249
UNITEXT for Physics Luca Salasnich Quantum Physics of Light and Matter Photons, Atoms, and Strongly Correlated Systems Second Edition

Quantum physics of light and matter : photons, atoms, and strongly correlated systems

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

UNITEXT for Physics

Luca Salasnich

Quantum Physics of Light and Matter Photons, Atoms, and Strongly Correlated Systems

Second Edition

Page 2: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

UNITEXT for Physics

Series editors

Michele Cini, Roma, ItalyAttilio Ferrari, Torino, ItalyStefano Forte, Milano, ItalyGuido Montagna, Pavia, ItalyOreste Nicrosini, Pavia, ItalyLuca Peliti, Napoli, ItalyAlberto Rotondi, Pavia, ItalyPaolo Biscari, Milano, ItalyNicola Manini, Milano, ItalyMorten Hjorth-Jensen, Oslo, Norway

Page 3: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

UNITEXT for Physics series, formerly UNITEXT Collana di Fisica e Astronomia,publishes textbooks and monographs in Physics and Astronomy, mainly in Englishlanguage, characterized of a didactic style and comprehensiveness. The bookspublished in UNITEXT for Physics series are addressed to graduate and advancedgraduate students, but also to scientists and researchers as important resources fortheir education, knowledge and teaching.

More information about this series at http://www.springer.com/series/13351

Page 4: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Luca Salasnich

Quantum Physics of Lightand MatterPhotons, Atoms, and Strongly CorrelatedSystems

Second Edition

123

Page 5: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Luca SalasnichFisica e Astronomia “Galileo Galilei”Università di PadovaPaduaItaly

ISSN 2198-7882 ISSN 2198-7890 (electronic)UNITEXT for PhysicsISBN 978-3-319-52997-4 ISBN 978-3-319-52998-1 (eBook)DOI 10.1007/978-3-319-52998-1

Library of Congress Control Number: 2017932090

© Springer International Publishing AG 2014, 2017This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or partof the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmissionor information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilarmethodology now known or hereafter developed.The use of general descriptive names, registered names, trademarks, service marks, etc. in thispublication does not imply, even in the absence of a specific statement, that such names are exempt fromthe relevant protective laws and regulations and therefore free for general use.The publisher, the authors and the editors are safe to assume that the advice and information in thisbook are believed to be true and accurate at the date of publication. Neither the publisher nor theauthors or the editors give a warranty, express or implied, with respect to the material contained herein orfor any errors or omissions that may have been made. The publisher remains neutral with regard tojurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by Springer NatureThe registered company is Springer International Publishing AGThe registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Page 6: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Preface to the Second Edition

The second edition of this book has two additional chapters based on lecture notesof the course “Quantum Field Theory in Condensed Matter Physics” given by theauthor at the Doctoral School in Physics of the University of Padova. These twochapters are an introduction to the functional integration for bosonic and fermionicfields with applications to superfluids, superconductors, and ultracold atomic gases.

Chapter 8 considers the quantum statistical mechanics of bosons within the for-malism of functional integration for the bosonic field, introducing the equation ofstate of weakly interacting bosons and the dimensional regularization of Gaussianfluctuations. The time-dependent Gross-Pitaevskii equation is analyzed: its connec-tion to the equations of superfluid hydrodynamics and its topological and solitonicsolutions (quantized vortices, dark solitons, and bright solitons). Chapter 9 introducesthe formalism of functional integration for the fermionic field. The equation of state ofweakly interacting repulsive fermions and the paring of attractive fermions within themean-field Bardeen-Cooper-Schrieffer (BCS) theory of low-temperature supercon-ductivity is derived. A section is devoted to the phenomenological Ginzburg-Landautheory of superconductivity which is quite successful near the critical temperature.Finally, the BCS theory is extended to analyze, in the case of ultracold atomicgases, the BCS-BEC crossover from the weak-coupling BCS regime of Cooper pairsto the strong-coupling regime with Bose-Einstein condensation (BEC) of molecules.

The author acknowledges several students who have detected missprints anderrors in the first edition. The author thanks Giacomo Bighin, Luca Dell’Anna,Roberto Onofrio, Fabio Sattin, Flavio Toigo, and Giorgio Urso for many sugges-tions and enlightening discussions about this book.

Padua, Italy Luca SalasnichDecember 2016

v

Page 7: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Preface to the First Edition

This book contains lecture notes prepared for the one-semester course “Structure ofMatter” belonging to the Master of Science in Physics at the University of Padova.The course gives an introduction to the field quantization (second quantization) oflight and matter with applications to atomic physics.

The first chapter briefly reviews the origins of special relativity and quantummechanics and the basic notions of quantum information theory and quantumstatistical mechanics. The second chapter is devoted to the second quantizationof the electromagnetic field, while the third chapter shows the consequences of thelight field quantization in the description of electromagnetic transitions. In thefourth chapter, the spin of the electron, and in particular its derivation from theDirac equation, is analyzed, while the fifth chapter investigates the effects ofexternal electric and magnetic fields on the atomic spectra (Stark and Zeemaneffects). The sixth chapter describes the properties of systems composed by manyinteracting identical particles. The Fermi degeneracy and the Bose-Einstein con-densation, introducing the Hartree-Fock variational method, the density functionaltheory, and the Born-Oppenheimer approximation, are also discussed. Finally, inthe seventh chapter the second quantization of the non-relativistic matter field, i.e.,the Schrödinger field, which gives a powerful tool for the investigation offinite-temperature many-body problems and also atomic quantum optics isexplained. Moreover, in this last chapter fermionic Fock states and coherent statesare presented, and the Hamiltonians of Jaynes-Cummings and Bose-Hubbard areintroduced and investigated. Three appendices on the Dirac delta function, theFourier transform, and the Laplace transform complete the book.

It is important to stress that at the end of each chapter, there are solved problemswhich help the students to put into practice the things they learned.

Padua, Italy Luca SalasnichJanuary 2014

vii

Page 8: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Contents

1 The Origins of Modern Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Special Relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.1 Axioms of Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . 81.2.2 Quantum Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.3 Quantum Statistical Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131.3.1 Microcanonical Ensemble . . . . . . . . . . . . . . . . . . . . . . . . . . 131.3.2 Canonical Ensemble . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141.3.3 Grand Canonical Ensemble . . . . . . . . . . . . . . . . . . . . . . . . . 15

1.4 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2 Second Quantization of Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.1 Electromagnetic Waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.1.1 Photons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.1.2 Electromagnetic Potentials and Coulomb Gauge . . . . . . . . . 24

2.2 Second Quantization of Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272.2.1 Fock versus Coherent States for the Light Field . . . . . . . . . 302.2.2 Linear and Angular Momentum of the Radiation Field . . . . 332.2.3 Zero-Point Energy and the Casimir Effect . . . . . . . . . . . . . . 34

2.3 Quantum Radiation Field at Finite Temperature . . . . . . . . . . . . . . . 362.4 Phase Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382.5 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3 Electromagnetic Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513.1 Classical Electrodynamics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513.2 Quantum Electrodynamics in the Dipole Approximation . . . . . . . . 53

3.2.1 Spontaneous Emission. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553.2.2 Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583.2.3 Stimulated Emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

ix

Page 9: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.3 Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 603.4 Einstein Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3.4.1 Rate Equations for Two-Level and Three-LevelSystems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.5 Life-Time and Natural Line-Width . . . . . . . . . . . . . . . . . . . . . . . . . 663.5.1 Collisional Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673.5.2 Doppler Broadening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3.6 Minimal Coupling and Center of Mass. . . . . . . . . . . . . . . . . . . . . . 693.7 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4 The Spin of the Electron. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814.1 The Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814.2 The Pauli Equation and the Spin . . . . . . . . . . . . . . . . . . . . . . . . . . 854.3 Dirac Equation with a Central Potential . . . . . . . . . . . . . . . . . . . . . 87

4.3.1 Relativistic Hydrogen Atom and Fine Splitting. . . . . . . . . . 884.3.2 Relativistic Corrections to the Schrödinger

Hamiltonian. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 894.4 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5 Energy Splitting and Shift Due to External Fields . . . . . . . . . . . . . . . 995.1 Stark Effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 995.2 Zeeman Effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.2.1 Strong-Field Zeeman Effect. . . . . . . . . . . . . . . . . . . . . . . . . 1025.2.2 Weak-Field Zeeman Effect . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.3 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

6 Many-Body Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1156.1 Identical Quantum Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1156.2 Non-interacting Identical Particles . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.2.1 Uniform Gas of Non-interacting Fermions . . . . . . . . . . . . . 1196.2.2 Atomic Shell Structure and the Periodic

Table of the Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1216.3 Interacting Identical Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

6.3.1 Variational Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246.3.2 Hartree for Bosons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256.3.3 Hartree-Fock for Fermions . . . . . . . . . . . . . . . . . . . . . . . . . 1266.3.4 Mean-Field Approximation . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.4 Density Functional Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316.5 Molecules and the Born-Oppenheimer Approximation . . . . . . . . . . 1386.6 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7 Second Quantization of Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1457.1 Schrödinger Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1457.2 Second Quantization of the Schrödinger Field . . . . . . . . . . . . . . . . 147

7.2.1 Bosonic and Fermionic Matter Field . . . . . . . . . . . . . . . . . . 150

x Contents

Page 10: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.3 Connection Between First and Second Quantization. . . . . . . . . . . . 1537.4 Coherent States for Bosonic and Fermionic Matter Fields . . . . . . . 1567.5 Quantum Matter Field at Finite Temperature . . . . . . . . . . . . . . . . . 1607.6 Matter-Radiation Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.6.1 Cavity Quantum Electrodynamics . . . . . . . . . . . . . . . . . . . . 1637.7 Bosons in a Double-Well Potential . . . . . . . . . . . . . . . . . . . . . . . . . 166

7.7.1 Analytical Results with N ¼ 1 and N ¼ 2. . . . . . . . . . . . . . 1707.8 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

