinear optical experiment was optical second-harmonic generation reported by Franken et al. [66]. A Q-switched ruby laserwas used as the fundamental beam at a wavelength of 694.3 nm. When thelaser was focused onto a quartz crystal, a very small signal at half the originalwavelength was observed amid the strong background radiation of the fun-

Page 169: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

152 7 Nonlinear Optics

damental. This engendered the new field of nonlinear optics. Theoreticallyspeaking, the nonlinear effect should be a fundamental property of materials.With either a strong enough fundamental field and/or a sensitive enough har-monic detector, one may proceed to look for new nonlinear effects. Startingas a scientific curiosity, the field of nonlinear optics has now grown into a vasttechnology empire with many applications.

In optical parametric generation, it is used to produce a coherent tunableoutput. Optical second- and third-harmonic generation provide good probesfor the material properties. A second-harmonic generator, also referred to asa frequency doubler, is now a standard optical component one can purchaseoff-the-shelf. Even surface optical second harmonic generation can be appliedto study the microscopic properties at the material surface with submicrom-eter spatial resolution by near-field harmonic imaging. In the extreme highfield area, optical harmonic generation up to hundreds of orders provides acoherent beam in the extreme ultraviolet or soft X-ray region. In other words,nonlinear optics has become an inseparable part of everyday scientific phe-nomena and is used in many high-technology industries, such as optical com-munications.

There are many books written on the subject of nonlinear optics already.The purpose of this book is to provide readers with a sufficient backgroundof the essence of nonlinear optics. We will first provide a phenomenologicaldescription of nonlinear optics. Wave propagation in a nonlinear medium willbe discussed next. Some nonlinear phenomena will be discussed in detail. Thequantum theory of nonlinear optical susceptibility will be presented using adensity matrix formalism. Applications of nonlinear optics will be presentedat the end.

7.2Phenomenological Description of Nonlinear Optics

In Eqs. (7.1) and (7.2), χ(n) may be regarded as the response function of thematerial. The polarization, when expressed in the frequency domain, can bewritten as

Pl(ωi) = χ(1)lm (ωi) Em (ωi)

+ 2χ(2)lmn

(ωi; ωj, ωk

)Em

(ωj

)En (ωk) δ

(ωi; ωj + ωk

)+ 6χ

(3)lmno

(ωi; ωj, ωk, ωh

)Em

(ωj

)En (ωk) En (ωk) Eo (ωh)

× δ(ωi; ωj, ωk, ωh

)+ higher terms in E and other terms involving the

magnetic field B

+ . . . ,

(7.3)

Page 170: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 153

where the subscripts l, m, n, o indicate Cartesian components and ωi, ωj, ωk,etc., represent the frequencies of interest. The factor “2” in Eq. (7.3) is fromour definition of the electric field:

E(t) = Ee−iωt + c.c.

The notation used here follows that of Boyd [67]. The delta function assumesthe value of unity when the frequencies in its argument satisfy the relation1

ωi = ωj + ωk + ωh.

For example, from the term

P(2)l (ωi) = 2χ

(2)lmn

(ωi; ωj, ωk

)Em

(ωj

)En (ωk) δ

(ωi; ωj + ωk

)(7.4)

we may have one of the following nonlinear effects:

· optical three-wave mixing χ(2) (ωi; ωj, ωk,

), which includes sum- and

difference-frequency generation,

· χ(2) (2ω; ω, ω), optical second-harmonic generation (SHG) or frequencydoubling,

· χ(2)(ω; 0, ω), linear electrooptic or Pockels effects,

· χ(2)(0; ω,−ω), optical rectification,

· χ(2) (ωj + ωk; ωj, ωk

), frequency upconversion, and

· χ(2) (ωj − ωk; ωj,−ωk

), parametric oscillation.

From the term

Pl = 6χ(3)lmno

(ωi; ωj, ωk, ωh

)Em

(ωj

)En (ωk) E0 (ωh) δ

(ωi; ωj + ωk + ωh

)(7.5)

we have

· χ(3)lmno

(ωi; ωj, ωk, ωh

), optical four-wave mixing,

· χ(3)lmno (3ω; ω, ω, ω), optical third-harmonic generation (THG),

1) See Appendix B.1:

δ(ωi; ωj + ωk) =

1, ωi = ωj + ωk

0, ωi = ωj + ωk

δ(ωi; ωj + ωk + ωh) =

1, ωi = ωj + ωk + ωh

0, ωi = ωj + ωk + ωh

Page 171: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

154 7 Nonlinear Optics

· Reχ(3)lmno (ω; ω, ω,−ω), a nonlinear refractive index,

· Imχ(3)lmno (ω; ω, ω,−ω), two-photon absorption, and

· Imχ(3)lmno (ωs; ωl,−ωl , +ωs), stimulated Raman scattering.

χ generally reflects the symmetry of the material system. For example,

χ(2)lmno

(ωi; ωj, ωk

)vanishes in a medium with inversion symmetry.

7.2.1Optical Second-Harmonic Generation (SHG) χ(2)(2ω; ω, ω)

Historically, the most important nonlinear optical effect is SHG. The first ex-periment by Franken et al. [66] led to the birth of the field of nonlinear optics.

In a typical SHG experiment one sends into a nonlinear crystal a laser beamat frequency ω. In the output, one can detect a new optical field at frequency2ω. Figure 7.1 shows the experimental arrangement. In a medium that lacksa center of inversion, this process is allowed. Physically, this nonlinear pro-cess can be described as a two-step process. First, the medium responds tothe incident electric field of the laser beam in a nonlinear fashion, generat-ing nonlinear polarization, P(2) (2ω). This polarization serves as the drivingforce in the wave equation. The medium is driven by an ensemble of coherentoscillating dipoles, which is also referred to as PNLS [68]. We have

−2 E + µ0 σc∂E∂t

+1c2

∂2E∂t2 = − µ0

∂2PNLS

∂t2 . (7.6)

Here PNLS(2ω) is the source term that propagates through the medium, cre-ating a radiating field E(2ω, r1) at r1, an upstream position in the medium.The radiating field propagates in the medium as soon as it is generated. In themeantime, the laser beam with E(ω) also propagates through the medium,

Fig. 7.1 Experimental setup for the detection of SHG light.

Page 172: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 155

creating a new PNLS(2ω, r) wave downstream (r > r1). This propagatingPNLS will radiate E(2ω) as it propagates along. E(2ω, r), generated down-stream will interfere with the E(2ω, r1) generated earlier but which has prop-agated to the same downstream position r. Since all these are coherent pro-cesses, there is a definite phase relationship between the PNLS wave withpropagation vector 2k1 and the reradiating field E(2ω, r) with propagationvector k2. Normally, because of dispersion, the fundamental wave which cre-ates the PNLS(2ω) will be out of step (out of phase) with the E(2ω). Thisresults in an oscillatory amplitude of the E(2ω, r) as a function of positioninside the crystal, causing periodic destructive and constructive interferencebetween the E(2ω) generated at r and the E(2ω)r1 generated at an upstreamposition (r1) but which has propagated to the same location r. This is theso-called non-phase-matched condition. It always limits the SHG conversionefficiency. A configuration can be devised to make the PNLS(2ω) and E(2ω)waves propagate with the same phase velocity so that they always propagatein phase, leading only to constructive interference. This is the phase-matchingconfiguration in SHG. Nowadays, vendors can supply crystals that have beenproperly cut and oriented for the phase-matching condition. In principle, withthe most favored condition, SHG conversion efficiency can approach 100%.However, as a researcher or student in nonlinear optics, one still needs to un-derstand the principle of phase matching in order to use the nonlinear crystalproperly. We will discuss this in detail later. The interplay between PNLS(2ω)and E(2ω) waves can be analyzed by the formalism developed for the wavepropagation in a nonlinear medium [68].

Fig. 7.2 Quantum mechanical description of optical second-harmonicgeneration.

In the wave–particle duality picture photons are quanta. The SHG processcan be regarded as particle annihilation and generation: two fundamentalphotons are annihilated but at the same time a new particle at double the

Page 173: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

156 7 Nonlinear Optics

frequency is created as shown in Fig. 7.2. In this process energy is conserved:

hω2 = hω1 + hω1,

ω2 = 2ω1.

The process shown in Fig. 7.2 also does not disturb the population of theinitial and final states. It is referred to as the parametric process. Since, undernormal conditions, the conversion efficiency of SHG is always small, one usu-ally needs to focus the laser beam very tightly into the nonlinear medium, in-creasing the intensity of the laser beam and hence the electrical field strength.Despite this focusing effort, in many instances the second-harmonic signalsgenerated are still very weak as compared to the signal at the fundamentalfrequency. One needs good isolation to block out the intense fundamentalbeam before sending the second-harmonic beam into a photodetector.

7.2.2Electrooptic Effect, χ(2)(ω; 0, ω)

The linear electrooptic effect may be regarded as a special χ(2) process whenone of the electric fields is a DC field, e.g., χ(2)(ω; 0, ω). The refractive indexof the extraordinary ray (e-ray) can be controlled by applying a DC field toan electrooptic crystal. For a wave propagating in a birefringent crystal, theequation of the index ellipsoid in the presence of a DC field is(

1n2

)1

x2 +(

1n2

)2

y2 +(

1n2

)3

z2

+(

1n2

)4

yz +(

1n2

)5

xz +(

1n2

)6

xy = 1. (7.7)

The linear change in the coefficients(

1n2

)′i, due to the presence of an arbitrary

DC field E(Ex, Ey, Ez

), is expressed as

(

1n2

)i=

3

∑j=1

rijEj, (7.8)

where i = 1, 2, . . . , 6 and rij is the electrooptic coefficient, which is a third-ranktensor. The contraction of a third-rank tensor with the vector E results in asecond-rank tensor similar to χ(2). The linear electrooptic effect is the principleutilized in electrooptic modulators (Mach–Zehnder type) and Pockels cells. Arecent but challenging application is using the modulator for a broadband RFfiber link for transporting broadband microwave or millimeter-wave signalsover optical fibers, instead of conventional copper transmission lines [69]. Thekey component of this type of system is a modulator which is used to impress

Page 174: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 157

a RF subcarrier onto optical carrier waves [70]. A high-performance modu-lator requires high spurious free dynamic range, low third-order intermod-ulation distortion and a high degree of linearization. To achieve broadbandoperation, the modulator electrodes must be part of the transmission line.

7.2.3Optical Rectification χ(2)(0; ω, −ω)

Optical rectification is represented by χ(2)(0; ω,−ω). Two fundamental wavesbeat together to generate a zero-frequency (DC or carrierless) signal. Opticalrectification usually refers to the generation of a DC polarization component,which is not of particular interest. However, with the advent of femtosecondlasers, this effect becomes responsible for terahertz (THz) pulse generationfrom dielectrics. The output from the process of

P2i (0) = ∑

j,kχ(2)(0; ω,−ω)Ej(ω)E∗

k (ω) (7.9)

is an envelope of the E field with its carrier stripped off. This effect becomesparticularly interesting when the envelope function of the optical wave lastsless than 1 picosecond (ps). The picosecond electrical pulse radiates a shortTHz impulse, consisting of only one and one-half cycles (a Mexican hat shapepulse). Pulsed THz sources have been used for time-resolved far-infrared(FIR) imaging of a concealed object [71].

In order to generate THz pulses by optical rectification processes, subpi-cosecond optical pulses are required to excite the nonlinear crystal. Materialscapable of generating THz pulses include semiconductors (GaAs, ZnSe, ZnTe,InP, CdTe, etc.), electrooptic materials (LiNbO3, LiTaO3, BaTiO3, KTP, ZnO,etc.), and organic crystals (dimethyl amino 4-N-methylstilbazolium tosylate orDAST, etc.). Detecting the THz pulses requires special techniques. To achievethe time resolution required, a physical effect with a quick response must beemployed. The electrooptic effect is such an effect. When a THz pulse im-pinges onto an electrooptic crystal it will induce an instantaneous electroopticeffect, causing the polarization of the probe pulse to rotate. By carefully mea-suring the amount of rotation as a function of time delay between the pump,which generates the THz pulse, and the probe pulse, one can map out the tem-poral shape of the THz pulse. This technique turns out to be the only way tomeasure the waveform of the THz pulses, since no other detecting system canprovide the necessary bandwidth to resolve the waveform [72].

Page 175: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

158 7 Nonlinear Optics

7.2.4Parametric Generation χ(2) (

ωs; ωp, −ωi)

The second-order nonlinear optical susceptibility, χ(2) (ωs; ωp,−ωi

), can be

used to generate a coherent output by a strong pump beam at frequency ωp

such that

hωp ⇒ hωs + hωi, (7.10)

i.e., the pump photon breaks up into two photons, one at a signal frequency,ωs, and the other at an idler frequency ωi, respectively.2 Signal and idler fieldsare coupled together through the pump field. The theory of parametric gen-eration will be presented in more detail in the next chapter.

Any parametric interaction satisfies energy and momentum conservationconditions

hkp ⇒ hks + hki (7.11)

and will emerge as the dominant process, generating a coherent output at ωs

and ωi, respectively [73]. From Eq. (7.10) it is clear that as long as the con-dition hωp ⇒ hωs + hωi is met, there will be an infinite number of values ofωs satisfying the energy-conservation condition, leading to a tunable coher-ent output. Two commercial optical devices, the optical parametric oscillator(OPO) and the optical parametric amplifier (OPA), have been built accord-ing to the principle of parametric generation discussed here. Using intensesecond-harmonic pulses from a commercial Ti:sapphire regenerative ampli-fier as the pump pulses, one can generate a tunable pulse in the infrared.

7.2.5Third-Order Nonlinear Effect

With the third-order χ(3) nonlinear effect the most important process is four-

wave mixing χ(3)lmno

(ωi; ωj, ωk, ωh

).

Energy conservation gives

hωi = h(ωj + ωk + ωh), (7.12)

2) In the parametric generation process theenergy taken away from the pump beamis fed into two beams of lower frequency.Consequently, the process can be used asan amplifier: a weak signal beam is madeto interact with a strong, high-frequencypump beam, and both the original signal,known as the “signal”, and the generated

difference frequency, known as the “idler”,are amplified. Note that the denominations“signal” and “idler” have specific mean-ings only for the parametric amplifier. Inthe oscillator, either of the two lower fre-quencies can be called the “signal” or the“idler”.

Page 176: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 159

where the j, k, h indices stand for 1, 2 or 3 respectively and the frequenciesmay take positive or negative values. For example, ωi = ω1 + ω1 + ω1 = 3ω1represents third-harmonic generation, ω4 = ω1 + ω2 − ω3 represents four-wave mixing and ω1 = ω1 + ω1 − ω1 represents a nonlinear refractive-indexchange.

Let us consider

E(t) = E1 e−iω1t + E2 e−iω2t + E3 e−iω3t + c.c.

The χ(3) nonlinear term gives rise to the nonlinear polarization

PNL (ω1) = χ(3) (3E1E∗1 + 6E2E∗

2 + 6E3E∗3) E1, (7.13)

PNL (ω1 + ω2 − ω3) = 6χ(3)E1E2E∗3 and (7.14)

PNL (2ω1 − ω2) = 3χ(3)E21E∗

2 , etc. (7.15)

Of particular interest is the degenerate case, where ω1 = ω2 = ω3 = ω and

PNL(ω1) = 3χ(3) | E |2 E. (7.16)

This is the nonlinear polarization at the same frequency of the fundamentalfield. This term will give rise to the intensity-dependent refractive index

n = n0 + n2 I(r, t), (7.17)

where I(r, t) = (noc/2π) | E(ω, r, t) |2 is the intensity of the optical beam. Fora short-pulse laser, the intensity of the laser is a function of position (spatialGaussian distribution) and time (short-pulse envelope function) at the carrierfrequency ω. The intensity-dependent refractive-index nonlinearity is respon-sible for phenomena such as self-focusing [74], self-phase modulation andsoliton formation [75].

7.2.6Nonlinear d coefficient

We have seen previously with Eq. (7.4) that the nonlinear polarization for aχ(2) effect is given by

PNL (ωn + ωm, ωn, ωm) = ∑jk

∑nm

χ(2)ijk (ωn + ωm, ωn, ωm)

× Ej (ωn) Ek (ωm) e−i(ωn+ωm)t,

where the indices i, j, k are referred to as the Cartesian components of the ana-lytical E field. Therefore, i, j, k may represent x, y or z while n and m representthe specific frequency of the component field.

Page 177: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

160 7 Nonlinear Optics

Experimentally, one can measure the electric field, which is a real quantity:

E = ε cos(ωt − φ), (7.18)

where ε is a real amplitude of the electric field with an angular frequency ω

and phase angle φ. Analytically, this electric field can be expressed as

E = E(ω)e−iωt + E∗(ω)e+iωt. (7.19)

The reality of the electric field requires that

E(ω) = E∗(ω) = E(−ω) =12

ε,

and thus

E =12

ε eiφe−iωt +12

ε e−iφeiωt

= ε cos(ωt − φ) = 2E(ω) cos(ωt − φ).(7.20)

We see that the analytical amplitude of the electric field is

E(w) =12

ε. (7.21)

We have used a script letter to denote a real quantity. With the usual second-order (χ(2) process) effect the real nonlinear polarization may be expressed as

p = 2dε2, (7.22)

where d is a nonlinear coefficient used by experimentalists; d is in general athird-rank tensor. The d tensors for various nonlinear materials can be foundin table form in many nonlinear optics books. The factor “2” in Eq. (7.22) isused so that the d coefficients conform with those listed in the other books.The application of Eq. (7.22) to a nonlinear crystal with two fields present isnow illustrated below:

ε = ε1(z, t) cos(ω1t − k1z) + ε2(z, t) cos(ω2t − k2z);

for simplicity we assume that these two fields at ω1 and ω2 are propagatingcollinearly,

p = 2d [ε1 cos(ω1t − k1z) + ε2 cos(ω2t − k2z)]2

= d(

ε21 + ε2

2

)+ d ε2

1 cos [2(ω1t − k1z)]

+ d ε22 cos [2(ω2t − k2z)]

+ 2d ε1ε2 cos [(ω1t + ω2)t − (k1 + k2)z]+ 2d ε1ε2 cos [(ω1 − ω2)t − (k1 − k2)z]

= pdc + p(2ω1) + p(2ω2) + p(ω1 + ω2) + p(ω1 − ω2).

(7.23)

Page 178: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 161

Equation (7.23) defines the real nonlinear polarization,

pdc = d(

ε21 + ε2

2

),

p(2ω1) = d ε21 cos [2(ω1t − k1z)]

= d ε21

12

[e−i2(ω1t−k1z) + e−i2(ω1t−k1z)

]= P(2ω1) + P∗(2ω1),

where P(2ω1) and P∗(2ω1) are the analytical amplitudes of the nonlinear po-larization. Note that

P∗(2ω1) = P(−2ω1).

On the other hand, we have

P(2ω1) = χ(2)(2ω1, ω1, ω1) E(ω1) E(ω1)e−i2(ω1t−k1z),

since E(ω1) = 12 ε1.

By comparing these two expressions, one has

12

d ε21 = χ(2) E(ω1) E(ω1),

d =12

χ(2). (7.24)

Equation (7.24) enables one to write the nonlinear polarization either in real-amplitude form or as an analytical expression. For the nondegenerative three-wave-mixing case,

Pi(ω1 + ω2) = ∑j,k

(2)ijk (ω1 + ω2, ω1, ω2)Ej(ω1)Ek(ω2)e−i(ω1+ω2)t

+ χ(2)ijk (ω2 + ω1, ω2, ω1)Ej(ω2)Ek(ω1)e−i(ω2+ω1)t

].

(7.25)

When the summation is carried out for all j and k values, there are a totalof 18 terms in Eq. (7.25). However, the difference between the first and sec-ond terms is merely an effect of interchanging the sequence of fields in theequation. The sequence of the fields only affects the appearance of the math-ematical expressions. It does not give any physically discernible difference.Mathematically, for a given index i (x, y or z) there are only six distinct termsin Eq. (7.25).

Passing to the d notation, one should realize that there are only six inde-pendent d coefficients corresponding to a given index, i. One can devise a

Page 179: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

162 7 Nonlinear Optics

contracted notation with a single running index to represent the double in-dices jk as follows:

d ixx︸︷︷︸(1)

d ixy︸︷︷︸(6)

←− d ixz︸︷︷︸(5)

↑d iyy︸︷︷︸

(2)

d iyz︸︷︷︸(4)

↑d izz︸︷︷︸

(3)

or jk : xx yy zz yz xz xy

: 1 2 3 4 5 6 (7.26)

For example,dyyz → d24, etc.,

dixx → di1.

In general, di has 18 independent terms with index “i” representing x, y or zand “” representing 1, 2, . . . , 6.

With the contracted notation given in Eq. (7.26), one can use matrix multi-plication to represent the second-harmonic generation of Eq. (7.25):

Pi(2ω1) = 2 diF = ∑

2diF,

where F =(

1 − 12

δjk

) [Ej(ω1)Ek(ω2) + Ek(ω1)Ej(ω2),

]or F1 =

(1 − 1

12

δxx

)[Ex(ω1)Ex(ω2) + Ex(ω1)Ex(ω2)]

=12

2Ex(ω1)Ex(ω2) = E2x(ω1),

F4 =(

1 − 12

δyz

) [Ey(ω1)Ez(ω2) + Ez(ω1)Ey(ω2)

]= Ey(ω1)Ez(ω2) + Ez(ω1)Ey(ω2).

When ω1 = ω2,

F4 = 2 Ey(ω1) Ez(ω1), etc.

Page 180: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

7.2 Phenomenological Description 163

or

Px(2ω)Py(2ω)Pz(2ω)

= 2

d11 d12 d13 d14 d15 d16d21 · · · · d26d31 · · · · d36

E2x(ω)

E2y(ω)

E2z(ω)

2Ey(ω)Ez(ω)2Ex(ω)Ez(ω)2Ex(ω)Ey(ω)

,

(7.27)

Px(ω1 + ω2)py(ω1 + ω2)Pz(ω1 + ω2)

= 4

d11 . . . d16

. . .d31 . . . d36

Ex(ω1)Ex(ω2)Ey(ω1)Ey(ω2)Ez(ω1)Ez(ω2)

Ey(ω1)Ez(ω2) + Ez(ω1)Ey(ω2)Ex(ω1)Ez(ω2) + Ez(ω1)Ex(ω2)Ex(ω1)Ey(ω2) + Ey(ω1)Ex(ω2)

.

(7.28)

The extra factor of two results when one sums over ω1 and ω2 sequentially.It should be noted that, depending on the symmetry of the nonlinear crystal,many d coefficients vanish and some may be equal to other elements. Forexample, all d coefficients vanish if the crystal has inversion symmetry (aninversion center). This can be illustrated as follows:

P(2ω, r) = 2dE(r)E(r).

For a medium with inversion symmetry, as shown in Fig. 7.3,

E(−r) = −E(r),

P(2ω,−r) = −P(2ω, r).

Applying an inversion operation on the equation

−P(2ω, +r) = 2d(−E)(−E) = 2dEE,

or P(2ω, r) = −2dEE = 2dEE, for arbitrary E,

we have

d ≡ 0.

Page 181: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

164 7 Nonlinear Optics

Similarly for a crystal with a 222 symmetry class, by repeatedly applyinga two-fold rotation symmetry operation about x-, y- and z-axes, it can beshown that the only nonzero terms in the d matrix are dxyz(d14), dyxz(d25) anddzxy(d36), i.e. when all three indices are different. The form of the dil matrixfor all 21 crystal classes can be found in a number of books. For example,

d =

. . . d14 · ·. . . · d14 ·. . . · · · d36

,

showing only two nonzero independent elements, d14 and d36. The values ofthese d coefficients can also be found in many books, for example, KTiOPO4(KTP) belongs to the point group mm2, with

d31 = d15 = 3.7 pm/Vand d33 = 14.6 pm/V,

while d36 for AgGaSe2 is 49.3 pm/V.3

It should be pointed out that different books use different units for the d co-efficient; when comparing the d coefficients listed from different sources oneneeds to make sure that they are in the same unit. Since the reader of this bookonly needs to understand some of the principles of nonlinear optics, we neednot concern ourselves with how to derive the entire d matrix.

Fig. 7.3 Eω(→r ) and

→P2ω (

→r ) vectors in a medium with inversion sym-

metry. d = 0 in this case.

3) Data are from Table 9.3 of J. M. Liu, Photonic Devices, CambridgeUniversity Press, 2005.

Page 182: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Further Reading 165

Further Reading

1 R. W. Boyd, Nonlinear Optics, Academic, SanDiego, CA, 1992

2 N. Bloembergen, Nonlinear Optics, 4th edn.,World Scientific, Singapore, 1996

3 F. Zernike and J. E. Midwinter, Applied Non-linear Optics, Wiley, New York, NY, 1973

4 R. H. Pantell and H. E. Puthoff, Fundamentalsof Quantum Electronics, Wiley, New York, NY,1969

5 Y. R. Shen, The Principles of Nonlinear Optics,Wiley, New York, NY, 1984

6 J. M. Liu, Photonic Devices, Cambridge Uni-versity Press, Cambridge, UK, 2005

Problems

7.1 The electric flux density (displacement vector) D is expressed as

D = ε0E + P = ε0(1 + χ)E

in the MKS system of units, and

D = E + 4πP = (1 + 4πχ)E

in the CGS Gaussian system of units.If one includes nonlinear effects, the polarization P can be expressed as

P(t) = χ(1)E(t) + χ(2)E2(t) + χ(3)E3(t) + . . .

in the Gaussian system of units, and

P(t) = ε0[χ(1)E(t) + χ(2)E2(t) + χ(3)E3(t) + . . .]

in the MKS system, where ε0 = 8.85 × 10−12 F/m denotes the permittivity offree space.

In some of the literature, the second-order nonlinear term is also expressedas

P(2)(t) = χ(2)E2(t) = 2d(E)2(t)

in CGS and

P2(t) = ε0χ(2)E2(t) = 2dE2(t)

in MKS. The d coefficient in CGS is in the unit of “esu” and in MKS, C/V2.

a) Show that χMKS = 4πχCGS.

b) Derive the ratio dMKS/dCGS.

Page 183: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

166 Further Reading

c) dMKS/ε0 is also referred to as the nonlinear coefficient in m/V.

The dil’s for the following materials are given as

Quartz, d11 = 0.96 × 10−9 esu,Proustite, d22 = 68 × 10−9 esu,KDP, d36 = 1.1 × 10−9 esu.

Calculate the corresponding nonlinear coefficient dil/ε0 for these mate-rials in m/V.

d) The following materials have the dil coefficient given in the unit of(1/9) × 10−22 C/V2.

LiIO3, d15 = 4.4,GaAs, d14 = 72,CdGeAs2, d36 = 360.

Calculate the corresponding dil in esu and dil/ε0 in m/V.

7.2 Find the d matrix for a crystal with 42m, 43m and 3m symmetry.

Page 184: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

167

8Wave Propagation in Nonlinear Media

8.1Nonlinear Wave Equation

For simplicity, let us use optical second-harmonic generation as an exam-ple. We have seen from the preceding chapter that when a nonlinear opticalmedium is illuminated by a strong laser beam, an induced polarization at afrequency twice the laser frequency is created:

PNL(2ω, r, t) = χ(2)(2ω) E(ω, r) E(ω, r)ei2ωt + c.c. (8.1)

Since E(ω, r, t) = E(ω)ei(ωt−k.r) is a propagating wave, the PNL term in-duced by this nonlinear interaction is also a propagating wave. It propa-gates collinearly with the fundamental beam with propagation constant 2k.This PNL serves as the driving term (source) for generating second-harmonicwaves according to the nonlinear wave equation

∇× (∇× E(2ω, r, t)) +ηAc

∂E(2ω, r, t)∂t

+η2

c2∂2E(2ω, r, t)

∂t2 = −µ0∂2PNL

∂t2 ,(8.2)

where

E(2ω, r, t) =12

E(2ω)ei(2ωt−k2·r) + c.c. (8.3)

is the second-harmonic field generated by this nonlinear process. For simplic-ity, let us assume that the light wave is propagating along the z direction, andfurther assume that the medium is transparent to ω and 2ω waves, e.g., A = 0.Figure 8.1 shows the collinear second-harmonic generation process.

For a uniform plane wave approximation, we have

∇× (∇× E) = −∇2E = − d2

dz2 E(2ω, z),

where

E(2ω, z) =12

E2ω(z)ei(2ωt−k2z) + c.c.

Page 185: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

168 8 Wave Propagation

Fig. 8.1 Second-harmonic wave generated by laser beam in a nonlin-ear crystal.

In the small-signal approximation, the fundamental field does not depreciate,but the second-harmonic field will grow; we have

− d2

dz2 E(2ω, z) =−12

[(d2E2ω(z)

dz2 − ik2dE2ω(z)

dz

)e−ik2z

−ik2

(dE2ω(z)

dz− ik2 E2ω(z)

)e−ik2z

].

(8.4)

For a slowly varying amplitude approximation, we have

d2E2w(z)dz2 k2

dE2w(z)dz

.

The nonlinear wave equation becomes

−12

[−i2k2

dE2ω(z)dz

− k22 E2ω(z) +

n22ω

c2 4ω2 E2ω(z)

]ei(2ωt−k2z)

=12

(4µ0ω2 P(2ω)

)ei(2ωt−2kz). (8.5)

Using the relations

ω =kcnω

, 2ω =k2cn2ω

,

we can substitute

k22 =

n22ω

c2 × 4ω2

and realize that the second and third terms on the left-hand side of the equa-tion cancel each other. We finally have

dE2ω(z)dz

=4µ0ω2P(2ω)

i2k2ei(k2−2k)z, (8.6)

Page 186: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.1 Nonlinear Wave Equation 169

which governs the growth of the→E2ω field as a function of z. Introducing

∆k = k2 − 2k,

the growth of the second-harmonic field can be obtained by integrating Eq.(8.6) with respect to z over the length of the nonlinear medium,

E2ω(z = ) =∫

0

[−i

2µ0ω2P(2ω)k2

]ei∆kzdz

=2µ0ω2P(2ω)

k2∆k

(1 − ei∆k)

)= A

(1 − ei∆k

),

(8.7)

where A =2µ0w2P(2w)

k2∆k. (8.8)

The intensity of the second-harmonic wave at the exit of the nonlinear crystalis given by

I2ω() =12

n2ω c ε0 | E2ω() |2

=12

n2ωcε0 | A |2(

1 − ei∆k) (

1 − e−i∆k)

=2n2ωcε0µ2

0ω4χ2(2ω)2E40(ω)

k20

sin2 ∆k2

( ∆k2 )2

.

(8.9)

In general, one can define a coherence length, c, for second-harmonic gener-ation as

sin∆kc

2= 1,

∆kc

2=

π

2or c =

π

∆k=

π

k2 − 2k

c =λ

4 (n2ω − nω).

(8.10)

In a nonlinear crystal with normal dispersion one usually has n2ω = nω. Themaximum second-harmonic signal one can expect to obtain from such a crys-tal is when the nonlinear crystal length is such that

sin∆k

2= sin(2m + 1)

π

2,

Page 187: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

170 8 Wave Propagation

or = (2m + 1)c. In this situation the maximum harmonic intensity one canexpect is

Imax(2ω, ) =2n2ωcε0µ2

0ω4χ2(2ω)E40(ω)

k20

(∆k2

)2 . (8.11)

The harmonic intensity is limited by the k-vector mismatch. Since p = hk rep-resents the photon momentum, the condition ∆kh = hk2 − 2hk is also referredto as the momentum mismatch. If one is clever, one may choose a nonlinearmedium in which one has n2ω = nω. In this case c = ∞ or ∆k = 0. Thisis the situation of momentum matching or phase matching. Since with phasematching the indices of the fundamental and second-harmonic waves must beequal, it is also referred to as index matching in the literature. Under this con-dition the phase velocity of the fundamental and second harmonic are equal;the two light waves always propagate in phase. Under the phase-matchingcondition

sin2 ∆k2(

∆k2

)2 → 1,

I2ω() ∝ 2. The longer the nonlinear crystal, the stronger the second-harmonic intensity. In practice, the second-harmonic intensity will alwaysbe less than the intensity of the fundamental wave. As the second-harmonicintensity grows stronger, the small-signal approximation breaks down. In re-ality, when the second-harmonic intensity becomes comparable to the funda-mental, it can act as the pump field, generating a new wave at the fundamentalfrequency by the nonlinear process

PNL(ω, 2ω,−ω) = χ(2)(ω, 2ω,−ω)E(2ω)E∗(ω). (8.12)

In this case, full coupled-wave equations between the ω and 2ω waves mustbe solved. If a proper phasing condition is maintained, it is possible to convertall fundamental energy into second harmonics [68].

8.2Phase Matching in Second-Harmonic Generation

In a crystal with normal dispersion, the indices of refraction of the fundamen-tal wave and the second harmonic are always different, and therefore phasematching is not possible. However, in a birefringent nonlinear crystal the in-dex of refraction of the ordinary ray (o-ray) and that of the extraordinary ray(e-ray) are different at a given wavelength. With a clever arrangement it may

Page 188: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.2 Phase Matching in SHG 171

Fig. 8.2 Normal (index) surfaces for the ordinary and extraordinaryrays in a negative (ne < no) uniaxial crystal.

be possible to obtain n2ω = nω by forcing the fundamental wave to propagateas an o-ray and the second harmonic as the e-ray or vice versa, depending onwhether the crystal is positive uniaxial or negative uniaxial. This is the usualtrick nonlinear optical engineers utilize to achieve index matching. In the earlydays researchers in nonlinear optics had to know the crystal optics so that theycould figure out how to manipulate the crystal orientation, laser polarization,etc., to achieve index matching. Nowadays, vendors of nonlinear optical ma-terials can usually provide a crystal that is already cut for the index-matchingconfiguration. A researcher after receiving the nonlinear crystal is merely re-quired to follow the instructions of the crystal supplier to perform the phase-matched frequency-doubling process. In this text we only show one exam-ple of how to achieve phase matching in potassium dihydrogen phosphate(KDP). Many new nonlinear crystals have been grown with various propertiesfor a variety of different applications, and these crystals and their propertiescan readily be found via various vendors’ online resources.1 Researchers orgraduate students, however, should know the fundamental principle of phasematching as illustrated in the following example.

8.2.1Phase Matching of Second-Harmonic Generation in KDP

Potassium dihydrogen phosphate (KDP) is a negative uniaxial crystal; the nor-mal (index) surface (note that it is not the index ellipsoidal surface) is shownin Fig. 8.2. A normal (index) surface reads the refractive-index value along the

1) See for example, Cleveland Crystal Inc.,www.clevelandcrystals.com.

Page 189: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

172 8 Wave Propagation

Fig. 8.3 Using the normal surface to illustrate the phase-matchingcondition for a negative uniaxial crystal, such as KDP. The fundamentalwave is propagating as an o-ray and the second harmonic as an e-ray.At the phase-matching angle θm, n2w

e (θm) = nwo .

direction of the k vector for the e- and o-rays. This is much more convenient touse in illustrating the index-matching process. Figure 8.3 shows the index ofthe e- and o-rays of KDP as a function of frequency. According to normal dis-persion, the index of the higher-frequency light wave is higher than that of thelower frequency. The index surface at 2ω is larger than that at ω. For a nega-tive uniaxial crystal, the e-ray surface is within the o-ray surface. Therefore, ifthe 2ω wave propagates as an e-ray and the ω wave as an o-ray their normal(index) surfaces will intersect at an angle θm. This means that the fundamen-tal laser beam should propagate along the θm direction, the phase-matchingdirection, as an o-ray, i.e., with its electric field polarized perpendicular toboth the optical axis (z-axis) and the k vector. The generated second-harmonicwave will propagate along exactly the same direction, k2ω, but polarized as ane-ray. This configuration is called collinear phase matching.

In order to find the phase-matching angle θm, one needs to use the equationfor the index ellipsoid of KDP (negative uniaxial crystal) and thus go back tothe index of an ellipsoidal surface

x2

n20

+y2

n20

+z2

n2e

= 1. (8.13)

Page 190: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.2 Phase Matching in SHG 173

Fig. 8.4 Index ellipsoidal surface of KDP, ne < no, at the x = 0 plane.

Figure 8.4 represents the index ellipsoid of KDP at the x = 0 plane. We have

| 0A |= ne(θ), z =| 0A | sin θ = ne(θ) sin θ,

y =| 0A | cos θ = ne(θ) cos θ,

n2e (θ) cos2 θ

n20

+n2

e (θ) sin2 θ

n2e

= 1, (8.14)

1n2

e (θ)=

cos2 θ

n20

+sin2 θ

n2e

.

At the phase-matching angle θm, we have

nω0 = n2ω

e (θm)

1(nω

0)2 =

1

[n2ωe (θm)]2

=cos2 θm(

n2ω0)2 +

sin2 θm

(n2ωe )2

and, solving for θm,

sin2 θm =

1

(nω0 )2 − 1

(n2ω0 )2

1

(n2ωe )2 − 1

(n2ω0 )2

. (8.15)

The calculated phase-matching angle, θm, means that the frequency-doublingprocess will have ∆k = 0 when the laser is propagating along this direction.From Eq. (8.9),

I2ω() =2n2ωc ε0µ2

0ω4χ22 E40(ω)

k20

∝ χ2E40(ω)2. (8.16)

Page 191: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

174 8 Wave Propagation

Under this condition the coherence length approaches infinity. I2ω is limited,among other things, by the length of the nonlinear crystal. Actually, the aboveequation is an oversimplified case since we have neglected the crystal ori-entation and the polarization of the optical field. To analyze the frequency-doubling process properly one needs to take these into account. We have forKDP

PNL

x (2ω)pNL

y (2ω)PNL

z (2ω)

=

0 0 0 d14 0 00 0 0 0 d14 00 0 0 0 0 d36

E2x(ω)

E2y(ω)

E2z(ω)

2Ey(ω)Ez(ω)2Ex(ω)Ez(ω)2Ex(ω)Ey(ω)

, (8.17)

where the d coefficient instead of χ is used. In general,

dijk =12

χ(2)ijk

and, furthermore, using the contracted notation for the third-rank tensor, wehave d14 = d123 = d132, d36 = d321 = d312, etc. The optical fields are expressedas the component fields along the crystallographic axes, x, y and z:

PNLx (2ω) = 2d14Ey(ω)Ez(ω)

PNLy (2ω) = 2d14Ex(ω)Ez(ω)

PNLz (2ω) = 2d36Ex(ω)Ey(ω).

(8.18)

The phase matching for this negative uniaxial crystal requires that the fun-damental laser beam must propagate as an o-ray, i.e., EL⊥ z-axis, or Ez(ω) =0. The only nonzero PNL(2ω) is along the z-axis, or

PNLz (2ω) = 2d36 Ex(ω) Ey(ω).

Fork(ω) in the (θ, φ) direction, EL(ω), which must be perpendicular to boththe z-axis and k(ω), lies in the x–y plane only; from Fig. 8.5

Ex(ω) = EL sin φ

Ey(ω) = −EL cos φ,

PNLz (2ω) = −2d36E2

L sin φ cos φ. (8.19)

The second-harmonic polarization has a z component; thus, it will create anoptical field that propagates as an e-ray. When the second-harmonic field ra-diates, only the component perpendicular to the direction of propagation mat-

ters, i.e., the effective PNL is that perpendicular to→k (2w), which is collinear

Page 192: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.2 Phase Matching in SHG 175

Fig. 8.5 Relationship of (a)→k (ω), (b)

→E L (ω) and (c) pNL

e f f (2ω) and→k (2ω) with both the fundamental and the second-harmonic wavespropagating along the (θ, φ) direction.

with→k (w):

PNL⊥ (2w) = PNL

e f f (2w)

= PNLz (2w) sin θ

= −2d36 E2L(w) sin θ sin φ cos φ

= de f f E2L = −d36 sin θ sin 2φE2

L,

de f f = −d36 sin θ sin 2φ. (8.20)

For the phase-matching case θ = θm one still needs to optimize de f f , leading toφ = 45 for this simple example. Alternatively, the de f f formula suggests thatthe polarization of the EL field not only should lie in the x–y plane to qualifyas an o-ray, it should not be polarized along the x- or y-axis. In other words,PNL

z (2ω) will be maximum when φ = 45 or when |Ex| = |Ey| = (1/√

2)EL.This simple fact suggests that one still needs to know what one is doing byorientating the crystal (by rotating it with respect to the crystal axis) and/orpolarizing the laser beam correctly.

The phase-matching example just described is the manifestation of the en-ergy and momentum conservation condition for the fundamental and second-harmonic photons:

hω + hω = h(2ω)

h→k L + h

→k L= h

→k (2ω).

(8.21)

In this example, two fundamental photons are both polarized as an o-ray. Thisis called type I phase matching. In a different case, two fundamental photons

Page 193: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

176 8 Wave Propagation

are polarized orthogonal to each other, i.e., one o-ray and one e-ray. This isknown as type II phase matching. The de f f ’s for various crystal classes forboth type I and type II phase matching have been previously calculated andthey can be found in Ref. [76].

8.2.2Noncollinear Momentum Matching

The index-matching example described in the previous section is a special caseof momentum matching when the fundamental and second-harmonic beamsare propagating collinearly. To describe the general concept of momentum

matching it is more instructive to use the→k representation. One can modify

the index-matching diagram using a normal (index) surface by a→k vector

diagram in the collinear case first:

kω =ω

cnω

k2ω =2ω

cn2ω.

Under the index-matching condition nω = n2ω, we have

2 | kω | = | k2ω |

or→k (ω)+

→k (ω) =

→k (2ω),

∆→k =

→k (2ω)−

→k (ω)−

→k (ω) = 0.

Fig. 8.6 Momentum matching in the collinear case.

Page 194: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.2 Phase Matching in SHG 177

Fig. 8.7 Noncollinear propagating fundamental waves showing mo-

mentum mismatch with ∆→k = 0.

This is depicted in Fig. 8.6. When the two→k (ω)’s are not collinear but make

a small angle with respect to each other (Fig. 8.7),

→k 1 (ω)+

→k 2 (ω) + ∆

→k =

→k (2ω) with ∆

→k = 0.

It may be possible, as shown in Fig. 8.8, that→k 1 (ω)+

→k 2 (ω) =

→k (2ω)

as in a true noncollinear situation. As long as the triangle closes, one hasa momentum-matching condition. Although the noncollinear momentum-matching condition satisfies mathematically the momentum-matching condi-tion, the process in general is less efficient in generating second harmonics,particularly when focused beams are used. In this case, all beams cannot re-main overlapped over a long propagating distance, resulting in the “walk-

Fig. 8.8 One example of noncollinear momentum matching.

Page 195: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

178 8 Wave Propagation

off” effect where the fundamental and the second harmonic beams walk offfrom each other. In other words, the interaction length between the variousbeams becomes a limiting factor, instead of the physical length of the non-linear crystal. However, nonlinear momentum-matching configurations havebeen used for ultrashort optical pulse measurement, since in this configurationthe second-harmonic and fundamental beams are spatially separated, result-ing in a background-free second-harmonic intensity autocorrelator [77].

8.2.3Experimental Arrangement for Phase-Matched Second-Harmonic Generation

From the discussion in the previous section, at phase matching, the intensityof the generated second harmonic I2ω is proportional to

I2ω ∝ d2e f f I2

L(ω)2. (8.22)

The goal of phase-matching SHG is to convert maximum laser light to that ofthe SH. To achieve this, the SHG is done in a transmission arrangement bychoosing an appropriate optically polished nonlinear crystal [78]. The crystalis cut and optically polished to ensure maximum transmission of light. Theorientation of the crystal is such that the crystal axis (the direction perpendic-ular to the end faces) is along the phase-matching direction, and that the z-axisof the crystal makes an angle θm with respect to the direction perpendicular tothe end faces. As mentioned earlier, nowadays these considerations are han-dled by the vendor. The user only needs to know what polarization the laserbeam needs to be in order to satisfy the particular type of phase-matching con-dition (i.e., whether the laser should be polarized as an o-ray or an e-ray, etc.).Since the transmission arrangement is used, the crystal has to be transparentto both the fundamental and second-harmonic wavelengths. One then focusesthe laser beam into the crystal as shown in Fig. 8.9.

Fig. 8.9 Experimental arrangement for detecting phase-matchedSHG. The laser beam is focused into the nonlinear crystal, NLC, oflength , and recollimated by another lens, f. A second-harmonic, 2ω,filter blocks the fundamental and passes the second-harmonic light,which is detected by the photodetector, PD.

Page 196: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.3 Parametric Interaction 179

In the focused arrangement with a Gaussian beam, it is not always optimumto focus as tight as possible since this will limit the effective crystal length tothe Rayleigh range of the lens, which may be much shorter than the length ofthe nonlinear crystal. Too tight a focus may also cause damage. Usually, theoptimal situation is to match the Rayleigh range to the length of the crystal.The filter shown in Fig. 8.9 is used to block the fundamental but let throughthe second harmonic. In practice, the filter used is a high-pass step functiontype filter which passes all light with a wavelength shorter than a certain valuebut blocks all light longer than this wavelength. Since the second harmonic isin the pass band and the fundamental in the stop band of the filter, the fil-ter provides a very effective means of separating the second harmonic fromthe fundamental wave. This type of pass-band filter usually is not expen-sive. Some colored glass filters may be used. Another way to discriminate thesecond harmonic from the fundamental beam is by spatial dispersion. For ex-ample, in the noncollinear phase-matching case the second harmonic is in thedirection bisecting the two noncollinear fundamental beams. Another simplemethod to spatially separate the second harmonic from the fundamental is tolet it pass through a dispersive prism. The second harmonic and fundamentalbeam will propagate along different directions after emerging from the prism.

8.3Parametric Interaction

8.3.1Coupled Equations for Parametric Interaction

We have seen how two photons at the fundamental frequency ω can be com-bined to generate a second-harmonic photon at 2ω frequency. In a more gen-eral sense, two photons at frequencies ω1 and ω2 can be combined to generatea sum-frequency photon at ω1 + ω2:

hω3 + hω2 =⇒ h(ω1 + ω2) = hω3. (8.23)

The arrow in the above equation is pointing from ω1 + ω2 toward ω3. Non-linear optical processes can occur in the reverse direction, namely, a higher-frequency photon at ω3 can be broken up into photons at ω1 and ω2, respec-tively. In other words, the arrow in the above equation is pointing as follows:

hω3 =⇒ hω1 + hω2. (8.24)

This is called parametric interaction. The nonlinear polarization for the sumfrequency generation process is

PNLi (ω3 = ω1 + ω2) = 2dijk(ω3 = ω1 + ω2)Ej(ω1)Ek(ω2), (8.25)

Page 197: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

180 8 Wave Propagation

where

Ej(ω1, z, t) =12

[E1j(z)ei(ω1t−k1z) + c.c.

],

Ek(ω2, z, t) =12

[E2k(z)ei(ω2t−k2z) + c.c.

],

Ei(ω3, z, t) =12

[E3i(z)ei(ω3t−k3z) + c.c.

].

(8.26)

One can write down the equation for the parametric process, PNLj (ω1 = ω3 −

ω2) and PNLk (ω2 = ω3 − ω1), in a similar manner:

PNLk (ω2 = ω3 − ω1) = 2dkijEi(ω3)E∗

j (ω1), (8.27)

where E∗j (ω1, z, t) =

12

[E∗

1je−i(ω1t−k1z) + c.c.

], (8.28)

and where ω3 = ω1 + ω2 is assumed. The wave equation is

−∇2→E +nAc

∂→E

∂t+

n2

c2∂2→E∂t2 = −µ0

∂2→

PNL

∂t2 , (8.29)

where n is the refractive index; nAc = µ0σ, σ being the conductivity of the

medium. We have chosen the light wave at various frequencies to propagatealong the z direction and also assumed that there is no x and y variation in thefield expression, i.e., we still deal with infinite plane waves for the purposeof keeping the mathematics simple. The wave equation under this specialsituation reduces to

−d2→Edz2 +

nAc

∂→E

∂t+

n2

c2∂2

→E

∂t2 = −µ0∂2

→PNL

∂t2 . (8.30)

The field and polarization in the nonlinear wave equation can represent thetotal field or polarization

→E=

→E1 (ω1, z, t)+

→E2 (ω2, z, t)+

→E3 (ω3, z, t),

→P

NL=

→P

NL

1 (ω1, z, t)+→P

NL

2 (ω2, z, t)+→P

NL

3 (ω3, z, t),

or the quantity at an individual frequency. Let us examine the ith componentof the E field at ω3. It satisfies the nonlinear Maxwell equation

− d2Ei(ω3, z, t)dz2 +

nAc

∂Ei(ω3, z, t)∂t

+n2

c2∂2Ei(ω3, z, t)

∂t2

= −µ0∂2PNL

i (ω3, z, t)∂t2 . (8.31)

Page 198: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.3 Parametric Interaction 181

Differentiating the first term twice and assuming the slowly varying envelopeapproximation,

k3dE3i(z)

dz=

λ3

dE3i(z)dz

d2E3i(z)dz2 , (8.32)

i.e., the change in the curvature of the amplitude as a function of z is muchless than the derivative over a distance of a wavelength. One can drop thesecond-derivative term; the nonlinear Maxwell equation then becomes

2ik3dE3i(z)

dz+ k2

3E3i(z) +n3 A

ciω3E3i(z)− ω2

3n23

c2 E3i(z)

= −µodijk(ω3 = ω1 + ω2)(−ω23)E1j(z)E2k(z)ei[k3−(k1+k2)]z. (8.33)

Dividing the equation by 2ik3 and realizing that

ω3

k3=

cn3

,

the equation becomes

dE3i(z)dz

= −12

AE3i(z) − ω3icµo

n3

12

dijk E1j(z)E2k(z) ei[k3−(k1+k2)]z,

since

nAc

= µ0σ,

or

A =cµ0

nσ =

√ε0

ε

õ0

ε0=

õ0

εσ.

Finally, we can express the equation as

dE3i(z)dz

= −12

õ0

εσ3E3i(z)− iω3

2

õ0

ε3dijk(ω3 = ω1 + ω2)

× E1j(z)E2k(z) ei[k3−(k1+k2)]z.(8.34)

Similarly, one can obtain

dE1j(z)dz

= −12

õ0

ε1σ1E1j(z) − iω1

2

õ0

ε1djik(ω1 = ω3 − ω2)

× E3i(z)E∗2k(z) ei[k1−(k3−k2)]z,

(8.35)

dE2k(z)dz

= −12

õ0

ε1σ2E2k(z) − iω2

2

õ0

ε2dkij(ω2 = ω3 − ω1)

× E3i(z)E∗1j(z) ei[k2−(k3−k1)]z

(8.36)

Page 199: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

182 8 Wave Propagation

or, if one takes the complex conjugate of the above equation,

dE∗2k(z)dz

= −12

õ0

ε2σ2E∗

2k(z) +iω2

2

õ0

ε2dkij(ω2 = ω3 − ω1)

× E∗3i(z)E1j(z) ei[−k2+(k3−k1)]z.

(8.37)

Defining ∆k = k3 − (k1 + k2) and also noting that the symmetry of the d coef-ficient satisfies

dijk(ω3, ω1, ω2) = djik(ω1, ω3 − ω2)

= d,

etc., we finally have a set of coupled wave equations:

dE3i(z)dz

= −12

õ0

ε3σ3E3i(z)− iω3

2

õ0

ε3dE1j(z)E2k(z) ei∆kz, (8.38)

dE1j(z)dz

= −12

õ0

ε1σ1E1j(z) − iω1

2

õ0

ε1dE3i(z)E∗

2k(z) e−i∆kz, (8.39)

dE∗2k(z)dz

= −12

õ0

ε2σ2E∗

2k(z) +iω2

2

õ0

ε2dE∗

3i(z)E1j(z) ei∆kz. (8.40)

This set of coupled wave equations governs the growth of the amplitude ofthe specific optical field. The first equation describes the sum-frequency gen-eration. If we denote ω3 = ωp, the frequency of the pump wave, ω1 = ωs,the signal wave and ω2 = ωi, the idler wave, then the second and third equa-tions describe the parametric process. It is instructive to express the intensityin terms of photon density from the relation

I(ω) =12

nω c0ε0 | E(ω) |2 =12

√ε0

µ0| E(ω) |2 . (8.41)

Let us introduce

A =√

µ

ωE,

I =12

n

√ε0

µ0

ω

n| A |2 =

12

√ε0

µ0ω | A |2 . (8.42)

On the other hand, the intensity can also be expressed as the product of thephoton density q, (#/cm3)× hω (energy per photon) ×c:

I = q hω c =12

√ε0

µ0ω | A |2

Page 200: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.3 Parametric Interaction 183

or

q =1hc

12

√ε0

µ0| A |2 . (8.43)

Substituting E’s everywhere in the three coupled equations by A’s, the equa-tion can be written as

dA3i(z)dz

=−12

õ0

ε3σ3

√ω3

n3A3i(z)

− iω3

2

õ0

ε3d√

ω1

n1

√ω2

n2A1j(z)A2k(z) ei∆kz.

Substituting√

1εi

= 1ni

1√ε0

, i = 1, 2 or 3, one can write the equation as

dA3i(z)dz

=−12

õ0

ε3σ3A3i(z) − i

2

õ0

ε0d√

ω1ω2ω3

n1n2n3A1j(z)A2k(z)ei∆kz.

(8.44)

Since the loss is

αi =√

µ0

εiσi, (8.45)

and defining a coupling coefficient λ,

λ =√

µ0

ε0d√

ω1ω2ω3

n1n2n3, (8.46)

we finally arrive at

dA3i(z)dz

=−12

α3 A3i(z)− i2

λA1j(z)A2k(z)ei∆kz. (8.47)

Similarly,

dA1j(z)dz

= −12

α1A1z(z)− i2

λA3i(z)A∗2k(z)e−i∆kz, (8.48)

dA∗2k(z)dz

= −12

α2 A∗2k(z) +

i2

λA∗3i(z)A1j(z)ei∆kz. (8.49)

Equations (8.47)–(8.49) are referred to as the coupled wave equations (or sim-ply as coupled equations). Note that there is only a single coupling coefficientλ involved in all equations. This set of three equations forms the basis foranalyzing parametric interaction as will be discussed in the following section.

Page 201: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

184 8 Wave Propagation

Fig. 8.10 Schematics of an optical parametric amplifier. Ap(0) ≈Ap(), As() As(0).

8.3.2Parametric Amplification

A single-pass parametric interaction may lead to the amplification of a signalwave when the nonlinear crystal is pumped by a strong pump beam. This isthe basis for the optical parametric amplifier (OPA) [79]. The simple schemat-ics of a parametric amplifier are shown in Fig. 8.10.

The nonlinear crystal cut correctly with proper orientation is pumped by astrong pump beam at ω3 = ωp, the signal wave is at frequency ω1 = ωs andthe idler is ω2 = ωi, where

ωp = ωs + ωi,

so that energy-conservation relation is satisfied. At the entrance face, we as-sume that Ap(0), As(0) and Ai(0) are known (in fact, either As(0) or Ai(0)may be zero). We further assume that the pump beam is not depleted. Thismeans that Ap(z) = Ap(0), independent of z. We thus only need to deal withtwo coupled equations. Let us further assume, for simplicity, that there is noloss at ωs and ωi, αs = αi = 0. We can also assume that the pump beam is areal field, Ap(0) = A∗

p(0). Let us also assume that the polarizations of Ap, As

and Ai are consistent with the crystallographic orientation and that we do notneed to worry about them and that we can drop the i, j, k indices denotingCartesian components. Then

dAs(z)dz

= − i2

λ Ap A∗i (z)e−i∆kz

dAi(z)dz

=i2

λ Ap As(z)ei∆kz.(8.50)

Page 202: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.3 Parametric Interaction 185

Also, if we denote

λAp = g, (8.51)

the equations become

dAs(z)dz

= − i2

g A∗i (z) e−i∆kz, (8.52a)

dA∗i (z)

dz=

i2

g As ei∆kz. (8.52b)

Differentiating Eq. (8.52a) with respect to z,

d2As(z)dz(2)

= − i2

g[

dA∗i (z)

dze−i∆kz − i∆kA∗

i (z)e−i∆kz]

and, substituting dA∗i (z)/dz from Eq. (8.52b), we have

d2As(z)dz2 =

14

g2 As(z) − i∆k[−i

2g A∗

i (z)e−i∆kz]

.

Then, substituting the second term on the right-hand side of the above equa-tion, we arrive at

d2As(z)dz2 + i∆k

dAs(z)dz

− g2

4As(z) = 0. (8.53)

The solution for As(z) takes the following form:

As(z) = A e[− i∆k

2 + 12

√g2−∆k2

]z + B e

[− i∆k

2 − 12

√g2−∆k2

]z. (8.54)

Substituting the expression of As(z) back into Eq. (8.52a), we can obtain theexpression for A∗

i (z):

A∗i (z) =

2ig

[A(− i∆k

2+

12

√g2 − ∆k2

)e[

i∆k2 + 1

2

√g2−∆k2

]z]

+ B(− i∆k

2− 1

2

√g2 − ∆k2

)e(

i i∆k2 − 1

2

√g2−∆k2

)z.

(8.55)

The undetermined coefficients A and B can be fixed by applying the boundaryconditions.

A. Phase-Matching Case, ∆k = 0

One of the important applications of the parametric nonlinear process is para-metric amplification. A parametric amplifier is a nonlinear optical device thatis designed to generate a maximum-intensity light wave at signal frequency.

Page 203: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

186 8 Wave Propagation

Therefore, the nonlinear optical crystal is already cut and the experimentalconfiguration has been set such that the phase-matching condition is satisfied:

∆k = kp − (ks − ki) = 0.

For such a situation kp,ks and ki are collinear to have maximum interactionlength. Therefore, the vector k relation becomes the scalar relation kp = ks + ki.This leads to the solution of parametric amplification for the phase-matchingcase of ∆k = 0:

As(0) = A + B,

A∗i (0) =

2ig

[A

12

g − B12

g]

= i(A − B),

when A and B are expressed as

A =12

[As(0)− iA∗i (0)] ,

B =12

[As(0) + iA∗i (0)] ,

and are substituted back into the expressions for As(z) and A∗i (z) to give

As(z) = As(0)12

[e

12 gz + e

−12 gz

]+ iA∗

i (0)12

[e−12 gz − e

12 gz]

= As(0) cos hgz2

− iA∗i (0) sin h

gz2

.(8.56)

Following a similar procedure, we can obtain

A∗i (z) = iAs(0) sin h

gz2

+ A∗i (0) cos h

gz2

. (8.57)

B. Non-Phase-Matching Case ∆k = 0

In the general case when there is no phase matching, i.e., ∆k = 0, one canstill obtain a mathematical expression for the solution of As(z) and A∗

i (z).Defining

b =12

√g2 − ∆k2,

As(z) ei∆kz

2 = As(0)[

cos h(bz) +i∆k2b

sin h(bz)]

− ig

2bA∗

i (0) sin h(bz)(8.58)

Page 204: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.3 Parametric Interaction 187

and

A∗i (z) e

−i∆kz2 = A∗

i (0)[

cos h(bz) − i∆k2b

sin h(bz)]

+ ig

2bAs(0) sin h(bz).

(8.59)

From the solution shown in Eqs. (8.56–8.57) (∆k = 0) and Eqs. (8.58–8.59)(∆k = 0), it is clear that initially only one wave (either signal or idler wave)needs to be present in order to generate a growing signal or idler wave. Whenthe process is not phase matched, in order for the signal or idler wave to grow,it must have a real b value; b is a real quantity only if g > ∆k. When thiscondition is satisfied the amplitude of the signal or idler wave will grow ex-ponentially according to ebz. Otherwise, b is imaginary and cos h(bz) becomesan oscillatory function of z. Therefore, the mismatch ∆k must be smaller than

g = d√

µ0ω1ω2ω3

ε0n1n2n3

(õ3

ω3E3(0)

).

In the case of phase matching, ∆k = 0 and also assuming that Ai(0) = 0,

As(z) = As(0) cos hgz2

.

When gz 1,| As(z) |2| As(0) |2 → 1

4egz;

the signal wave grows exponentially as a function of z. This is the theoreticalbasis for the commercial parametric amplifier. To achieve gz 1, one shouldtry to use a long crystal. In this case, the crystal length L limits the interactionlength.

gL =√

µ0ω1ω2ω3

ε0 n1n2n3

õ3

ω3dE3(0)L 1.

It is obvious that we need: (1) a crystal with a large d coefficient, (2) a long crys-tal (provided that the phase-matchable interaction can be maintained over theentire length of the crystal) and (3) a strong pump field. Strong pump fieldscan always be accomplished by focusing. However, focusing too strongly willshorten the phase-matchable interaction length (this is generally the main lim-iting length factor in the parametric amplification process, not the physicallength of the crystal).

Based on these guiding principles, a commercial parametric amplifier canbe built. The wavelength tuning of the OPA is achieved by angular rotationof the OPA crystal. One essentially lets the phase-matching process select apair of signal and idler wavelengths so that both energy and momentum areconserved.

Page 205: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

188 8 Wave Propagation

8.4Parametric Oscillation

The parametric amplification process is essentially a single-path processwhich does not require feedback or recycling of the signal wave throughthe pumped medium. In contrast to the single-path process, one can builda cavity which allows both the signal and idler waves to recirculate throughthe nonlinear crystal. With proper feedback control one can construct anoscillator based on the parametric process. The optical parametric oscillator(OPO) in every aspect looks like a laser. An OPO cavity is formed by twofeedback reflectors in a Fabry–Perot cavity arrangement as shown in Fig. 8.11.To achieve the best result, the cavity should be designed to provide modematching between the pump and signal waves.

Fig. 8.11 Schematic diagram of optical parametric oscillator. Thepump wave at ωp gives rise to oscillations at ωs and ωi in an opticalcavity that contains the nonlinear crystal, NLC, and resonates at ωsand ωi.

The signal and idler waves are expressed as

Es = Es(z, t)ei(ωst−ks·r),

Ei = Ei(z, t)ei(ωit−ki·r),

respectively. The associated “photon” field through the relation of As =√nsω

E, etc., is given by

As = As(z, t)ei(ωst−ks·r)

and

Ai = Ai(z, t)ei(ωit−ki·r). (8.60)

In the OPA case, one is interested in investigating the amplitude growth asa function of position (inside the nonlinear crystal), i.e., As(z) and Ai(z). In

Page 206: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.4 Parametric Oscillation 189

the OPO case, one wants to examine how the amplitude of As and Ai growas a function of time. It would be clearer if the spatially dependent functionand differential equation are converted into temporally dependent ones. Letus recall the coupled equations for parametric interaction, Eqs. (8.47)–(8.49),

dAs(z)dz

=−12

αs As(z)− i2

λAp A∗i (z)e−i∆kz,

dA∗i (z)

dz=

−12

αi A∗i (z) +

i2

λAp As(z)ei∆kz, (8.61)

where we have taken A∗p = Ap. To convert this set of equations from z depen-

dence to time dependence, let us recall that

z = vt =cn

t,

ddz

=nc

ddt

.

Also, realizing that the only interesting case for an OPO is when ∆k = 0, Eq.(8.61) becomes

ns

cdAs(t)

dt= −1

2αs As(t) − i

2λ Ap A∗

i (t), (8.62)

ni

cdA∗

i (t)dt

=−12

αi A∗i (t) +

i2

λ Ap As(t). (8.63)

We are interested in investigating how the amplitude function of As(t) andA∗

i (t) grows at a fixed point inside the nonlinear crystal as it is pumped by

a cw pump Ap ei(ωpt−kp·r). If the device leads to oscillation, As(t) and A∗i (t)

will reach a constant value. Therefore,

dAs(t)dt

= 0,dA∗

i (t)dt

= 0

at oscillation. These conditions convert the set of differential equations inEqs. (8.62)–(8.63) into a set of simultaneous algebraic equations:−1

2 αs As − i2 λAp A∗

i = 0,−12 αi A∗

i + i2 λAp As = 0.

The necessary condition for a nontrivial solution of As and A∗i to exist is for

the secular equation to be zero, or

∣∣∣∣∣ −12 αs

−i2 λAp,th

i2 λ Ap,th

−12 αi

∣∣∣∣∣ = 0,

Page 207: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

190 8 Wave Propagation

where we have denoted the Ap that satisfies the secular equation as the thresh-old pump wave amplitude Ap,th with a threshold value of

λ2 | Ap,th |2 = g2 = αiαs. (8.64)

We reach essentially the same conclusion with an OPO as in a laser, i.e., thegain equal to the loss sets the threshold condition for lasing.

Remember that λ2 =µo

εod2 ωpωsωi

npnsni.

We have the threshold pump beam intensity

Ith(ωp) =12

√εo

µoωp | Ap,th |2

=12

(εo

µo

) 32 npnsni

d2αiαs

ωiωs.

(8.65)

Since the losses αi and αs are distributed loss parameters, if there is no suchloss in the OPO, one may define the total loss from the OPO as

αiαs 2R = (1 − Ri)(1 − Rs),

where R is the length of the OPO cavity and Ri and Rs are the reflectivities ofthe idler and signal waves, respectively. The threshold pumping intensity isthen

Ith(ωp) =12

(εo

µO

) 32 npnsni(1 − Ri)(1− Rs)

ωsωi 2Rd2

. (8.66)

In a singly resonated OPO, one may have Ri = 0 to suppress the oscillation ofthe idler wave at the expense of high threshold pumping intensity.

8.4.1Tuning of OPO

The output wavelength of the OPO is governed by the phase-matching condi-tion (momentum matching)

ωpnp

c=

nsωs

c+

niωi

c(8.67)

and the energy-conservation condition

ωp = ωs + ωi. (8.68)

Page 208: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.4 Parametric Oscillation 191

For the sake of discussion, let us assume that the nonlinear crystal is bire-fringent and the pump beam is propagating as an e-ray. Initially along the θ

direction one has

ωpnep(θ) = ω0

s n0s + ω0

i n0i , (8.69)

where nep(θ) is a function of θ. We may tune the angle from θ to θ + ∆θ, in

which case

nep → ne

p + ∆np

n0s → n0

s +∂ns

∂ωs| ∆ωs

n0i → n0

i +∂ni

∂ωi| ∆ωi.

The energy-conservation condition leads to

∆ωs = −∆ωi,

since ωp is fixed. The differential relation of Eq. (8.68) is

ωp∆nep + ∆ωpne

p = ω0s ∆n0

s + ∆ω0s n0

s + ω0i ∆n0

i + ∆ω0i n0

i ,

ωp

(ne

p +∂ne

p

∂θ∆θ

)= ω0

s

(n0

s +∂n0

s∂ω

∆ωs

)+ ∆ωsn0

s + ω0i

(n0

i +∂ni

∂ω∆ωi

)+ ∆ωin

0i ,

from which one obtains

∆ωs

∆θ=

ωp∂ne

p∂θ(

n0s − n0

i

)+[

ω0s

∂n0s

∂ω − ω0i

∂n0i

∂ω

] , (8.70)

since np(θ) satisfies

1n2

p(θ)=

cos2(θ)n2

0(ωp)+

sin2 θ

n2e (ωp)

.

After some calculation, it is possible to show that

∂np(θ)∂θ

=n3

p(θ) sin 2θ

2

[1

n20(ωp)

− 1n0

e (ωp)

], (8.71)

or

∆ωs

∆θ=

12 ωpn2

p(θ0) sin 2θ

[1

n20(ωp)

− 1n2

e (ωp)

](n0

s − n0i

)+[

ω0s

∂n0s

∂ω − ω0i

∂n0i

∂ω

] , (8.72)

Page 209: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

192 8 Wave Propagation

as the angular tuning relationship for the signal wave. This relation showsthat the birefringence, dispersion and initial angle all play important roles infrequency-tuning characteristics of the OPO [80].

8.5The Manley–Rowe Relations

In the parametric interaction, photons at higher energy (pump photons)are converted to photons of lesser energy (signal and idler photons). Therules that govern the photon number density at various frequencies are theManley–Rowe relations [81]. They can be derived from the coupled equations,Eqs. (8.47)–(8.49).