8 Functional Integration for the Bosonic Field . . . . . . . . . . . . . . . . . . . . 1778.1 Partition Function of Interacting Bosons. . . . . . . . . . . . . . . . . . . . . 177

8.1.1 Semiclassical Approximation and Imaginary Time . . . . . . . 1798.2 Broken Symmetry and Order Parameter . . . . . . . . . . . . . . . . . . . . . 182

8.2.1 Ideal Bose Gas: Critical Temperature and CondensateFraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

8.2.2 Interacting Bose Gas: Bogoliubov Spectrum . . . . . . . . . . . . 1868.2.3 Dimensional Regularization of Gaussian Fluctuations . . . . . 190

8.3 Gross-Pitaevskii Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1918.3.1 Superfluid Hydrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . 1928.3.2 Bright and Dark Solitons . . . . . . . . . . . . . . . . . . . . . . . . . . 194

8.4 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

9 Functional Integration for the Fermionic Field . . . . . . . . . . . . . . . . . . 2039.1 Partition Function of Interacting Fermions . . . . . . . . . . . . . . . . . . . 2039.2 Ideal Fermi Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2059.3 Repulsive Fermions: Hartree-Fock Approximation . . . . . . . . . . . . . 2089.4 Attractive Fermions: BCS Approximation . . . . . . . . . . . . . . . . . . . 209

9.4.1 Condensate Fraction of Cooper Pairs . . . . . . . . . . . . . . . . . 2129.4.2 Ginzburg-Landau Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 2149.4.3 BCS-BEC Crossover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

9.5 Solved Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222

Appendix A: Dirac Delta Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

Appendix B: Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

Appendix C: Laplace Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239

Contents xi

Page 11: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 1The Origins of Modern Physics

In this chapter we review themain results of earlymodern physics (whichwe supposethe reader learned in previous introductory courses), namely the special relativity ofEinstein and the old quantum mechanics due to Planck, Bohr, Schrödinger, andothers. The chapter gives also a brief description of the relevant axioms of quan-tum mechanics, historically introduced by Dirac and Von Neumann, and elementarynotions of quantum information. The chapter ends with elements of quantum statis-tical mechanics.

1.1 Special Relativity

In 1887 Albert Michelson and Edward Morley made a break-thought experiment ofoptical interferometry showing that the speed of light in the vacuum is

c = 3 × 108 m/s , (1.1)

independently on the relative motion of the observer (here we have reported anapproximated value of c which is correct within three digits). Two years later, HenryPoincaré suggested that the speed of light is the maximum possible value for anykind of velocity. On the basis of previous ideas of George Francis FitzGerald, in1904 Hendrik Lorentz found that the Maxwell equations of electromagnetism areinvariant with respect to this kind of space-time transformations

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_1

1

Page 12: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2 1 The Origins of Modern Physics

x ′ = x − vt√1 − v2

c2

(1.2)

y′ = y (1.3)

z′ = z (1.4)

t ′ = t − vx/c2√1 − v2

c2

, (1.5)

which are called Lorentz (or Lorentz-FitzGerald) transformations. This researchactivity on light and invariant transformations was summarized in 1905 by AlbertEinstein, who decided to adopt two striking postulates:

(i) the law of physics are the same for all inertial frames;(ii) the speed of light in the vacuum is the same in all inertial frames.

From these two postulates Einstein deduced that the laws of physics are invariant withrespect to Lorentz transformations but the law of Newtonian mechanics (which arenot) must be modified. In this way Einstein developed a new mechanics, the specialrelativisticmechanics, which reduces to theNewtonianmechanics when the involvedvelocity v is much smaller than the speed of light c. One of the amazing results ofrelativistic kinematics is the length contraction: the length L of a rod measured byan observer which moves at velocity v with respect to the rod is given by

L = L0

√1 − v2

c2, (1.6)

where L0 is the proper length of the rod. Another astonishing result is the timedilatation: the time interval T of a clock measured by an observer which moves atvelocity v with respect to the clock is given by

T = T0√1 − v2

c2

, (1.7)

where T0 is the proper time interval of the clock.We conclude this section by observing that, according to the relativisticmechanics

of Einstein, the energy E of a particle of rest mass m and linear momentum p = |p|is given by

E =√p2c2 + m2c4 . (1.8)

If the particle has zero linear momentum p, i.e. p = 0, then

E = mc2 , (1.9)

Page 13: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.1 Special Relativity 3

which is the rest energy of the particle. Instead, if the linear momentum p is finitebut the condition pc/(mc2) � 1 holds one can expand the square root finding

E = mc2 + p2

2m+ O(p4) , (1.10)

which shows that the energy E is approximated by the sum of two contributions: therest energy mc2 and the familiar non-relativistic kinetic energy p2/(2m). In the caseof a particle with zero rest mass, i.e. m = 0, the energy is given by

E = pc , (1.11)

which is indeed the energyof light photons.We stress that all the predictions of specialrelativistic mechanics have been confirmed by experiments. Also many predictionsof general relativity, which generalizes the special relativity taking into account thetheory of gravitation, have been verified experimentally and also used in applications(for instance, someglobal positioning system (GPS) devices includegeneral relativitycorrections).

1.2 Quantum Mechanics

Historically the beginning of quantum mechanics is set in the year 1900 when MaxPlanck found that the only way to explain the experimental results of the electro-magnetic spectrum emitted by hot solid bodies is to assume that the energy E of theradiation with frequency ν emitted from the walls of the body is quantized accordingto

E = hν n , (1.12)

where n = 0, 1, 2, . . . is an integer quantum number. With the help of statisticalmechanics Planck derived the following expression

ρ(ν) = 8π2

c3ν2 hν

eβhν − 1(1.13)

for the electromagnetic energy density per unit of frequency ρ(ν) emitted by the hotbody at the temperature T , with kB = 1.38 × 10−23 J/K the Boltzmann constant,c = 3 × 108 m/s the speed of light in the vacuum. This formula, known as Plancklaw of the black-body radiation, is in very good agreement with experimental data.Note that a hot solid body can be indeed approximated by the so-called black body,that is an idealized physical body that absorbs all incident electromagnetic radiationand it is also the best possible emitter of thermal radiation. The constant h derivedby Plank from the interpolation of experimental data of radiation reads

Page 14: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4 1 The Origins of Modern Physics

h = 6.63 × 10−34 J s . (1.14)

This parameter is called Planck constant. Notice that often one uses instead thereduced Planck constant

� = h

2π= 1.06 × 10−34 J s . (1.15)

Few years later the black-body formulation of Planck, in 1905, Albert Einsteinsuggested that not only the electromagnetic radiation is emitted by hot bodies ina quantized form, as found by Planck, but that indeed electromagnetic radiation isalways composed of light quanta, called photons, with discrete energy

ε = hν . (1.16)

Einstein used the concept of photon to explain the photoelectric effect and predictedthat the kinetic energy of an electron emitted by the surface of a metal after beingirradiated is given by

1

2mv2 = hν − W , (1.17)

where W is work function of the metal (i.e. the minimum energy to extract theelectron from the surface of the metal), m is the mass of the electron, and v is theemission velocity of the electron. This prediction clearly implies that the minimalradiation frequency to extract electrons from a metal is νmin = W/h. Subsequentexperiments confirmed the Einstein’s formula and gave a complementary measureof the Planck constant h.

In 1913Niels Bohr was able to explain the discrete frequencies of electromagneticemission of hydrogen atom under the hypothesis that the energy of the electronorbiting around the nucleus is quantized according to the formula

En = − me4

2ε20h2

1

n2= −13.6 eV

1

n2, (1.18)

where n = 1, 2, 3, . . . is the principal integer quantumnumber, e is the electric chargeof the electron and ε0 the dielectric constant of the vacuum. This expression showsthat the quantum states of the system are characterized by the quantum number n andthe ground-state (n = 1) has the energy −13.6 eV, which is the ionization energyof the hydrogen. According to the theory of Bohr, the electromagnetic radiation isemitted or absorbed when one electron has a transition from one energy level En toanother Em . In addition, the frequency ν of the radiation is related to the energies ofthe two states involved in the transition by

hν = En − Em . (1.19)

Page 15: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.2 Quantum Mechanics 5

Thus any electromagnetic transitionbetween twoquantumstates implies the emissionor the absorption of one photon with an energy hν equal to the energy difference ofthe involved states.

In 1922 Arthur Compton noted that the diffusion of X rays with electrons (Comp-ton effect) is a process of scattering between photons of X rays and electrons. Thephoton of X rays has the familiar energy

ε = hν = hc

λ, (1.20)

with λ wavelength of the photon, because the speed of light is given by c = λν. Byusing the connection relativistic between energy ε and linear momentum p = |p| fora particle with zero rest mass, Compton found

ε = pc , (1.21)

where the linear momentum of the photon can be then written as

p = h

λ. (1.22)

By applying the conservation of energy and linear momentum to the scatteringprocess Compton got the following expression

λ′ − λ = h

mc(1 − cos(θ)) (1.23)

for the wavelength λ′ of the diffused photon, with θ the angle between the incomingphoton and the outcoming one. This formula is in full agreement with experimentalresults.

Inspired by the wave-particle behavior exhibited by the light, in 1924 Louis deBroglie suggested that also the matter, and in particular the electron, has wave-likeproperties. He postulated that the relationship

λ = h

p(1.24)

applies not only to photons but also to material particles. In general, p is the linearmomentum (particle property) and λ the wavelength (wave property) of the quantumentity, that is usually called quantum particle.