From the expression of

ddz

| Ap(z) |2= Ap(z)dA∗

p(z)dz

+dAp(z)

dzA∗

p(z),

one substitutes from the coupled equation, Eq. (8.47),

dAp(z)dz

=−i2

λAs(z)Ai(z)ei∆kz

and its complex conjugate to obtain

d | Ap(z) |2dz

=−i2

λAs(z)Ai(z)A∗p(z) ei∆kz +

i2

λA∗s A∗

i Ap e−i∆kz. (8.73)

Similarly, one can obtain

d | As(z) |2dz

=i2

λAs A∗p Ai ei∆kz − i

2λAp A∗

i A∗s e−i∆kz (8.74)

and

d | Ai(z) |2dz

=i2

λAi A∗p Asei∆kz − i

2λA∗

i Ap A∗s e−i∆kz. (8.75)

We observe that the right-hand side of Eq. (8.74) is the same as that of Eq.(8.75) and has opposite sign to the right-hand side of Eq. (8.73); we then have

d | As(z) |2dz

=d | Ai |2

dz= −d | Ap |2

dz.

Since

I =12

√ε0

µω | A |2 =

P(ω)s

= q hωc,

Page 210: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.5 The Manley–Rowe Relations 193

where P(ω) is the power at ω and s the cross-sectional area of the beam, weobtain

dqs(ωs)dz

=dqi(ωi)

dz= −dqp(ωp)

dz. (8.76)

We have

ddz

Ps(z)ωs

=ddz

Pi(z)ωi

= − ddz

Pp(z)ωp

. (8.77)

Equation (8.76) states that the signal photon density increase per unit lengthin the medium is exactly the same as the increase in idler photon density andat the expense of the decreasing pump photon density. In other words, in thisparametric process the one pump photon is split into one signal and one idlerphoton.

By integrating Eq. (8.77) and rearranging, we arrive at

(Ps)out − (Ps)in(Pp)

in −(

Pp)

out=

ωs

ωp, (8.78)

(Pi)out − (Pi)in(Pp)

in −(

Pp)

out=

ωi

ωp, (8.79)

where the frequency relation is shown in Fig. 8.12. Note that ωs/ωp < 1 andωi/ωp < 1.

Fig. 8.12 Manley–Rowe relation showing that one pump photon gen-erates one signal and one idler photon.

The equations (8.76–8.79) are referred to as the Manley–Rowe relations.They mean that the net power gain in signal power is always less than thepower loss in pump photons according to the ratio in frequency. In order tohave maximum power conversion from pump to signal wave, the frequencyof the signal wave should be close to that of the pump wave.

Page 211: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

194 8 Wave Propagation

8.6Parametric Upconversion

An inverse process to parametric generation is parametric upconversion,2

where

ωin + ωp → ωco. (8.80)

This relation indicates that an initial photon at frequency ωin is upconvertedto a photon at higher frequency ωco with the aid of a strong pump beam atfrequency ωp. One application of the parametric upconversion process is toconvert a photon in a spectral region where there is no sensitive photodetectoravailable into a photon at higher frequency where a sensitive photodetectoris readily available. Since this conversion process is coherent, the convertedphoton keeps all the characteristics of the original photon. Assuming that∆k = 0,

dAco(z)dz

=−i2

λAin(z)Ap(z)

dAin

dz= − i

2λAco A∗

p(z).(8.81)

This set of equations differs from the set describing parametric interaction inthat the signs of the terms on the right-hand side are both negative in thecurrent case, while the signs are opposite in the parametric interaction case.Let us assume that Ap is real and define the gain similar to Eq. (8.51),

g = λ Ap;

we then have

d2Aco(z)dz2 = −

( g2

)2Aco,

which leads to the oscillatory solution

Aco(z) = Aco(0) cosgz2

− i Ain(0) singz2

Ain(z) = Ain(0) cosgz2

− i Aco singz2

.(8.82)

The Manley–Rowe relation becomes

Pin(z)− Pin(0)ωin

=Pco(0)− Pco(z)

ωco.

2) Parametric amplification and oscillation are parametric downcon-version processes, where a pump photon is split into two photonsof lower frequencies. In parametric upconversion, the process isreversed, i.e., a lower-frequency photon is converted to a higher-frequency photon with the aid of a pump photon.

Page 212: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

8.6 Parametric Upconversion 195

Since

Pco(0) = 0,

this can be written as

Pin(0)ωin

=Pco(z)

ωco+

Pin(z)ωin

(8.83)

or

nin(0) = nco(z) + nin(z). (8.84)

In other words,nco(z) ≤ nin(0).

In this upconversion process, the maximum number of upconverted photonsis limited by the number of initial photons at the original frequency ωin. Theabove relation can also be obtained by examining the expressions of Ain(z)and Aco(z), realizing that Aco(0) = 0,

| Ain(z) |2 + | Aco(z) |2=| Ain(0) |2

are equivalent to the expression seen in Eq. (8.84). The upconversion effi-ciency is defined as

η =Pco()Pin(0)

=ωco

ωinsin2

(g

2

).

For g 1 and from Eqs. (8.46) and (8.51),

g = λAp =√

µ0

εod

√ωinωpωco

ninnpncoAp.

We have

η =wco

win

g22

4=

w2co 2d2

2ηinηpηw

(µ0

ε0

) 32

Ip, (8.85)

where Ip is the intensity of the pump. The upconversion process is sometimesused to obtain a visible image of an original infrared image [82]. If Ip is froman ultrashort laser pulse, it can be used as the gate pulse to gate an opticalevent taking place at the infrared spectral region. This effect is used to obtaintime-resolved detection of infrared waves [83].

Page 213: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

196 8 Wave Propagation

Problems

8.1 Derive de f f for 622, 422 and 3m crystals.

8.2 The first nonlinear optical experiment was done in crystalline quartz,which belongs to the trigonal symmetry group of class 32.

a) Show that in order to generate optical second harmonics, the incidentlaser beam cannot propagate along the x direction.

b) In a standard optical harmonic generation experiment, a z-cut quartzcrystal normal to the platelet (z-cut means that the z-axis of the quartzcrystal is normal to the platelet) is used as a reference to monitor thesquare of the laser intensity, [I(w)]2. The z-cut quartz crystal is chosenmainly because of the convenience in not having to worry about theorientation of the laser polarization with respect to the crystalline axis.Show explicitly with a formula that this is the case. The nonlinear dtensor of the 32 class is d11 −d11 0 d14 0 0

0 0 0 0 −d14 −d140 0 0 0 0 0

.

8.3 How to cut a rectangular LiIO3 crystal for type I SHG (λ1 = 1.06 µm)?

n0(ω) = 1.8517, n0(2ω) = 1.8976,

ne(ω) = 1.7168, ne(2ω) = 1.7473.

8.4 In a birefringent crystal, the direction of electromagnetic wave propaga-tion is different from the direction of energy transport. Discuss the implicationof this phenomenon on the phase matching of optical harmonic generation,assuming that the laser beam is a very thin pencil of light and the nonlinearcrystal is very long.

8.5 a) The nonlinear polarization for the third-harmonic generation in anisotropic medium is given by

Pi(3)(3ω) = 3CEi(ω)(E(ω) · E(ω)),

where C can be regarded as a constant. Show that a circularly polar-ized fundamental beam cannot generate a third-harmonic wave in thismedium while a linearly polarized beam can.

b) Using the effect described in Part 8.5, devise a method to measure thepulse duration of a picosecond optical pulse. (Assume that you have aseries of identical pulses emitted at a regular time interval.)

c) Is it possible to achieve phase matching for the process shown in Part8.5? How?

Page 214: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Problems 197

8.6 Use a normal (index) surface to make a sketch to illustrate

a) Collinear type I phase matching.

b) A noncollinear type I phase-matching process so that the second-harmonic wave is propagating along the direction normal to the opticalaxis.

c) Phase matching for the sum-frequency generation process.

8.7 The LiNbO3 crystal has a crystal symmetry of 3m. The electrooptic coeffi-cients are of the form

0 −γ22 γ130 γ22 γ130 0 γ330 γ51 0

γ51 0 0−γ22 0 0

.

The index of an ellipsoid of a LiNbO3 crystal is of the form

x2

n2o

+y2

n2o

+z2

n2e

= 1.

If we now apply a DC electric field along the c-axis (or z-axis) of the crystaland send in a linearly polarized optical beam propagating along the a-axis(or x-axis), derive an expression for the phase retardation of the LiNbO3 platewith thickness 1 assuming that the polarization of the light field is 45 withrespect to the c-axis.

8.8 Assume that a nonlinear crystal can be fabricated in such a way that (seeFig. 8.13) it consists of alternating sections each with the thickness equal tothe coherence length but the inversion image of the other. It is known that asingle such section can convert the fundamental to the second harmonic with1% conversion efficiency. Assuming that all the other experimental parame-ters remain the same, what would be the expected conversion efficiency if thecrystal contains five (5) sections?

Fig. 8.13 Nonlinear crystal with alternating sections.

Page 215: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 216: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

199

9Quantum Theory of Nonlinear Processes

9.1Introduction

Up to this point we have treated nonlinear interactions in optical media phe-nomenologically. We assumed that there was a nonlinear susceptibility as-sociated with a particular nonlinear process. For readers who are interestedin exploring the theoretical analysis from fundamental physical principles, itis pointed out here that nonlinear susceptibilities and multiphoton transitionprobabilities can be calculated using quantum mechanical analysis to obtainexplicit expressions. The objective for the presentation in this chapter is to pro-vide readers with an overall perspective of how nonlinear coefficients may bederived from basic quantum principles. Examples of calculations are limitedto some specific, and usually simple, cases. Readers who are interested in amore advanced theoretical treatment should consult Refs. [67, 84, 85].

We begin the theoretical discussion by reviewing the density operator for-malism and modifying it for nonlinear processes. For a nonparametric processonly the two-photon absorption cross section is presented. For the parametricprocess only χ(2)(2ω) and χ(3) will be calculated.

9.2Density Matrix Formalism for Calculation of Nonlinear Coefficient

The equation of motion for the density operator ρ is [86]

∂ρ

∂t=

1ih

[H, ρ]. (9.1)

The equation for the off-diagonal matrix elements of ρ is

∂ρij

∂t=

1ih

[H, ρ]ij, (9.2)

where

ρij =< ui | ρ | uj >

Page 217: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

200 9 Quantum Theory

and where the | ui >’s are the eigenstates of the unperturbed Hamilton-ian H0

1:

H0 | ui > = Ei | ui > (9.3)

and

H = H0 + H1. (9.4)

H1 is the perturbing Hamiltonian; Eq. (9.1) becomes

ih∂ρij

∂t= (Ei − Ej)ρij + [H1, ρ]ij, (9.5)

where [H1, ρ]ij is the off-diagonal matrix element of the commutator of theperturbing Hamiltonian H1 and the density operator ρ. H1 has two contribu-tions: H′, the external perturbation and Hr, the randomized internal pertur-bation which restores the system back to the equilibrium condition after it wasperturbed:

H1 = H′ + Hr. (9.6)

Substituting back, one has

ih∂ρij

∂t= hωij ρij + [H′, ρ]ij + [Hr , ρ]ij, (9.7)

where hωij = Ei − Ej.In the absence of an external perturbation, H′, ρij(t → ∞) → 0 for i = j.

It is customary to introduce a phenomenological relaxation time for the off-diagonal density operator elements, such that

[Hr, ρ]ij → ihτij

ρij (9.8)

and

ρij(t) = ρij(0) e−t/τij e−iwijt. (9.9)

The diagonal elements ρjj approach a time-independent equilibrium value ρejj

as H′ → 0. The equation of motion for the diagonal element ρjj is

ih∂ρjj

∂t= [H′, ρ]jj + ih Σk

(ρkk Wkj − ρjj Wjk

), (9.10)

where the terms ρkkWkj in Eq. (9.10) represent the per-unit-time increase in theprobability of occupation of the state j resulting from a transition from state

1) For a review of quantum mechanics, please see Appendix B.8.

Page 218: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.3 Perturbation Method 201

k as a result of a randomized perturbation. The terms −ρjjWjk represent theper-unit-time decrease in the probability of occupation of the j state resultingfrom a transition from the j state to the k state. With no external perturbation,H′ = 0, one demands that ∂ρjj/∂t = 0 when the system is in equilibrium.Therefore,

Σk

(ρe

kkWkj − ρejjWjk

)= 0.

When the system does not have any net emission or absorption of energy atany of the transition frequencies, the principle of detailed balance requires that

ρekk Wkj = ρe

jj Wjk (9.11)

or

ρekk

Wjk=

ρejj

Wkj.

We define the relaxation times Tjk = ρekk/Wjk and Tkj = ρe

jj/Wkj. From theprinciple of detailed balance it is obvious that Tkj = Tjk = Ti, the longitudinalrelaxation time. We have

ih∂ρjj

∂t= [H′, ρ]jj + ihΣk

(ρkk

ρejj

T1− ρjj

ρekk

T1

),

ih∂ρjj

∂t= [H′, ρ]jj +

ihT1

(ρe

jj − ρij

). (9.12)

We note that

ρejj =

e−Ej/kT

Σm e−Em/kT (9.13)

follows a Boltzmann distribution.For a two-level system, one has

ih∂ρ11

∂t= [H′, ρ]11 +

ihT1

(ρe11 − ρ11) , (9.14)

ih∂ρ22

∂t= [H′, ρ]22 +

ihT1

(ρe22 − ρ22) , (9.15)

ih∂ρ12

∂t= hw12 ρ12 + [H′, ρ]12 − ih

T2ρ12. (9.16)

9.3Perturbation Method in Density Operator Formalism

We will now apply the perturbation method to the quantum system. Let uschoose a perturbation parameter λ such that

H = H0 + λH′, (9.17)

Page 219: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

202 9 Quantum Theory

where H0 is the unperturbed Hamiltonian and H′ represents the external per-turbation. With λ = 1 the perturbation is fully turned on and, with λ = 0,there is no perturbation. One can also expand ρ in terms of powers of λ fol-lowing the justification given below.

From

∂ρ

∂t= −ih−1[H, ρ],∫ t

0dρ(t) = −ih−1

∫ t

0[H, ρ(t′)]dt′,

ρ(t) = ρ(0)− ih−1∫ t

0[H, ρ(0)]dt′,

for the first-order approximation, where we simply substitute ρ(t′) by ρ(0). Togo to higher orders we need to use the recursion formula for ρ(t) repeatedly.Carrying out this procedure to the second order in ρ, we have

ρ(t)− ρ(0) = −ih−1∫ t

0[H, ρ(0)]dt′︸ ︷︷ ︸

ρ(1)

− ih−2∫ t

0dt′

∫ t′

0dt′′ [H, [H, ρ(0)]]︸ ︷︷ ︸

ρ(2)

, (9.18)

where the first integral may be defined as ρ(1) and the second, ρ(2), etc. Onemay write

ρ = Σp λpρ(p) and

ρij = Σp λpρ(p)ij .

Substituting into the equation

ih∂ρjj

∂t= [λH′, ρ]jj,

we have

ihΣpλp∂ρ

(p)jj

∂t= Σpλp+1 [H′, ρ(p)]jj

and, equating λp terms on both sides, we have the recursion formula for ρ(p):

ih∂

∂tρ

(p)jj = [H′, ρ(p−1)]jj,

for the case excluding a damping term. When the damping terms are in-cluded, the equation becomes

ih∂

∂tρ

(p)jj = [H′, ρ(p−1)]jj + ih Σk

(p)kk Wkj − ρ

(p)jj Wjk

). (9.19)

Page 220: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.4 Transition Probability 203

Similarly, one can derive the recursion formula for the off-diagonal ele-ments, ρij,

ih

(∂

∂t+ iωij +

1τij

(p)ij = [H′, ρ(p−1)]ij. (9.20)

9.4Transition Probability in Electric Dipole Interaction

In a nonparametric process, the quantum system will be at a different stateafter it interacts with the photons. The physical picture of this interactionis shown in Fig. 9.1. The transition probability for single and/or multiple-photon absorption will be evaluated by the perturbation procedure intro-duced in the previous section. Assuming that the ground state is state | 1 >and the system is in its ground state at t = 0, we have

ρij(t = 0) = δi1δj1. (9.21)

The only nonzero matrix element for ρ is ρ11 = 1 at t = 0. The transitionprobability from the ground state | 1 > to any | k > is by definition

W =∂ρkk

∂t,

where the overbar indicates a time average over a time much longer than theperiod of the perturbation.

Fig. 9.1 Interaction of photons with quantum systems. (a) Single-photon, (b) two-photon and (c) three-photon interaction.

Page 221: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

204 9 Quantum Theory

In the perturbation procedure outlined above,

W =∂

∂t

[Σpρ

(p)kk

]=

∂ρ(0)kk

∂t+

∂ρ(1)kk

∂t+

∂ρ(2)kk

∂t+ · · · .

(9.22)

If the time average for the lowest order ∂ρ(p)kk /∂t is nonzero, one does not

need to evaluate the higher-order quantity. In order to calculate the multiple-photon transition probability, one should begin by calculating the single-photon transition probability first. Assuming that the system is in thermal

equilibrium with the reservoir, then ρ(0)11 , ρ

(0)22 , ρ

(0)33 , . . . , etc., will follow a Boltz-

mann distribution. If the states are electronic states, and E2 − E1 kT, thenwe can assume that ρij(t = 0) = δi1δj1 as in Eq. (9.21). ρ

(0)kk = constant = 0

before the perturbation H′ is turned on. To zeroth order ∂ρ(0)kk /∂t = 0. One

needs to go the next order to calculate ∂ρ(1)kk /∂t :

ih∂ρ

(1)kk

∂t= [H′, ρ(0)]kk = Σj

(H′

kjρ(0)jk − ρ

(0)kj H′

jk

), (9.23)

where

the ρ(0)′jk s are ρ

(0)1k , ρ

(0)2k , ρ

(0)3k , . . . , etc.,

and

the ρ(0)kj s are ρ

(0)k1 , ρ

(0)k2 , ρ

(0)k3 , . . . , etc.

One needs to solve for ρ(0)kj (t) or ρ

(0)jk (t) from

ih

(∂

∂t+ iωjk +

1τjk

(0)jk (t) = 0. (9.24)

Since the zeroth-order ρ(0)jk (t) is the diagonal matrix element of ρ before the

perturbation is turned on, the solution is

ρ(0)jk (t) = ρ

(0)jk (0)e−iωjkt−t/τjk

= δj1δk1 e−iωjkt−t/τjk.(9.25)

Substituting back into Eq. (9.23), we obtain

ih∂ρ

(1)kk

∂t= 0

Page 222: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.4 Transition Probability 205

and hence

∂ρ(1)kk

∂t= 0.

Proceeding to higher order,

ih∂ρ

(2)kk

∂t=

[H′, ρ(1)

]kk

= Σj

(H′

kjρ(1)jk − ρ

(1)kj H′

jk

) (9.26)

and

ih

(∂

∂t+ iωjk +

1τjk

(1)jk (t) = Σi

(H′

jiρ(0)ik − ρ

(0)ji H′

ik

)= H′

j1δk1 − δj1H′1k.

(9.27)

The only nonzero term is ρ(1)ji , which satisfies

ih

(∂

∂t+ iωji +

1τji

(1)ji = H′

ji. (9.28)

In the electric dipole approximation,

H′ = −µ · E, (9.29)

E =E0

2

(eiωt + e−iωt

), (9.30)

H′ = −µ · E0

2

(eiωt + e−iωt

)=

H′

2

(eiωt + e−iωt

), (9.31)

where

H′ = −µ · E0.

Solving for ρ(1)j1 (t) from Eq. (9.27),

ih

(∂

∂t+ iωj1 +

1τj1

(1)j1 (t) =

H′j1

2

(eiωt + e−iωt

), (9.32)

the steady-state solution is of the form

ρ(1)j1 (t) = A+eiωt + A−e−iωt.

Page 223: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

206 9 Quantum Theory

Substituting back into the equation, we obtain

A+ =H′

j1

21

ih(iω + iωj1 + 1τj1

),

A− =H′

j1

21

ih(−iω + iωj1 + 1τj1

),

or

ρ(1)j1 (t) =

H′i1 eiωt

2ih(iω + iωj1 + 1τj1

)+

H′i1 e−iωt

2ih(−iω + iωj1 + 1τj1

). (9.33)

Since ρ(1)1j (t) =

(1)j1 (t)

]∗, we have

ρ(1)1j (t) =

H′∗j1 e−iωt

−2ih(−iω − iωj1 + 1τj1

)+

H′∗j1 eiωt

−2ih(iω − iωj1 + 1τj1

). (9.34)

All other ρ(1)ij (t) terms with i = 1 or j = 1 are zero. We can then proceed to

calculate

ih∂ρ

(2)kk

∂t= Σj

(H′

kjρ(1)jk − ρ

(1)kj H′

jk

)= H′

k1ρ(1)1k − ρ

(1)k1 H′

jk

=H′∗

k12

(eiωt + e−iωt

) [H′∗

k1e−iωt

−2ih(−iω − iωk1 + 1τk1

)+

H′∗k1eiωt

−2ih(iω − iωk1 + 1τk1

)

]

− H′

k1eiωt

2ih(iω + iωk1 + 1τk1

)+

H′k1e−iωt

2ih(−iω + iωk1 + 1

τk1

) H′

1k2

(eiωt + e−iωt

).

(9.35)

When the time average is taken, only DC terms contribute. There are four DCterms:

| H′k1 |2

4ih(iω + iωk1 − 1τk1

)− | H′

k1 |2−4ih(−iω − iωk1 − 1

τk1)

+| H′

k1 |24ih(−iω + iωk1 − 1

τk1)− | H′

k1 |2−4ih(iω − iωk1 − 1

τk1)

.

Page 224: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.4 Transition Probability 207

We notice that the second term is the complex conjugate of the first term andthe fourth is the complex conjugate of the third. Finally, we have

∂ρ(2)kk

∂t=

| H′k1 |2

2h2τk1

[1

(ω − ωk1)2 + ( 1τk1

)2+

1(ω + ωkl)2 + ( 1

τk1)2

]. (9.36)

In the case of single-photon absorption as shown in Fig. 9.1, ω ≈ ωk1 ≈ 1014–1015 Hz, and the quantity 1/τk1 ∼ 108–1012 Hz. It is clear at resonance foroptical frequencies, i.e., ω = ωk1, that the second term can be neglected,

W(1) =∂ρ

(2)kk

∂t=

π | H′k1 |2

2h21π

1τk1

(ω − ωk1)2 + ( 1τk1

)2

=π | H′

k1 |22h2 gL(ω − ωk1),

(9.37)

where W(1) is the single-photon transition probability rate, and

gL(ω − ωk1) =1π

1τk1

(ω − ωk1)2 + ( 1τk1

)2(9.38)

is a Lorentzian line shape function. In the transition probability just derived,we are assuming that there is a single photon interacting with a single quan-tum system which may be an ion or a molecule. In a real system there area large number of identical photons interacting with a large number of iden-tical quantum systems, e.g., molecules. Since all particles in this ensembleare identical (homogeneous system), the transition probability for the wholeensemble will have exactly the same line shape described by the Lorentzianline gL(w − wk1). To such a system, the Lorentzian line is called a homoge-neously broadened line with line width (FWHM) equal to 2/τk1. τk1 = T2is also referred to as the transverse relaxation time. If the molecules are in agaseous phase, they will be moving. Because of the Doppler effect the mole-cules with different velocities with respect to the propagation vector of thephoton will cause a shift in the resonant frequency. Such a system is calledan inhomogeneously broadened system. In such a system, the homogeneousline associated with a specific velocity should be integrated over its velocitydistribution. The resultant line is a Gaussian line shape function. The linewidth is further broadened. This type of broadening is called inhomogeneousbroadening. The Gaussian line function is given by

gG (ω − ωkl) =2

∆WG

(ln 2π

) 12

exp

[−4(ln 2)

(ω − ωk1)2

∆WG

], (9.39)

where ∆wG is the line width of the Gaussian line.

Page 225: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

208 9 Quantum Theory

9.5Two-Photon Absorption Transition Probability

In the single-photon transition case if ω = 12 ωj1 the transition probability rate

W1 will be zero. However, if the energy-resonant denominator has the follow-ing expression:

1(2ω − ωj1)2 + ( 1

τj1)2

,

one will have two-photon absorption as shown in Fig. 9.1b.We follow a similar procedure as outlined in the previous section for the

single-photon case but carry it to higher order (p > 2) until it contains a DCterm containing the factor shown above. Since

ih∂ρ

(p)jj

∂t= [H′, ρ(p−1)]jj for p > 2,

we look for a term that has this factor in

ρ(p−1)ij , ρ

(p−2)ij , . . . or ρ

(p−n)ij .

From the equation

ih

(∂

∂t+ iωij +

1τij

(p−n)ij =

[H′, ρ(p−n−1)

]ij

, (9.40)

we want ρ(p−n)ij to vary as ei2ωt. Recall that

ρ(0)11 = 1, ρ

(0)ij = δi1 δj1, ρ

(1)jj = 0, ρ

(1)ij = 0 except for ρ

(1)i1 ,

ρ(1)i1 =

H′i1 eiωt

2ih(iω + iωi1 + 1τi1

)+

H′i1 e−iωt

2ih(−iω + iωi1 + 1τij

).

(9.41)

Substituting into Eq. (9.40) for p = 2, n = 0 and realizing that H′11 = 0 and

ω1j = −ωj1,

ih

(∂

∂t− iωj1 +

1τj1

(2)1j = −Σ ρ

(1)1 H′

j

= −Σ

[H′∗

1 e−iωt

−2ih(−iω − iω1 + 1τi

)+

H′∗1eiωt

−2ih(iω − iω1 + 1τ1

)

]

×H′

j

2

(eiωt + e−iωt

)+ c.c.

Page 226: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.5 Two-Photon Absorption 209

We only need to keep the terms that go with ei2ωt:

= −Σ

H′∗1H′

j

2ei2ωt

−2ih(iω − iω1 − 1τ1

)+ c.c. + other terms

= −Σ

H′∗1H′

j

4ihei2ωt

iω − iω1 + 1τ1

+ other terms. (9.42)

We are looking for the complex amplitude of ρ(2)1j (t), defined as

ρ(2)1j = ρ

(2)1j ei2ωt,

where ρ(2)1j is the complex amplitude of ρ

(2)1j (t).

Substituting it into (Eq. 9.41), we find that

ρ(2)1j

=1

4h2(

2ω − ωj1 − iτj1

) [Σ

H′1H′

j

ω − ωj

]. (9.43)

In the case of two-photon resonance with 2ω = ωj1, ρ(2)1j (t) will be the only

dominant term. We may now proceed to evaluate

ih∂ρ

(3)jj (t)

∂t=

[H′, ρ(2)

]jj

= H′i1 ρ

(2)1j − ρ

(2)j1 H′

1j. (9.44)

We notice that the terms on the right-hand side of Eq. (9.44) do not contain aDC term; therefore, the time average of

∂ρ(3)jj (t)

∂t= 0. (9.45)

We have to go to ∂ρ(4)jj /∂t and ρ

(3)1j (t), etc.

In

ih∂ρ

(4)jj (t)

∂t= Σ

(H′

jρ(3)j − ρ

(3)j H′

j

)we look for DC terms. This means that we are only interested in finding ρ(3)

that varies as e(±iωt). This indeed is possible by solving

ih

(∂

∂t+ iωij +

1τij

(3)ij (t) =

[H′, ρ(2)

]ij

= H′i1ρ

(2)1j , (9.46)

Page 227: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

210 9 Quantum Theory

since only the |j > state is resonant with 2ω. Again defining a complex ampli-

tude of ρ(3)ij as ρ

(3)ij (t) = ρ

(3)ij eiωt, we find that

ρ(3)ij

=H′

i12h(ω − ωi1)

ρ(2)ij

, (9.47)

where we have dropped the term 1/τij since it is much smaller than the ω +ωij term. We also use the following relation:

1ω + ωij

=1

ω + ωi − ωj − ω1 + ω1=

1ω + (ωi − ω1) − (ωj − ω1)

=1

ω + ωi1 − ωj1=

1ω + ωi1 − 2ω

=−1

ω − ωi1.

The expression of ρ(3)ji can be obtained from

ρ(3)ji (t) =

(3)ij (t)

]∗=

(3)ij

]∗e−iωt

=H′∗

i12h(ω − ωi1)

(2)ij

)∗e−iωt.

Substituting ρ(3) back into

ih∂ρ

(4)jj

∂t= Σ

(H′

jρ(3)j

)− Σk

(3)jk H′

kj

)= Σ

Hj

2

(eiωt + e−iωt

)ρ(3)

jeiωt

− Σkρ(3)jk

e−iωt Hkj

2

(eiωt + e−iωt

),

(9.48)

we need to keep the DC terms

= Σ

H′j

2H′

12h(ω − ω1)

ρ(2)ij − Σk

H′∗k1

2h(ω − ωk1)ρ

(2)j1

H′kj

2.

Page 228: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.6 Scattering Cross Section 211

Taking the time average, we have the two-photon transition probability rate,

W(2) =∂ρ

(4)jj (t)

∂t

=1

16ih4

Σ

H′jH′

1

ω − ω1

1(2ω − ωj1 − i

τj1

) ΣkH′

1kHkj

ω − ωk1

− Σk1(

2ω − ωj1 + iτj1

) Σ

H′∗1H′∗

j

ω − ω1

H′1kH′

kj

ω − ωk1

=

1

16ih4

1(2ω − ωj1 − i

τj1

) − 1(2ω − ωj1 + i

τj1

)∣∣∣∣∣Σ

H′1H′

j

ω − ω1

∣∣∣∣∣2

8h4

∣∣∣∣∣Σ

H′1H′

j

ω − ω1

∣∣∣∣∣2

1/τj1

(2ω − ωj1)2 + ( 1τj1

)2.

(9.49)

One can define the Lorentzian line shape function for two-photon absorptionas

gL(2w) =1π

1/τj1

(2ω − ωj1)2 + ( 1τj1

)2. (9.50)

Since H′ ∝ E, we note that

W(2) ∝| E |4= I2

and

W(1) ∝| E |2= I.

Therefore, the single-photon transition rate is proportional to the intensity ofthe beam while the two-photon transition rate is proportional to the inten-sity squared. One can define a scattering cross section from these transitionprobabilities, which we do in the next section.

9.6Scattering Cross Section

The scattering cross section is defined as the effective cross section of an ab-sorbing atom, i.e.,

σc =areaatom

=power absorbed per atom

power per unit area

Page 229: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

212 9 Quantum Theory

=P/N

I=

Nhω21 × W/NI

=2hω21

ncε0 | E |2 W. (9.51)

For the single-photon cross section,

W(1) =π | H′

j1 |22h2 gL(w − w21),

where | H′ |2 =| µ12 |2

3| E |2,

we have

σc(1) =ω21π

3hncε0| µ12 |2 gL(ω − ω21), (9.52)

W(2) =π

8h4 | ΣH′

1H′1

ω − ω1|2 gL(2ω − ω21),

σc(2) =ω21π

36ncε0h3 | µ12 |4 | E |2 gL(2ω − ω21).(9.53)

Note that σc(1) is independent of I while σc(2) is linearly dependent on I. Atypical σc(1) is about 3 × 10−16 cm2 while σc(2) ∼ 10−35 | E |2 with E in V/cm.For a laser beam of intensity 10 MW/cm2, E ∼ 105 V/cm, σc(2) ∼ 10−25 cm2,much smaller than σc(1). This shows that the two-photon process only be-comes significant at high intensity. In absorption spectroscopy, single photonsonly connect states of opposite parity. On the other hand, the two-photon tran-sition connects states of the same parity. Therefore, two-photon spectroscopycan yield complementary information on atomic states.

9.7Three-Photon Absorption

The transition probability for three-photon absorption can be evaluated in asimilar manner by going to ∂ρ(6)/∂t. We look for DC terms in the expressionof ∂ρ(6)(t)/∂t, which has an energy denominator that shows three-photon res-onance:

gL(3ω) =1π

1τj1

(3ω − ωj1)2 + ( 1τj1

)2. (9.54)

It can be shown that [86]

W(3) =πgL(3ω)

32h6

∣∣∣∣∣Σg.k

H′jgH′

gkH′k1

(2ω − ωq1)(ω − ωk1)

∣∣∣∣∣2

. (9.55)

Page 230: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.8 Doppler-Free Two-Photon Absorption 213

Fig. 9.2 Interaction of moving atom with velocity→v with counterprop-

agating laser beams. (a) In the laboratory frame, (b) in the atomicframe.

Following a similar procedure, one can even calculate the m-photon transitionprobability, which is given in Ref. [86].

9.8Doppler-Free Two-Photon Absorption

In a gaseous medium the single-photon transition line shape function willbe inhomogeneously broadened by the Doppler effect. The line shape willbe a Gaussian, instead of a Lorentzian. For two-photon absorption, even inan inhomogeneous Doppler broadened medium one can obtain Doppler-freetwo-photon absorption. This effect, which was first observed by Levensonand Bloembergen [87] and by Biraben et al. [88], may be understood by con-sidering the two-photon absorption with counterpropagating laser beams asshown in Fig. 9.2. If the atomic system interacts with two photons, one eachfrom the counterpropagating laser beams, it will sense one beam at frequencyw (1 − vz/c) and the other at w (1 + vz/c); the transition line shape functionwill be

gL =1π

1τ[

ω(1 − vz

c

)+ ω

(1 + vz

c

)− ωj1]2 +

(1τ

)2

=1π

1τ(

2ω − ωj1)2 +

(1τ

)2 = gL(2ω).

The Doppler effect is balanced out in the two-photon absorption process. Ev-ery atom in the ensemble participates, regardless of its velocity. The two-photon absorption transition coming from the atoms absorbing two photons

Page 231: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

214 9 Quantum Theory

from the same beam, on the other hand, will experience a Doppler shift, giv-ing rise to Doppler broadening. In this case, only atoms with the same velocitycontribute to the absorption at this specific frequency. Since the Doppler-freetwo-photon absorption involves all atoms in the medium, it will be the domi-nant absorption, giving rise to the Doppler-free absorption line shape with thetransition probability given as

W(2) =1

16h4

∣∣∣∣∣Σn(µjn · E+)(µn1 · E−) + (µjn · E−)(µn1 · E+)

ωn1 − ω

∣∣∣∣∣2

× 1π

1τ(

2ω − ωj1)2 +

(1τ

)2

+

∣∣∣∣∣Σn(µjn · E+)(µn1 · E+)

ωn1 − ω

∣∣∣∣∣2

+

∣∣∣∣∣Σn(µjn · E−)(µn1 · E−)

ωn1 − ω

∣∣∣∣∣2

× exp

[−(2ω − ωj1)2/Ω2

0

2√

πΩ0

].

(9.56)

The composite line shape is sketched in Fig. 9.3. Using this technique, theZeeman effect for the two-photon 3S–5S transition in a Na23 atom has beenobserved [87].

Fig. 9.3 The line-shape function of the Doppler-free two-photon ab-sorption of an ensemble of moving atoms. The Doppler-free line issitting on top of a Doppler pedestal.

Page 232: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.9 Calculation of Susceptibility 215

9.9Calculation of Linear and Nonlinear Susceptibility

Linear susceptibility is the linear response function to the external appliedfield. An induced polarization results,

P(ω) = χ(1)(ω)E(ω) = N < µ >= N Tr(ρµ). (9.57)

In an anisotropic medium, χ(1) becomes a tensor; the relation between P, χ

and E takes the tensorial form

Pi(ω) = χijEj(ω). (9.58)

For simplicity, without the loss of other physical principles, we will limit our-selves to the nontensorial relation. The relation between P and E can be ex-tended to the nonlinear region, where one can express

P = P(0) + P(1) + P(2) + · · · ,

P(n) ∝ En.

We will extend the calculation to the optical second-harmonic nonlinear po-larization,

P(2)x = N < µx(2ω) >= χ

(2)xxx E2 e−i2ωt

= −Ne Σn,n′xn′nρ(2)nn′(2ω).

(9.59)

We proceed to calculate χ(1) first. Using

ih(

∂t+ iωn′n +

1τn′n

(1)n′n = Σk

[H′

n′kρ(0)kn − ρ

(0)n′kH′

kn

],

H′nk = −µE = −µE0 e−iωt,

The e−iωt component of ρ(1)n′n(ω) satisfies

ih(−iω + iωn′n +

1τn′n

(1)n′n(ω) = −E0 Σk

[µn′kρ

(0)kn − ρ

(0)n′kµkn

]= −E0

(µn′nρ

(0)nn − µn′nρ

(0)n′n

); with µn′n = −exn′n, (9.60)

ρ(1)n′n(ω) =

eE0

hxn′ne−iωt

ω − ωn′n + iτn′n

(0)nn − ρ

(0)n′n′

). (9.61)

We have

P(1) = χ(1)(ω)E0 e−iωt = N Tr(ρµ)

= −NΣnn′ ρ(1)nn′ exn′n

= −NeE0

hΣnn′

exn′n xn′nω − ωn′n + iτn′n

(0)nn − ρ

(0)n′n′

)e−iωt,

(9.62)

Page 233: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

216 9 Quantum Theory

χ(1)(ω) =Ne2

hΣnn′

xn′n xn′nω − ωn′n + iτn′n

(0)n′n′ − ρ

(0)nn

). (9.63)

To calculate the nonlinear susceptibility for second-harmonic generation, onefollows a similar procedure, looking for ρ(2) in PNL(2ω) which goes with the2ω frequency,

PNLx (2ω) = χ

(2)xx ExEx e−i2ωt = −Ne Σn,n′ xn′n ρ

(2)nn′(2ω). (9.64)

From

ih∂

∂tρ

(p)jj =

[H′, ρ(p−1)

]jj

+ ih Σk

(p)kk Wkj − ρ

(p)jj Wjk

),

and the off-diagonal element,

ih

(∂

∂t+ iωij +

1τij

(p)ij =

[H′, ρ(p−1)

]ij

, i = j,

we have ρ(0)nj = 0 for n = j, ρ

(1)jj (ω) = 0 and ρ

(1)ij (ω) = ρ

(1)ij e−iωt, where the

amplitude function ρ(1)ij (ω) satisfies

ih

(−iω + iωij +

1τij

(1)ij (ω) = Σn

(H′

in(ω)ρ(0)nj − ρ

(0)in H′

nj

)= H′

ij(ω) ρ(0)jj − ρ

(0)ii H′

ij = H′ij

(0)jj − ρ

(0)ii

),

ρ(1)ij (ω) =

H′ij

ih(−iω + iωij + 1

τij

) (ρ

(0)jj − ρ

(0)ii

). (9.65)

Where H′ij = H′

ij e−iωt = −exij E0 e−iωt,

ρ(1)ij (ω) =

−exijE0e−iωt

h(ω − ωij + i/τij)

(0)jj − ρ

(0)ii

).

Substituting this into the equation for ρ(2), we have

ih(−i2ω + iωij +

1τij

(2)ij (2ω) = Σn

[H′

in(ω)ρ(1)nj (ω)− ρ

(1)in (ω)H′

nj(ω)]

= Σne2xin xnj E2

0 e−i2ωt

h(ω − ωnj + i/τnj)

(0)jj − (ρ

(0)nn

)+ Σn

e2xin xnj E20 e−i2ωt

h(ω − ωin + i/τin)

(0)ii − (ρ

(0)nn

). (9.66)

Page 234: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.9 Calculation of Susceptibility 217

Solving for ρ(2)ij (2ω) and substituting into

PNLx (2ω) = χ

(2)xxx(2ω)ExEx e−i2ωt

= −NeΣn,n′ xn′nρ(2)nn′(2ω) = −NeΣi,jxjiρ

(2)ij

= −NeΣi,jxjiΣn

e2xin xnj E20 e−i2ωt

(0)jj − ρ

(0)nn

)h2

(2ω − ωij + i

τij

) (ω − ωnj + i

τnj

)−NeΣi,jxjiΣn

e2xinxnjE20e−i2ωt

h2(

2ω − ωij + iτij

) (ω − ωin + i

τin

) (ρ

(0)ii − ρ

(0)nn

), (9.67)

we finally have

χ(2)xxx(2ω) = NΣi,j,n

e3xin xnj xji

(0)nn − ρ

(0)jj

)h2

(2ω − ωij + i

τij

) (ω − ωnj + i

τij

)+

e3xin xnj xji

(0)nn − ρ

(0)ii

)h2

(2ω − ωij + i

τij

) (ω − ωin + i

τin

) .

(9.68)

If there is no resonance, one can drop the i/τij term, which is much smallerthan the (2ω − ωij) or (2ω − ωnj) terms:

χ(2)xxx(2ω) = NΣiΣjΣn

e3xin xnj xji

h2

(0)nn

2ω − ωij

[1

ω − ωnj+

1ω − ωin

]

−ρ

(0)jj

(2ω − ωij)(ω − ωnj)− ρ

(0)ii

(2ω − ωij)(ω − ωin)

).

Since

1ω − ωnj

+1

ω − ωin=

ω − ωi + ωn + ω − ωn + ωj

(ω − ωnj)(ω − ωin)

=2ω − ωij

(ω − ωnj)(ω − ωin),

we have

χ(2)xxx(2ω) = NΣiΣjΣn

e3xin xnj xji

h2

(0)nn

(2ω − ωnj)(ω − ωin)

−ρ

(0)jj

(2ω − ωji)(ω − ωnj)− ρ

(0)ii

(2ω + ωji)(ω + ωni)

].

Page 235: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

218 9 Quantum Theory

Exchanging the n, j indices in the second term and the i, n indices in the thirdsummation, we finally have

χ(2)xxx(2ω) = NΣiΣjΣn

e3xin xnj xji

h2 ρ(0)nn

[1

(ω + ωjn)(ω − ωin)

− 1(2ω − ωin)(ω − ωjn)

− 1(2ω + ωjn)(ω + ωin)

].

(9.69)

In most cases, only the ground state is populated, ρ(0)nn = 0 for n = g, ρ

(0)gg = 1.

Then

χ(2)xxx(2ω) = NΣiΣj

e3xgj xji xig ρ(0)gg

h2

[1

(ω + ωjg)(ω − ωig)

− 1(2ω − ωig)(ω − ωjg)

− 1(2ω + ωjg)(ω + ωig)

],

(9.70)

where the i and j states are usually not resonant with either ω or 2ω. Thesecond-harmonic generation process due to the interaction of material withphotons is presented in Fig. 9.4.

Fig. 9.4 Quantum mechanical representation of optical second-harmonic generation process.

9.10Third-Order Nonlinear Susceptibility χ(3)

Following the procedure outlined in this chapter, one can calculate the third-order nonlinear susceptibility ρ(3). Bloembergen et al. [89] have worked outthe detailed form of all 48 terms which occur in the general case. Here wewill outline the procedure of the calculation using the density operator to the

Page 236: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.10 Third-Order Nonlinear Susceptibility 219

third-order term with the recursion formula. We start by first integrating the

equation for ρ(p)ij (t),

ρ(p)ij (t) =

∫ t

−∞

−ih

[H′, ρ(p−1)

]ij

e(iωij+γij)(t′−t)dt′, (9.71)

where γij = 1/τij is the damping constant:

ρ(3)nm(t) =

∫ t

−∞

−ih

[H′, (t′), ρ(2)

]nm

e(iωnm+γnm)(t′−t)dt′. (9.72)

We need to first calculate ρ(2)(t) and ρ(1)(t):

ρ(1)nm(t) =

−ih

e−(iωnm+γnm)t∫ t

−∞[H′(t′), ρ(0)] e(iωnm+γnm(t′−t)dt′, (9.73)

ρ(0)nm = 0 for n = m. (9.74)

Therefore,[H′(t′), ρ(0)

]nm

= Σν H′(t′)nν ρ(0)νm − ρ

(0)nν H′(t′)νm

= Σν [µnνρ(0)νm − ρ

(0)nν µνm] · E(t′)

= −(

ρ(0)mm − ρ

(0)nn

)µmm · E(t′),

(9.75)

so we can rewrite this by using E(t′) = ΣjE(ωj) e−iωjt′ ,

ρ(1)nn =

−ih

µnm · Σj E(ωj)(

ρ(0)mm − ρ

(0)nn

)e−(iωnm+γnm)t

×∫ t

−∞e[i(ωnm+ωj)+γnm]t′dt′

(0)mm − ρ0

nnh ∑

j

µnm · E(ωj)e−iωjt

(ωnm − ωj) − iγnm.

This expression for ρ(1)nm is used to calculate[

H′(t′), ρ(1)]

nm= −Σν

(µnνρ

(1)νm − ρ

(1)nν µνm

)· E(t′) (9.76)

and

ρ(2)nm =

−ih

e−(iωnm+γnm)t∫ t

−∞[H′(t′), ρ(1)]nm e(iωnm+γnm)t′dt′. (9.77)

Page 237: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

220 9 Quantum Theory

A similar procedure is used to calculate the higher-order terms. We have

ρ(2)nm(t) =

−i

h2 Σνjq

×(

ρ(0)mm − ρ

(0)νν

) [µnν · E(ωq)][µνm · E(ωj)][ωnm − ωj − ωq − iγnm][(ωνm − ωj) − iγνm]

−(

ρ(0)νν − ρ

(0)nn

) [µnν · E(ωj)][µνm · E(ωq)][ωnm − ωj − ωq − iγnm][(ωνm − ωj) − iγnν]

× e−i(ωj+ωq)t (9.78)

and

(ρ(3)nm(t) =

−i

h3 Σν Σjgγ(µnν · E(ωγ)) Ωnν − (µnν · E(ωγ)) Ωνm

ωnm − ωj − ωq − ωγ − iγnr

× e−i(ωj+ωg+ωr)t,

(9.79)

where

Ωνm =(

ρ(0)mm − ρ

(0)

) [µν · E(ωq)][µm · E(ωj)][(ωνm − ωj − ωq) − iγνm][ωm − ωj − iγm]

−(

ρ(0) − ρ

(0)νν

) [µν · E(ωj)][µm · E(ωq)][(ωnν − ωj − ωq)− iγnν][ωn − ωj − iγn]

.

(9.80)

We can now use this result for ρ(3) to calculate the third-order nonlinear sus-ceptibility by combining the following relationships:

P(ωj + ωq + ωr) = N < µ(ωj + ωq + ωr) >

= N Tr(ρµ)

= N Σnm ρ(3)nm µmn

and

Pk(ωj + ωq + ωr) = Σxyz Σjgr χ(3)kzyx(ωj + ωq + ωr, ωr, ωq, ωj)

× Ey(ωr)Ex(ωq)Ez(ωj).(9.81)

Combining Eqs. (9.79), (9.80) and (9.81) and performing the summation overthe input frequencies, together with the Cartesian coordinates, we finally ob-

Page 238: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

9.10 Third-Order Nonlinear Susceptibility 221

tain

χ(3)kijh(ωj + ωq + ωr, ωj, ωq, ωr) = Σνnm

N

h(3)ρ(0)

×

µkνµ

jνnµi

nmµhm

[(ων − ωj − ωq − ωr)− iγν][(ωn − ωj − ωq)− iγn][(ωm − ωj) − γm]

+µk

vµiνnµ

jnmµh

m

[(ων − ωj − ωr − ωq)− iγν][(ωn − ωj − ωr) − iγn][ωm − ωj − iγm]

+µk

νµhνnµ

jnmµi

m

[(ων − ωq − ωr − ωj)− iγν][(ωn − ωq − ωr) − iγn][(ωn − ωq)− iγn]

+µk

νµjνnµh

nmµim

[(ων − ωq − ωj − ωr)− iγν][(ωn − ωq − ωr) − iγn][(ωn − ωq)− iγn]

+µk

νµiνnµh

nmµjm

[(ων − ωr − ωj − ωq)− iγν][(ωn − ωr − ωj) − iγn][(ωn − ωr)− iγn]

+µk

νµhνnµi

nmµjm

[(ων − ωj − ωr − ωq)− iγν][(ωn − ωj − ωr) − iγn][(ωn − ωj) − iγn]+ · · ·

,

(9.82)

plus another 42 terms of similar type with different permutations of the an-gular frequencies ωj, ωq and ωr, along with the permutations of the Cartesiancomponents of the dipole matrix elements. Notice that each frequency cantake on a “+” or a “–” sign. The complete 48 terms have been derived byBloembergen et al. [89]. This general expression of χ(3) contains all possi-ble third-order effects, including four-wave mixing. χ(3) effects take place inall media while the χ(2) effect vanishes in media with central symmetry. Ifthere is a resonance with a particular energy level, the term representing thisresonance effect will dominate. For example, in coherent anti-Stokes Ramanscattering (CARS) one vibrational state | t > is in resonance with the differ-ent frequencies of the laser photon, ωa, and the Stokes photon, ωe, such thatωtg = ωa − ωe. The anti-Stokes photon is scattered from the coherently ex-cited phonon ωA = ωa − ωc + ωa = 2ωa − ωe. This situation is shown inFig. 9.5.

In this situation, χ(3) can be grouped into two terms, one resonantly en-hanced and the remaining one a “nonresonant” background term,

χ(3) = χ(3)res + χ

(3)NR.

Page 239: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

222 9 Quantum Theory

Fig. 9.5 Quantum mechanical representation of coherent anti-StokesRaman scattering (CARS).

Problems

9.1 In the text we see examples of calculating the single-photon transitionprobability W(1) and the linear susceptibility χ(1). Show that the expressionsfor W(1) and χ(1) are consistent with each other by considering the power ab-sorbed per unit volume.

9.2 Derive the expression of the three-photon transition probability W(3).

9.3 Derive Eq. (9.82). You may consult the paper by N. Bloembergen,H. Lotem and R. T. Lynch, Jr, Lineshapes in coherent resonant Raman scatter-ing, Indian J. Pure Appl. Phys., 16:151, 1978 for the complete 48 terms.

Page 240: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

223

10Applications of Nonlinear Optical Effects

In the last three chapters a survey of nonlinear phenomena and a simple the-ory of nonlinear optical effects were given. Nonlinear wave equations were il-lustrated mostly using only harmonic generation as examples. There are manynonlinear effects such as intensity-dependent refractive-index change, stimu-lated Raman and Brillouin scattering, soliton formation and propagation inoptical fibers that are not addressed. These phenomena have been describedextensively in the standard text books of nonlinear optics [67, 84–86, 90, 91].Readers are referred to these references for further studies. In this chapter weshall present several nonlinear phenomena that are not commonly found instandard nonlinear optics books.

10.1Optical Harmonic Generation at Oblique Incidence

Optical harmonic generation is commonly used for generating coherent lightat twice the fundamental wave frequency. In this application, one needs toconvert as much of the fundamental light waves into harmonics as possible.The conventional practice is to purchase a nonlinear optical crystal alreadyoriented and cut for maximum frequency-conversion efficiency in the phase-matching configuration. The second-harmonic generation (SHG) is done intransmission with a normally incident laser beam. From an academic point ofview there are more interesting phenomena associated with the optical har-monic generation with obliquely incident laser light. One can extend the lawof optics to the nonlinear region. In the transmission experiment at normalincidence, the boundary conditions at an interface are straightforward. How-ever, at oblique incidence they become interesting.

Extension of the fundamental optical phenomena from the linear to the non-linear regime will not be complete, however, if the basic nonlinear interactionat the boundary between a linear and a nonlinear medium is not fully inves-tigated [92]. This is the area that has not been fully exploited. Only a limitedamount of work has been reported, in particular, when the fundamental beamis at total internal reflection [93–96]. In order to cover nonlinear counterparts

Page 241: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

224 10 Applications

of total reflection, we shall assume that a light wave is incident onto a nonlin-ear medium from an optically denser linear medium. This situation was inves-tigated in Refs. [93–96], where potassium dihydrogen phosphate (KDP) wasimmersed in 1-bromonaphthalene. In addition to SHG near total reflection,there have been a number of device applications employing nonlinear opticalinteractions or the electrooptic effect with total internal reflection [95,97]. Theelectrooptic effect with total internal reflection can be regarded as a nonlinearinteraction between the low-frequency microwave and the high-frequency op-tical wave at the boundary of the nonlinear medium. This theoretical opticalmodel explains critical angles well. The theoretical formalism for the elec-trooptic effect is essentially the same as SHG described in this book. Thus, itis important to have a good understanding of the nonlinear phenomena tak-ing place at the boundary of the nonlinear medium, in particular, when thefundamental beam is obliquely incident. Recently, obliquely incident beamsproducing nonlinear optical phenomena have also gained attention becausethey can be used as a tool to probe solid–solid interfaces [85, 98]. This is dueto the fact that traditional optical spectroscopies lack interface specificity andmost surface diagnostic techniques are inconvenient for studying buried in-terfaces. Although the work described here is for a liquid–solid interface, itcan easily be extended to study buried solid–solid interfaces.

The reflection of laser light at the interface between linear and nonlinear me-dia at an oblique angle of incidence, θi, near the critical angle, θcr, can provideinteresting nonlinear optical effects as predicted by the Bloembergen–Pershan(BP) theory [92]. In the experimental work of Ref. [93] the investigation ofsecond-harmonic generation in the vicinity of the critical angle, θcr(ω), hasbeen performed at total reflection. This leads to the verification of nonlin-ear optical effects [93, 96, 99]. Furthermore, using the same crystallographiccut of KDP [96], the work was extended theoretically to a nonlinear Brewsterangle in ammonium dihydrogen phosphate (ADP) [100].

10.1.1Theory of Light Wave at Boundary of Nonlinear Media

The general oblique incidence geometrical situation is shown in Fig. 10.1. Forthe situation just before total reflection occurs, the fundamental beam is trans-mitted almost parallel to the surface in the nonlinear KDP crystal. Accordingto the BP theory [92], there are reflected harmonic beams and two transmit-ted harmonic beams. The driven polarization wave propagates in the samedirection as the transmitted laser beam. It has a wave vectorks = 2kL(ω) andrepresents the particular solution of the inhomogeneous solution. In addition,there is the homogeneous solution with wave vector kT(2ω). In the nonlinearKDP crystal, the two transmitted harmonic beams are spatially distinct and

Page 242: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 225

readily observed separately. The relative magnitude of θS and θT in Fig. 10.1depends on the magnitude of the ordinary ray index of refraction n0(ω) andthe extraordinary ray index of refraction ne(2ω), respectively. If at a particu-lar crystallographic orientation of the KDP crystal nω

0 > n2ωe , then θT > θS,

and vice versa. As the angle of incidence, θi, in Fig. 10.1 is increased, it isintuitively clear that the beam with wave vector kS(2ω) will disappear at thesame time as the transmitted fundamental. The transmitted harmonic beamwith wave vector kT(2ω) will persist.

Fig. 10.1 K vectors of fundamental and second-harmonic waves in thevicinity of total reflection.

As θi becomes increasingly larger than the critical angle, this second-harmonic wave will eventually disappear and only the reflected harmonicwave remains. According to Fig. 10.1, the angles θR, θS and θT of the reflected,inhomogeneous and homogeneous transmitted waves are given, respectively,by

nliq(ω) sin θi = nliq(2ω) sin θR = n(ω) sin θS = n(2ω) sin θT . (10.1)

The refractive indices without subscripts refer to the KDP crystal. The com-ponents of the nonlinear source polarization PNLS along the cubic axes of thenonlinear KDP crystal are given in terms of the fundamental field componentsat each point inside the crystal by

PNLSz (2ω) = χNL

36 ETx (ω)ET

y (ω). (10.2)

PNLSx (2ω), PNLS

y (2ω) can be obtained by cyclic permutation of Eq. (10.2).

Page 243: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

226 10 Applications

If the incident fundamental field is polarized perpendicular to the plane ofincidence and along the [110] direction with respect to the crystallographicaxes of KDP, PNLS will be along the [001] direction and lies in the plane ofincidence. Equation (10.2) can be expressed in terms of the amplitude E0 ofthe fundamental wave by

PNLSz (2ω) = χNL

36 η(FLT E0)2, (10.3)

where η is a geometrical factor that depends on the orientation of the funda-mental field vector and nonlinear polarization component with respect to thecrystallographic axes of KDP. The linear Fresnel factor FL

T describes the changeof the amplitude of the fundamental wave on transmission at the crystal sur-face. If the linear polarization is perpendicular to the plane of incidence, it isgiven by

FLT =

2 cos θi

cos θi + sin θcr(ω) cos θS, (10.4)

where the critical angle for total reflection θcr(ω) is defined by

sin θcr(ω) = n(ω)/nlig(ω) < 1.

The nonlinear polarization is the source of the three harmonic waves. Theelectric field amplitudes of the reflected and transmitted harmonic waves aregiven by

ER(2ω) = 4πPNLSFNLR , (10.5a)

ES(2ω) = 4πPNLSFNLS , (10.5b)

ET(2ω) = 4πPNLSFNLT . (10.5c)

The nonlinear Fresnel factors FNLR , FNL

S and FNLT have been calculated by the

BP theory. For the case of second-harmonic polarization with PNLS(2ω) par-allel to the plane of incidence, the nonlinear Fresnel factors, according to theBP theory, becomes

FNLR,|| =

sin θS sin2 θT sin(a + θS + θT)εR(2ω) sin θR sin(θT + θR) cos(θT − θR) sin(θS + θT)

, (10.6)

FNLS,|| =

sin α

εs − εT, (10.7)

FNLT,|| =

−ε1/2S sin α

ε1/2T (εS − εT)

+ε1/2

R

ε1/2T

FNLR,|| , (10.8)

where ε1/2R (2ω) = nliq(2ω), ε1/2

S = nKDP(ω) and ε1/2T = nKDP(2ω). The

angle α is the angle between the nonlinear polarization PNLS(2ω) in the plane

Page 244: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 227

of incidence and the direction of the source vector KS. It is emphasized thatthese expressions remain valid in the case of total reflection.

The time-averaged second-harmonic power carried by the harmonic beamis given by the real part of the Poynting vector times the cross-sectional areaA of the respective beams:

IR,S,T(2ω) = (c/8π)ε1/2R,S,T(2ω) | ER,S,T(2ω) |2 AR,S,T. (10.9)

IR,S,T(2ω) in Eq. (10.9) is the intensity integrated over the beam cross section,or power. This is the experimentally observed quantity. In the remainderof this chapter intensity and power will be used interchangeably in a similarmanner as in Refs. [93,96,99]. In Eq. (10.9), AR,S,T is the cross-sectional area ofthe beam and is given by

AR,S,T = dd′ cos θR,S,T/ cos θi, (10.10)

where dd′ is the rectangular slit which defines the size of the incident laserbeam. By substitution of the relevant expression into Eq. (10.9), we finallyobtain

IR(2ω) = (c/8π)ε1/2R | Eo |4 dd′(4πχNL

36 )2η2 | FL |4| FNLR,|| |2

× cos θR(cos θi)−1,(10.11)

IS(2ω) = (c/8π)ε1/2S | Eo |4 dd′(4πχNL

36 )2η2 | FL |4| FNLS,|| |2

× cos θS(cos θi)−1,(10.12)

IR(2ω) = (c/8π)ε1/2T | Eo |4 dd′(4πχNL

36 )2η2 | FL |4| FNLT,|| |2

× cos θT(cos θi)−1,(10.13)

for the intensities of reflected, transmitted source and transmitted homoge-neous harmonic waves, respectively. The expression of Eqs. (10.12) and (10.13)will facilitate us in obtaining the theoretical curves of the intensity of the har-monic waves and their comparison to experimental data.

When the two transmitted second-harmonic beams are not spatially re-

solved, the sum of the homogeneous and inhomogeneous intensities I(2ω)total =

IS(2ω) + IT(2ω) is observed. The total transmitted second-harmonic inten-sity I(2ω)

total is equal to the average intensity of the interference pattern of thetwo transmitted beams, which is observed in the more common geometry thatthe nonlinear crystal is a plane-parallel platelet and that the light beams arenearly at normal incidence. However, for a wedge-shaped sample, as in thecase discussed here, a spatial average is taken over the interference pattern inthe direction normal to the surface of entry, where the two transmitted beamsoverlap. Therefore, the total transmitted harmonic intensity is equal to thesum of the intensities of the separated harmonic beams.

Page 245: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

228 10 Applications

The uniaxial KDP crystal is employed for transmitted SHG at phase-matching conditions by birefringence. The phase-matching angle, θm, isthe result of making the birefringence, (n2ω

0 − n2ωe ), equal to the dispersion,

(n2ω0 − nω

0 ), at the phase-matching angle, θm, which as a consequence will given2ω

e = nω0 for this condition. Furthermore, the determination of n2ω

e (θ) for aspecific value of θ can be obtained from the equation of the index ellipsoid,given by

1[n2ω

e (θ)]2=

cos2 θ

[n2ω0 ]2

+sin2 θ

[n2ωe (π/2)]2

. (10.14)

10.1.2Criteria of Null Transmitted SHG

According to the BP theory for the case of PNLS(2ω) lying in the plane of in-cidence, the nonlinear Fresnel factors in reflection and transmission are givenby Eqs. (10.6), (10.7) and (10.8), respectively. As a consequence, the intensi-ties of the reflected, transmitted inhomogeneous and homogeneous harmonicbeams are given in Eqs. (10.11), (10.12) and (10.13), respectively. Analysis ofEqs. (10.6), (10.7) and (10.8) derived from the BP theory leads us to the con-ditions of null transmitted second-harmonic intensities. These conditions canbe summarized as follows.

1. For the case of the angle of incidence, θi, in the neighborhood of the crit-

ical angles θ(2ω)cr and θ

(ω)cr , one can expect that the transmitted homoge-

neous harmonic intensity, IT(2ω), becomes zero first when θi > θ(2ω)cr

(64.76). After further increasing the angle of incidence to equal the

value of θ(ω)cr (66.78), the transmitted inhomogeneous harmonic inten-

sity becomes zero as well. This situation has been confirmed by experi-mental results as indicated in Fig. 10.2 [101]. A more detailed descriptionof the experimental arrangement for the observation of SHG in obliqueincidence will be presented in the next section.

2. Furthermore, the condition of IS(2ω) = 0 and IT(2ω) = 0 can occuraside from the total reflection case. The situation can be achieved whenα = 0, θS = 0 and θT = 0. Under this condition the fundamentalbeam has the angle of normal incidence (θi = 0) and PNLS(2w) mustbe along the face normal of the crystal, which is the optical axis in the Zdirection of [001] as indicated in the inset of Fig. 10.3.

When a = 0, θS = 0 and θT = 0, the nonlinear Fresnel factors FNLR,|| , FNL

S,||and FNL

T,|| in Eqs. (10.6)–(10.8) become zero and as a consequence IR(2ω),IS(2ω) and IT(2ω) also become zero. It is interesting to notice that under

Page 246: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 229

Fig. 10.2 The intensities of inhomogeneous and homogeneous trans-mitted second-harmonic waves in the neighborhood of the critical an-gles θcr(2ω) and θcr(ω).

Page 247: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

230 10 Applications

Fig. 10.3 The total transmitted second-harmonic intensity Itotal(2ω) =Is(2ω) + IT(2ω) as a function of the angle of incidence θi. Is(2ω) =IT(2ω) = 0 at θi = 0 and the phase-matching angle θm

i = 37.27 .

Page 248: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 231

this condition not only do the inhomogeneous and homogeneous harmonicintensities become zero, but also the reflected harmonic intensities. There-fore, no second-harmonic generation occurs at normal incidence for the crys-tal that has a particular crystallographic orientation as indicated in Fig. 10.3.The physical interpretation of the null transmitted second-harmonic inhomo-geneous, homogeneous and reflected intensities is that the nonlinear polariza-tion, PNLS(2ω), cannot radiate inside the medium in the direction which oth-erwise would yield transmitted inhomogeneous homogeneous, and reflectedharmonic rays. This situation reflects an uncommon phenomenon in whichone cannot obtain transmitted second-harmonic light with transmission froma nonlinear optical medium as compared to the simple normal incidence case.It has been demonstrated for the null of IS(2ω) and IT(2ω) and they are ingood agreement with the BP theory.

10.1.3Experimental Observations of the Second-Harmonic Generation at Oblique In-cidence

The experimental arrangement to verify the points mentioned in the previoussection is shown in Fig. 10.4. The KDP crystal with various cuts for differentexperimental purposes was immersed in an optically dense fluid, 1-bromo-napthalene, which has larger indices than the KDP crystal at both frequenciesω and 2ω. The wavelength for the fundamental wave is 1.05 µm. The in-dices for KDP are nω

0 = 1.4943, n2ω0 = 1.5130 and n2ω

e = 1.4707; for 1-bromo-napthalene, nliq(ω) = 1.6262 and n2ω

liq = 1.6701. The fluid is transparent in thewavelength range of 0.4–1.6 µm. The critical angles are θω

cr(ω) = 66.78 andθ2ω

cr (2ω) = 64.76, respectively.

A. Transmitted Homogeneous and Inhomogeneous SHG in the Neighborhoodof Critical Angles

According to the BP theory, there are two transmitted harmonic beams. Thedriven polarization wave propagates in the same direction as the transmit-ted laser beam. It has a wave vector KS = 2 Klaser(ω). Furthermore, thereis a homogeneous harmonic wave with a wave vector KT(2ω). In the ex-periments, the KDP crystal has right-angled corners and the two transmittedharmonic beams are spatially distinct and readily observed separately. Ac-cording to the KDP crystallographic orientation as indicated in Fig. 10.2 andalso the investigation of transmitted SHG in the neighborhood of critical an-gles, the KDP crystal employed in the experiment under this situation has thevalue nω

0 > n2ωe (θ). Therefore, in the neighborhood of the critical angle, for a

given angle of incidence, θi, there exists θS and θT which can be determined byEq. (10.1), such that θT > θS. Therefore, from the experiment it is found that

Page 249: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

232 10 Applications

Fig. 10.4 Experimental arrangement for the oblique incidence SHGexperiment. To detect SHG in transmission, the detector arm needs toswing in the direction of transmission.

θcr(ω) > θcr(2ω). It is clear that the beam with wave vector KT(2ω) will dis-appear at θi = θcr(2ω) and the ray with wave vector KS will disappear at thesame time as the fundamental beam. As the angle of incidence, θi, becomeslarger than θcr(ω), there will be no transmitted second-harmonic beam.

The inhomogeneous and homogeneous harmonic intensities are given byEqs. (10.11) and (10.12), respectively. The solid curves drawn in Fig. 10.2 aretheoretical curves predicted by the BP theory and are calculated from the lastfour factors, respectively,

| FL |4| FNLS,T |2 cos θS,T(cos θi)−1.

The experimental dotted points are in excellent agreement with the theoret-ical prediction that the homogeneous and inhomogeneous harmonic intensi-ties will be terminated at θcr(2ω) and θcr(ω), respectively. The reason that thehomogeneous and inhomogeneous harmonic intensities vanish at θcr(2ω) andθcr(ω) is that when θi is greater than θcr(2ω) and θcr(ω) the values of cos θTand cos θS become purely imaginary. As a consequence, the two harmonic in-tensities will become imaginary, which is not physically allowed. It is worth-while to notice that, from Fig. 10.2, there is an interval θcr(2ω) < θi < θcr(ω)that is about 2.02. In this region, there exists only an inhomogeneous har-monic intensity that has a direct association with the nonlinear polarization,PNL(2ω), and with the nonlinear susceptibility, χNLS. The knowledge of inho-mogeneous second-harmonic generation in this particular region will directlyfacilitate the study of PNLS(2ω) and χNL of a nonlinear medium.

Page 250: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 233

B. Second-Harmonic Generation in Reflection Near the Critical Angle.

The orientation of KDP is the same as the one shown for the transmissionexperiment. The polarization of the fundamental beam is along the [110]direction with respect to the crystallographic axes of the crystal. The re-flected second-harmonic intensity generated from the KDP crystal immersedin 1-bromonapthalene was observed as a function of the angle of incidence, θi.The theoretical curve was calculated from the expression

| FLT |4 FNL

R,11 |2 cos θR(cos θi)−1

of Eq. (10.11). The result of the experiment is shown in Fig. 10.5. The solidcurve represents the result of the theoretical calculation. The figure showsexcellent agreement between the experimental data and the result predictedby the BP theory. There are two cusps at θcr(ω) and θcr(2ω). The theoreticalcurve reflects the influence of the linear and nonlinear Fresnel factors given byEqs. (10.4) and (10.6) when θi approaches θcr(ω) and θcr(2ω). Nonanalyticalsingularities occur at θi = θcr(ω) and θi = θcr(2ω), as predicted by the BPtheory. The reason for the appearance of two cusps at critical angles θcr(ω) andθcr(2ω) is that at these critical angles the values of cos θS and cos θT changefrom real to imaginary. This experimental feature was not clearly observedin earlier experiments [93] of this sort. We attribute the success of the work inRef. [96] to the fact that in the experiment, a mode-locked laser beam was usedand, therefore, multimode effects in the nonlinear process were minimized.