Two years later the proposal of de Broglie, in 1926, Erwin Schrödinger went tothe extremes of the wave-particle duality and introduced the followingwave equationfor one electron under the action of an external potential U (r)

i�∂

∂tψ(r, t) =

[− �

2

2m∇2 +U (r)

]ψ(r, t) , (1.25)

Page 16: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6 1 The Origins of Modern Physics

where ∇2 = ∂2

∂x2 + ∂2

∂y2 + ∂2

∂z2 is the Laplacian operator. This equation is now knownas the Schrödinger equation. In the case of the hydrogen atom, where

U (r) = − e2

4πε0 r2, (1.26)

Schrödinger showed that setting

ψ(r, t) = Rn(r) e−i En t/� (1.27)

one finds [− �

2

2m∇2 − e2

4πε0 r2

]Rn(r) = EnRn(r) (1.28)

the stationary Schrödinger equation for the radial eigenfunction Rn(r) with eigen-value En given exactly by the Bohr formula (1.18). Soon after, it was understood thatthe stationary Schrödinger equation of the hydrogen atom satisfies a more generaleigenvalue problem, namely

[− �

2

2m∇2 − e2

4πε0 r2

]ψnlml (r) = Enψnlml (r) , (1.29)

where ψnlml (r) is a generic eigenfunction of the problem, which depends on threequantum numbers: the principal quantum number n = 1, 2, 3, . . ., the angular quan-tum number l = 0, 1, 2, . . . , n − 1, and third-component angular quantum numberml = −l,−l + 1, . . . , l − 1, l. The generic eigenfunction of the electron in thehydrogen atom in spherical coordinates is given by

ψnlml (r, θ,φ) = Rnl(r) Ylml (θ,φ) ,

where Rnl(r) is the radial wavefunction while Ylml (θ,φ) is the angular wavefunction.Notice that inEq. (1.28)we have Rn(r) = Rn0(r). It is important to stress that initiallySchrödinger thought that ψ(r) was a matter wave, such that |ψ(r, t)|2 gives the localdensity of electrons in the position r at time t . It was Max Born that correctlyinterpreted ψ(r, t) as a probability field, where |ψ(r, t)|2 is the local probabilitydensity of finding one electron in the position r at time t , with the normalizationcondition ∫

d3r |ψ(r, t)|2 = 1 . (1.30)

In the case of N particles the probabilistic interpretation of Born becomes cru-cial, with �(r1, r2, . . . , rN , t) the many-body wavefunction of the system, such that|�(r1, r2, . . . , rN , t)|2 is the probability of finding at time t one particle in the posi-tion r1, another particle in the position r2, and so on.

Page 17: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.2 Quantum Mechanics 7

In the same year of the discovery of the Schrödinger equation Max Born, PasqualJordan and Werner Heisenberg introduced the matrix mechanics. According to thistheory the position r and the linear momentum p of an elementary particle are notvectors composed of numbers but instead vector composed of matrices (operators)which satisfies strange commutation rules, i.e.

r = (x, y, z) , p = ( px , py, pz) , (1.31)

such that[r, p] = i� , (1.32)

where the hat symbol is introduced to denote operators and [ A, B] = A B − B Ais the commutator of generic operators A and B. By using their theory Born, Jor-dan and Heisenberg were able to obtain the energy spectrum of the hydrogen atomand also to calculate the transition probabilities between two energy levels. Soonafter, Schrödinger realized that the matrix mechanics is equivalent to his wave-likeformulation: introducing the quantization rules

r = r , p = −i�∇ , (1.33)

for any function f (r) one finds immediately the commutation rule

(r · p − p · r) f (r) = (−i�r · ∇ + i�∇ · r) f (r) = i� f (r) . (1.34)

Moreover, starting from classical Hamiltonian

H = p2

2m+U (r) , (1.35)

the quantization rules give immediately the quantum Hamiltonian operator

H = − �2

2m∇2 +U (r) (1.36)

from which one can write the time-dependent Schrödinger equation as

i�∂

∂tψ(r, t) = Hψ(r, t) . (1.37)

In 1927 the wave-like behavior of electrons was eventually demonstrated by Clin-ton Davisson and Lester Germer, who observed the diffraction of a beam of electronsacross a solid crystal. The diffracted beam shows intensity maxima when the follow-ing relationship is satisfied

2 d sin(φ) = n λ , (1.38)

Page 18: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8 1 The Origins of Modern Physics

where n is an integer number, λ is the de Broglie wavelength of electrons, d is theseparation distance of crystal planes and φ is the angle between the incident beam ofelectrons and the surface of the solid crystal. The condition for a maximum diffractedbeam observed in this experiment corresponds indeed to the constructive interferenceof waves, which is well know in optics (Bragg condition).

1.2.1 Axioms of Quantum Mechanics

The axiomatic formulation of quantum mechanics was set up by Paul Maurice Diracin 1930 and John von Neumann in 1932. The basic axioms are the following:Axiom1. The state of a quantumsystem is describedby aunitary vector |ψ〉belongingto a separable complex Hilbert space.Axiom 2. Any observable of a quantum system is described by a self-adjoint linearoperator F acting on the Hilbert space of state vectors.Axiom 3. The possible measurable values of an observable F are its eigenvalues f ,such that

F | f 〉 = f | f 〉

with | f 〉 the corresponding eigenstate. Note that the observable F admits the spectralresolution

F =∑f

| f 〉 f 〈 f | ,

which is quite useful in applications, as well as the spectral resolution of the identity

I =∑f

| f 〉〈 f | .

Axiom 4. The probability p of finding the state |ψ〉 in the state | f 〉 is given by

p = |〈 f |ψ〉|2 ,

where the complex probability amplitude 〈 f |ψ〉 denotes the scalar product of thetwo vectors. This probability p is also the probability of measuring the value f ofthe observable F when the system in the quantum state |ψ〉. Notice that both |ψ〉and | f 〉 must be normalized to one. Often it is useful to introduce the expectationvalue (mean value or average value) of an observable F with respect to a state |ψ〉as 〈ψ|F |ψ〉. Moreover, from this Axiom 4 it follows that the wavefunction ψ(r) canbe interpreted as

ψ(r) = 〈r|ψ〉 ,

that is the probability amplitude of finding the state |ψ〉 in the position state |r〉.

Page 19: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.2 Quantum Mechanics 9

Axiom 5. The time evolution of states and observables of a quantum system withHamiltonian H is determined by the unitary operator

U (t) = exp (−i H t/�) ,

such that |ψ(t)〉 = U (t)|ψ〉 is the time-evolved state |ψ〉 at time t and F(t) =U−1(t)FU (t) is the time-evolved observable F at time t . From this axiom one findsimmediately the Schrödinger equation

i�∂

∂t|ψ(t)〉 = H |ψ(t)〉

for the state |ψ(t)〉, and the Heisenberg equation

i�∂

∂tF(t) = [F(t), H ]

for the observable F(t).

1.2.2 Quantum Information

The qubit, or quantum bit, is the quantum analogue of a classical bit. It is the unitof quantrum information, namely a tw-level quantum system. There are many two-level system which can be used as a physical realization of the qubit. For instance:horizontal and vertical polarizations of light, ground and excited states of atoms ormolecules or nuclei, left and right wells of a double-well potential, up and downspins of a particle.

The two basis states of the qubit are usually denoted as |0〉 and |1〉. They can bewritten as

|0〉 =(10

), |1〉 =

(01

), (1.39)

by using a vector representation, where clearly

〈0| = (1, 0) , 〈1| = (0, 1) , (1.40)

and also

〈0|0〉 = (1, 0)

(10

)= 1 , (1.41)

〈0|1〉 = (1, 0)

(01

)= 0 , (1.42)

Page 20: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

10 1 The Origins of Modern Physics

〈1|0〉 = (0, 1)

(10

)= 0 , (1.43)

〈1|1〉 = (0, 1)

(01

)= 1 , (1.44)

while

|0〉〈0| =(10

)(1, 0) =

(1 00 0

), (1.45)

|0〉〈1| =(10

)(0, 1) =

(0 10 0

), (1.46)

|1〉〈0| =(01

)(1, 0) =

(0 01 0

), (1.47)

|1〉〈1| =(01

)(0, 1) =

(0 00 1

). (1.48)

Of course, instead of |0〉 and |1〉 one can choose other basis states. For instance:

|+〉 = 1√2

(|0〉 + |1〉) = 1√2

(11

), |−〉 = 1√

2(|0〉 − |1〉) = 1√

2

(1

−1

),

(1.49)but also

|i〉 = 1√2

(|0〉 + i |1〉) = 1√2

(1i

), | − i〉 = 1√

2(|0〉 − i |1〉) = 1√

2

(1−i

).

(1.50)

A pure qubit |ψ〉 is a linear superposition (superposition state) of the basis states|0〉 and |1〉, i.e.

|ψ〉 = α |0〉 + β |1〉 , (1.51)

where α and β are the probability amplitudes, usually complex numbers, such that

|α|2 + |β|2 = 1 . (1.52)

A quantum gate (or quantum logic gate) is a unitary operator U acting on qubits.Among the quantum gates acting on a single qubit there are: the identity gate

I =(1 00 1

), (1.53)

the Hadamard gate

H = 1√2

(1 11 −1

), (1.54)

Page 21: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.2 Quantum Mechanics 11

the not gate (also called Pauli X gate)

X =(0 11 0

)= σ1 , (1.55)

the Pauli Y gate

Y =(0 −ii 0

)= σ2 , (1.56)

and the Pauli Z gate

Z =(1 00 −1

)= σ3 . (1.57)

It is strightforward to deduce the properties of the quantum gates. For example, oneeasily finds

H |0〉 = |+〉 , H |1〉 = |−〉 , (1.58)

but alsoH |+〉 = |0〉 , H |−〉 = |1〉 . (1.59)

A N -qubit |�N 〉, also called a quantum register, is a quantum state characterizedby

|�N 〉 =∑

α1...αN

cα1...αN |α1〉 ⊗ . . . ⊗ |αN 〉 , (1.60)

where |αi 〉 is a single qubit (with αi = 0, 1), and

∑α1...αN

|cα1...αN |2 = 1 . (1.61)

In practice, the N -qubit describes N two-state configurations. The more general2-qubit |�2〉 is consequently given by

|�2〉 = c00|00〉 + c01|01〉 + c10|10〉 + c11|11〉 , (1.62)

where we use |00〉 = |0〉⊗|0〉, |01〉 = |0〉⊗|1〉, |10〉 = |1〉⊗|0〉, and |11〉 = |1〉⊗|1〉to simplify the notation.