The enhancement of the reflected second-harmonic intensity in the vicin-ity of the critical angles θcr(ω) and θcr(2ω) arises mainly from two sources.First, the linear Fresnel factor FL near the critical angle θcr(ω) is larger thanit is away from this angle, by a factor of approximately two. This gives anincrease by a factor of 16 in the reflected second-harmonic intensity IR(2ω) inEq. (10.9). Second, the nonlinear Fresnel factor FNL

R,11, defined in Eq. (10.6), isdominated in the vicinity of θcr(ω) and θcr(2ω) by the term [sin (θS + θT)]−1,and it is larger than that away from these points by a factor of three. There-fore, after all factors are accounted for, the additional enhancement of IR(2ω)in the vicinity of the critical angle will be about two orders of magnitude, asindicated in Fig. 10.5.

It is interesting to consider the limiting case when θcr(ω) and θcr(2ω) coa-lesce into a single value [96]. Under this condition, the wave vectors KS and KTwill propagate along the same direction, i.e., along the crystal surface, becauseof total internal reflection. Since the crystal was cut with the phase-matchingdirection along the crystal surface, the two critical angles θcr(ω) and θcr(2ω)merged into one and the additional enhancement of IR(2ω) occurs as a resultof phase matching, as indicated in Fig. 10.6. One can consider the curve shownin Fig. 10.6 as a limiting case of that of Fig. 10.5. When the phase-matching

Page 251: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

234 10 Applications

Fig. 10.5 Reflected second-harmonic intensity (SHI) from KDP crystalof the non-phase-matchable orientation in the neighborhood of thecritical angle for total reflection.

condition of total reflection prevails, the two cusps in Fig. 10.5 collapsed intoa single maximum peak of IR(2ω) at θcr(ω).

C. Nonlinear Brewster Angle

According to the BP theory, a nonlinear Brewster angle is predicted for second-harmonic generation from a nonlinear medium. The nature of a nonlinearBrewster angle is considered as a counterpart to the Brewster angle in the

Page 252: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.1 Optical Harmonic Generation 235

Fig. 10.6 Reflected second-harmonic intensity (SHI) from a KDP crys-tal with phase matching at total reflection.

linear optical case. When the fundamental beam is incident upon a nonlin-ear medium of a specific crystallographic orientation, the reflected second-harmonic intensity IR(2ω) vanishes at a particular angle of incidence calledthe nonlinear Brewster angle. The origin and physical interpretation of thenonlinear Brewster angle [92,93] can be understood in terms of classical dipoleradiation. At the nonlinear Brewster angle, the fundamental beam will cre-ate a nonlinear polarization PNLS(2ω) inside the medium in the direction ofpropagation of the reflected second-harmonic wave. According to classicaldipole radiation theory, there is no radiation in this direction. This nonradia-tive wave upon refraction back into the linear medium would otherwise giverise to the reflected second-harmonic wave in the direction of KR. The re-flected second-harmonic intensity IR(2ω) vanishes at the nonlinear Brewster

Page 253: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

236 10 Applications

angle when the nonlinear polarization source PNLS(2ω) lies in the plane of in-cidence. Since the reflected second-harmonic intensity IR(2ω) depends on thecomponents of PNLS(2ω) in the plane of incidence, measurement of the non-linear Brewster angle, at which IR(2ω) = 0, will directly yield the ratio of thecomponents of PNLS(2ω) in the plane of incidence, which could be helpful indetermining the relative signs of various components of the χ(2) tensor [102].Hence, measurement of the nonlinear Brewster angle could be applicable as a“null” method, which is similar to the work of Heinz et al. [103] for the precisemeasurement of the nonlinear optical susceptibility tensor. It is interesting topoint out that if the orientation of the KDP crystal as indicated in Fig. 10.5is employed for studying the reflected second-harmonic intensity, IR(2ω), forthe range 20 ≤ θi < 50, one can obtain a dip of IR(2ω) at θNB

i = 42.83,which corresponds to the nonlinear Brewster angle of KDP that was observedexperimentally [95] as shown in Fig. 10.7.

Fig. 10.7 Reflected second-harmonic intensity from KDP, showingnonlinear Brewster angle of KDP at 42.83.

Page 254: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.2 SHG due to Reflection from Media 237

10.2Optical Second-Harmonic Generation due to Reflection from Media with Inver-sion Symmetry

The third-rank χ(2) tensor vanishes in media with inversion symmetry, suchas Ag, Au, Si and Ge. However, second-harmonic generation in reflection hasbeen observed in these media. In early experiments, interests were in iden-tifying the physical mechanism responsible for the observed effects. It wasfound that both conduction and bound electrons contribute to the observedeffect. From the beginning, it was pointed out that such phenomena may beutilized to probe the surface property of materials. Surface optical harmonicgeneration indeed becomes a powerful tool for probing interfaces betweenmedia with inversion symmetry. Second-harmonic generation processes areforbidden by symmetry in media with an inversion center but are allowedon the surface because of a lack of inversion symmetry of the surface layers.They are therefore highly surface specific and can be used as surface probes.In comparison with other surface techniques, such as coherent anti-Stokes Ra-man scattering (CARS) and stimulated Raman gain, surface SHG has the ad-vantages of being much simpler in the experimental arrangement and yieldsmuch stronger signals.

The theory of SHG at an interface between linear and nonlinear mediawith inversion symmetry follows from the work of Bloembergen and Peshan,which was outlined in Section 10.1.1 with the following nonlinear polariza-tion:

PNLS (2ω) = (δ − β)(E(ω) · ∇)E(ω) + βE(ω)(∇ · E(ω))

+ αE(ω)× H(ω) + χ(2)S δ(z) : E(ω)E(ω)

+ χ(3)b : E(ω)E(ω)Edc(z).

The first two terms on the right are of quadrupole character. The third termis a magnetic dipole. Following the work of Shen, a surface dipole layer has

been added with surface nonlinear susceptibility χ(2)S , which can model the

submonolayer of adsorbate at the interface between two dense media. Finally,a DC-field-induced SHG (EFISHG) was added. The last term is of techno-logical significance since it can be utilized to probe the solid interface in Simicroelectronic materials and devices. For a given sample, it may only havea contribution from a few terms in PNL

S (2ω). In rare cases there are contribu-tions from all terms. Using the appropriate term for PNL

S as the driving sourcefor the nonlinear wave equation, one can solve for the solution of the reflectedSH field. In the case of quadrupolar source terms only, the solution has beenworked out by Bloembergen et al. [104].

In general, the reflected harmonic wave has components polarized in theplane of incidence (ER

2‖) as well as perpendicular to the plane of incidence

Page 255: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

238 10 Applications

(ER2⊥). In terms of the incident field E0, one obtains

| ER2⊥ | = (32πω/c)E2

0 sin θ cos2 θ sin φ cos φ,

×∣∣∣∣ [ε(ω)− 1]β[ε(ω)]1/2g0(ω)g1(ω)g1(2ω)

∣∣∣∣ (10.15)

and

| ER2‖ | = (32πω/c)E2

0 sin θ cos2 θ

× | [ε(2ω)]1/2ε(ω)g0(2ω)g20(ω)−1 |

×∣∣∣[−γε(ω)[cos2 φ + sin2 φg2

0(ω)/g21(ω)]

− 12

δ[ε2(ω)− 1]ε(2ω) cos2 φ

+ β[ε(ω)− 1][ε(ω)− sin2 θ]1/2

× [ε(2ω)− sin2 θ]1/2 cos2 φ]∣∣∣,

(10.16)

where

g0(ω) = [ε(ω)]1/2 cos θ + [1 − ε−1(ω) sin2 θ]1/2

g0(2ω) = [ε(2ω)]1/2 cos θ + [1 − ε−1(2ω) sin2 θ]1/2,

g1(ω) = cos θ + [ε(ω)− sin2 θ]1/2,

g1(2ω) = cos θ + [ε(2ω)− sin2 θ]1/2.

These equations give the SH amplitude and polarization in terms of the in-cident fundamental amplitude E0, as a function of the angle of incidence θ

and of the direction of the incident polarization φ. For semiconductors andinsulators one expects that β = −2γ = 2β. For materials with a high dielec-tric constant, the dominant term in the expression of ER

2‖ is the term involv-

ing β. The drawn curve in Fig. 10.8 describes the theoretical dependence of thesquare of this term on the angle of incidence, θ. The behavior is largely deter-mined by the function sin2 θ cos4 θ. The behavior is common to silicon, wherethe contribution due to valence electrons dominates. However, in metals bothconduction and bound electrons contribute, leading to a more complicated an-gular dependence. The dominant nonlinear surface terms are also responsiblefor the dependence of the harmonic intensity on φ. The last two terms in theexpression of ER

2‖ predict a cos4 φ dependence for both Ag and Si as shownin Fig. 10.9. For metals, the conduction electron (plasma) contribution to thenonlinearity is usually larger than that arising from interband transitions of

Page 256: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.2 SHG due to Reflection from Media 239

Fig. 10.8 The variation of the SH intensity in Si and Ag as a functionof the incidence angle. The fundamental electric field vector lies in theplane of incidence, φ = 0. The drawn curves are derived from theory.

the valence electrons. In this case α, β, δ, γ should be replaced by

βpl =e

δπm∗ω2 , γpl =c

2iωαpl = βpl

(ω2

p

4ω2

)

δpl = βpl + 2γpl,

with ω2p = 4πn0e2/m∗ being the plasma frequency.

From earlier studies, it has been shown that SHG due to reflection from me-dia with inversion symmetry is described rather well by the quadrupole-typenonlinear properties calculated for the homogeneous bulk material with anabrupt discontinuity at the boundary. The order of magnitude can be correctly

Page 257: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

240 10 Applications

Fig. 10.9 Variation of the SH intensity as a function of angle φ be-tween the fundamental electric vector and the plane of incidence. Theangle of incidence is 45. (a) Data for Ag; (b) data for Ge. The solidline representing a cos4 φ dependence is drawn for comparison.

related to the linear dielectric constant in insulators. There is a marked trendof increasing nonlinearity with increasing dielectric constant. The bound elec-trons also contribute significantly to the nonlinearity in metals. The direc-tional and polarization properties are well described by a combination of lin-

Page 258: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.2 SHG due to Reflection from Media 241

ear and nonlinear Fresnel factors. In materials with a high dielectric constant,the dominant contribution comes from the first layer of atoms in the bulk ma-terial.

The second-harmonic generation from a submonolayer of adsorbates can bedescribed by the surface susceptibility term,

χ(2)S = χ

(2)SA + χ

(2)SS , (10.17)

where χ(2)SA denotes the contribution from adsorbed atoms or molecules, and

χ(2)SS that from the surface layer of the adjoining medium. If | χ

(2)SA || χ

(2)SS , the

surface SHG can be used to probe the adsorbates. This effect has been utilizedby Shen [85] to study the nonlinear spectroscopy of dye molecules on a fused-quartz substrate. Using a tunable laser, resonant-enhanced SHG was observedfrom rhodamine 6G and rhodamine 110 dyes, which show distinct peaks ofthe SH signal when the second-harmonic frequency 2ω is in resonance withthe S0 → S2 transition. The resonant-enhanced SH signal was very strong,several orders of magnitude stronger than the SH signal from the fused-quartzsubstrate. With input and output beam polarizations and geometries properly

chosen, it is possible to measure a particular tensor element of χ(2)SA selectively.

The average orientation of p-nitrobenzoic acid (PNBA) molecules on fusedquartz has been determined in this fashion at solid–air and solid–ethanol in-terfaces to be 70 and 40, respectively [85].

The SHG technique offers a number of advantages as a surface-specificprobe: simple experimental setup, in situ probing of the interface and time-resolved dynamic properties of interfaces obtainable with a pump–probemethod using a femtosecond laser. There is a major advantage in using a highrepetition rate femtosecond laser for the surface SHG experiment. In these ex-periments, a photon-counting technique is used to detect the SH photons. Thedata-acquisition rate and signal-to-noise ratio depend on the photon-countingrate, i.e., photons/s. It has been shown [105] that the photon-counting rate, C,is given by

C = S(2ω) frep = 10−9 F2 ATp

frep,

where S(2ω) is SH photons/pulse, frep the pulse-repetition rate, F the fluencein mJ/cm2, A the area of the laser spot in cm2 and Tp the pulse duration inseconds. For a Ti:sapphire laser with A ∼ 10−4 cm2, tp ∼ 100 fs, F ∼ 0.1mJ/cm2 and frep ∼ 106/s the counting rate is about 106 SH photons/s. Withthis counting rate, near “real-time” monitoring of subsecond surface kineticprocesses, e.g., CVD growth, can be obtained using a SHG surface probe [105].

With the improved sensitivity due to the use of a femtosecond high repeti-tion rate laser, it is now possible to study the nonlinear electroreflectance or

Page 259: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

242 10 Applications

electric-field-induced second-harmonic generation (EFISHG) at silicon–soliddielectric interfaces [106].

10.3Nonlinear Electroreflectance

Nonlinear electroreflectance (NER) was first observed in 1967 [107]. It hasbeen shown that the reflected optical second-harmonic generation dependsquadratically on the externally applied bias voltage perpendicular to the in-terface. Its precise physical mechanism remains uncertain due to the lack ofsensitivity in detecting SHG signals from these early experiments. Thanks tothe advent of high repetition rate femtosecond lasers, sensitive gated photoncounters and data-acquisition instruments, systematic and detailed studies ofNER or electric-field-induced SHG (EFISHG) are now possible. This techniquehas been applied to silicon microelectronic devices operating at multiple giga-hertz frequencies. The internal electric field distribution at the solid–siliconinterface can be probed and mapped out with this new technique. We shallreview some of these works in this section.

Optical SHG was widely observed from interface-specific probes of cen-trosymmetric materials, because χ(2) vanishes in the bulk of these materials.Si(001) interfaces are among the most important and the simplest for suchnonlinear optical analysis. Their importance derives from their ubiquitouspresence in metal–oxide–semiconductor (MOS) devices. Their simplicity de-rives from the small number of tensor elements required to describe the in-terfacial nonlinear response. Downer and coworkers [108] have systemati-cally studied the SH spectroscopy in which they have varied and indepen-dently measured various parameters affecting SHG: subsurface DC electricfields, carrier-induced screening of DC EFISHG, surface hydrogen coverageand doping concentration.

By varying the DC E fields, one can distinguish the surface, bulk and DC-field-induced contributions due to SHG and validate the theoretical frame-work for NER in semiconductors with unprecedented completeness. Phe-nomenological description of NER for Si(001)/SiO2 interfaces is much sim-pler than that for the Si(111)/SiO2 interfaces. For a p-polarized fundamentallight wave and a p-polarized SH field (Pin–Pout) configuration, the second-harmonic intensity from a Si(001) interface can be written as [108]

Ipp(2ω) ∼| PSisotropic + PB

iso + PBaniso cos(4Ψ) + PBE(φ − φ f b) |2,

where Ψ is the azimuthal angle between the incident plane and the < 110 >direction of Si. PS

is0 and PBis0 are due to the isotropic surface and the bulk non-

linear polarizations. PBanis0 and PBE are the anisotropic and field-induced bulk

polarizations. Also, φ f b is the flat-band voltage and φ is the applied voltage. If

Page 260: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.3 Nonlinear Electroreflectance 243

we substitute 2PBE × PBaniso ∼ a4, | PBE |2∼ aFD

0 , | PBaniso |2 /2 ∼ a8, Ipp(2ω, Ψ)

can be expressed as [108]

Ipp(2ω, Ψ) = aFI0 + aFD

0 (φ − φ0)2 + a4(φ − φ0) cos(4Ψ)

+ a8 cos(8Ψ),(10.18)

where aFI0 is a field-independent parameter, aFD

0 is the coefficient for the field-dependent term, cos(4Ψ) reflects the four-fold symmetry of the Si(001) sur-face and φ0 is the applied voltage that yields a minimum SH intensity whenφ0 ∼ φ f b. By measuring Ipp as a function of the applied bias voltage, az-imuthal angle Ψ and second-harmonic photon energy, all a parameters in Eq.(10.18) can be extracted from experimental data. Since EFISHG is strongly re-lated to the intrinsic potential at the silicon surface, carrier-induced screeningof the DC field at the Si(001)–SiO2 interface plays an important role in deter-mining the SH intensity. Downer’s group [108] studied this screening effectusing a pump beam to illuminate the area which generates the EFISHG. It wasobserved that the SH signal amplitude decreases as the pump laser fluence in-creases. This can be understood by rewriting

P(2ω) ∼ χ(3)B : E(ω)E(ω)EDC(z) +χ

(2)S : E(ω)E(ω) +χ

(2)QB : E(ω)∇E(ω),

where χ(2)QB is the four-fold anisotropic bulk quadrupole susceptibility, EDC is

the space-charge field normal to the interface which is sensitive to screening

and χ(2)S is the isotropic surface dipole susceptibility. It should be noted that

χ(2)S and χ

(3)B are frequency dependent. χ

(2)B dominates near the 2hω ∼ 3.26 eV

resonance, and is responsible for the field-independent SHG; while χ(3)B dom-

inates near the 2hω ∼ 3.4 eV resonance associated with the E1 critical point.Fluence-independent SHG was observed at 2hω ∼ 3.26 eV, while the SH in-tensity at 3.4 eV decreases quadratically with increasing fluence, illustratingthe carrier-induced screening of the DC electric field. Since doping densityaffects the band bending near the surface, SH signals show a strong influencefrom doping concentration and hydrogen coverage. This effect has been uti-lized for in situ “real-time” monitoring of the silicon epitaxial growth rate andgrowth condition on Si(001).

Since Si is the dominant semiconductor for advanced microelectronic de-vices that are fabricated on a single chip, there is a strong demand for newnoninvasive measurement techniques with high temporal and spatial resolu-tion. The EFISHG method described above can conveniently fill this need.It can be used to map out the electrical field distribution at various internalnodes on the chip and at the same time measure the high-frequency electricalwaveform or electrical impulses propagating through each node. Broadbandelectrical impulse measurement with 1-ps time resolution with the EFISHGtechnique has been reported [106].

Page 261: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

244 10 Applications

10.4Near-Field Second-Harmonic Microscopy

In the previous section we have shown examples of harmonic generationin reflection. These are all far-field effects. We have learned that opticalsecond-harmonic generation is an extremely sensitive technique for character-ization and investigation of a wide variety of material surfaces and absorbates.Second-harmonic generation is known to be affected by crystal structure, mag-netic and ferroelectric order, mechanical strain, etc. Second-harmonic genera-tion is especially useful for recovery of the symmetry properties of the sampleunder investigation. Far-field SH microscopy has already demonstrated greatpotential in such fields as magnetic imaging and characterization of metal–oxide–semiconductor (MOS) devices [106]. In many applications, spatial res-olution below the diffraction limit of far-field optics is required. For example,better than 100-nm resolution is required for characterization of thin ferroelec-tric films such as BaTiO3 and PbZrxTi1−xO3 (PZT), which form the basis of anew thin-film technology for data storage. As a result, many attempts havebeen made recently to implement various near-field optical techniques for SHimaging. In this section we shall review these topics.

The first example to be discussed is the near-field SH imaging of BaTiO3and PZT. They are ferroelectric materials. The interest of these materials liesin their ability to switch from one stable polarization state to another. Thisproperty forms the basis of a new thin-film technology for data storage. ThinPZT films are used in prototype nonvolatile ferroelectric random-access mem-ory (NVFRAM) and dynamic random-access memory (DRAM). PZT materi-als also have numerous applications in actuators, transducers, resonators andsensors. Crucial parameters of piezoceramic performance in different appli-cations are hysteresis, nonlinearity, polarization retention (or the loss of polar-ization), etc. For proper application of the material, it is imperative to under-stand this nonlinear behavior on a microscopic level. Optical microscopy ofSHG in piezoceramic PZT is a natural tool to study this local nonlinearity.

Far-field observations of SHG in PZT ceramic near the ferroelectric phasetransition demonstrate a sharp drop in SHG in the paraelectric state. The SHintensity vs temperature curve corresponds qualitatively to the temperaturedependence of the spontaneous polarization, which is an order parameter inthe transition to the polar phase. PbZrxTi1−xO3 (PZT) ceramic is a stronglyscattering medium consisting of individual submicrometer-size crystallites.The elastic scattering lengths of fundamental and second-harmonic light inPZT are of the order of the characteristic size of a crystallite. Theory of SHGin strongly scattering media shows that generation of SH light in PZT shouldoccur in a thin layer (of the order of the elastic scattering length) near theilluminated surface. Thus near-field second-harmonic microscopy is ideallysuited for studies of PZT nonlinearity and poling properties at the microscopiclevel (the level of individual crystallites and crystallite boundaries). Its main

Page 262: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.4 Near-Field Second-Harmonic Microscopy 245

Fig. 10.10 Schematic of the near-field second-harmonic microscopysetup. p- and s-polarization directions of the fundamental light areindicated.

advantages in comparison with other scanning probe techniques such as therecently developed piezo response atomic force microscopy are the possibilityof fast time-resolved measurements and substantially smaller perturbation ofthe sample under investigation caused by the optical probe. It is a noninva-sive probe technique. The experimental setup for near-field SH imaging ofthe surface of PZT ceramic is described in Fig. 10.10. Typically, a Ti–sapphirelaser system consisting of an oscillator and a regenerative amplifier at 810 nmis used as the light source. The laser has a repetition rate up to 250 kHz and apulse duration of 100 fs with a pulse energy of 10 µJ. The SH from the surfaceof a PZT sample is excited at an angle of incidence around 60 by a weakly fo-cused beam of the Ti–sapphire laser. The excitation power should be kept wellbelow the ablation threshold. The local SHG is collected using either a coatedor an uncoated adiabatically tapered fiber tip. The fiber tip is drawn at the endof a single-mode fiber by a standard heating and pulling procedure. The fibertip is scanned over the sample surface with a constant tip–surface distance of afew nanometers using a shear force feedback control system [109]. Therefore,surface topography images can be obtained with a resolution on the nanome-ter scale, while simultaneously recording the SH near-field image. The SHsignal is measured with a gated photon-counting photodetector. The charac-teristic SH photon counting rate is about one photon count per 30 laser pulses.With a laser repetition rate of 250 kHz, a typical SH signal at every point of theimage requires averaging over 30–100 ms, or 7500 to 25 000 pulses. The imageacquisition time for an image size of 1× 3 µm2 is about 20–30 min. This showsa clear advantage for using high repetition rate, high peak power Ti–sapphireregenerative amplifier lasers. In comparison, a similar image would require200–300 min to acquire if a Q-switched kilohertz repetition rate Nd:YAG laserwere used.

Page 263: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

246 10 Applications

Fig. 10.11 Orientation of the BaTiO3 crystal with respect to the illumi-nating fundamental light for different configurations.

Smolyaninov et al. [110] have recently demonstrated the near-field SHimaging of the c/c/a/a polydomain structure of epitaxial PZT thin films. Inorder to investigate the symmetry property of the nonlinear effect, they firstperformed an experiment on a poled single crystal of BaTiO3. Since BaTiO3and PZT have a similar perovskite structure and belong to the same tetragonalsymmetry class, symmetry characteristics of BaTiO3 can be applied directly tothe crystallites of the PZT thin film. Furthermore, phase-matched SHG is pro-hibited in the bulk of BaTiO3 because of its strong dispersion. As a result, themeasured SH signal originates at the surface, which makes the experimentalsituation look very similar to the case of a thin PZT film. Near-field SH signaldependences on the polarization of the fundamental light for three differentpoling (optical axis) directions of the BaTiO3 crystal can be analyzed accord-ing to the configuration shown in Fig. 10.11. Let us consider fundamentallight illuminating the tip of the microscope located near the BaTiO3 crystalsurface. Three different cases of the poling vector orientation are shown inFig. 10.11, where α is the incident angle, and the z-axis is chosen to coincidewith the poling direction. All three cases may be considered in a similar way,so we will concentrate on case (b) and only present the final results for thecases (a) and (c). Let us consider the fundamental optical field distributionin the tip–sample region at distances much smaller than the wavelength λ

of the fundamental light. In this region the quasielectrostatic approximationmay be used. The tip shape may be approximated as an ellipsoid with a veryhigh aspect ratio. For p-polarized excitation light along the y component (thecomponent perpendicular to the sample surface) of the crystal the exact ana-lytic solution of the Laplace equation [111] gives Etip(ω) = εEy0(ω), where ε

is the dielectric constant of the tip, Ey0 is the incoming field and Etip is the ycomponent of the field just below the tip apex. As a result, for the incidenceangle α around π/2 the local SHG under the tip is enhanced by a factor of n8

(since I(2ω)tip = E2tip(2ω) = E4

tip(ω) = n8E4y0(ω)), where n is the tip refrac-

tive index. Assuming that n = 1.6, the enhancement factor equals about 40. It

Page 264: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.4 Near-Field Second-Harmonic Microscopy 247

may be even higher (of the order of (2ε)4) in the vicinity of a sample surfacewith a high dielectric constant, such as a metal or a ferroelectric, due to theimage potential. In our discussion to follow, we will not consider any particu-lar value of the field enhancement factor. Instead, we denote the fundamentalfield enhancement factor by γ. The fundamental optical field componentsunder the tip apex at a polarization angle φ can be written as

Ex = E cos φ cos α, Ey = (γ/εb)E cos φ sin α, Ez = E sin φ, (10.19)

where εb is the dielectric constant of BaTiO3 and φ = 0 and φ = 90 corre-spond to the p-polarized and s-polarized excitation light, respectively. Tak-ing into account the nonzero components of the second-harmonic suscepti-

bility tensor for BaTiO3: χ(2)yyz = χ

(2)xxz = χ

(2)yzy = χ

(2)xzx = 17.7 × 10−12 m/V,

χ(2)zyy = χ

(2)zxx = −18.8 × 10−12 m/V and χ

(2)zzz = −7.1 × 10−12 m/V (where z

is the poling direction [76]), the second-harmonic field components at the tipapex may be written as

D(2)x = χ

(2)xzxE2 sin2 φ cos α, D(2)

y = (γ/εb)χ(2)xzxE2 sin2 φ cos α,

D(2)z = (γ2/ε2

b)χ(2)zxxE2 cos2 φ sin2 α

+ χ(2)zxxE2 cos2 φ cos2 α + χ

(2)zzzE2 sin2 φ.

(10.20)

If we assume that the microscope tip collects only dipole radiation, the second-harmonic optical signal will be proportional to

I(2) = D(2)2z + D(2)2

x . (10.21)

(Second-harmonic dipoles that oscillate along the y direction do not radiatetoward the tip.) Similar calculations of the field at the tip apex are easy to per-form for cases (a) and (c). Let us point out that within this model the angulardependences of the field E for the points in space around the tip apex shouldbe approximately the same as in Eq. (10.19). Only the enhancement factor γ

is different for these points (its value changes from a maximum just under thetip apex to γ = 1 far from the apex). So, in the final result one must replace thefactors γ4 and γ2 by the average values < γ4 > and < γ2 > taken over all thearea around the tip apex where the SH signal is collected.

The final results of the calculations are presented in Fig. 10.12. They shouldbe compared to the experimental result shown in Fig. 10.13. The SH signal ex-hibits four-fold symmetry with respect to a polarization angle rotation in thecase (a) of the poling (optical axis) direction located in the incidence plane ofthe fundamental light parallel to the crystal surface. The symmetry becomestwo-fold in case (b) when the optical axis is perpendicular to the incidenceplane. In case (c) when the poling direction is perpendicular to the crystal sur-face, the SH signal is much weaker than in cases (a) and (b). Big differences

Page 265: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

248 10 Applications

Fig. 10.12 Theoretical near-field SH signal dependences on the po-larization angle of fundamental light corresponding to the geometryshown in Fig. 10.11.

Fig. 10.13 Near-field SH signal dependences on the polarization ofthe fundamental light for the three different poling directions of theBATiO3 crystal (shown by the arrows) with respect to the incomingfundamental light (shown by the straight lines) for the geometry inFig. 10.11.

between these dependences indicate that the near-field SHG may be used torecover the local poling direction in the thin ferroelectric films. The best agree-ment with experimentally measured symmetry properties of the near-fieldSH polarization curves was achieved in the γ = 5.7 range for the field en-hancement factor. This is reasonably close to the expected value of γ = 2εtip.A much smaller value of the SH signal in case (c) is a consequence of the factthat in the cases (a) and (b), I(2) contains terms proportional to < γ4 >, but in

Page 266: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.4 Near-Field Second-Harmonic Microscopy 249

case (c) the SH signal is proportional to < γ2 >:

I(2) = D(2)2x + D(2)2

y

= (< γ2 > /ε2b)χ

(2)2xzx E4(sin2 2φ sin2 α + cos4 φ sin2 2α).

(10.22)

The transition from the near-field to the far-field mode of the microscope op-eration (which corresponds to an increase in the tip–sample separation from afew nanometers to a distance of a few wavelengths of light) can be describedmathematically as a transition from γ = 5–7 to γ = 1 (no field enhancementby the tip). Results of numerical calculations for such a transition performedusing Eq. (10.21) are shown in Fig. 10.14. Experimental data shown in thesame figure are in good qualitative agreement with the calculations: both the-oretical and experimental data have the same symmetry of the SH polarizationdependences in the near-field and far-field zones. This indicates that this sim-ple model provides an adequate description of the essential physics involvedin near-field observation of SHG from ferroelectric samples.

Fig. 10.14 Theoretical (open circles and triangles) and experimentallymeasured (filled circles and triangles) data for the transition from thenear-field to the far-field behavior of the SH signal polarization depen-dency for geometry (b) in Fig. 10.11.

Based on the developed understanding of the SHG contrast due to differ-ences in ferroelectric poling directions, one can calculate the spatial distribu-tion of the local SHG from a typical PbZrxTi1−xO3 film showing c/a/c/a poly-domain structure assuming 100-nm spatial resolution of the microscope. Suchcalculations have been carried out by Smolyaninov et al. [110]. Their experi-mental observation agreed qualitatively with the theoretical analysis.

Page 267: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

250 10 Applications

10.5Terahertz Pulse Generation by Optical Rectification

Most examples given so far has been on second-harmonic generation, whichis the degenerate case for the sum-frequency generation. To generate low-frequency or even DC waves one can use difference-frequency generationwith χ(2)(ω1 − ω2). When ω1 = ω2 this leads to an optical rectification effectwhich was first observed in 1962. Interest in difference-frequency generationwas primarily centered on the generation of coherent far-infrared waves. Withthe advent of the femtosecond optical pulse, optical rectification provides aconvenient means of generating ultrashort electrical impulses by the opticalrectification process [85]

PNL(0; ω,−ω) = χ(2)(0; ω,−ω)E(ω)E∗(ω).

The beat wave has zero frequency. The nonlinear process essentially stripsoff the high-frequency component, leaving only the envelope function of thepulse. In principle, the electric impulse thus generated should follow the en-velope function of the original optical pulse. For an optical pulse of a fewhundred femtoseconds in duration, the electric impulse will last about 1 ps.When such pulses are radiated into free space, they are referred to as terahertzradiation since their central frequency lies in the terahertz spectral range. Apure DC electric signal cannot radiate; however, carrierless short electricalwaves can radiate, generating ultrawideband signals with a small number ofRF cycles. Previous work with this type of terahertz radiation has reportedthat the radiation typically consists of the one and one-half cycles – Mexicanhat – pulse shape [112].

Pulsed terahertz radiation emits ultrawideband signals covering the spec-tral region from 300 GHz to 12 THz (or 10–400 cm−1). This spectral regionfills the so-called “gap” (or void) between microwave and mid-infrared. Thedifficulties of conventional far-IR spectroscopy lie in the lack of efficient ra-diation sources and detectors. Pulsed terahertz radiation is a clean coherentTHz source enabling scientists to perform THz spectroscopy by using a time-domain spectroscopy technique.

By sampling the precise waveform emitted from the radiator and by mea-suring the transmitted or scattered signal from the sample, one can extract allnecessary information about the sample. This is possible because the changein the waveform is the result of interaction between the pulsed terahertz ra-diation and the sample. The time-domain measurement provides completeinformation on the amplitude and phase (time delay or variation) changeof the pulsed waveform as it interacts with the sample. Thus, by perform-ing a Fourier transform of the time-domain result one obtains the frequency-domain information just like that obtained in standard spectroscopy.

Page 268: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.5 Terahertz Pulse Generation 251

Time-domain spectroscopy (TDS) techniques can yield both the real andimaginary parts of the dielectric constant of the material as a function of fre-quency without resorting to the Kramers–Kronig relations. THz emissionfrom materials can also be used to study the physical properties of the materi-als, such as semiconductor interfaces, lattice and carrier dynamics via coher-ent phonon oscillation, velocity overshoot and the coherent charge oscillator.Optical pump-THz probe spectroscopy techniques have also been applied tostudy carrier transport in semiconductors. Strong THz pulses have been usedto ionize Rydberg atoms for atomic study.

THz impulse radiation can also be generated by exciting a photoconduc-tive switch with femtosecond optical pulses [113]. In this case the photocon-ductive semiconducting material is transformed instantaneously from a semi-insulating state (assume that a high dark resistivity semiconductor is used)to a quasimetallic state when the femtosecond pulses illuminate the semi-conductor with a photon energy greater than the semiconductor band-gapenergy. If a voltage is applied between the two electrodes of the dipole an-tenna, a transient current will suddenly appear when the semiconducting gapis illuminated, causing radiation of an ultrawideband electrical impulse sig-nal. Among the two methods of generating THz pulses, the photoconductivedipole antenna is in general more efficient. However, optical rectification en-ables one to study nonlinear optical characteristics of the sample. It serves asthe only technique to generate THz pulses from a dielectric. In this section wewill limit our discussion to the optical rectification method.

Fig. 10.15 Experimental setup for THz pulse generation and detec-tion.

Page 269: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

252 10 Applications

Fig. 10.16 Electrooptic THz field sensor in a balanced detectionscheme.

Fig. 10.17 THz signal detected by TiTaO3 EO sensor.

Fig. 10.18 Fourier transform of the waveform shown in Fig. 10.17.

Page 270: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.5 Terahertz Pulse Generation 253

A typical THz experimental setup is shown in Fig. 10.15 [114]. It consists of alaser-driven THz emitter and a laser-actuated, gated THz detector (sampler orsensor). THz detectors can either be a PDA (photoconductive dipole antenna)type or an electrooptic sensor type. Impulse THz is generated in the nonlin-ear optical crystal by the optical rectification effect. It radiates in the forwarddirection. THz radiation is then collected by an off-axis paraboloidal mirroror THz lens onto a THz detector or sensor (EO sensor, for example). The tem-poral waveforms of the THz radiation from the emitter without the sampleand the waveforms after transmission through the sampler are obtained by apump and probe technique. The EO sensor to detect the free-space THz radi-ation is shown in detail in Fig. 10.16 [115]. The THz beam and optical probebeam interact in the EO sensor, which should be properly designed and cutwith velocity matching between THz pulses and probe optical pulses, yield-ing the maximum EO modulation effect. Without THz radiation illuminatingthe EO sensor, the probe laser beam polarization, EO crystal orientation, com-pensator and Wollaston prism are adjusted so that the balanced detector giveszero signal. When the THz pulse overlaps with the probe laser pulse in the EOcrystal, the transient electric field associated with the THz beam induces an EOeffect and the two beams emerging from the Wollaston prism become unequalin amplitude; consequently, the detector becomes unbalanced and registers asignal. The system is set up and calibrated such that the signal from the bal-anced detector is proportional to the amplitude of the electric field of the THzbeam. Since the EO crystal senses the electric field, the signal changes polaritywhen the electric field polarity is changed. To map out the THz waveform theprobe pulse is temporally delayed by an optical delay line driven by a step-ping motor. The THz field induced intensity modulation of the probe beamis picked up by the balanced detector and is mapped out as a function of thedelay between the pump and probe pulses. To detect a small change in modu-lation a phase sensitive detection technique as typically employed in a lock-inamplifier is usually required.

A typical THz signal generated by a LiTaO3 crystal is shown in Fig. 10.17; aMexican hat shape temporal waveform is clearly displayed. The Fourier trans-form of the waveform in Fig. 10.18 renders frequency-domain information.

The coherent THz detection technique can be used to study the mid-infraredproperty of the EO crystal. Using a 10-fs laser pulse, this technique is capableof reaching 44 THz. Recently, this has been demonstrated by Wu [114] andWu and Zhang [115]. They studied the mid-IR optical property using well-characterized EO sensors of 10 and 30 µm in thickness. Each sensor has beentested to operate between 10 and 50 THz. For example, the mid-IR frequencyresponses of 10- and 30-µm ZnTe sensors are shown in Fig. 10.19. These EOsensors have been used to study the mid-IR radiation from a GaAs emitter

Page 271: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

254 10 Applications

Fig. 10.19 Mid-IR frequency response functions of ZnTe sensors withthicknesses of 10 µm and 30 µm, respectively.

by optical rectification. Figure 10.20 shows a typical waveform from a GaAsemitter. The EO sensor is a 30-µm-thick ZnTe crystal. A Fourier transform ofthe waveform is shown in Fig. 10.21. The measured frequency characteristicsof the THz pulses emitted from GaAs are the convolution of the frequencyresponses of the emitter and detector (sensor). A spectral modulation periodof about 2 THz is due to the multiple reflections from the ZnTe sensor (30 µm)and the dips at 17 THz and 37 THz are also due to the response of the 30-µmZnTe sensor. It would be more desirable to use a thinner sensor (10 µm) toavoid these big dips.

The coherent THz detection technique can also be applied to determine theratios between various nonlinear optical coefficients. The THz field due tooptical rectification is

Erd(t) ∼ ∂2

∂t2 P(t)

∼ ddcijk

∫ ∞

0exp(iΩt)Ω2 dΩ

∫ ω0+∆ω/2

ω0−∆ω/2Ej(ω + Ω)E∗

k (ω) dω.

When the detection of this field is done with a dipolar photoconductive an-tenna (DPA), the antenna will measure the component of the radiated electricfield along its axis. The signal measured depends on the crystalline symme-try, the orientation of the sample crystal DPA and the polarization of the laserand THz fields. Ma and Zhang have figured out a way to sort out the variousconfigurations [117]. They have derived theoretically the THz field as a func-tion of angle, θ, between the axis of the detector and the x′-axis of the crystalsample for various crystal symmetries and geometrical orientations.

Page 272: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

10.5 Terahertz Pulse Generation 255

Fig. 10.20 (a) Temporal waveform of the THz radiation from a GaAsemitter measured by a 30-µm ZnTe sensor. (b) Expanded view of thefirst wave packet.

For the cubic structure GaAs with point group 43m, there is only one in-dependent nonvanishing second-order nonlinear optical coefficient, namelyd14 = d25 = d36. It can be shown and experimentally confirmed that thereis no optical rectification fields when a GaAs(100) sample is incident nor-mally. However, there are rectification fields with three-fold rotational sym-metry from a GaAs(111) sample at normal incidence. The experimental resultis shown in Fig. 10.22. The experimental curve can be explained by

Ep = −Es =A√

6cos(3θ),

where Ep is the component of the rectified field parallel to the polarization ofthe incident laser light and Es is the perpendicular component. For crystals

Page 273: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

256 10 Applications

Fig. 10.21 Spectrum of the waveform in Fig. 10.20.

Fig. 10.22 Rectification field vs GaAs(111) crystal rotation angle,showing three-fold rotation symmetry of both Ep and Es.

with lower symmetry, such as LiTaO3, (trigonal structure), there will be moreindependent d coefficients; one has

Ep(θ) = A[(3d33 + d31 + 2d15) cos θ

+ (d33 − d31 − 2d15) cos 3θ],

Es(θ) = A′[(3d31 + d33 − 2d15) cos θ

− (d33 − d31 − 2d15) cos 3θ].

Similar expressions can be derived for monoclinic structures such as the or-ganic crystal dimethyl amino 4-N-methylstilbazolium tosylate or DAST. By

Page 274: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Problems 257

measuring Ep (θ) and Es (θ) of the rectified fields as a function of the az-imuthal angle, θ, the ratio of the d coefficients can be determined. For example,d31 = 0.21d33 for LiTaO3 and d12 = −0.4d11 = 295d26 for DAST.

Problems

10.1 Derive Eqs. (10.11)–(10.13).

10.2 Terahertz (THz) pulses can be generated by optical rectification in apoled polymer. If the optical pulses used for THz generation have a pulsewidth of 1 ps, calculate the bandwidth of the THz pulses.

Page 275: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 276: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

259

Appendix AAtomic Physics Definitions

In this appendix we present the most recent accepted values for the funda-mental constants used in this volume. All of these can be obtained from aweb site maintained by the National Institute for Standards and Technology(NIST). At the writing of this text, the web address is

http://physics.nist.gov/cuu/Constants/.

It must be understood that no physical quantity can be measured with ab-solute certainty; there will always be uncertainties! The uncertainties in thelast digits are given in parentheses and are the standard uncertainties.

Uncertainties come in two principal flavors, statistical and systematic. Sys-tematic uncertainties can be difficult to ascertain, requiring intimate knowl-edge of the instrumentation and the technique used in the measurement aswell as a reliable way to determine the deviation a priori. Often one can correctfor systematic uncertainties. Statistical uncertainties are not correctable per se,and are thus handled differently. Consider a measurement that yields a valuey. The uncertainty in the measurement is statistical when the distribution thatcharacterizes the probability for measuring a specific value y is approximatelynormal (Gaussian). We can define a standard deviation, σ, such that if we wereto measure y many many times, approximately 68% of the distribution wouldlie in a range

y − σ to y + σ,

where y is the mean value of the measurements. By definition, the standarduncertainty is the estimated standard deviation. Typically, this uncertainty isexpressed as

y = 1234.56789± 0.00011 [units].

A more concise form used in this appendix, as well as the NIST tables, is

y = 1234.56789(11) [units].

Here, (11) is understood to be the numerical value for the standard uncertaintyreferred to the last two digits of the quoted result.

Page 277: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

260 A Atomic Physics Definitions

A.1Air and Vacuum Wavelengths

The index of refraction of a medium changes the wavelength of light travers-ing through it. The difference between the vacuum and air wavelengths near600 nm, for example, is approximately 0.16 nm in dry air. Given the air wave-length, the vacuum wavelength is calculated by

λair =λvac

nair. (A.1)

At 288 K and 1 atm, the index of refraction of air can be approximated withthe Cauchy formula given by

nair = 1 + 2.72643× 10−4 +(

1.2288 +3.555× 104

λ2

)1

λ2 , (A.2)

where λ is in nanometers.

A.2Wavenumber

The wavenumber, ν, is a convenient way to specify the energy of an electro-magnetic wave as well as the transition energy between states. Recall that themodulus of the wave vector,

−→k , of an electromagnetic wave is given by∣∣∣−→k ∣∣∣ =

λair=

2πnair

λvac=

2πnairν

c, (A.3)

ν =1

λvac=

1nairλair

=

∣∣∣−→k ∣∣∣2πnair

, (A.4)

where c is the speed of light in vacuum and ν is its frequency. Typically, ν hasthe units of cm−1, which is sometimes referred to as a kayser. One can thinkof the wavenumber as an energy unit with h = c = 1. Note that ν is usuallythought of as independent of the index of refraction. An example where thisis not true is standing waves generated by a high-finesse, fixed-length opticalcavity. When the index of refraction is changed in this case, the wavelengthremains fixed and so the frequency changes.

A.3Fine-Structure Constant

α =1

4πεo

e2

hc=

1137.035 999 76(50)

= 7.297 352 533(27)× 10−3. (A.5)

Recently, Gabrielse et al. 1 reported a new value for α−1 = 137.035 999 710(96).

1) G. Gabrielse, D. Hanneke, T. Kinoshita, M. Nio, and B. Odom. Phys.Rev. Lett. 97, 030802 (2006).

Page 278: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

A.4 Atomic Energy Unit (Hartree) 261

A.4Atomic Energy Unit (Hartree)

The atomic energy unit (a.u.), also known as a Hartree, is e times the electricpotential associated with two elementary charges (e) separated by the Bohrradius (ao),

1 a.u. = e(

e4πεoao

)=

αhcao

= 4.35974381(34)× 10−18 J. (A.6)

In Eq. (A.6), α(≡ e2/4πεo hc ∼ 1/137

)is the fine-structure constant, h is

Planck’s constant and c is the speed of light. These units are equivalent tosetting e = h = me = 1, where me is the electron rest mass.

A.5Rydberg Energy Unit

The Rydberg energy unit is given by

1 Ry =12

Hartree = 13.60569172(53) eV, (A.7)

and is the binding energy of the hydrogen atom with an infinitely massivenucleus. This unit is equivalent to setting h = e2/2 = 2me = 1. The atomicenergy unit (a.u.) and Rydberg energy unit (cm−1) are also related:

1 a.u. = 4πhcR∞, (A.8)

where R∞ is the Rydberg constant:

R∞ ≡ 12

e2

4πεoaohc=

α

4πao= 109737.31568549(83) cm−1. (A.9)

A.6eV Energy Unit

Since the quantity in parentheses in Eq. (A.6) has units of volts, we can alsodefine a new energy unit called an electron volt, eV. An atomic unit of energyis related to the eV through

1 a.u. = 1 Hartree = 27.2113834(11) eV. (A.10)

The conversion between Rydberg energy units and electron-volt units (eV) is

1eV = 8065.54477(32) cm−1. (A.11)

Page 279: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

262 A Atomic Physics Definitions

It is helpful to know that the NIST web site also offers a conversion factorapplet that was used here to convert from a.u. to eV.

A.7Mass

The atomic mass, as found in the periodic table, is typically expressed inatomic mass units (amu) and based on the definition of carbon 12 being ex-actly 12 amu. Consequently,

1 amu = 1.66053873(13)× 10−24 g. (A.12)

Another convenient mass unit is the rest mass, mc2, which is expressed as eV.The conversion between amu and eV is given by

1 amu · c2 = 931.494013(37)MeV. (A.13)

The mass of the electron is then

me = 9.10938188(72)× 10−28 g

= 5.485799110(12)× 10−4 amu (A.14)

= 0.510998902(21)MeV,

while the mass of the proton is

mp = 1.67262158(13)× 10−24 g

= 1.00727646688(13) amu (A.15)

= 938.271998(38)MeV.

A.8Length

There are several lengths that characterize atomic systems.

• The classical electron radius,

re ≡ e2

4πεomec2 = 2.817940285(31)× 1013 cm, (A.16)

is the radius at which the electrostatic potential is equal to the electronrest mass.

• The Compton wavelength,

λc ≡ re

α=

hmec

= 386.1592642(28)× 10−11 cm, (A.17)

is the wavelength where the photon energy is equal to the electron restmass.

Page 280: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

A.9 Atomic Velocity and Momentum 263

• The Bohr radius,

ao ≡ 4πεoh2

mee2 =re

α2 = 5.291772083(19)× 10−9 cm, (A.18)

is the smallest orbit in the hydrogen atom. The typical atomic size fora 1s electron in a hydrogen-like ion is given by ao/Z, where Ze is thecharge of the nucleus.

A.9Atomic Velocity and Momentum

In Section 1.1 we developed the Bohr energy spectrum by equating theCoulomb potential with the centripetal acceleration (Eq. (1.4a)). Solvingfor v in Eq. (1.4a) with r = ao, the Bohr radius, and Z = 1 defines the atomicvelocity; the velocity with which the electron moves in the first Bohr orbit.Thus,

vo =

√(e2

4πεo

)1

meao

= αc = 2.1876912529(80)× 108 cm/s,

(A.19)

where α is the fine-structure constant and c is the speed of light. The last stepwas made with the aid of Eq. (A.18). Multiplying this expression by the massof the electron defines the atomic momentum

po = meαc = 1.99285151(16)× 10−19 g cm/s, (A.20)

which can also be written as

po =hao

= 1.2438× 10−7 eV s/cm. (A.21)

A.10Atomic Time Scale

The scale is set by the electron’s period in the Bohr orbit,

τao ≡ao

αc= 2.418884326500(18)× 10−17 s, (A.22)

where αc is the atomic velocity (Eq. (A.19)).

Page 281: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

264 A Atomic Physics Definitions

A.11Atomic Field Strength

The characteristic electric field strength is the atomic field associated with thefield produced by the proton at the Bohr distance,

Fao ≡e

4πεoa2o

= 5.142 206 24(20)× 109 V/cm.

A.12Atomic Unit of Dipole Moment

eao = 8.47835267(33)× 10−28 C cm.

A.13Magnetic Moments

Magnetic moments play a central role in the details of atomic structure. Wewill summarize the key features for the electron and the nucleons in this ap-pendix.

A.13.1Electron Magnetic Moment

The magnetic moment of the electron has two contributions – orbital angularmomentum and spin angular momentum. The contribution due to its orbitalmotion about the nucleus is given by

µl = −µBge,ll, (A.23)

with an expectation value in a state of the form |γlml〉 of

〈µl〉 = −µBge,l

√l (l + 1). (A.24)

The z component is given by

µl,z = −µB ge,llz, (A.25)

〈µl,z〉 = −µB ge,lml . (A.26)

In these expressions µB is the Bohr magneton,

µB =eh

2me= 5.788 381 804(39)× 10−5 eV/T, (A.27)

and ge,l is known as the orbital g-factor for the electron. In the flavor of the os-cillator strength, the g-factor is a measure of the degree to which the quantum

Page 282: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

A.13 Magnetic Moments 265

mechanical orbital dipole moment is determined by the mechanical (classical)orbital angular momentum – ge,l = 1.2

The spin contribution to the magnetic moment is expressed similarly as

µs = −µBge,ss, (A.28)

〈µs〉 = −µBge,s

√s (s + 1). (A.29)

Its z component is given by

µs,z = −µB ge,ssz, (A.30)

〈µs,z〉 = −µB ge,sms, (A.31)

with ge,s ≈ −2, given by the Dirac equation. The total magnetic moment isgiven by

µ = µl +µs

= −µB [ge,ll + ge,ss]

= −µB [ge,lj + (ge,s − ge,l)s]. (A.32)

Since ge,s/ge,l ≈ 2, µ will not point in the same direction asj in general and,thus, will not always be a constant of the motion. A more useful quantity thatis a constant of the motion is the projection of µ ontoj,

µj ≡ µ · uj = −µB[ge,lj · uj + (ge,s − ge,l)s · uj] = −µBgjj · uj, (A.33)

where uj =j/|j| is a unit vector in thej direction. With the aid of Eq. (E.23), itis possible to show that the matrix element of µj in the vector |γlsjmj〉 is

〈µj〉 = −µB

[ge,l

√l (l + 1)

j (j + 1) + l (l + 1) − s (s + 1)2√

j (j + 1)√

l (l + 1)

+ge,s

√s (s + 1)

j (j + 1) + s (s + 1) − l (l + 1)2√

j (j + 1)√

s (s + 1)

]. (A.34)

Exploiting Eqs. (E.26) and (E.27) allows us to write this as

〈µj〉 = −µB

[ge,l〈cos(l,j)〉

√l(l + 1) + ge,s〈cos(s,j)〉

√s(s + 1)

]. (A.35)

This can be written as

〈µj〉 = −µB gj

√j(j + 1), (A.36)

2) We point out that caution is required when comparing equationsinvolving the g-factor of the electron, as some authors define thisquantity as being negative. See also, footnote 5 on p. 60.

Page 283: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

266 A Atomic Physics Definitions

from which we define µj as

µj = −µBgjj. (A.37)

It is straightforward to show that gj is given by

gj =32

+12

[s (s + 1)− l (l + 1)

j (j + 1)

]. (A.38)

Before we leave this section, a few words about the anomalous value ofge,s are in order. Both experiment and quantum electrodynamics require thespin g factor to be anomalous, g = 2(1 + a), where a > 0 is called the g-factor anomaly. The best measured values of the anomaly for free electronsand positrons are

ae− = 1.159621884(43)× 10−3, (A.39)

ae+ = 1.159621879(43)× 10−3. (A.40)

These results were obtained by comparing the spin-precession and cyclotronfrequencies of isolated charges in a Penning trap, which contains both elec-tric and magnetic fields [118]. Theoretically, Schwinger showed that the firstcorrection term contributing to the anomaly is 0.5α/π [119]. Additional termsof higher powers of α/π contribute to a as well. Hughes and Kinoshita, forexample, gave [120]

ae =12

( α

π

)− 0.328478965

( α

π

)2+ 1.181241456

( α

π

)3

− 1.5098( α

π

)4+ 4.393 × 10−12. (A.41)

The current best theoretical value is 1.159652201(27)× 10−3. It is important tonote that the coefficient of the second term in Eq. (A.41) is given erroneouslyas 2.973 in Bethe and Salpeter.3 The best accepted value for the anomalousspin g factor is

ge,s = 2.002 319 304 3718(75). (A.42)

Recently, Odom et al.4 reported a new value ge,s/2 = 1.001 159 652 180 85(76).

A.13.2Proton Magnetic Moment

Since the proton is a spin-1/2 particle with positive charge we would expectgyromagnetic ratios of gp,l = 1 and gp,s = 2. Although the orbital portion is

3) See p. 93, Eq. (18.5) of Ref. [3].4) B. Odom, D. Hanneke, B. D’Urso and G. Gabrielse Phys. Rev. Lett.

97, 030801 (2006).

Page 284: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

A.14 Quadrupole Moment of the Nucleus 267

correct, experimentally gp,s = 5.585 694 701(56). One must remember that theproton is not a simple point charge. The magnetic moment of the proton thenbecomes

µp = µN

(l + gp,ss

), (A.43)

where µN is the nuclear magneton and has the value

µN ≡ eh2mp

= µBme

mp= 3.152 451 259(21)× 10−8 eV/T. (A.44)

A.13.3Neutron Magnetic Moment

Even though the neutron has no charge, its magnetic moment is not zero.Its contribution from orbital motion is zero but its spin contribution leads togn,s = 3.826 085, 46(90) and µn = −gn,sµNs.

A.13.4Magnetic Moment of the Nucleus

It is convenient to write a g factor for the entire nucleus parallel to the totalnuclear spin I as

µI = −gIµNI.

The value of gI depends on how the spins are coupled.

A.14Quadrupole Moment of the Nucleus

In electrodynamics the quadrupole moment tensor is given by

Qij =∫

ρ(

3rirj − δijr2)

d3r, (A.45)

with units of C m2. It is convenient to classify the strength of this momentquantum mechanically by the expectation value of the Qzz component in astate |γIM〉,

Q ≡ 〈γI I|Qzz|γI I〉. (A.46)

Converting to spherical tensors, where Qzz = 2Q2,0, we have

Q = 2〈γI||Q2||γI〉(

I 2 I−I 0 I

). (A.47)

Page 285: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

268 A Atomic Physics Definitions

The reduced matrix element is evaluated with the aid of the Wigner–Eckarttheorem, Eq. (E.14). Values for Q/e (i.e., in units of cm2) for a few nuclei canbe found in Ref. [27].

A.15Frequently Used AMO Quantities

Many of the quantities given in this appendix are used frequently in calcula-tions in AMO5 science. For quick and dirty estimations, the reader will find itconvenient to commit several of the following to memory:

Quantity Value

Atomic momentum, po = h/ao 1.2438× 10−7 eV s/cm1 Bohr, ao 0.0529 nmDistance light travels in 1 ns 1 footElectron mass, me 0.511 MeVFine-structure constant, α 1/137Refractive index of glass (visible light) 1.5Proton mass, mp 937 MeVSpeed of light in vacuum, c 3 × 108 m/s

≡ 1/√

εoµo

Vacuum permeability, µo ≡ 4π × 10−7 N/A2

kBT at room temperature 1/40 eVhc 197 eV nm

Further Reading

The National Institute of Standards and Technology maintains a comprehen-sive web source.

1. For the most up to date values, seehttp://physics.nist.gov/cuu/Constants/.

2. For a detailed discussion of units and uncertainties, seehttp://physics.nist.gov/cuu/.

5) Atomic, Molecular and Optical

Page 286: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

269

Appendix BMathematics Related to AMO Calculations

B.1Kronecker Delta, δij

δij =

1, i = j,0, i = j,

(B.1)

where

∑i

fiδij = fj (B.2)

and

∑i

δij = 1. (B.3)

B.2Dirac Delta Function, δ (x − xo)

The formal definition of the delta function introduced by Dirac is∫ ∞

−∞dx f (x) δ (x − xo) = f (xo) , (B.4)

where∫ ∞

−∞dx δ (x − xo) = 1 (B.5)

and

δ (x − xo) =

∞, x = xo,0, x = xo.

(B.6)

A very useful representation of the delta function is

δ (x − xo) =1

∫ ∞

−∞dkei(x−xo)k. (B.7)

Page 287: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

270 B Mathematics Related to AMO Calculations

However, it must be understood that the left-hand side of Eq. (B.7) is onlywell defined under the integral sign. There are several approximations fordelta functions. One particularly useful one is

δ(x − xo) =1

2πlim

n→∞

∫ n/2

−n/2dkei(x−xo)k. (B.8)

The nth derivative of the delta function, (dn/dxn)δ (x − xo), leads to∫ ∞

−∞dx f (x)

dn

dxn δ (x − xo) = (−1)n dn

dxn f (x)∣∣∣∣x=xo

, (B.9)

while

(x − xo)d

dxδ (x − xo) = δ (x − xo) . (B.10)

B.3Hypergeometric Series

The hypergeometric functions, which are sometimes called Gauss functions,are defined by1

F(α, β, γ; x) = 1 +αβ

γ

x1!

+α(α + 1)β(β + 1)

γ(γ + 1)x2

2!+ · · · , γ = 0,−1,−2, . . .

=Γ(γ)

Γ(α)Γ(β)

∑k=0

Γ(α + k)Γ(β + k)Γ(γ + k)

xk

k!. (B.11)

The hypergeometric functions obey

F(α, β, γ; x) = (1 − x)−γF(

α, γ − β, γ;−x

1 − x

), (B.12)

and are related to the Legendre and associated Legendre functions (see Sec-tion B.6):

Pn(x) = F(−n, n + 1, 1;

1 − x2

), (B.13)

Pmn (x) =

(n + m)!(n − m)!

(1 − x2)m/2

2mm!

× F(

m − n, m + n + 1, m + 1;1 − x

2

). (B.14)

1) For integer n, Γ(n + 1) = n!.

Page 288: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.4 Confluent Hypergeometric Series 271

B.4Confluent Hypergeometric Series

In the limit as β → ∞ in Eq. (B.11), the confluent hypergeometric functions,which are also known as Kummer functions, are defined as

F (α, γ; x) = 1 +α

γ

x1!

+α(α + 1)γ(γ + 1)

x2

2!+ · · · , γ = 0,−1,−2, . . .

=Γ(γ)Γ(α)

∑k=0

Γ(α + k)Γ(γ + k)

xk

k!. (B.15)

The confluent hypergeometric functions obey

F(α, γ; x) = exF(γ − α, γ;−x), (B.16)

and have an asymptotic behavior for large x > 0 given by

F (α, γ; x) → Γ (γ)Γ (α)

exxα−γ. (B.17)

B.5Associated Laguerre Polynomials

The Laguerre polynomials are defined as

Lmn (x) = ex x−m

n!dn

dxn e−xxn+m

=n

∑k=0

Γ (n + m + 1)Γ (k + m + 1)

(−x)k

k! (n − k)!, (B.18)

and are associated with the confluent hypergeometric functions

L2l+1n−l−1(x) =

(n + l)!(n − l − 1)!(2l + 1)!

F(l + 1 − n, 2l + 2; x). (B.19)

B.6Legendre and Associated Legendre Functions

The Legendre functions Pl(x) are the solutions to Legendre’s equation,(1 − x2

) d2Pl

dx2 − 2xdPl

dx+ l(l + 1)Pl = 0, (B.20)

Page 289: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

272 B Mathematics Related to AMO Calculations

which are finite at the two singular points, x = ±1.2 The first three Legendrefunctions are

P0 (x) = 1, P1 (x) = x and P2 (x) = 12

(3x2 − 1

), (B.21)

and can be generated from Rodrigues’ formula,

Pl =1

2l l!dl

dxl

(xl − 1

)l. (B.22)

The Legendre functions obey∫ +1

−1Pl(x)Pl ′(x)dx =

22n + 1

δll ′ . (B.23)

Thus, any well-behaved function, f (x), can be expanded in terms of them,

f (x) =∞

∑l=0

al Pl(x). (B.24)

Finally, the parity of the functions depends on the index, l,

Pl(−x) = (−1)lPl(x). (B.25)

The associated Legendre functions, Pml (x),3 are solutions to the associated

Legendre equation(1 − x2

) d2Pml

dx2 − 2xdPl

dx+

[l(l + 1)− m2

1 − x2

]Pm

l = 0. (B.26)

It is straightforward, via direct substitution, to show that Pml (x) and Pl are

related,

Pml (x) =

(1 − x2

)m/2 dm

dxm Pl(x). (B.27)

The associated Legendre functions obey

P−ml (x) = (−1)m (l − m)!

(l + m)!Pm

l (x), (B.28)

Pml (−x) = (−1)l+mPm

l (x) (B.29)

and

P0l (x) = Pl(x). (B.30)

2) The Pl(x) functions are solutions of the first kind. Solutions of thesecond kind, Ql(z), are discussed, for example, in Refs. [1] and [121].

3) See Ref. [121] for a discussion of solutions that are singular atx = ±1.

Page 290: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.7 Spherical Harmonics 273

Thus, m can be positive or negative with a range −l m l. A more usefulexpression for Pm

l (x) is then

Pml (x) =

12l l!

(1 − x2

)m/2 dl+m

dxl+m

(x2 − 1

)l. (B.31)

Two associated Legendre functions with the same m obey the following or-thogonality property:∫

Pml (x)Pm

l ′ (x)dx =(l + m)!(l − m)!

22l + 1

δll ′ . (B.32)

Finally, the first three associated Legendre functions for m = 0 are

P11 (x) =

(1 − x2)1/2 = sin ϑ,

P12 (x) = 3x

(1 − x2)1/2 = 3 cos ϑ sin ϑ,

P22 (x) = 3

(1 − x2) = 3 sin2 ϑ,

(B.33)

where in the last column it is assumed that x = cos ϑ.

B.7Spherical Harmonics

It is customary to combine the associated Legendre functions of the previoussection with the azimuthal phase, eimϕ, to define the spherical harmonics,

Ylm (ϑ, ϕ) = (−1)(m+|m|)/2

√(2l + 1)

(l − |m|)!(l + |m|)!

P|m|l (cos ϑ) eimϕ, (B.34)

for all positive and negative values of m with Pml (cos ϑ) given by Eq. (B.27).

The spherical harmonics can also be written as

Ylm (ϑ, ϕ) = (−1)m

√(2l + 1)

(l − m)!(l + m)!

Pml (cos ϑ) eimϕ, (B.35)

with Pml (cos ϑ) given by Eq. (B.31). In either case,

Yl,−m (ϑ, ϕ) = (−1)mY∗lm (ϑ, ϕ) (B.36)

and ∫Y∗

lm (ϑ, ϕ) Yl ′m′ (ϑ, ϕ) sin ϑdϑdφ = δll ′δmm′ . (B.37)

Page 291: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

274 B Mathematics Related to AMO Calculations

B.8Mathematical Formalism of Quantum Mechanics

Extracting measurable quantities in quantum mechanics requires a mathemat-ical formalism that allows us to track the state of a system and to perform var-ious calculations. This formalism, introduced by Dirac in the late 1920s, anddiscussed in detail in his book The Principles of Quantum Mechanics [122] as wellas elsewhere (see for example Condon and Odabasi [123]), involves linear al-gebra, abstract vectors and an infinite-dimensional complex vector space.4 Inthis appendix we will outline some of the key principles through worked ex-amples that are used in the main body of the book. We end the appendix witha summary of the main issues of the algebra necessary for quantum mechan-ics.

The complex vector space, which we will call the system space, includes lin-ear operators, vectors and inner products. To distinguish vectors in physicalspace, symbolized by r or r, from the state vectors in the system space, wesymbolize the latter by |S〉; Dirac called these kets. Every state vector in thisabstract vector space has a so-called dual, 〈S|, which Dirac called a bra andexists in a corresponding vector space. The bras and kets are related by theirHermitian adjoint,

〈S| ≡ |S〉†, (B.38)

where the Hermitian adjoint of the state vector is the complex conjugate of itstranspose. Just as a vector is used to describe the state of a point in physicalspace – its location in space – the state vectors describe the state of the systemin the system space, for a one-electron atom this might include the values forn, l, ml , etc.

To write an expression for a vector requires a complete basis – a set of lin-early independent vectors. For example, in three-dimensional Cartesian spacethe basis set is the three unit vectors x, y, z. These vectors are linearly inde-pendent because no linear combination of any two will give the third. Theset is complete because any vector in the space can be expressed by a linearcombination of just these three vectors,

V = Vxx + Vyy + Vzz, (B.39)

where Vx, Vy and Vz are real numbers. It is often convenient to let x → x1,y → x2 and z → x3 and Vx → V1, Vy → V2 and Vz → V3, so that the vector is

4) This complex vector space is often referredto in quantum mechanics texts as a Hilbertspace [124, 125]. A discussion of a Hilbertspace is beyond the scope of this text butwe do point out that technically, as Diracmentioned, the abstract space of quantum

mechanics is more general than the stan-dard Hilbert space. Extensions, such as therigged Hilbert space, have been shown toinclude continuous spectra, for example,which are not included in the standardHilbert space.

Page 292: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.8 Mathematical Formalism of Quantum Mechanics 275

expressed as5

V =3

∑i=1

Vixi. (B.40)

In the system space, an arbitrary state vector is similarly given by a linearcombination of the basis states (vectors),

|S〉 =∫∑

isi|σi〉, (B.41)

where si are complex numbers and |σi〉 are the basis vectors. A similar ex-pression exists for a state vector 〈S| involving bras. The compound symbolcomposed of a summation sign with the superimposed integral sign means tosum over discrete parameters and integrate over continuous parameters. Forexample, the energy of the bound electron in the hydrogen atom is discreteand given by the value of n. However, when the atom is ionized, the energyof the electron is determined by a continuous parameter. The complete state,then, must span both the bound and continuous parts of the spectrum becausethese are all the possible states of the electron. In this text we will not use thiscompound symbol. Rather, we will apply the sum to discrete parameters andan integration to continuous parameters. For the general case like Eq. (B.41)that could include both types of parameters, we will simply write a summa-tion sign: but it should be understood that this will mean that we sum overthe discrete parameters and integrate over the continuous parameters.

The inner product in this algebra is defined between a bra and a ket. Thus,the inner product between |A〉 and |B〉 is

〈A|B〉. (B.42)

The inner product in this complex vector space, like in physical space, gener-ates a number, albeit complex in this case, so that

〈A|B〉 = 〈B|A〉∗. (B.43)

Below, we will see that such an inner product can also correspond to an am-plitude of the wavefunction in physical space. It follows that 〈A|A〉 is a realnumber and is called the norm of |A〉. Setting 〈A|A〉 = 1 is convenient be-cause this means that the state vector is of unit length. However, the statevector is not fully specified when normalized because it can still be multipliedby a phase, eiφ, without changing its length,(

〈A|e−iφ) (

eiφ|A〉)

= 〈A|A〉. (B.44)

5) Einstein introduced the more compact convention of summing overrepeated indices, ∑3

i=1 Vixi → Vixi , thus allowing the summationsymbol to be dropped.

Page 293: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

276 B Mathematics Related to AMO Calculations

Thus, only the direction of the state vector is important, not its length! When〈A|B〉 = 0 (= 〈B|A〉), the two states are said to be orthogonal. The basis vectorsin this space are orthogonal and normalized to 1, so that

〈σi|σj〉 = δij. (B.45)

It is straightforward to see that the inner product is tantamount to ma-trix multiplication of a row and a column vector of the coefficients when|A〉 and |B〉 are expressed in terms of an orthonormal set of basis vectors,|A〉 = a1|σ1〉+ a2|σ2〉+ · · ·+ an|σn〉 and |B〉 = b1|σ1〉+ b2|σ2〉+ · · ·+ bn|σn〉.6In this case, the inner product can be written as

〈A|B〉 =(

a∗1 a∗2 · · · a∗n)

b1b2...

bn

=n

∑i=1

a∗i bi, (B.46)

where Eq. (B.45) has been used. If |σi〉 represents a continuous parameter, thesum would be an integration.