For the basis states of 2-qubits we one introduce the following vector representa-tion

|00〉 =

⎛⎜⎜⎝

1000

⎞⎟⎟⎠ , |01〉 =

⎛⎜⎜⎝

0100

⎞⎟⎟⎠ , |10〉 =

⎛⎜⎜⎝

0010

⎞⎟⎟⎠ , |11〉 =

⎛⎜⎜⎝

0001

⎞⎟⎟⎠ . (1.63)

Page 22: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

12 1 The Origins of Modern Physics

Moreover, quantum gates can be introduced also for 2-qubits. Among the quantumgates acting on a 2-qubit there are: the identity gate

I =

⎛⎜⎜⎝

1 0 0 00 1 0 00 0 1 00 0 0 1

⎞⎟⎟⎠ (1.64)

and the CNOT (also called controlled-NOT) gate

ˆCNOT =

⎛⎜⎜⎝

1 0 0 00 1 0 00 0 0 10 0 1 0

⎞⎟⎟⎠ . (1.65)

Notice that the action of the CNOT gate on a state |αβ〉 = |α〉|β〉 is such that the“target” qubit |β〉 changes only if the “control” qubit |α〉 is |1〉.

The 2-qubit |�2〉 is called separable if it can be written as the tensor product oftwo generic single qubits |ψA〉 and |ψB〉, i.e.

|�2〉 = |ψA〉 ⊗ |ψB〉 . (1.66)

If this is not possible, the state |�2〉 is called entangled. Examples of separable statesare

|00〉 = |0〉 ⊗ |0〉 , (1.67)

or|01〉 = |0〉 ⊗ |1〉 , (1.68)

but also1√2

(|01〉 + |11〉) = 1√2

(|0〉 + |1〉) ⊗ |1〉 . (1.69)

Examples of entangled states are instead

1√2

(|01〉 ± |10〉) . (1.70)

but also1√2

(|00〉 ± |11〉) , (1.71)

which are called Bell states.Remarkably, a CNOT gate acting on a separable state can produce an entagled

state, and viceversa. For instance:

Page 23: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.2 Quantum Mechanics 13

ˆCNOT1√2

(|00〉 + |10〉) = 1√2

(|00〉 + |11〉) , (1.72)

whileˆCNOT

1√2

(|00〉 + |11〉) = 1√2

(|00〉 + |10〉) . (1.73)

1.3 Quantum Statistical Mechanics

Statistical mechanics aims to describe macroscopic properties of complex systemsstarting from their microscopic components by using statistical averages. Statisticalmechanics must reproduce the general results of Thermodynamics both at equilib-rium and out of equilibrium. Clearly, the problem is strongly simplified if the systemis at thermal equilibrium. Here we discuss only quantum systems at thermal equilib-rium and consider a many-body quantum system of identical particles characterizedby the Hamiltonian H such that

H |E (N )i 〉 = E (N )

i |E (N )i 〉 , (1.74)

where |E (N )i 〉 are the eigenstates of H for a fixed number N of identical particles and

ENi are the corresponding eigenenergies.

1.3.1 Microcanonical Ensemble

In the microcanonical ensemble the quantum many-body system in a volume V hasa fixed number N of particles and also a fixed energy E . In this case the HamiltonianH admits the spectral decomposition

H =∑i

E (N )i |E (N )

i 〉〈E (N )i | , (1.75)

and one defines the microcanonical density operator as

ρ = δ(E − H) , (1.76)

where δ(x) is the Dirac delta function. This microcanonical density operator ρ hasthe spectral decomposition

ρ =∑i

δ(E − E (N )i )|E (N )

i 〉〈E (N )i | . (1.77)

Page 24: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

14 1 The Origins of Modern Physics

The key quantity in the microcanonical ensemble is the density of states (or micro-canonical volume) W given by

W = Tr [ρ] = Tr [δ(E − H)] , (1.78)

namelyW =

∑i

δ(E − E (N )i ) . (1.79)

The ensemble average of an observable described by the self-adjunct operator A isdefined as

〈 A〉 = Tr [ A ρ]Tr [ρ] = 1

W

∑i

A(N )i i δ(E − E (N )

i ) , (1.80)

where A(N )i i = 〈E (N )

i | A|E (N )i 〉. The connection with Equilibrium Thermodynamics

is given by the formulaS = kB ln (W ) , (1.81)

which introduces the entropy S as a function of energy E , volume V and number Nof particles. Note that Eq. (1.81) was discovered by Ludwig Boltzmann in 1872 andkB = 1.38× 10−23 J/K is the Boltzmann constant. From the entropy S(E, V, N ) theabsolute temperature T , the pressure P and the chemical potential μ are obtained as

1

T=

(∂S

∂E

)

V,N

, P = T

(∂S

∂V

)

E,N

, μ = −T

(∂S

∂N

)

E,V

, (1.82)

which are familiar relationships of Equilibrium Thermodynamics such that

dS = 1

TdE + P

TdV − μ

TdN . (1.83)

In fact, in the microcanonical ensemble the independent thermodynamic variablesare E , N and V , while T , P and μ are dependent thermodynamic variables.

1.3.2 Canonical Ensemble

In the canonical ensemble the quantum system in a volume V has a fixed number Nof particles and a fixed temperature T . In this case one defines the canonical densityoperator as

ρ = e−β H , (1.84)

with β = 1/(kBT ). Here ρ has the spectral decomposition

Page 25: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.3 Quantum Statistical Mechanics 15

ρ =∑i

e−βE (N )i |E (N )

i 〉〈E (N )i | . (1.85)

The key quantity in the canonical ensemble is the canonical partition function (orcanonical volume) ZN given by

ZN = Tr [ρ] = Tr [e−β H ] , (1.86)

namely

ZN =∑i

e−βE (N )i . (1.87)

The ensemble average of an observable A is defined as

〈 A〉 = Tr [ A ρ]Tr [ρ] = 1

ZN

∑i

A(N )i i e−βE (N )

i , (1.88)

where A(N )i i = 〈E (N )

i | A|E (N )i 〉. Notice that the definition of canonical-ensemble aver-

age is the same of the microcanonical-ensemble average but the density of state ρ isdifferent in the two ensembles. The connection with Equilibrium Thermodynamicsis given by the formula

ZN = e−βF , (1.89)

which introduces theHelmholtz free energy F as a function of temperature T , volumeV and number N of particles. From the Helholtz free energy F(T, V, N ) the entropyS, the pressure P and the chemical potential μ are obtained as

S = −(

∂F

∂T

)

V,N

, P = −(

∂F

∂V

)

T,N

, μ =(

∂F

∂N

)

T,V

, (1.90)

which are familiar relationships of Equilibrium Thermodynamics such that

dF = −SdT − PdV + μdN . (1.91)

In fact, in the canonical ensemble the independent thermodynamic variables are T ,N and V , while S, P and μ are dependent thermodynamic variables.

1.3.3 Grand Canonical Ensemble

In the grand canonical ensemble the quantum system in a volume V has a fixedtemperature T and a fixed chemical potential μ. In this case the Hamiltonian H hasthe spectral decomposition

Page 26: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

16 1 The Origins of Modern Physics

H =∞∑N=0

∑i

E (N )i |E (N )

i 〉〈E (N )i | , (1.92)

which is a generalization of Eq. (1.75), and one introduces the total number operatorN such that

N |E (N )i 〉 = N |E (N )

i 〉 , (1.93)

and consequently N has the spectral decomposition

N =∞∑N=0

∑i

N |E (N )i 〉〈E (N )

i | . (1.94)

For the grand canonical ensemble one defines the grand canonical density operatoras

ρ = e−β(H−μN ) , (1.95)

with β = 1/(kBT ) and μ the chemical potential. Here ρ has the spectral decompo-sition

ρ =∞∑N=0

∑i

e−β(E (N )i −μN )|E (N )

i 〉〈E (N )i | =

∞∑N=0

zN∑i

e−βE (N )i |E (N )

i 〉〈E (N )i | , (1.96)

where z = eβμ is the fugacity. The key quantity in the grand canonical ensemble isthe grand canonical partition function (or grand canonical volume) Z given by

Z = Tr [ρ] = Tr [e−β(H−μN )] , (1.97)

namely

Z =∞∑N=0

∑i

e−β(E (N )i −μN ) =

∞∑N=0

zN ZN . (1.98)

The ensemble average of an observable A is defined as

〈 A〉 = Tr [ A ρ]Tr [ρ] = 1

Z∞∑N=0

∑i

A(N )i i e−(βE (N )

i −μN ) , (1.99)

where A(N )i i = 〈E (N )

i | A|E (N )i 〉. Notice that the definition of grand canonical-ensemble

average is the same of both microcanonical-ensemble average and canonical-ensemble average but the density of state ρ is different in the three ensembles. Theconnection with Equilibrium Thermodynamics is given by the formula

Z = e−β� , (1.100)

Page 27: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.3 Quantum Statistical Mechanics 17

which introduces the grand potential � as a function of temperature T , volume Vand chemical potential μ. From the grand potential �(T, V,μ) the entropy S, thepressure P and the average number N of particles are obtained as

S = −(

∂�

∂T

)

V,μ

, P = −(

∂�

∂V

)

T,μ

, N = −(

∂�

∂μ

)

V,T

, (1.101)

which are familiar relationships of Equilibrium Thermodynamics such that

d� = −S dT − P dV − N dμ . (1.102)

In fact, in the grand canonical ensemble the independent thermodynamic variablesare T , μ and V , while S, P and N are dependent thermodynamic variables.

To conclude this section we observe that in the grand canonical ensemble, insteadof working with eigenstates |E (N )

i 〉 of H at fixed number N of particles, one canwork with multi-mode Fock states

|n0 n1 n2 . . . n∞〉 = |n0〉 ⊗ |n1〉 ⊗ |n2〉 ⊗ . . . ⊗ |n∞〉 , (1.103)

where |nα〉 is the single-mode Fock state which describes nα particles in the single-mode state |α〉 with α = 0, 1, 2, . . .. The trace Tr which appears in Eq. (1.97) isindeed independent on the basis representation. We analyze this issue in Chaps. 2and 7.

1.4 Solved Problems

Problem 1.1The electromagnetic radiation of the black body has the following energy densityper unit of frequency

ρ(ν) = 8π2

c3ν2 hν

eβhν − 1,

where c is the speed of light in vacuum, h is the Planck constant, and β = 1/(kBT )

with T absolute temperature and kB Boltzmann constant. Determine the correspond-ing energy density per unit of wavelength.

SolutionThe energy density ρ(ν) is such that

E =∫ ∞

0ρ(ν) dν

Page 28: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

18 1 The Origins of Modern Physics

represents the energy of the radiation per unit of volume, i.e. the energy density. Thelinear frequency ν is related to the wavelength λ by the expression

λν = c .

In practice, ν = cλ−1, from which

dν = −cλ−2dλ .

By changing variable the energy density becomes

E =∫ ∞

0

8π2

λ5

hc

eβhc/λ − 1dλ ,

and consequently the energy density per unit of wavelength reads

ρ(λ) = 8π2

λ5

hc

eβhc/λ − 1,

such that

E =∫ ∞

0ρ(λ) dλ .

Problem 1.2Calculate the number of photon emitted in 4 s by a lamp of 10 W which radiates1% of its energy as monochromatic light with wavelength 6000 × 10−10 m (orangelight).

SolutionThe energy of one photon of wavelength λ and linear frequency ν is given by

ε = hν = hc

λ,

whereh = 6.63 × 10−34 J s

is the Planck constant. In our problem one gets

ε = hc

λ= 6.62 × 10−34 J s

3 × 108 m/s

6 × 103 × 10−10 m= 3.3 × 10−19 J .

During the period �t = 4 s the energy of the lamp with power P = 10 W is

E = P �t = 10 J/s × 4 s = 40 J .

Page 29: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

1.4 Solved Problems 19

The radiation energy is instead

Erad = E × 1% = E × 1

100= 40

100J = 0.4 J .

The number of emitted photons is then

N = Erad

ε= 0.4 J

3.3 × 10−19 J= 1.2 × 1018 .

Problem 1.3On a photoelectric cell it arrives a beam of light with wavelength 6500 × 10−10

m and energy 106 erg per second [1 erg = 10−7 J]. This energy is entirely used toproduce photoelectrons. Calculate the intensity of the electric current which flowsin the electric circuit connected to the photoelectric cell.

SolutionThe wavelength can be written as

λ = 6.5 × 103 × 10−10 m = 6.5 × 10−7 m .

The energy of the beam of light can be written as

E = 106 erg = 106 × 10−7 J = 10−1J .

The time interval is�t = 1 s .

The energy of a single photon reads

ε = hc

λ= 6.6 × 10−34 3 × 108

6.5 × 10−7J = 3 × 10−19 J ,

where h is the Planck constant and c the speed of light in the vacuum.The number of photons is thus given by

N = E

ε= 10−1 J

3 × 10−19 J= 3.3 × 1017 .

If one photon produces one electron with electric charge e = −1.6 × 10−19 C, theintensity of electric current is easily obtained:

I = |e|N�t

= 1.6 × 10−19 C × 3.3 × 1017

1 s= 5.5. × 10−2 A .

Page 30: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

20 1 The Origins of Modern Physics

To conclude we observe that with wavelength 6500 × 10−10 m the photoelectriceffect is possible only if the work function of the sample is reduced, for instance byusing an external electric field.

Problem 1.4Prove that the simple 2-qubit state

|�2〉 = c00|01〉 + c10|10〉

is separable if and only if c01 = 0 or c10 = 0.

SolutionLet us consider two generic 1-qubits, given by

|ψA〉 = αA|0〉 + βA|1〉 , |ψB〉 = αB |0〉 + βB |1〉 .

Then|ψA〉|ψB〉 = αAαB |00〉 + αAβB |01〉 + βAαB |10〉 + βAβB |11〉 .

If the 2-qubit state |�2〉 is separable, it can be written as

|�2〉 = |ψA〉 ⊗ |ψB〉 ,

namely

c00|01〉 + c10|10〉 = αAαB |00〉 + αAβB |01〉 + βAαB |10〉 + βAβB |11〉 .

It follows

αAαB = βAβB = 0 , αAβB = c01 , βAαB = c10 .

These conditions are all satisfied only if c01 = 0 or c10 = 0, or both.

Further Reading

For special relativity:W. Rindler, Introduction to Special Relativity (Oxford Univ. Press, Oxford, 1991).For quantum mechanics:C. Cohen-Tannoudji, B. Dui, and F. Laloe, Quantum Mechanics (Wiley, New York,1991).P.A.M. Dirac, The Principles of Quantum Mechanics (Oxford University Press,Oxford, 1982).For quantum information:M. Le Bellac, A Short Introduction to Quantum Information and Quantum Compu-tation (Cambridge Univ. Press, Cambridge, 2006).For quantum statistical mechanics:K. Huang, Statistical Mechanics (Wiley, New York, 1987).

Page 31: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 2Second Quantization of Light

In this chapter we discuss the quantization of electromagnetic waves, which we alsodenoted as light (visible or invisible to human eyes). After reviewing classical andquantum properties of the light in the vacuum, we discuss the so-called second quan-tization of the light field showing that this electromagnetic field can be expressed asa infinite sum of harmonic oscillators. These oscillators, which describe the possiblefrequencies of the radiation field, are quantized by introducing creation and annihi-lation operators acting on the Fock space of number representation. We analyze theFock states of the radiation field and compare them with the coherent states. Finally,we consider two enlightening applications: the Casimir effect and the radiation fieldat finite temperature.

2.1 Electromagnetic Waves

The light is an electromagnetic field characterized by the coexisting presence of anelectric field E(r, t) and a magnetic field B(r, t). From the equations of James ClerkMaxwell in vacuum and in the absence of sources, given by

∇ · E = 0 , (2.1)

∇ · B = 0 , (2.2)

∇ ∧ E = −∂B∂t

, (2.3)

∇ ∧ B = ε0μ0∂E∂t

, (2.4)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_2

21

Page 32: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

22 2 Second Quantization of Light

one finds that the coupled electric and magnetic fields satisfy the wave equations

(1

c2∂2

∂t2− ∇2

)E = 0 , (2.5)

(1

c2∂2

∂t2− ∇2

)B = 0 , (2.6)

where

c = 1√ε0μ0

= 3 × 108 m/s (2.7)

is the speed of light in the vacuum. Note that the dielectric constant (electric per-mittivity) ε0 and the magnetic constant (magnetic permeability) μ0 are respectivelyε0 = 8.85 × 10−12 C2/(Nm2) and μ0 = 4π × 10−7 Vs/(Am). Equations (2.5) and(2.6), which are fully confirmed by experiments, admit monochromatic complexplane wave solutions

E(r, t) = E0 ei(k·r−ωt) , (2.8)

B(r, t) = B0 ei(k·r−ωt) , (2.9)

where k is the wavevector and ω the angular frequency, such that

ω = c k , (2.10)

is the dispersion relation, with k = |k| is thewavenumber. FromMaxwell’s equationsone finds that the vectors E and B are mutually orthogonal and such that

E = cB , (2.11)

where E = |E| and B = |B|. In addition they are transverse fields, i.e. orthogonal tothe wavevector k, which gives the direction of propagation of the wave. Notice thatEq. (2.11) holds formonochromatic planewaves but not in general. For completeness,let us remind that the wavelength λ is given by

λ = 2π

k, (2.12)

and that the linear frequency ν and the angular frequency ω = 2πν are related to thewavelength λ and to the wavenumber k by the formulas (Fig. 2.1)

λ ν = ω

k= c . (2.13)

Page 33: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.1 Electromagnetic Waves 23

Fig. 2.1 Plot of a sinusoidal electromagnetic wave moving along the x axis

2.1.1 Photons

At the beginning of quantum mechanics Satyendra Nath Bose and Albert Einsteinsuggested that the light can be described as a gas of photons. A single photon of amonochromatic wave has the energy

ε = hν = �ω , (2.14)

where h = 6.63×10−34 J s is the Planck constant and � = h/(2π) = 1.05×10−34 J sis the reduced Planck constant. The linear momentum of the photon is given by thede Broglie relations

p = h

λn = �k , (2.15)

where n is a unit vector in the direction of k. Clearly, the energy of the photon canbe written also as

ε = p c , (2.16)

which is the energy one obtains for a relativistic particle of energy

ε =√m2c4 + p2c2 , (2.17)

setting to zero the rest mass, i.e.m = 0. The total energy H of monochromatic waveis given by

H =∑s

�ω ns , (2.18)

where ns is the number of photons with angular frequency ω and polarization s in themonochromatic electromagnetic wave. Note that in general there are two possiblepolarizations: s = 1, 2, corresponding to two linearly independent orthogonal unitvectors ε1 and ε2 in the plane perpendicular to the wavevector k.

A generic electromagnetic field is the superposition of many monochromaticelectromagneticwaves. Callingωk the angular frequency of themonochromaticwavewith wavenumber k, the total energy H of the electromagnetic field is

H =∑k

∑s

�ωk nks , (2.19)

Page 34: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

24 2 Second Quantization of Light

where nks is the number of photons with wavevector k and polarization s. The resultsderived here for the electromagnetic field, in particular Eq. (2.19), are called semi-classical because they do not take into account the so-called “quantum fluctuationsof vacuum”, i.e. the following remarkable experimental fact: photons can emergefrom the vacuum of the electromagnetic field. To justify this property of the electro-magnetic field one must perform the so-called second quantization of the field.