The inner product is not the only possible product between state vectors;|A〉|B〉 or 〈A|〈B|, for example, defines a separable state in the spirit of the ra-dial and angular parts of the wavefunctions (see Eq. (1.21)) and can be writtenas |AB〉 or 〈AB|. The product |A〉〈B|, called the outer product, is an operatorbecause when it acts on a bra or a ket, it changes the state of the system. Forexample, when |A〉〈B| is applied to |C〉,

(|A〉〈B|) |C〉 = |A〉 · 〈B|C〉, (B.47)

it changes the state to |A〉 since 〈B|C〉 is just a complex number. The specialouter product

|A〉〈A| (B.48)

is known as the projection operator, because it will extract the contribution of|A〉 from an arbitrary state vector. When a projection operator is formed fromthe basis vectors, |σi〉, it obeys a property called completeness, where7

∑all states

|σi〉〈σi| = 1. (B.49)

6) The Gram–Schmidt procedure can be used to construct an orthonor-mal set of basis vectors from an arbitrary initial set (see Ref. [121]).

7) When |σi〉 represents a discrete parameter then this would be anactual sum. If, however, it were continuous then the sum wouldreally be an integral. Clearly, if |σi〉 were a mixture we would haveto sum and integrate over the appropriate parts.

Page 294: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.8 Mathematical Formalism of Quantum Mechanics 277

Now we are ready to discuss the extraction of physics from this abstractvector space. Of chief concern is the measurement of physical quantities andexplicit representations of operators and wavefunctions. Measurable quan-tities, dynamical parameters like position, energy, quadrupole moment, etc.,are called observables and exist in physical space. These quantities, which arereal numbers and variables in classical mechanics, become tensor operators inquantum mechanics because they can modify the state of the system, in gen-eral. If the system were in an eigenstate of the operator, however, the operatorsimply extracts the eigenvalue, the observable, while leaving the system inthe same state. Consequently, in quantum mechanics the extraction of observ-ables can be reduced to eigenvalue problems. What makes quantum physicsso interesting is the fact that not all eigenvalues are extractable simultaneously.

To begin, we will look at an example of a simple eigenvalue problem in theabstract space. Consider an operator A acting on a state |α〉 with eigenvaluea. The relationship between the operator, eigenstate and eigenvalue can bewritten as8

A|α〉 = a|α〉. (B.50)

Applying 〈α| to the left, defines the expectation value of A

〈α|A|α〉 = a〈α|α〉 = a, (B.51)

where we have assumed that |α〉 is normalized and we used the fact that a,a number, commutes with the state vector. The expectation value, a, can bethought of as the average value that can be measured by experiment. Thisaverage is in the statistical sense of repeated measurement. When there areseveral eigenvalues of the operator, the problem can be solved with matrices.In general, we can define a matrix element,

〈α2|A|α1〉 = A21

= a2〈α2|α1〉= a2δ21. (B.52)

In the second step we used the facts that 〈α2|A = a2 and that operators as-sociated with the physical observables are Hermitian – eigenvalues are real.If α were to represent a continuous observable, the Kronecker delta would

8) We will use a hat over a variable to indi-cate an operator. Quantities without hatswill simply be complex numbers, whichin some cases will be purely real. John vonNeumann is given credit for inventingoperator algebra in his 1932 book Mathe-

matische Grundlagen der Quantenmechanik(Mathematical Foundations of Quantum Me-chanics). His work allowed the Schrödingerand Heisenberg approaches to quantummechanics to be shown to be mathemati-cally equivalent.

Page 295: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

278 B Mathematics Related to AMO Calculations

have to be replaced by the Dirac delta function (Appendix B.7). Since |α〉 isan eigenstate of A, only diagonal terms exist. In general, for an arbitrary statevector off-diagonal elements will exist as well. The off-diagonal elements areassociated with transitions between eigenstates. A state of the system can alsobe an eigenstate of two or more observables simultaneously. In this case, theoperators commute.

To make this more concrete, let us apply this to the motion of a free, spinlesspoint particle. Observables that we can extract include position (x), momen-tum (p) and energy (E). The corresponding operators are x, p and H = p2/2m.For this problem there are two convenient sets of basis vectors, |x1x2x3〉 ≡ |x〉and |p1 p2 p3〉 ≡ |p〉,9 that can be used to describe the state of the particle. Ingeneral, the explicit expression for an operator depends on the basis. To seethis, we will find the expectation value of the energy operator in both bases.In the momentum basis, for example, H is diagonal since it can operate on |p〉without changing the state of the system:

〈p′|H|p〉 =1

2m〈p′|p2|p〉

=1

2m

∫d3 p′′〈p′|p|p′′〉 · 〈p′′|p|p〉 (B.53)

=1

2m

∫d3 p′′p′′ · pδ(p′ − p′′)δ(p′′ − p)

=p2

2mδ(p′ − p). (B.54)

Since the momentum of a free particle is a continuous variable, the Kroneckerdelta in Eq. (B.52) has been replaced by a delta function. The delta functionmeans that H indeed does not alter the state of the system so that p′ = p. Thisexpectation value is precisely what we expected.

Now we will do the problem in the position basis. The problem is a bit moreinvolved because these basis vectors are not eigenstates of H and so |x〉 will bemodified when H is applied. To solve this problem we can use completenesstwice to surround H by its eigenstates,

〈x′|H|x〉 =∫∫

d3 p′d3 p〈x′|p′〉〈p′|H|p〉〈p|x〉

=∫∫

d3 p′d3 p(

p′2

2mδ(p′ − p)

)〈x′|p′〉〈p|x〉

=∫

d3 p(

p2

2m

)〈x′|p〉〈p|x〉.

To go further, we must digress and discuss the meaning of 〈x′|p〉 and 〈p|x〉.From Eq. (B.43) we know that both of these are complex numbers or an am-

9) It is typical to use just one symbol to represent several states of thesystem that are of the same nature.

Page 296: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.8 Mathematical Formalism of Quantum Mechanics 279

plitude. But, they are more than this. Consider the arbitrary state |ψ〉 andthe basis states |x〉. The set of complex numbers 〈x|ψ〉, which are essentiallythe coefficients in Eq. (B.41), can be thought of as the projection of the stateinto the position representation. These coefficients provide a way to obtain anexplicit representation of the state in physical space. That is,

〈x|ψ〉 = ψ(x), (B.55)

where ψ(x) is the representation of the state in physical space. We then rec-ognize 〈x′|p〉 as the physical space wavefunction of the particle. For a freeparticle the Schrödinger solution is a plane wave,

ψ(x) = 〈x|p〉 = eip·x/h/√

2π.

It follows then that the matrix element of H in the position basis is

〈x′|H|x〉 =1

∫d3 peip·(x′−x)/h

(p2

2m

). (B.56)

The expectation value of the energy operator in position eigenstates yields theFourier transform of the momentum space solution and takes the system from|x〉 to |x′〉.

Before we summarize, we point out that it is possible to determine the rep-resentation of p in the position representation if we go back to Eq. (B.53) andinsert a complete set of position basis states in 〈p′|p|p′′〉. This leads to

12m

∫d3 p′′〈p′|p|p′′〉 ·

∫d3x〈p′′|p|x〉〈x|p〉,

12m

∫d3 p′′〈p′|p|p′′〉 ·

∫d3x〈p′′|p

(eip·x/h/

√2π

)|x〉.

If we let p → −ih∇, perform the differentiation and write 〈p′′|x〉 in terms ofthe plane wave again, we will have

p2m

· 〈p′|p|p〉.

Repeating this process with the remaining matrix element leads to Eq. (B.54).Consequently, p takes a different form in different representations.

B.8.1Summary

Here is a summary of some rules associated with the algebra of bras andkets used in quantum mechanical calculations. Most of these can be foundin Dirac’s text [122] and many are found elsewhere such as in Condon andOdabasi [123], and Messiah [4] as well as standard linear algebra texts. Thereader is encouraged to read these texts for a more thorough understanding.

Page 297: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

280 B Mathematics Related to AMO Calculations

1. Any linear operator can be represented by a matrix.

2. The unit operator is represented by the unit matrix.

3. A real linear operator is represented by a Hermitian matrix, A = A†.

4. Eigenvalues are represented by diagonal matrices.

5. The matrix representing the product of two linear operators is the prod-uct of the matrices representing the two factors.

6. A state can be an eigenstate of two observables (operators) simultane-ously, if the two observables commute.

7. A linear operator that commutes with an observable also commutes withany function of the observable.

B.9Schrödinger’s Equation in Parabolic Coordinates

In this appendix we define parabolic coordinates and use them to express theSchrödinger equation along with its solutions for the hydrogen atom. Thethree parabolic coordinates, ξ, η and φ, are related to the more familiar spher-ical and rectangular coordinates through

ξ = r(1 + cos θ),

η = r(1 − cos θ),

φ = φ,

x =√

ξη cos φ,

y =√

ξη sin φ and

z = (ξ − η)/2.

The coordinate surfaces for ξ and η are paraboloids with their foci at the ori-gin. Those for constant ξ open toward z → −∞ while those for constant η

open toward z → ∞; both take values from 0 to ∞. The coordinate surfacesfor φ are the meridian planes and range, as usual, from 0 to 2π. Figure B.1shows a cross section of the coordinate system in the yz-plane.

The length in spherical coordinates is related to that in parabolic coordinatesthrough

r = (ξ + η)/2,

while the differential volume element is given by

dV =14(ξ + η)dξdηdφ.

Page 298: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.9 Schrödinger’s Equation in Parabolic Coordinates 281

Fig. B.1 A cross section of the parabolic coordinate system. The coor-dinate surfaces for η and ξ are paraboloids with the z-axis as the axisof rotation. The surfaces for φ are the meridian planes, one of which isthe plane of the page.

The wave equation can be written for E < 0 as

1ξ + η

[∂

∂ξ

∂ξψ(ξ, η, φ)

)+

∂η

∂ηψ(ξ, η, φ)

)]+

1ξη

∂2

∂φ2

+2µ

h2

(E +

2ξ + η

Ze2

4πεo

)ψ(ξ, η, φ) = 0. (B.57)

The wavefunction is separable, so that ψ(ξ, η, φ) = f (ξ)g(η)Φ(φ). As withspherical coordinates, we can let

Φ(φ) =1√2

eimφ, where m = 0,±1,±2, . . . .

If we let ν be the separation constant, f and g are solutions to

∂ξ

∂ f∂ξ

)−

(m2

4ξ+

µ|E|ξ2h2 − µ

2h2Ze2

4πεo+ ν

)f = 0, (B.58)

∂η

∂g∂η

)−

(m2

4η+

µ|E|η2h2 − ν

)g = 0, (B.59)

respectively. We can simplify these equations if we let

a2 =µ|E|2h2 ,

b1 =1a

h2Ze2

4πεo− ν

), (B.60)

b2 =ν

a.

Page 299: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

282 B Mathematics Related to AMO Calculations

If we let ρ = aξ in Eq. (B.58) and ρ = aη in Eq. (B.59) both equations take thesame form:

∂ρ

∂u∂ρ

)+

(bρ− 1

4− m2

4ρ2

)u = 0, (B.61)

where b = b1(or b2) when u = f (or g).The approach we can take to solve Eq. (B.61) is the same we took for the

hydrogen atom problem in spherical coordinates in Section 1.2.2. Namely, forlarge ρ,

u(ρ) e−ρ/2.

Seeking a power-series solution that must be truncated for m > 0 leads to

u(ρ) = e−ρ/2ρm/2U(ρ).

The equation to be solved becomes

ρd2Udρ2 + (m + 1 − ρ)

dUdρ

+[

b − 12(m + 1)

]u = 0. (B.62)

The solutions can be written in terms of associated Laguerre polynomials orconfluent hypergeometric functions as we had before:

Lmn+m(ρ), where n = b − m + 1

2.

Since m can be positive or negative, our solutions become

f (ρ) = e−ρ/2ρ|m|/2L|m|n1+|m|(ρ), n1 = b1 − |m|+ 1

2, (B.63)

g(ρ) = e−ρ/2ρ|m|/2L|m|n2+|m|(ρ), n2 = b2 − |m|+ 1

2, (B.64)

respectively, where both n1 and n2 must be positive integers or zero. Theprincipal quantum number from our spherical solution is given by

n = b1 + b2 = n1 + n2 + |m|+ 1 =µ

h2a

Ze2

4πεo,

which leads to

|En| =12

(Ze2

4πεo

)2 4µ

h2n2,

as it must.Each stationary state of the discrete spectrum (E < 0) is determined by the

three integers in parabolic coordinates, the parabolic quantum numbers n1, n2

Page 300: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

B.10 Voigt Line Profile 283

and m. For a given n, |m| can range from 0 to n − 1. For a given n and |m|,n1 and n2 can be chosen in n − |m| ways such that n1 + n2 = n − |m| − 1 – n1and n2 sort of play the role of l.10 As in spherical coordinates, the energy is n2

degenerate.The total wavefunction is given by

ψn1n2m(ξ, η, φ) = Ce−(ξ+η)/2an1n2m(ξη)|m|/2eimφ

× L|m|n1+|m|(ξ/an1n2m)L|m|

n2+|m|(η/an1n2m), (B.65)

where

an1n2m =4πεo

Ze2h2(n1 + n2 + |m|+ 1)

µ.

If we look at the ground state we will immediately see how Eq. (B.65)is related to R10 of Eq. (1.45). The ground state in parabolic coordinates,n1 = n2 = m = 0, takes the form

ψ000 ∝ e−(ξ+η)/2a000.

In spherical coordinates, n = 1 and l = m = 0, leading to

ψ100 ∝ e−r/ao .

Clearly, these are the same if we recognize a000 as the Bohr radius. In this case,the charge distribution is symmetric. For other choices, the charge distributionis asymmetric with respect to a plane through z = 0. For example, whenn1 > n2 the distribution is skewed to positive z.

B.10Voigt Line Profile

In cases were the Gaussian and Lorentzian widths of the line are of the sameorder of magnitude, we must treat them on an equal footing. The appropriateline shape, S(ω, v), is given by

S(ω, v) =2Γo

(ωo − ω + vω/c)2 + Γ2o

. (B.66)

To determine the line shape as a function of ω, S(ω), we must integrate overthe normalized velocity distribution, f (v).

S(ω) =∫ ∞

−∞dvS(ω, v) f (v). (B.67)

10) In spherical coordinates we had |m| ≤ l < n − 1.

Page 301: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

284 Further Reading

For atoms or molecules in thermal equilibrium with a temperature bath,

f (v) = (M/2πRT)1/2 e−Mv2/2RT,

where R is the gas constant (Boltzmann constant per mole). SubstitutingS(ω, v) and f (v) into Eq. (B.67) and changing variables, we can write the lineshape in the well-known Voigt profile form as

S(ω) =1

π3/2b2

Γo

∫ ∞

−∞

dye−y2

(x + y)2 + b2 , (B.68)

where

x =ωo − ω

ΓD

√4 ln 2,

b =Γo

ΓD

√4 ln 2

and ΓD is the Doppler width. When nonthermal velocity distributions aremore appropriate it will be necessary to redo the integrations to find a newexpression for S(ω).

Further Reading

1 G. Arfken and H. Weber, Mathematical Meth-ods for Physicists, Academic, New York, NY,2000.

2 J. D. Jackson, Classical Electrodynamics, Wiley,New York, NY, 1999.

3 J. Mathews and R. L. Walker, MathematicalMethods of Physics, Benjamin Cummings, SanFrancisco, CA, 1970.

4 N. N. Lebedev, Special Functions and TheirApplications, Dover, New York, NY, 1972.

5 P. A. M. Dirac, The Principles of Quantum Me-chanics, Oxford University Press, Oxford,UK, 1981.

6 E. U. Condon and H. Odabasi, Atomic Struc-ture, Cambridge University Press, New York,NY, 1980.

Page 302: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

285

Appendix CAtomic and Molecular Data

C.1NIST Online Data

The National Institute of Standards and Technology (NIST) maintains a veryvaluable online compilation of atomic and molecular data. The general link tothis portal is

http://physics.nist.gov/products.html.

Below, we give the links to some specific key information.

C.1.1NIST Online Atomic Data

The link to the Physical Reference Data page is

http://physics.nist.gov/PhysRefData/contents.html.

Fundamental constants can be obtained from

http://physics.nist.gov/PhysRefData/contents-constants.html.

C.1.2NSIT Online Molecular Data

Chemical data, including constants for diatomic molecules, can be accessedthrough

http://webbook.nist.gov/.

C.2Molecular Constants

Another source for constants for diatomic molecules is the compilation by Hu-ber and Herzberg [45].

Page 303: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

286 C Atomic and Molecular Data

C.3Filling Subshells

The table below shows the order in which subshells are filled. Energy in-creases down and to the left.

1s2s 2p

3s 3p 3d4s 4p 4d 4 f

5s 5p 5d6s 6p

7s

Page 304: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

C.4 Electronic Configurations 287

C.4Electronic Configurations

The following table shows the ground-state configurations for the currentlyknown elements in the periodic table.

H − He 1s − 1s2

Li − Be 1s22s − 1s22s2

B − Ne 1s22s22p − 1s22s22p6

Na − Mg [Ne]3s − [Ne]3s2

Al − Ar [Ne]3s23p − [Ne]3s23p6

K − Ca [Ar]4s − [Ar]4s2

Sc − Zn [Ar]4s23d − [Ar]4s23d10

Ga − Kr [Ar]4s23d104p − [Ar]4s23d104p6

Rb − Sr [Kr]5s − [Kr]5s2

Y − Cd [Kr]5s24d − [Kr]5s24d10

In − Xe [Kr]5s24d105p − [Kr]5s24d105p6

Cs − Ba [Xe]6s − [Xe]6s2

La [Xe]6s25dCe − Lu [Xe]6s25d4 f − [Xe]6s25d4 f 14

Hf − Hg [Xe]6s24 f 145d2 − [Xe]6s24 f 145d10

Tl − Rn [Xe]6s24 f 145d106p − [Xe]6s24 f 145d106p6

Fr − Ra [Rn]7s − [Rn]7s2

Ac [Rn]7s26dTh − Lr [Rn]7s26d5 f − [Rn]7s26d5 f 14

Rf − [Rn]7s25 f 146d2(?) −

Page 305: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics
Page 306: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

289

Appendix DCoupling Angular Momenta

D.1Two Angular Momenta and 3-j Symbols

Consider a system described by two angular momenta j1 and j2. We can de-fine all states of the system by combining two state vectors |j1m1〉 and |j2m2〉.If the states of the system these vectors represent not interact then we can writethe total state vector as

|j1m1 j2m2〉 = |j1m1〉 |j2m2〉 . (D.1)

We could equally well have coupled j1 and j2 to give a total angular momen-tum J and have written a total state vector in terms of the total angular mo-mentum as

|JM〉 = |(j1 j2)JM〉 , (D.2)

where in (j1 j2)J we coupled the two angular momenta within the parenthesesto give the angular momentum just to the right outside the parentheses. De-pending upon the problem, one representation is often more convenient thanthe other. Since either representation can be used to completely define thesystem, we can write one representation in terms of the other. Using the factthat

1 = ∑m1,m2

|j1m1 j2m2〉 〈j1m1 j2m2| , (D.3)

often called the sum over states, we can write

|JM〉 = ∑m1,m2

|j1m1 j2m2〉 〈j1m1 j2m2|JM〉 , (D.4)

which is subject to the constraint M = m1 + m2. In Eq. (D.4), 〈j1m1 j2m2|JM〉is a real number, called a Clebsch–Gordon coefficient, that can be rewritten interms of 3-j symbols:

〈j1m1 j2m2|JM〉 = (−1)−j1+j2−M √2J + 1

(j1 j2 J

m1 m2 −M

). (D.5)

Page 307: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

290 D Coupling Angular Momenta

A consequence of Eqs. (D.2), (D.4) and (D.5) allows us to write |JM〉 as

|JM〉 = (−1)j2−j1−M√

2J + 1 ∑m1,m2

|j1m1 j2m2〉(

j1 j2 Jj1 m2 −M

). (D.6)

D.2Properties of 3-j Symbols

In this section we summarize a few of the symmetry properties and sumrules of 3-j symbols that are useful in manipulating equations. Additionalproperties and numerical values for specific 3-j symbols can be found in Ed-monds [126] and Appendix C.8 of Cowan [2]. An algebraic expression for the3-j symbol can be found in Chapter 5 of Cowan [2] and Appendix 2 of [126].

The 3-j symbol is a real number and has the property that the three elementsin the upper row must add as vectors. That is, all permutations of the elementsobey j1 + j2 ≥ J. At the same time, the sum of the three elements in thelower row must equal zero. If these conditions are not met, the 3-j symbol isidentically zero. The value of the 3-j symbol is unchanged upon any cyclic oreven permutation of columns. Any odd permutation, however, or a change ofall the signs of the magnetic quantum numbers, leads to(

j1 J j2m1 −M m2

)=

(J j2 j1

−M m2 m1

)= (−1)j1+j2+J

(j1 j2 J

m1 m2 −M

)= (−1)j1+j2+J

(j1 J j2

−m1 M −m2

). (D.7)

The 3-j symbols are orthogonal and obey the sum rules

∑j,m

(2J + 1)(

j1 j2 Jm1 m2 −M

) (j1 j2 J

m′1 m′

2 −M

)= δm1,m′

1δm2,m′

2(D.8)

and

∑m1,m2

(j1 j2 J

m1 m2 −M

) (j1 j2 J′

m1 m2 −M′)

=δJ,J′δM,M′

(2J + 1). (D.9)

When j1 = j2 = j, m1 = −m2 = m and M = 0, we have

∑m

(−1)j−m(

j j Jm −m 0

)= δJ,0

√(2j + 1). (D.10)

Page 308: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

D.3 Three Angular Momenta and 6-j Symbols 291

D.3Three Angular Momenta and 6-j Symbols

Given three angular momenta j1, j2 and j3, there are three ways to couple themto give a total angular momentum of J:

|(j1 j2) j12, j3; JM〉 ,

|(j1 j3) j13, j2; JM〉 , (D.11)

|j1, (j2 j3) j23; JM〉 ,

where we couple the two angular momenta outside the parentheses to give J.Applying a procedure similar to that used to couple two angular momenta, wecan write each of these in terms of the state vector |j1m1 j2m2 j3m3〉. Recallingthat this state vector is separable, we can insert a sum over states twice tocouple the angular momenta pairwise. This gives for the first entry

|(j1 j2) j12, j3; JM〉 (D.12)

= ∑m1,m2,m12,m3

|j1m1 j2m2 j3m3〉 〈j1m1 j2m2| j12m12〉 〈j12m12 j3m3|JM〉 .

In Eq. (D.12), we have two Clebsch–Gordon coefficients, 〈j1m1 j2m2| j12m12〉and 〈j12m12 j3m3|JM〉, that can be evaluated with Eq. (D.5). A similar re-lationship holds for the other two coupling schemes. In general, we will beinterested in the transformation between two of these coupling schemes. Thatis, we want to evaluate matrix elements of the type

〈(j1 j2) j12, j3; JM|j1, (j2 j3) j23; JM〉 ,

that represents the projection of one coupling scheme onto another. This ma-trix element is just a number that can be evaluated in terms of 6-j symbols.The projection of the first scheme onto the third leads to

〈(j1 j2) j12, j3; J|j1, (j2 j3) j23; J〉 (D.13)

= (−1)j1+j2+j3+J√

(2j12 + 1) (2j23 + 1)

j1 j2 j12j3 J j23

,

where we have dropped M from the left-hand side because the matrix elementdoes not depend on M. The property of the 6-j symbols is that the followingtriples add as vectors: (j1, j2, j12), (j1, J, j23), (j2, j3, j23) and (j12, j3, J); other-wise, the 6-j symbol is identically zero. Numerical values for the 6-j symbolsare tabulated in Appendix D of Cowan [2] while an analytical expression isgiven in Chapter 5 of Cowan [2] and Appendix 2 of [126]. It is also possible to

Page 309: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

292 D Coupling Angular Momenta

evaluate the 6-j symbols in terms of a sum of the products of four 3-j symbols:j1 j2 j12j3 J j23

= (2J + 1) ∑

m1,m2,m12,m23,M

(−1)κ(

j1 j2 j12m1 m2 m12

)

×(

j1 J j23−m1 M −m23

) (j3 j2 j23

−m3 −m2 m23

(j3 J j12

m3 −M −m12

), (D.14)

where κ is the scalar sum κ = j1 + j2 + j12 + j3 + j23 + J.

D.4Four Angular Momenta and 9-j Symbols

Coupling four angular momenta is quite important because it allows us totransform between LS and jj coupling schemes. The general matrix elementof concern has the form

〈(j1 j2) j12, (j3 j4) j34; J| (j1 j3) j13, (j2 j4) j24; J〉 (D.15)

=√

(2j12 + 1) (2j34 + 1) (2j13 + 1) (2j24 + 1)

j1 j2 j12j3 j4 j34j13 j24 J

.

The 9-j symbol can be rewritten in terms of a sum over 6-j symbols for evalu-ation. Specifically,

j1 j2 j12j3 j4 j34j13 j24 J

(D.16)

= ∑k

(−1)2k (2k + 1)

j1 j3 j13j24 J k

j2 j4 j24j3 k j34

j12 j34 Jk j1 j2

.

Page 310: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

293

Appendix ETensor Algebra

In this appendix we will present the basic properties associated with sphericaltensors. This is not intended to be a tutorial on the subject. The reader isdirected to the references at the end of this appendix for a more detailed study.

E.1Spherical Tensors

A three-dimensional Cartesian vector, V, has components(Vx, Vy, Vz

), which

can be written as Vi with i = 1, 2, 3, where V1 = Vx , V2 = Vy and V3 = Vz.The latter notation is the usual designation for a three-dimensional Cartesiantensor of rank 1. It is more convenient, when working with atoms, to usespherical coordinates (r, θ, φ) instead of x, y, z, where the conversion betweenthe two systems of components is given by

x = r sin θ cos φ, (E.1)

y = r sin θ sin φ, (E.2)

z = r cos θ. (E.3)

We will now encounter spherical vectors and higher-ranked spherical tensors.The usual notation for a spherical tensor of rank k is Tk,q. A rank-1 spheri-cal tensor in three dimensions will have k = 1 and q = −1, 0, 1. Similar tothe Cartesian tensor of rank 1, the spherical tensor of rank 1 is related to thecomponents of a spherical vector. In spherical coordinates, V has componentsV+1, V0 and V−1. The quantity Tk,q is an irreducible tensor operator that isHermitian if

T†k,q = (−1)qTk,−q. (E.4)

When Tk,q acts only on spatial coordinates (i.e., does not act on spins) then

Tk,q = CkYkq(Ω), (E.5)

where Ykq are the spherical harmonics (see Appendix B.7) and Ck is a func-tion that is independent of the angular coordinates. It is often convenient to

Page 311: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

294 E Tensor Algebra

convert between Cartesian and spherical vectors. As an example, we will con-sider the spherical vector Vq = f (r)Y1q, where we have taken C1 to be purely afunction of r and have dropped the index 1 from V. From Eq. (1.25), we writethe components of V as of a spherical vector in the three spherical harmonicsY10 and Y1,±1, from which it is clear that

V±1 = f (r)Y1,±1 = ∓ f (r)√

38π

sin θe±iφ, (E.6)

V0 = f (r)Y1,0 = f (r)√

34π

cos θ. (E.7)

Expanding the exponential and recognizing sin θ cos φ, etc., as unit vectors inthe x, etc., direction, we can write

V±1 = ∓ 1√2

(Vx ± iVy

), (E.8)

V0 = Vz, (E.9)

where the magnitude of V is√

3/4π f (r).

E.2Commutation Relations

There are two commutation relations that are frequently encountered whenworking with spherical tensors:[

J±, Tk,q

]=

√(k ∓ q) (k ± q + 1)Tk,q±1, (E.10)[

Jz, Tk,q

]= qTk,q. (E.11)

As an example, the following is how the orbital angular momentum L−1 com-mutes with L± and Lz:[

L±, L1,−1

]=

√2L1,0, (E.12)[

Lz, L1,−1

]= −L1,−1. (E.13)

E.3Reduced Matrix Elements

When an operator is represented as a spherical tensor, such as angular mo-mentum (L) or an electric dipole (er), we need to find matrix elements of the

Page 312: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

E.3 Reduced Matrix Elements 295

form 〈γjm| Tk,q |γ′ j′m′〉. The Wigner–Eckart theorem,

〈γjm| Tk,q∣∣γ′ j′m′⟩ = (−1)j−m

(j k j′

−m q m′)〈γj‖Tk‖γ′ j′〉, (E.14)

allows us to relate this matrix element to 3-j symbols and a reduced matrixelement. Values for 3-j symbols are tabulated in several places1 or easily cal-culated in programs such as MATHEMATICA®. The reduced or double-barmatrix element is equivalent to the line strength (see Section 4.3.4.1). Specifi-cally, we relate the square of the reduced matrix element to the sum over thematrix elements,

|〈γj‖Tk‖γ′ j′〉|2 = ∑q,m,m′

|〈γjm|Tk,q|γ′ j′m′〉|2. (E.15)

Clearly, the reduced matrix element is independent of angle (the orientation ofthe operator) and the magnetic quantum numbers. Consequently, neither themagnetic quantum number, m, nor q appear in the reduced matrix element.Its independence of q allows us to write

|〈γj‖Tk‖γ′ j′〉|2 = (2k + 1) ∑m,m′

|〈γjm|Tk,q|γ′ j′m′〉|2, (E.16)

with 2k + 1 being the number of possible orientations of the operator. ForHermitian operators (Eq. (E.5)), the reduced matrix elements obey

〈γj‖Tk,q‖γ′ j′〉 = (−1)j−j′ 〈γ′ j′‖Tk,q‖γj〉∗. (E.17)

One evaluates the reduced matrix element by selecting an m state for whichthe left-hand side of Eq. (E.14) can be determined most conveniently. For ex-ample, the reduced matrix element for the spin operator, T = S with k = 1,for an electron in a state |γsms〉, can be found as follows:

〈γs||S||γs〉 = (−1)ms−s 〈γsms|S|γsms〉(s k s

−ms q ms

) . (E.18)

The most convenient orientation ofS will be along the z-axis where q = 0 andS → Sz. The numerator on the right-hand side of the expression will just bems. Evaluating the 3-j symbol leads to

〈γs||S||γs〉 =√

s(s + 1)(2s + 1) =

√32

. (E.19)

1) See, for example, A. R. Edmonds, Angular Momentum in QuantumMechanics [126].

Page 313: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

296 E Tensor Algebra

The last step is made by setting s = 1/2. As promised, the reduced matrixelement is independent of ms. One can show that it is also independent of qby letting q = ±1 and showing that the right-hand side gives the same answer.Once we have a value for the reduced matrix element, we are able to evaluatespecific matrix elements in terms of it. For example, the matrix element forS±, where q = ±1, is

〈γsms|S±|γ′s′m′s〉 =

√(s ∓ ms)(s ± ms + 1)δs,s′δm′

s,ms±1

=

√(12∓ ms

) (32± ms

)δs,s′δm′

s,ms±1.(E.20)

We recognize S± as the raising and lowering operators.

E.4Matrix Elements of Products of Operators

Consider an operator Uk2,q2 that operates in a space spanned by the basis set|γJM〉. If |ψ〉 is a function in this space, then we can write

|ψ〉 = Uk2,q2

∣∣γ′ J′M′⟩= ∑

γ′′,J′′,M′′

∣∣γ′′ J′′M′′⟩ ⟨γ′′ J′′M′′∣∣ Uk2,q2

∣∣γ′ J′M′⟩= ∑

γ′′,J′′,M′′

∣∣γ′′ J′′M′′⟩ (−1)J′′−M′′(

J′′ k2 J′−M′′ q2 M′

)× ⟨

γ′′ J′′∥∥Uk2

∥∥ γ′ J′⟩

. (E.21)

Product of tensor operators. Often, we need to find matrix elements oftensor products of operators. If, for example, we have another operator, Tk1,q1

,which also acts in this space, the matrix element of the product can be writtenas

〈γJM| Tk1,q1Uk2,q2

∣∣γ′ J′M′⟩= ∑

γ′′,J′′,M′′〈γJM| Tk1,q1

∣∣γ′′ J′′M′′⟩ (−1)J′′−M′′

×(

J′′ k2 J′−M′′ q2 M′

) ⟨γ′′ J′′

∥∥Uk2

∥∥ γ′ J′⟩

.

Page 314: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

E.4 Matrix Elements of Products of Operators 297

Applying the Wigner–Eckart theorem a second time leads to

〈γJM| Tk1,q1Uk2,q2

∣∣γ′ J′M′⟩= ∑

γ′′,J′′,M′′(−1)J−M

(J k1 J′′

−M q1 M′′)〈γJ‖ Tk1

∥∥γ′′ J′′M′′⟩× (−1)J′′−M′′

(J′′ k2 J′

−M′′ q2 M′) ⟨

γ′′ J′′∥∥Uk2

∥∥ γ′ J′⟩

.

Scalar product of two tensors. The preceding expression is an expressionfor a general product of two tensors when they both act in the same space.When the product is a scalar product, the expression can be simplified. Usingthe shorthand ∑q (−1)q Tk,−qUk,q ≡ TkUk ≡ T · U, the previous expressionreduces to⟨

γJM|TkUk|γ′ J′M′⟩ = ∑γ′′ J′′M′′

⟨γJ‖Tk‖γ′′ J′′

⟩ ⟨γJ‖Uk‖γ′′ J′′

⟩× (−1) J+J′′ ∑

q(−1)q−M−M′′

(J k J′′

−M −q M′′)

×(

J′′ k J′−M′′ q M′

)=

δJ,J′δM,M′

2J + 1 ∑γ′′ J′′

(−1) J′′−J

× 〈γJ‖ Tk∥∥γ′′ J′′

⟩ ⟨γ′′ J′′

∥∥ Uk∥∥γ′ J′

⟩. (E.22)

The last step can be shown using the symmetry and summation properties ofthe 3-j symbols in Appendix D.2 and recognizing the fact that 2J − 2M is even.

Scalar products of tensors acting in different spaces. It is also commonto need matrix elements of products when Tk,q and Uk,q commute and act indifferent spaces. Matrix elements of the scalar product between orbital andspin angular momenta (L ·S) in the state |γLSJM〉 is an example. It can beshown that in this case the matrix element takes the form

〈(γ1 j1γ2 j2) JM| TkUk∣∣(γ′

1 j′1γ′2 j′2

)J′M′⟩

= δj,j′δm,m′(−1)j′1+j2+J 〈γ1 j1‖ Tk∥∥γ′

1 j′1⟩ 〈γ2 j2‖Uk

∥∥γ′2 j′2

⟩×

j1 j2 Jj′2 j′1 k

. (E.23)

Reduced matrix element of L in the state |γLSJM〉. When a tensor op-erator acts in only part of the space, such as the orbital (L) or spin (S)angular momenta, we must evaluate reduced matrix elements of the form〈γLSJ ‖L‖ γLSJ〉. It is instructive to determine what such a reduced element

Page 315: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

298 E Tensor Algebra

means. To that end, we first look at the reduced matrix element of an operatorTk,q in a state |γj1 j2 JM〉. It is possible to show that

〈γj1 j2 J‖Tk‖γj′1 j′2 J′〉 = (−1)j1+j2+J+k

×√

(2J + 1)(2J′ + 1)

j′1 J′ j2J j1 k

〈γj1 ‖Tk‖ γ′ j′1〉.

The orbital angular momentum reduced matrix element then reduces to

〈γLSJ ‖L‖ γLSJ〉 = (−1)L+S+J+1(2J + 1)

×

L J SJ L 1

〈γL ‖L‖γL〉. (E.24)

From the Wigner–Eckart theorem, it is possible to show that 〈γL ‖L‖ γL〉 =√L(L + 1)(2L + 1). Evaluating the 6-j symbol leads to

〈γLSJ ‖L‖ γLSJ〉 =J(J + 1) + L(L + 1)− S(S + 1)

2J(J + 1)

×√

J(J + 1)(2J + 1). (E.25)

A semiclassical interpretation of Eq. (E.25) is possible if we define 〈cos(L,J)〉,where (L,J) denotes the angle between L andJ. (A similar expression is pos-sible forS as well.) From the semiclassical vector model (Fig. E.1) coupling L,S andJ and the law of cosines, it is possible to show that

〈cos(L,J)〉 =J(J + 1) + L(L + 1) − S(S + 1)

2√

J(J + 1)√

L(L + 1), (E.26)

〈cos(S,J)〉 =J(J + 1) + S(S + 1) − L(L + 1)

2√

J(J + 1)√

S(S + 1). (E.27)

Equation (E.26) allows us to rewrite Eq. (E.25) as

〈γLSJ‖L‖γLSJ〉 =√

L(L + 1)〈cos(L,J)〉 〈γJ‖J‖γJ〉√J(J + 1)

. (E.28)

Thus, the reduced matrix element can be viewed as the average projection ofL onto the average J axis.

One can write the left-hand side of Eq. (E.25) in a different way if we writeL ·J semiclassically as

S2 = (J − L)2

= J2 + L2 − 2(L · J).(E.29)

Page 316: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

E.4 Matrix Elements of Products of Operators 299

Fig. E.1 Semiclassical vector precession model.

Taking matrix elements with the state vector |γLSJM〉 ≡ |γJM〉 of both sidesof Eq. (E.29) and dividing by 〈γJM|J2|γJM〉 = J(J + 1) leads to

〈γJM|L ·J|γJM〉〈γJM|J2|γJM〉 =

J(J + 1) + L(L + 1)− S(S + 1)2J(J + 1)

. (E.30)

The matrix elements of L2 and S2 were found with the aid of Eqs. (D.5), (D.6)and (D.8). Comparing Eqs. (E.25) and (E.30) allows us to write

〈γLSJ‖L‖γLSJ〉 = 〈γJM|L · J|γJM〉√

J(J + 1)(2J + 1)J(J + 1)

. (E.31)

Finally, the matrix element of L is given by

〈γLSJM|L|γLSJM〉 = MJ(J + 1) + L(L + 1) − S(S + 1)

2J(J + 1), (E.32)

or

〈γLSJM|L|γLSJM〉 = M〈γLSJ‖L‖γLSJ〉

〈γJ‖J‖γJ〉 . (E.33)

Semiclassically, this corresponds to the projection of the average value of L,while it precesses aboutJ, onto the z-axis as shown in Fig. E.1.

Page 317: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

300 Further Reading

Tensor product of two tensors. Finally, we end by stating without proof thereduced matrix element of the tensor product of two operators acting in thesame and different spaces.2 If we let

XK,Q = Tk1,q1 × Uk2,q2 , (E.34)

then the reduced matrix element of Pk,q when both operators act in the samespace will be

〈γ′ j′‖XK‖γj〉 =√

2K + 1(−1)K+j+j′ ∑γ′′,j′′

k1 k2 Kj j′ j′′

× 〈γ′ j′‖Tk1‖γ′′ j′′〉〈γ′′ j′′‖Uk2‖γj〉. (E.35)

When the operators act in different spaces, we have

〈γ′ j′1 j′2 J′‖XK‖γj1 j2 J〉 =√

(2J + 1)(2J′ + 1)(2K + 1)

× ∑γ′′〈γ′ j′1‖Tk1‖γ′′ j1〉〈γ′′ j′2‖Uk2‖γj2〉.

j′1 j1 k1j′2 j2 k2J′ J K

. (E.36)

Further Reading

1 A. R. Edmonds, Angular Momentum in Quan-tum Mechanics, Princeton University Press,Princeton, NJ, 1974 [ISBN 0-691-07912-9].

2 R. D. Cowan, The Theory of Atomic Structureand Spectra, University of California Press,Berkeley, CA, 1981.

3 I. I. Sobelman, Atomic Spectra and RadiativeTransitions, Springer Series in ChemicalPhysics. Springer, New York, NY, 1979.

4 F. Yang and J. H. Hamilton, Modern Atomicand Nuclear Physics, McGraw-Hill, New York,NY, 1996 [ISBN 0-07-025881-3].

2) The derivation is discussed in Ref. [126].

Page 318: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

301

References

1 G. Arfken and H. Weber. MathematicalMethods for Physicists. Academic, NewYork, NY, 2000. ISBN 0-12-059825-6, Fifthedition.

2 R. D. Cowan. The Theory of Atomic Struc-ture and Spectra. University of CaliforniaPress, Berkeley, CA, 1981. ISBN 0-520-03821-5.

3 H. A. Bethe and E. E. Salpeter. QuantumMechanics of One- and Two-Electron Atoms.Plenum, New York, NY, 1977. ISBN 0-306-20022-8, First paperback printing.

4 Albert Messiah. Quantum Mechan-ics. Dover Publications, 2000. ISBN:0486409244.

5 K. Bockasten. Phys. Rev. A, 9:1087, 1974.

6 J. Weiner and P.-T. Ho. Light–MatterInteraction, Volume 1: Fundamentals andApplications. Wiley, Hoboken, NJ, 2003.ISBN 0-471-25377-4, First edition.

7 NIST. Basic atomic spectroscopic data, a.http://physics.nist.gov/PhysRefData/Handbook/periodictable.htm.

8 NIST. Physical reference data, b.http://physics.nist.gov/PhysRefData/contents.html.

9 NIST. Physics laboratory products andservices, c. http://physics.nist.gov/products.html.

10 J. Javanainen, J. H. Eberly, and Q. C. Su.Numerical simulations of multiphotonionization and above-threshold electronspectra. Phys. Rev. A, 38:3430, 1988.

11 Q. Su and J. H. Eberly. Model atom formultiphoton physics. Phys. Rev. A, 44:5997, 1991.

12 S. Augst, D. Strickland, D. D. Meyerhofer,S. L. Chin, and J. H. Eberly. Phys. Rev.Lett., 63:2212, 1989.

13 L. V. Keldysh. Sov. Phys. JETP, 20:1307,1965.

14 A. M. Perelomov, V. S. Popov, and M. V.Terent’ev. Sov. Phys. JETP, 23:924, 1966.

15 M. V. Ammosov, N. B. Delone, and V. P.Krainov. Sov. Phys. JETP, 64:1191, 1986.

16 T. D. G. Walsh, F. A. Ilkov, J. E. Dechker,and S. L. Chin. The tunnel ionization ofatoms, diatomic and triatomic moleculesusing intense 10.6 µm radiation. J. Phys.B: At. Mol. Opt. Phys., 27:3767, 1994.

17 S. Augst, D. D. Meyerhofer, D. Strickland,and S. L. Chin. J. Opt. Soc. Am. B, 8:858,1991.

18 F. A. Ilkov, J. E. Decker, and S. L. Chin. J.Phys. B: At. Mol. Opt. Phys., 25:4005, 1992.

19 B. Walker, B. Sheehy, L. F. DiMauro,P. Agostini, K. J. Schafer, and K. C. Ku-lander. Precision measurement of strongfield double ionization of helium. Phys.Rev. Lett., 73:1227, 1994.

20 X. M. Tong, Z. X. Zhao, and D. Lin. The-ory of molecular tunneling ionization.Phys. Rev. A, 66:033402, 2002.

21 E. Schrödinger. Ann. Phys., 80:457, 1926.

22 P. S. Epstein. Phys. Rev., 28:695, 1926.

23 I. Walker. Z. Phys., 38:635, 1926.

24 D. R. Inglis and E. Teller. Astrophys. J., 90:439, 1939.

25 G. Wentzel. Z. Phys., 38:518, 1926.

26 A. Khadjavi, A. Lurio, and W. Happer.Stark effect in the excited states of rb, cs,cd and hg. Phys. Rev., 167:128, 1968.

Page 319: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

302 References

27 I. I. Sobelman. Atomic Spectra and Ra-diative Transitions. Springer Series inChemical Physics. Springer, Berlin, 1979.ISBN 0-387-09082-7.

28 N. Ramsey. Nuclear Moments. Wiley, NewYork, NY, 1953.

29 H. Kopfermann. Nuclear Moments. Englishversion prepared from the second German edi-tion by E. E. Schneider. Pure and AppliedPhysics, Volume 2. Academic, New York,NY, 1958.

30 Landolt-Börnstein. Zahlenwert und Funk-tionen. Springer, Berlin, 1952. Sixthedition, Volume I, Part 5.

31 P. A. M. Dirac. Proc. R. Soc. Lond. Ser. A,112:661, 1926.

32 P. A. M. Dirac. Proc. R. Soc. Lond. Ser. A,114:243, 1927.

33 U. Fano and J. W. Cooper. Spectral distri-bution of atomic oscillator strengths. Rev.Mod. Phys., 40:441, 1968.

34 W. T. Hill III, J. Sugar, T. B. Lucatorto,and K. T. Cheng. Analysis of the5p6 → 5p5nl j = 1) rydberg series inba2+. Phys. Rev. A, 36:1200, 1987.

35 W. T. Hill III and C. L. Cromer. Laser-driven ionization and photoabsorptionspectroscopy of atomic ions. In W. M. Yenand M. D. Levenson, editors, Lasers, Spec-troscopy and New Ideas: A Tribute to ArthurL. Schawlow, volume 54 of Optical Sciences,page 183, New York, NY, 1987. Springer.

36 U. Fano. Phys. Rev. A, 2:353, 1970.

37 K. T. Lu. Phys. Rev. A, 4:579, 1971.

38 C.-M. Lee and K. T. Lu. Phys. Rev. A, 8:1241, 1973.

39 U. Fano. Effects of configuration interac-tion on intensities and phase shifts. Phys.Rev., 124:1866, 1961.

40 F. H. Faisal. Theory of Multiphoton Pro-cesses. Plenum, New York, NY, 1987.ISBN 0-306-42317-0.

41 P. Agostini and G. Petite. Photoelectriceffect under strong irradiation. Contemp.Phys., 29:57, 1988.

42 P. Agostini, F. Fabre, G. Mainfray, G. Pe-tite, and N. K. Rahman. Free–free tran-sitions following six-photon ionizationof xenon atoms. Phys. Rev. Lett., 42:1127,1979.

43 H. Lefevbre-Brion and R. W. Field. Per-turbations in the Spectra of Diatomic Mole-cules. Academic, Orlando, FL, 1986. ISBN0124426905.

44 G. Herzberg. Molecular Spectra and Molec-ular Structure I: Spectra of Diatomic Mole-cules. Van Nostrand Reinhold, New York,NY, 1950.

45 K. P. Huber and G. Herzberg. MolecularSpectra and Molecular Structure: Constantsof Diatomic Molecules. Van Nostrand Rein-hold, New York, NY, 1978.

46 NIST. Chemical data, d. http://webbook.nist.gov/.

47 J. L. Dunham. Phys. Rev., 41:721, 1932.48 T. E. Sharp. Potential-energy curves for

molecular hydrogen and its ions. At.Data, 2:119, 1971.

49 I. Kovács. Rotational Structure in the Spec-tra of Diatomic Molecules. Am. Elsevier,New York, 1969.

50 R. Rydberg. Z. Phys., 80:514, 1933.51 O. Klein. Z. Phys., 76:221, 1932.52 A. L. G. Rees. Proc. Phys. Soc. A, B, 59:

998, 1947.53 R. T. Birge and H. Sponer. Phys. Rev., 28:

259, 1926.54 R. J. LeRoy and R. B. Bernstein. J. Chem.

Phys., 52:3869, 1970.55 W. C. Stwalley. Chem. Phys. Lett., 6:241,

1970.56 NIST J. T. Hougen. Nbs monograph 115:

The calculation of rotational energy levelsand rotational line intensities in diatomicmolecules. http://physics.nist.gov/Pubs/Mono115/contents.html.

57 M. L. Ginter. Private communication.58 E. Wigner and E. E. Witmer. Z. Phys., 51:

859, 1928.59 D. D. Konowalow, M. E. Rosenkrantz,

and M. L. Olson. The molecular elec-tronic structure of the lowest 1Σ+

g , 3Σ+u ,

1Σ+u , 3Σ+

g , 1Πu, 1Πg , 3Πu, 3Πg states ofNa2. J. Chem. Phys., 72:2612, 1980.

60 Z. Vager, R. Naaman, and E. P. Kanter.Science, 244:426, 1989.

61 J. H. Posthumus, L. J. Frasinski, A. J.Giles, and K. Codling. J. Phys. B: At.Mol. Opt. Phys., 28:L349, 1995.

62 T. Zuo and A. D. Bandrauk. Phys. Rev. A,52:R2511, 1995.

Page 320: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

References 303

63 T. Seideman, M. Yu. Ivanov, and P. B.Corkum. The role of electron localiza-tion in intense-field molecular ionization.Phys. Rev. Lett., 75:2819, 1995.

64 R. S. Mulliken. Intensities of elec-tronic transitions in molecular spectraII: charge-transfer spectra. J. Chem. Phys.,7:20, 1939.

65 S. Chelkowski and A. D. Bandrauk. Two-step coulomb explosions of diatoms inintense laser fields. J. Phys. B: At. Mol.Opt. Phys., 28:L723, 1995.

66 P. A. Franken, A. E. Hill, C. W. Peters, andG. Weinreich. Phys. Rev. Lett, 7:118, 1961.

67 R. W. Boyd. Nonlinear Optics. Academic,San Diego, CA, 1992.

68 J. A. Armstrong, N. Bloembergen,J. Ducuing, and P. S. Pershan. Phys. Rev.,127:1918, 1962.

69 T. Kuri and K. I. Kitayama. J. LightwaveTechnol., 21:3167, 2003.

70 R. Krahenbuhl and W. K. Burns. IEEETrans. Microwave Theory Tech., 48:860,2000.

71 L. Xu, X. C. Zhang, and D. H. Auston.Appl. Phys. Lett., 61:1784, 1992.

72 Q. Wu and X. C. Zhang. Appl. Phys. Lett.,67:3523, 1995.

73 J. A. Giordmaine and R. C. Miller. Phys.Rev. Lett., 14:973, 1965.

74 J. E. Bjorkholm and A. Ashkin. Phys. Rev.Lett., 32:129, 1974.

75 L. F. Mollenauer, R. H. Stolen, and J. P.Gorden. Phys. Rev. Lett., 45:1095, 1980.

76 F. Zernike and J. E. Midwinter. In AppliedNonlinear Optics, New York, NY, 1973.Wiley.

77 E. P. Ipper and C. V. Shank. In S. L.Shapico, editor, Ultrashort Light Pulses,pages 83–122, Berlin and New York, 1977.Springer.

78 P. D. Maker, R. W. Terhune, M. Nisenoff,and C. M. Savage. Phys. Rev. Lett., 8:21,1962.

79 R. L. Byer and R. L. Herbst. In Y. R. Shen,editor, Tunable Infrared Generation, Berlin,1977. Springer.

80 D. Magde and H. Mahr. Phys. Rev. Lett,18:905, 1967.

81 J. M. Manley and H. E. Rowe. Proc. IRE,47:2115, 1959.

82 J. Warner. Appl. Phys. Lett., 12:222, 1968.83 J. Shah, T. C. Damen, B. Deve, and

D. Block. Appl. Phys. Lett., 50:1307, 1987.84 N. Bloembergen. Nonlinear Optics. World

Scientific, Singapore, 4th edition, 1996.85 Y. R. Shen. The Principles of Nonlinear

Optics. Wiley, New York, NY, 1984.86 R. H. Pantell and H. E. Puthoff. Funda-

mentals of Quantum Electronics. Wiley,New York, NY, 1969.

87 M. D. Levenson and N. Bloembergen.Observation of two-photon absorptionwithout doppler broadening on the 3s–5stransition in sodium vapor. Phys. Rev.Lett., 32:645, 1974.

88 F. Biraben, B. Cagnar, and G. Grynberg.Phys. Rev. Lett., 32:643, 1974.

89 N. Bloembergen, H. Lotem, andJr R. T. Lynch. Indian J. Pure Appl. Phys.,16:151, 1978.

90 M. D. Levenson and S. S. Kano. Intro-duction to Nonlinear Laser Spectroscopy.Academic, San Diego, CA, 1988.

91 G. P. Agrawal. Nonlinear Fiber Optics.Academic, San Diego, CA, 1989.

92 N. Bloembergen and P. S. Pershan. Phys.Rev., 128:606, 1962.

93 N. Bloembergen and C. H. Lee. Phys. Rev.Lett., 19:835, 1967.

94 C. H. Lee and V. Bhanthumnavin. Opt.Commun., 18:326, 1976.

95 H. J. Simon and C. H. Lee. Opt. Lett., 13:440, 1988.

96 V. Bhanthumnavin and C. H. Lee. Phys.Rev. A, 50:2579, 1994a.

97 D. L. Butler, G. L. Burdge, C. H. Lee, andH. J. Simon. Opt. Lett., 17:1125, 1992.

98 C. Ohlhoff, G. Lupke, C. Meyer, andH. Kurz. Phys. Rev. B, 55:4596, 1997.

99 N. Bloembergen, H. J. Simon, and C. H.Lee. Phys. Rev., 181:1261, 1969.

100 V. Bhanthumnavin and N. Ampole. Mi-crowave Opt. Technol. Lett., 3:239, 1990.

101 V. Bhanthumnavin and C. H. Lee. J. Appl.Phys., 75:3294, 1994.

102 B. Dick, A. Gierulski, G. Marowski, andG. A. Reider. Appl. Phys. B, 38:107, 1985.

103 T. F. Heinz, H. W. K. Tom, and Y. R. Shen.Phys. Rev. A, 28:1883, 1983.

Page 321: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

304 References

104 N. Bloembergen, R. K. Chang, S. S. Jha,and C. H. Lee. Phys. Rev., 174:813, 1968.

105 J. I. Dadap, X. F. Hu, N. M. Russell, J. G.Ekendt, J. K. Lowell, and M. C. Downer.IEEE J. Sel. Top. Quantum Electron., 1:1145,1995.

106 J. I. Dadap, P. T. Wilson, M. H. Anderson,and M. C. Downer. Opt. Lett., 22:1, 1997.

107 C. H. Lee, R. Chang, and N. Bloembergen.Phys. Rev. Lett., 18:167, 1967.

108 J. I. Dadap, X. F. Hu, M. H. Anderson,M. C. Downer, J. K. Lowell, and O. A.Aktsipetrov. Phys. Rev. B, 53:R7607, 1996.

109 I. I. Smolyaninov, A. V. Zayats, and C. C.Davis. Opt. Lett., 22:1592, 1997.

110 I. I. Smolyaninov, H. Y. Liang, C. H. Lee,and C. C. Davis. J. Appl. Phys., 89:206,2001.

111 J. I. Gersten and A. N. Tzan. J. Chem.Phys., 73:3023, 1980.

112 X. C. Zhang, Y. Jin, and X. F. Ma. Appl.Phys. Lett., 61:2764, 1992.

113 N. M. Froberg, B. B. Hu, X. C. Zhang, andD. H. Auston. Appl. Phys. Lett., 59:3207,1991.

114 Q. Wu, Ph. D thesis, Rensselaer Polytech-nic Institute, 1997.

115 Q. Wu and X. C. Zhang. Appl. Phys. Lett.,68:1604, 1996.

116 Q. Wu and X. C. Zhang. Appl. Phys. Lett.,71:1285, 1997.

117 X. F. Ma and X. C. Zhang. J. Opt. Soc. Am.B, 10:1175, 1993.

118 R. S. Van Dyck, P. B. Schwinberg, andH. G. Dehmelt. Phys. Rev. Lett., 59:26,1987.

119 J. Schwinger. On quantum-electro-dynamics and the magnetic moment ofthe electron. Phys. Rev., 73:416, 1948.

120 V. W. Hughes and T. Kinoshita. Rev. Mod.Phys., 71:S133, 1999.

121 J. Mathews and R. L. Walker. Mathemat-ical Methods of Physics. Benjamin Cum-mings, San Francisco, CA, 1970. ISBN0-8053-7002-1, Second edition.

122 P. A. M. Dirac. The Principles of Quan-tum Mechanics. Oxford University Press,Oxford, UK, 1981. ISBN 0-19-852011-5,Fourth edition.

123 E. U. Condon and H. Odabasi. AtomicStructure. Cambridge University Press,New York, NY, 1980. ISBN 0-521-21893-8,Fourth edition.

124 R. Courant and D. Hilbert. Methods ofMathematical Physics, Volume I. Inter-science, New York, NY, 1953. First edi-tion.

125 P. R. Halmos. Introduction to Hilbert Spaceand the Theory of Spectral Multiplicity.Interscience, Chelsea, NY, 1957. ISBN0828400822, Second edition.

126 A. R. Edmonds. Angular Momentum inQuantum Mechanics. Princeton UniversityPress, Princeton, NJ, 1974. ISBN 0-691-07912-9.

Page 322: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics.W. T. Hill and C. H. LeeCopyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-40661-6

305

Index

3-j symbol 2906-j symbol 2919-j symbol 292

aabsorption– Doppler-free 214– – two-photon 213– multiple-photon 203AC tunneling ionization– ADK rate, atomic 56– ADK rate, molecular 144– atomic 55, 56– molecular 142angle of incidence 240associated Laguerre polynomial 271associated Legendre functions 271atomic constants 268atomic core electrons 36atomic data, see data under atomic struc-

tureatomic degeneracy 20atomic field strength 264atomic mass unit 262atomic momentum 263atomic orbital 16atomic orbitals 34atomic selection rules 82– E1 82– E2 84– electric dipole 82– electric quadrupole 84– M1 84– magnetic dipole 84atomic spectra 85– above threshold ionization (ATI) 93– autoionization 88–91– bound states 88, 90– intense-field photoionization 92atomic structure– S, P, D, . . . 41– s, p, d, . . . 20, 35

– atomic coupling schemes 40– – LS coupling (Russell–Saunders) 40– – intermediate coupling 45– – jj-coupling 45– – pair coupling 45– autoionization 29, 89– Bohr atom 4– Bohr radius 6, 151, 263– bound states 6, 7, 11, 13, 18, 19– channel 30, 86– configurations 34–36– continuum states 10, 18, 19, 30– core electrons 36– coupling angular momenta 289– – recoupling 48– data– – atomic constants 268– – configurations 287– – electronic configuration 287– – filling subshells 286– – NIST web page 285– degeneracy 20, 78– fine structure 28, 49– – Hund’s rule 49– – Landé rule 50– hyperfine structure 61– – Al 62– – electric quadrupole interaction 61, 65,

66– – magnetic dipole interaction 61– – Zeeman effect 65– K-, M- and L-shell 35– Lu-Fano plot 88, 90– magnetic quantum number (m) 20– orbital angular momentum (l) 8, 19, 20– oscillator strength 86– principal quantum number (n) 19– relativistic correction 21, 26, 27– shell 20, 34– – closed 35– Stark effect 53– – linear Stark effect 57– – quadratic Stark effect 59

Page 323: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

306 Index

– subshell 20, 34, 36– – closed 35– term 36– Zeeman effect 60, 214– – Landé g factor 60atomic time scale 263atomic valence electrons 36atomic velocity 263atoms– multielectron atom– – effective potential 33– one-electron atom 3, 7– valence electrons 36autoionization, see atomic spectra and

structure

bBaTiO3 246birefringent crystal 156, 170Bohr atom 4Bohr magneton (µB) 26, 63, 264Bohr radius 6, 151, 263Boltzmann distribution 201Born–Oppenheimer approximation 99bound states, see atomic structurebras & kets, see quantum mechanics

cchannel 30, 86classical electron radius 262coherence length 169coherent anti-Stokes Raman scattering

221Compton wavelength 262configuration– atomic, see atomic structure– molecular 97, 109, 110, 128confluent hypergeometric series 271conservation– energy 158, 175– momentum 158, 175continuum states, see atomic structureCoulomb explosion 142– kinetic energy released 144Coulomb gauge 71coupled equations 192coupling angular momenta 40, 110, 289– atomic, see atomic structure– diatomic, see diatomic structurecritical angle 224, 226, 231, 233

dd-coefficients– off-diagonal element 216d coefficients 163, 164, 174, 187– nonlinear 159

density operator ρ 199diagonal matrix element 204diatomic data, see data under diatomic

structurediatomic selection rules 139– Hund’s case (a) 140– Hund’s case (b) 140– Hund’s case (c) 140– Hund’s case (d) 141diatomic structure 97– P, Q and R branches 139, 142, 146– anharmonicity 104– average electronic potential (U(R)) 100,

102– constants– – αe 106– – βe 106– – γe 106– – ωexe 104, 105– – ωeye 105– – ωe 103– – Be 101, 106– – Bv 105– – De 106– – Dv 105– – Yij 106– coupling angular momenta 110, 112,

289– – Hund’s case (a) 114– – Hund’s case (b) 114, 116– – Hund’s case (c) 114, 117– – Hund’s case (d) 114, 118– data– – molecular constants 285– – NIST web page 285– Dunham expression 106– electronic orbital angular momentum

111– electronic spin 111– electronic state labeling convention 128– electronic states 109– – gerade state (g) 122– – inversion 118– – Morse potential 107– – reflection 118– – symmetries 118– – ungerade state (u) 123– Franck–Condon principle 138– Hamiltonian 98– Herzberg bookkeeping diagram 123,

143– labeling nomenclature 113– molecular orbital (MO) theory 128– Morse potential 107– multiplet splitting 112– nomenclature 112

Page 324: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Index 307

– nuclear motion 99– nuclear spin 126– nuclear wavefunction 101– optical transitions 137– quantum numbers 112– – K vs R usage 101, 112, 115, 118, 123,

124– rotational energy 101– Rydberg–Klein–Rees method (RKR) 108– separated atom MO 132– total angular momentum 112– united atom MO 131– vertical transition 138– vibrating rotator 105– vibrational wavefunction 102dipole antenna 251Dirac delta function 269Dirac equation 23

eeffective potential 33effective quantum number 37EFISHG 237, 242, 243electro-optic– effect 224– sensor 253electron spin 22electronic states, see diatomic structureelectrons– bound 238– conduction 238electrooptic– coefficient 156– effect 157– extraordinary ray 156, 170– linear effect 153, 156– modulator 156– ordinary ray 170energy units– eV 261– Hartree 261– kayser 260– Rydberg 261– wavenunber 260

fferroelectric materials 244field ionization 54fine structure constant 260fine structure, see atomic structurefour-wave mixing 159, 221free-electron wavefunction 71frequency upconversion 153Fresnel factors– linear 226, 233– nonlinear 226, 228, 233, 241

hHamiltonian– radiation 24, 27– single particle– – nonrelativistic 4– – relativistic correction 21Herzberg bookkeeping diagram 123, 143homogeneously broadened line 207hyperfine structure, see atomic structurehypergeometric series 270

iindex ellipsoid 156, 172, 173inhomogeneously broadened 207inversion symmetry 239isoelectronic sequence 87

kK shell 35KDP 171, 228, 231Keldysh parameter 55Kramers–Kronig relations 251Kronecker delta 269

lL shell 35Landé g factor 60Legendre functions 271– Rodrigues’ formula 272level degeneracy 39line shape– Gaussian 207, 213– Lorentzian 207, 213linear susceptibility 215longitudinal relaxation time 201Lu–Fano plot 87

mM-shell 35magnetic dipole 237magnetic moment– Bohr magneton (µB) 26, 63, 264– electron 264– neutron 267– nuclear magneton (µN) 266– nucleus 61, 267– proton 266magnetic quantum number (m) 20Manley–Rowe relations 192, 193matching– index 171, 176– momentum 177– – noncollinear 176– phase 155, 170, 175, 178, 186, 233– – angle 172, 173– – noncollinear 179– – type I 176

Page 325: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

308 Index

– – type II 176molecular data, see data under diatomic

structuremolecular orbitals– antibonding orbital 134– bonding orbital 131– LCAO 134– linear combination of atomic orbitals

134– molecular orbital (MO) theory 128

nn-j symbols– 3-j 289, 290– 6-j 291– 9-j 292negative uniaxial crystal 171nonlinear optics 152– Brewster angle 224, 234, 236– effect 151– electroreflectance 241, 242– Maxwell equation 180– polarization 154, 179, 215– refractive index 154– – change 159– susceptibility 151, 199, 215, 216– wave equation 167, 237nonvolatile ferroelectric random-access

memory 244normal dispersion 172normal surface 171nuclear quadrupole tensor 65, 267

oone-electron atom 3, 7– nonrelativisitc energy 15optical four-wave mixing 153optical parametric generation 152, 158– amplification 184, 185– amplifier 187– interaction 179– oscillation 153, 188– process 182– upconversion 194optical parametric oscillator 188optical rectification 153, 157, 254optical second harmonic generation– near-field 248– oblique incidence 223optical second-harmonic generation 151,

154, 162, 167, 169, 170– collinear 167– near-field microscopy 244– oblique incidence 231– reflection 233

optical second-harmonic intensity autocor-relator 178

optical third-harmonic generation 152,153, 159

optical three-wave mixing 153orbital g-factor (ge,l) 264orbital angular momentum– atomic (l) 8, 19, 20– diatomic (L, Λ) 111oscillator strength density 81, 85over-the-barrier ionization– atomic 55– molecular 142

pparabolic coordinates 280parity 20, 41Pauli spin matrices 23Pauli’s exclusion principle 40photoconductive dipole antenna 251piezoceramic 244plasma frequency 239ponderomotive energy 72– vibration radius 72P, Q and R branches, see diatomic struc-

tureprincipal quantum number 19principle of detailed balance 201probability density 14, 16– moments 15PZT ceramic 244, 246

qquadrupole 237quantum defect 37quantum mechanics– bras & kets 42, 274– complete set 274– completeness 276– eigenstate 277– eigenvalue 277– Hermitian 280– Hermitian adjoint 274– inner product 275– mathematical formalism 274– maxtrix mechanics 279– outer product 276– projection operator 276quasibound state, see various autoioniza-

tion entries

rRabi frequency 77radiation gauge 71radiation Hamiltonian 69radiative transitions 72

Page 326: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Index 309

– arrow direction convention 142– cross section 80– degeneracy 78– Fermi’s Golden Rule 76– line strength 79– Lorentzian profile 78– matrix element 74, 77– one-photon 74– oscillator strength 80, 81– two-photon 75– Voigt profile 78, 283Rayleigh range 179reduced mass 7reflected second harmonic intensity 233,

235Rodrigues’ formula 272rovibrational energy 106Rydberg constant 6Rydberg series 29, 85– channel 30, 86

sscattering cross section 211Schrödinger equation 3– angular solution 8– effective potential 10– parabolic coordinates 280– radial solution 13– radial wavefunction 10– spherical coordinates 7– spin wavefunction 27– time independent 3self-focusing 159self-phase modulation 159separated atom 129single-photon transition probability rate

207, 211S, P, D, . . . 41spherical harmonics 8, 9, 273spherical tensors 293– reduced matrix element 294– tensor algebra 293– tensor products 296– – scalar 297

– – tensor 300Stark effect, see atomic structurestate vector 42Stokes photon 221surface nonlinear susceptibility 237surface optical harmonic generation 237susceptibility 151

tTerahertz– pulse 251– radiation 250terahertz– pulse 157term 36third-order nonlinear susceptibility 218,

220three-photon– absorption 212– resonance 212threshold pumping intensity 190time-domain spectroscopy 250, 251total internal reflection 224, 226total reflection 234two-photon– absorption 154, 211, 213– resonance 209– transition 212– – probability rate 211

uunited atom 129upconverted 194

vVoigt profile 283

wwavenumber 260Wigner–Eckart theorem 295Wollaston prism 253

zZeeman effect, see atomic structure

Page 327: Light-Matter Interaction: Atoms and Molecules in External Fields and Nonlinear Optics

Related Titles

Weiner, J., Ho, P.-T.

Light-Matter InteractionVolume 1: Fundamentals and Applications256 pages2003HardcoverISBN 0-471-25377-4

Demtröder, W.

Molecular PhysicsTheoretical Principles and Experimental Methods484 pages with 288 figures and 43 tablesSoftcoverISBN 3-527-40566-6

Bachor, H.-A., Ralph, T. C.

A Guide to Experiments in Quantum Optics434 pages with 195 figures2004SoftcoverISBN 3-527-40393-0

Cohen-Tannoudji, C., Diu, B., Laloe, F.

Quantum MechanicsVolume 1914 pages1977SoftcoverISBN 0-471-16433-X

Cohen-Tannoudji, C., Diu, B., Laloe, F.

Quantum MechanicsVolume 2640 pages1977SoftcoverISBN 0-471-16435-6