2.1.2 Electromagnetic Potentials and Coulomb Gauge

In full generality the electric field E(r, t) and the magnetic field B(r, t) can beexpressed in terms of a scalar potential φ(r, t) and a vector potential A(r, t) asfollows

E = −∇φ − ∂A∂t

, (2.20)

B = ∇ ∧ A . (2.21)

Actually these equations do not determine the electromagnetic potentials uniquely,since for an arbitrary scalar function �(r, t) the so-called “gauge transformation”

φ → φ′ = φ + ∂�

∂t, (2.22)

A → A′ = A − ∇� , (2.23)

leaves the fields E and B unaltered. There is thus an infinite number of differentelectromagnetic potentials that correspond to a given configuration of measurablefields. We use this remarkable property to choose a gauge transformation such that

∇ · A = 0 . (2.24)

This condition defines the Coulomb (or radiation) gauge, and the vector field A iscalled transverse field. For a complex monochromatic plane wave

A(r, t) = A0 ei(k·r−ωt) (2.25)

the Coulomb gauge (2.24) givesk · A = 0 , (2.26)

i.e. A is perpendicular (transverse) to the wavevector k. In the vacuum and withoutsources, from the first Maxwell equation (2.1) and Eq. (2.20) one immediately finds

∇2φ + ∂

∂t(∇ · A) = 0 , (2.27)

Page 35: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.1 Electromagnetic Waves 25

and under the Coulomb gauge (2.24) one gets

∇2φ = 0 . (2.28)

Imposing that the scalar potential is zero at infinity, this Laplace’s equation has theunique solution

φ(r, t) = 0 , (2.29)

and consequently

E = −∂A∂t

, (2.30)

B = ∇ ∧ A . (2.31)

Thus, in the Coulomb gauge one needs only the electromagnetic vector potentialA(r, t) to obtain the electromagnetic field if there are no charges and no currents.Notice that hereE andB are transverse fields likeA, which satisfyEqs. (2.5) and (2.6).The electromagnetic field described by these equations is often called the radiationfield, and also the vector potential satisfies the wave equation

(1

c2∂2

∂t2− ∇2

)A = 0 . (2.32)

We now expand the vector potential A(r, t) as a Fourier series of monochromaticplane waves. The vector potential is a real vector field, i.e.A = A∗ and consequentlywe write

A(r, t) =∑k

∑s

[Aks(t)

eik·r√V

+ A∗ks(t)

e−ik·r√V

]εks , (2.33)

where Aks(t) and A∗ks(t) are the dimensional complex conjugate coefficients of the

expansion, the complex plane waves eik·r/√V normalized in a volume V are the

basis functions of the expansion, and εk1 and εk2 are two mutually orthogonal realunit vectors of polarization which are also orthogonal to k (transverse polarizationvectors).

Taking into account Eqs. (2.30) and (2.31) we get

E(r, t) = −∑k

∑s

[Aks(t)

eik·r√V

+ A∗ks(t)

e−ik·r√V

]εks , (2.34)

B(r, t) =∑k

∑s

[Aks(t)

eik·r√V

− A∗ks(t)

e−ik·r√V

]ik ∧ εks , (2.35)

where both vector fields are explicitly real fields. Moreover, inserting Eq. (2.33)into Eq. (2.32) we recover familiar differential equations of decoupled harmonicoscillators:

Page 36: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

26 2 Second Quantization of Light

Aks(t) + ω2k Aks(t) = 0 , (2.36)

with ωk = ck, which have the complex solutions

Aks(t) = Aks(0) e−iωk t . (2.37)

These are the complex amplitudes of the infinite harmonic normal modes of theradiation field.

A familiar result of electromagnetism is that the classical energy of the electro-magnetic field in vacuum is given by

H =∫

d3r(

ε0

2E(r, t)2 + 1

2μ0B(r, t)2

), (2.38)

namely

H =∫

d3r(ε0

2

(∂A(r, t)∂t

)2 + 1

2μ0

(∇ ∧ A(r, t))2)

(2.39)

by using the Maxwell equations in the Coulomb gauge (2.30) and (2.31). Insertinginto this expression Eq. (2.33) or Eqs. (2.34) and (2.35) into Eq. (2.38) we find

H =∑k

∑s

ε0ω2k

(A∗ks Aks + Aks A

∗ks

). (2.40)

It is now convenient to introduce dimensionless complex coefficients aks(t) anda∗ks(t) related to the dimensional complex coefficients Aks(t) and A∗

ks(t) by

Aks(t) =√

2ε0ωkaks(t) . (2.41)

A∗ks(t) =

√�

2ε0ωka∗ks(t) . (2.42)

In this way the energy H reads

H =∑k

∑s

�ωk

2

(a∗ksaks + aksa

∗ks

). (2.43)

This energy is actually independent on time: the time dependence of the complexamplitudes a∗

ks(t) and aks(t) cancels due to Eq. (2.37). Instead of using the complexamplitudes a∗

ks(t) and aks(t) one can introduce the real variables

Page 37: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.1 Electromagnetic Waves 27

qks(t) =√2�

ωk

1

2

(aks(t) + a∗

ks(t))

(2.44)

pks(t) = √2�ωk

1

2i

(aks(t) − a∗

ks(t))

(2.45)

such that the energy of the radiation field reads

H =∑k

∑s

(p2k,s

2+ 1

2ω2k q

2ks

). (2.46)

This energy resembles that of infinitely many harmonic oscillators with unitary massand frequency ωk . It is written in terms of an infinite set of real harmonic oscillators:two oscillators (due to polarization) for each mode of wavevector k and angularfrequency ωk .

2.2 Second Quantization of Light

In 1927 Paul Dirac performed the quantization of the classical Hamiltonian (2.46)by promoting the real coordinates qks and the real momenta pks to operators:

qks → qks , (2.47)

pks → pks , (2.48)

satisfying the commutation relations

[qks, pk′s ′ ] = i� δk,k′ δs,s ′ , (2.49)

where [ A, B] = A B − B A. The quantum Hamiltonian is thus given by

H =∑k

∑s

(p2k,s

2+ 1

2ω2k q

2ks

). (2.50)

Following a standard approach for the canonical quantization of the Harmonicoscillator, we introduce annihilation and creation operators

aks =√

ωk

2�

(qks + i

ωkpks

), (2.51)

a+ks =

√ωk

2�

(qks − i

ωkpks

), (2.52)

Page 38: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

28 2 Second Quantization of Light

which satisfy the commutation relations

[aks, a+k′s ′ ] = δk,k′ δs,s ′ , (2.53)

and the quantum Hamiltonian (2.50) becomes

H =∑k

∑s

�ωk

(a+ks aks + 1

2

). (2.54)

Obviously this quantumHamiltonian can be directly obtained from the classical one,given by Eq. (2.43), by promoting the complex amplitudes aks and a∗

ks to operators:

aks → aks , (2.55)

a∗ks → a+

ks , (2.56)

satisfying the commutation relations (2.53).The operators aks and a+

ks act in the Fock space F , i.e. the infinite dimensionalHilbert space of “number representation” introduced in 1932 by Vladimir Fock. Ageneric state of this Fock space F is given by

| . . . nks . . . nk′s ′ . . . nk′′s ′′ . . . 〉 , (2.57)

meaning that there are nks photons with wavevector k and polarization s, nk′s ′ pho-tons with wavevector k′ and polarization s ′, nk′′s ′′ photons with wavevector k′′ andpolarization s ′′, et cetera. The Fock space F is given by

F = H0 ⊕ H1 ⊕ H2 ⊕ H3 ⊕ · · · ⊕ H∞ =∞⊕n=0

Hn , (2.58)

whereHn = H ⊗ H ⊗ · · · ⊗ H = H⊗n (2.59)

is the Hilbert space of n identical photons, which is n times the tensor product ⊗ ofthe single-photon Hilbert space H = H⊗1. Thus, F is the infinite direct sum ⊕ ofincreasing n-photon Hilbert statesHn , and we can formally write

F =∞⊕n=0

H⊗n . (2.60)

Notice that in the definition of the Fock space F one must include the space H0 =H⊗0, which is the Hilbert space of 0 photons, containing only the vacuum state

|0〉 = | . . . 0 . . . 0 . . . 0 . . . 〉 , (2.61)

Page 39: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.2 Second Quantization of Light 29

and its dilatations γ|0〉 with γ a generic complex number. The operators aks and a+ks

are called annihilation and creation operators because they respectively destroy andcreate one photon with wavevector k and polarization s, namely

aks | . . . nks . . . 〉 = √nks | . . . nks − 1 . . . 〉 , (2.62)

a+ks | . . . nks . . . 〉 = √

nks + 1 | . . . nks + 1 . . . 〉 . (2.63)

Note that these properties follow directly from the commutation relations (2.53).Consequently, for the vacuum state |0〉 one finds

aks |0〉 = 0F , (2.64)

a+ks |0〉 = |1ks〉 = |ks〉 , (2.65)

where 0F is the zero of the Fock space (usually indicated with 0), and |ks〉 is clearlythe state of one photon with wavevector k and polarization s, such that

〈r|ks〉 = eik·r√V

εks . (2.66)

From Eqs. (2.62) and (2.63) it follows immediately that

Nks = a+ks aks (2.67)

is the number operator which counts the number of photons in the single-particlestate |ks〉, i.e.

Nks | . . . nks . . . 〉 = nks | . . . nks . . . 〉 . (2.68)

Notice that the quantum Hamiltonian of the light can be written as

H =∑k

∑s

�ωk

(Nks + 1

2

), (2.69)

and this expression is very similar, but not equal, to the semiclassical formula (2.19).The differences are that Nks is a quantum number operator and that the energy Evac

of the vacuum state |0〉 is not zero but is instead given by

Evac =∑k

∑s

1

2�ωk . (2.70)

A quantum harmonic oscillator of frequency ωk has a finite minimal energy �ωk,which is called zero-point energy. In the case of the quantum electromagnetic fieldthere is an infinite number of harmonic oscillators and the total zero-point energy,given by Eq. (2.70), is clearly infinite. The infinite constant Evac is usually eliminatedby simply shifting to zero the energy associated to the vacuum state |0〉.

Page 40: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

30 2 Second Quantization of Light

The quantum electric and magnetic fields can be obtained from the classicalexpressions, Eqs. (2.34) and (2.35), taking into account the quantization of the clas-sical complex amplitudes aks and a∗

ks and their time-dependence, given by Eq. (2.37)and its complex conjugate. In this way we obtain

E(r, t) = i∑k

∑s

√�ωk

2ε0V

[aks e

i(k·r−ωk t) − a+ks e

−(ik·r−ωk t)]

εks , (2.71)

B(r, t) =∑k

∑s

√�

2ε0ωkV

[aks e

i(k·r−ωk t) − a+ks e

−i(k·r−ωk t)]ik|k| ∧ εks .

(2.72)

It is important to stress that the results of our canonical quantization of the radi-ation field suggest a remarkable philosophical idea: there is a unique quantum elec-tromagnetic field in the universe and all the photons we see are the massless particlesassociated to it. In fact, the quantization of the electromagnetic field is the first steptowards the so-called quantum field theory or second quantization of fields, where allparticles in the universe are associated to few quantum fields and their correspondingcreation and annihilation operators.

2.2.1 Fock versus Coherent States for the Light Field

Let us now consider for simplicity a linearly polarized monochromatic wave of theradiation field. For instance, let us suppose that the direction of polarization is givenby the vector ε. From Eq. (2.71) one finds immediately that the quantum electric fieldcan be then written in a simplified notation as

E(r, t) =√

�ω

2ε0Vi

[a ei(k·r−ωt) − a+ e−i(k·r−ωt)

]ε (2.73)

where ω = ωk = c|k|. Notice that, to simplify the notation, we have removed thesubscripts in the annihilation and creation operators a and a+. If there are exactly nphotons in this polarized monochromatic wave the Fock state of the system is givenby

|n〉 = 1√n!

(a+)n |0〉 . (2.74)

It is then straightforward to show, by using Eqs. (2.62) and (2.63), that

〈n|E(r, t)|n〉 = 0, (2.75)

Page 41: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.2 Second Quantization of Light 31

for all values of the photon number n, no matter how large. This result holds forall modes, which means then that the expectation value of the electric field in anymany-photon Fock state is zero. On the other hand, the expectation value of E(r, t)2

is given by

〈n|E(r, t)2|n〉 = �ω

ε0V

(n + 1

2

). (2.76)

For this case, as for the energy, the expectation value is nonvanishing even whenn = 0, with the result that for the total field (2.71), consisting of an infinite numberof modes, the expectation value of E(r, t)2 is infinite for all many-photon states,including the vacuum state. Obviously a similar reasoning applies for the magneticfield (2.72) and, as discussed in the previous section, the zero-point constant is usuallyremoved.

One must remember that the somehow strange result of Eq. (2.75) is due to thefact that the expectation value is performed with the Fock state |n〉, which meansthat the number of photons is fixed because

N |n〉 = n|n〉 . (2.77)

Nevertheless, usually the number of photons in the radiation field is not fixed, inother words the system is not in a pure Fock state. For example, the radiation field ofa well-stabilized laser device operating in a single mode is described by a coherentstate |α〉, such that

a|α〉 = α|α〉 , (2.78)

with〈α|α〉 = 1 . (2.79)

The coherent state |α〉, introduced in 1963 by Roy Glauber, is thus the eigenstate ofthe annihilation operator a with complex eigenvalue α = |α|eiθ. |α〉 does not havea fixed number of photons, i.e. it is not an eigenstate of the number operator N , andit is not difficult to show that |α〉 can be expanded in terms of number (Fock) states|n〉 as follows

|α〉 = e−|α|2/2∞∑n=0

αn

√n! |n〉 . (2.80)

From Eq. (2.78) one immediately finds

N = 〈α|N |α〉 = |α|2 , (2.81)

and it is natural to setα =

√N eiθ , (2.82)

Page 42: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

32 2 Second Quantization of Light

where N is the average number of photons in the coherent state, while θ is the phaseof the coherent state. For the sake of completeness, we observe that

〈α|N 2|α〉 = |α|2 + |α|4 = N + N 2 (2.83)

and consequently〈α|N 2|α〉 − 〈α|N |α〉2 = N , (2.84)

while〈n|N 2|n〉 = n2 (2.85)

and consequently〈n|N 2|n〉 − 〈n|N |n〉2 = 0 . (2.86)

The expectation value of the electric field E(r, t) of the linearly polarized mono-chromatic wave, Eq. (2.73), in the coherent state |α〉 reads

〈α|E(r, t)|α〉 = −√2N�ω

ε0Vsin(k · r − ωt + θ) ε , (2.87)

while the expectation value of E(r, t)2 is given by

〈α|E(r, t)2|α〉 = 2N�ω

ε0Vsin2(k · r − ωt + θ) . (2.88)

These results suggest that the coherent state is indeed a useful tool to investigate thecorrespondence between quantum field theory and classical field theory. Indeed, in1965 at the University ofMilan Fortunato Tito Arecchi experimentally verified that asingle-mode laser is in a coherent state with a definite but unknown phase. Thus, thecoherent state gives photocount statistics that are in accord with laser experimentsand has coherence properties similar to those of a classical field, which are usefulfor explaining interference effects.

To conclude this subsectionwe observe that we started the quantization of the lightfield by expanding the vector potential A(r, t) as a Fourier series of monochromaticplane waves. But this is not the unique choice. Indeed, one can expand the vec-tor potentialA(r, t) by using any orthonormal set of wavefunctions (spatio-temporalmodes) which satisfy theMaxwell equations. After quantization, the expansion coef-ficients of the single mode play the role of creation and annihilation operators of thatmode.

Page 43: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.2 Second Quantization of Light 33

2.2.2 Linear and Angular Momentum of the Radiation Field

In addition to the total energy there are other interesting conserved quantities whichcharacterize the classical radiation field. They are the total linear momentum

P =∫

d3r ε0 E(r, t) ∧ B(r, t) (2.89)

and the total angular momentum

J =∫

d3r ε0 r ∧ (E(r, t) ∧ B(r, t)) . (2.90)

By using the canonical quantization one easily finds that the quantum total linearmomentum operator is given by

P =∑k

∑s

�k(Nks + 1

2

), (2.91)

where �k is the linear momentum of a photon of wavevector k.The quantization of the total angular momentum J is a more intricate problem.

In fact, the linearly polarized states |ks〉 (with s = 1, 2) are eigenstates of H andP, but they are not eigenstates of J, nor of J 2, and nor of Jz . The eigenstates ofJ 2 and Jz do not have a fixed wavevector k but only a fixed wavenumber k = |k|.These states, which can be indicated as |k jm j s〉, are associated to vector sphericalharmonics Y j,m j (θ,φ), with θ and φ the spherical angles of the wavevector k. Thesestates |k jm j s〉 are not eigenstates of P but they are eigenstates of H , J 2 and Jz ,namely

H |k jm j s〉 = ck|k jm j s〉 (2.92)

J 2|k jm j 〉 = j ( j + 1)�2|k jm j s〉 (2.93)

Jz|k jm j s〉 = m j�|k jm j s〉 (2.94)

with m j = − j,− j + 1, . . . , j − 1, j and j = 1, 2, 3, . . .. In problems where thecharge distributionwhich emits the electromagnetic radiation has spherical symmetryit is indeed more useful to expand the radiation field in vector spherical harmonicsthan in plane waves. Clearly, one can write

|ks〉 =∑jm j

c jm j (θ,φ)|k jm j s〉 , (2.95)

where the coefficients c jm j (θ,φ) of this expansion give the amplitude probability offinding a photon of fixed wavenumber k = |k| and orbital quantum numbers j andm j with spherical angles θ and φ.

Page 44: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

34 2 Second Quantization of Light

2.2.3 Zero-Point Energy and the Casimir Effect

There are situations in which the zero-point energy Evac of the electromagnetic fieldoscillators give rise to a remarkable quantum phenomenon: the Casimir effect.

The zero-point energy (2.70) of the electromagnetic field in a region of volumeV can be written as

Evac =∑k

∑s

1

2�c

√k2x + k2y + k2z = V

∫d3k

(2π)3�c

√k2x + k2y + k2z (2.96)

by using the dispersion relation ωk = ck = c√k2y + k2y + k2z . In particular, consid-

ering a region with the shape of a parallelepiped of length L along both x and y andlength a along z, the volume V is given by V = L2a and the vacuum energy Evac inthe region is

Evac = �c∫ +∞

−∞L dkx2π

∫ +∞

−∞L dky2π

∫ +∞

−∞a dkz2π

√k2x + k2y + k2z

= �c

2πL2

∫ ∞

0dk‖k‖

[ ∫ ∞

0dn

√k2‖ + n2π2

a2

], (2.97)

where the second expression is obtained setting k‖ =√k2x + k2y and n = (a/π)kz .

Let us now consider the presence of two perfect metallic plates with the shapeof a square of length L having parallel faces lying in the (x, y) plane at distance a.Along the z axis the stationary standing waves of the electromagnetic field vanisheson the metal plates and the kz component of the wavevector k is no more a continuumvariabile but it is quantized via

kz = nπ

a, (2.98)

where nown = 0, 1, 2, . . . is an integer number, andnot a real number as inEq. (2.97).In this case the zero-point energy in the volume V = L2a between the two platesreads

E ′vac = �c

2πL2

∫ ∞

0dk‖k‖

[k‖2

+∞∑n=1

√k2‖ + n2π2

a2

], (2.99)

where for kz = 0 the state |(kz, ky, 0)1〉 with polarization εk1 parallel to the z axis(and orthogonal to the metallic plates) is bound, while the state |(kz, ky, 0)2〉 withpolarization εk2 orthogonal to the z axis (and parallel to the metallic plates) is notbound and it does not contribute to the discrete summation (Fig. 2.2).

Page 45: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.2 Second Quantization of Light 35

Fig. 2.2 Graphicalrepresentation of the parallelplates in the Casimir effect

The difference between E ′vac and Evac divided by L2 gives the net energy per unit

surface area E , namely

E = E ′vac − Evac

L2= �c

a

)3 [12A(0) +

∞∑n=1

A(n) −∫ ∞

0dn A(n)

], (2.100)

where we have defined

A(n) =∫ +∞

0dζ ζ

√ζ2 + n2 = 1

3

[(n2 + ∞)3/2 − n2

], (2.101)

with ζ = (a/π)k‖. Notice that A(n) is clearly divergent but Eq. (2.100) is not diver-gent due to the cancellation of divergences with opposite sign. In fact, by using theEuler-MacLaurin formula for the difference of infinite series and integrals

1

2A(0) +

∞∑n=1

A(n) −∫ ∞

0dn A(n) = − 1

6 · 2!d A

dn(0) + 1

30 · 4!d3A

dn3(0) − 1

42 · 6!d5A

dn5(0) + · · ·(2.102)

and Eq. (2.101) from which

d A

dn(0) = 0 ,

d3A

dn3(0) = −2 ,

d5A

dn5(0) = 0 (2.103)

and all higher derivatives of A(n) are zero, one eventually obtains

E = − π2

720

�c

a3. (2.104)

From this energy difference E one deduces that there is an attractive force per unitarea F between the two plates, given by

Page 46: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

36 2 Second Quantization of Light

F = −dEda

= − π2

240

�c

a4. (2.105)

Numerically this result, predicted in 1948 by Hendrik Casimir during his researchactivity at the Philips Physics Laboratory in Eindhoven, is very small

F = −1.30 × 10−27N m2

a4. (2.106)

Nevertheless, it has been experimentally verified by Steven Lamoreaux in 1997 atthe University of Washington and by Giacomo Bressi, Gianni Carugno, RobertoOnofrio, and Giuseppe Ruoso in 2002 at the University of Padua.

2.3 Quantum Radiation Field at Finite Temperature

Let us consider the quantum radiation field in thermal equilibrium with a bath at thetemperature T . The relevant quantity to calculate all thermodynamical properties ofthe system is the grand-canonical partition function Z , given by

Z = Tr [e−β(H−μN )] (2.107)

where β = 1/(kBT ) with kB = 1.38 × 10−23 J/K the Boltzmann constant,

H =∑k

∑s

�ωk Nks , (2.108)

is the quantum Hamiltonian without the zero-point energy,

N =∑k

∑s

Nks (2.109)

is the total number operator, andμ is the chemical potential, fixed by the conservationof the particle number. For photons μ = 0 and consequently the number of photonsis not fixed. This implies that

Z =∑{nks }

〈 . . . nks . . . |e−β H | . . . nks . . . 〉 =∑{nks }

〈 . . . nks . . . |e−β∑

ks �ωk Nks | . . . nks . . . 〉

=∑{nks }

e−β∑

ks �ωknks =∑{nks }

∏ks

e−β�ωknks =∏ks

∑nks

e−β�ωknks =∏ks

∞∑n=0

e−β�ωk n

=∏ks

1

1 − e−β�ωk. (2.110)

Page 47: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.3 Quantum Radiation Field at Finite Temperature 37

Quantum statistical mechanics dictates that the thermal average of any operator A isobtained as

〈 A〉T = 1

Z Tr [ A e−β(H−μN )] . (2.111)

In our case the calculations are simplified becauseμ = 0. Let us suppose that A = H ,it is then quite easy to show that

〈H〉T = 1

Z Tr [H e−β H ] = − ∂

∂βln

(Tr [e−β H ]

)= − ∂

∂βln(Z) . (2.112)

By using Eq. (2.110) we immediately obtain

ln(Z) = −∑k

∑s

ln(1 − e−β�ωk

), (2.113)

and finally from Eq. (2.112) we get

〈H〉T =∑k

∑s

�ωk

eβ�ωk − 1=

∑k

∑s

�ωk 〈Nks〉T . (2.114)

In the continuum limit, where

∑k

→ V∫

d3k(2π)3

, (2.115)

with V the volume, and taking into account that ωk = ck, one can write the energydensity E = 〈H〉T /V as

E = 2∫

d3k(2π)3

c�k

eβc�k − 1= c�

π2

∫ ∞

0dk

k3

eβc�k − 1, (2.116)

where the factor 2 is due to the two possible polarizations (s = 1, 2). By usingω = ck instead of k as integration variable one gets

E = �

π2c3

∫ ∞

0dω

ω3

eβ�ω − 1=

∫ ∞

0dω ρ(ω) , (2.117)

where

ρ(ω) = �

π2c3ω3

eβ�ω − 1(2.118)

is the energy density per frequency, i.e. the familiar formula of the black-body radi-ation, obtained for the first time in 1900 by Max Planck. The previous integral canbe explicitly calculated and it gives

Page 48: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

38 2 Second Quantization of Light

E = π2k4B15c3�3

T 4 , (2.119)

which is nothing but the Stefan-Boltzmann law. In an similar way one determinesthe average number density of photons:

n = 〈N 〉TV

= 1

π2c3

∫ ∞

0dω

ω2

eβ�ω − 1= 2ζ(3)k3B

π2c3�3T 3 . (2.120)

where ζ(3) � 1.202. Notice that both energy density E and number density n ofphotons go to zero as the temperature T goes to zero. To conclude this section, westress that these results are obtained at thermal equilibrium and under the condition ofa vanishing chemical potential, meaning that the number of photons is not conservedwhen the temperature is varied.

2.4 Phase Operators

We have seen that the ladder operators a and a+ of a single mode of the electromag-netic field satisfy the fundamental relations

a|n〉 = √n |n − 1〉 , (2.121)

a+|n〉 = √n + 1 |n + 1〉 , (2.122)

where |n〉 is the Fock state, eigenstate of the number operator N = a+a and describ-ing n photons in the mode. Ramarkably, these operator can be expressed in terms ofthe Fock states as follows

a =∞∑n=0

√n + 1|n〉〈n + 1| = |0〉〈1| + √

2|1〉〈2| + √3|2〉〈3| + · · · , (2.123)

a+ =∞∑n=0

√n + 1|n + 1〉〈n| = |1〉〈0| + √

2|2〉〈1| + √3|3〉〈2| + · · · . (2.124)

It is in fact straightforward to verify that the expressions (2.123) and (2.124) implyEqs. (2.121) and (2.122).

We now introduce the phase operators

f = a N−1/2 , f + = N−1/2 a+ , (2.125)

Page 49: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.4 Phase Operators 39

which can be expressed as

f =∞∑n=0

|n〉〈n + 1| = |0〉〈1| + |1〉〈2| + |2〉〈3| + · · · , (2.126)

f + =∞∑n=0

|n + 1〉〈n| = |1〉〈0| + |2〉〈1| + |3〉〈2| + · · · , (2.127)

and satisfy the very nice formulas

f |n〉 = |n − 1〉 , (2.128)

f +|n〉 = |n + 1〉 , (2.129)

showing that the phase operators act as lowering and raising operators of Fock stateswithout the complication of coefficients in front of the obtained states.

The shifting property of the phase operators on the Fock states |n〉 resembles thatof the unitary operator ei p/� acting on the position state |x〉 of a 1D particle, with pthe 1D linear momentumwhich is canonically conjugated to the 1D position operatorx . In particular, one has

ei p/�|x〉 = |x − 1〉 , (2.130)

e−i p/�|x〉 = |x + 1〉 . (2.131)

Due to these formal analogies, in 1927 Fritz London and Paul Dirac independentlysuggested that the phase operator f can be written as

f = ei� , (2.132)

where � is the angle operator related to the number operator N = a+a. But, contraryto ei p/�, the operator f is not unitary because

f f + = 1 , f + f = 1 − |0〉〈0| . (2.133)

This implies that the angle operator � is not Hermitian and

f + = e−i�+. (2.134)

In addition, a coherent state |α〉 is not an eigenstate of the phase operator f . However,one easily finds

〈α| f |α〉 = αe−|α|2∞∑n=0

|α|2n√n!(n + 1)! , (2.135)

and observing that for x � 1 one gets∑∞

n=0xn√

n!(n+1)! � ex√xthis implies

Page 50: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

40 2 Second Quantization of Light

〈α| f |α〉 � α

|α| = eiθ , (2.136)

for N = |α|2 � 1, with α =√Neiθ. Thus, for a large average number N of photons

the expectation value of the phase operator f on the coherent state |α〉 gives thephase factor eiθ of the complex eigenvalue α of the coherent state.

2.5 Solved Problems

Problem 2.1Show that the eigenvalues n of the operator N = a+a are non negative.

SolutionThe eigenvalue equation of N reads

a+a|n〉 = n|n〉 .

We can then write〈n|a+a|n〉 = n〈n|n〉 = n ,

because the eigenstate |n〉 is normalized to one. On the other hand, we have also

〈n|a+a|n〉 = (a|n〉)+(a|n〉) = |(a|n〉)|2

Consequently, we getn = |(a|n〉)|2 ≥ 0 .

We stress that we have not used the commutations relations of a and a+. Thus wehave actually proved that any operator N given by the factorization of a genericoperator a with its self-adjunct a+ has a non negative spectrum.

Problem 2.2Consider the operator N = a+a, where a and a+ satisfy the commutation ruleaa+ − a+a = 1. Show that if |n〉 is an eigenstate of N with eigenvalue n then a|n〉is an eigenstate of N with eigenvalue n − 1 and a+|n〉 is an eigenstate of N witheigenvalue n + 1.

SolutionWe have

N a|n〉 = (a+a)a|n〉 .

Page 51: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.5 Solved Problems 41

The commutation relation between a and a+ can be written as

a+a = aa+ − 1 .

This implies that

N a|n〉 = (aa+ − 1)a|n〉 = a(a+a − 1)|n〉 = a(N − 1)|n〉 = (n − 1)a|n〉 .

Finally, we obtain

N a+|n〉 = (a+a)a+|n〉 = a+(aa+)|n〉 = a+(a+a + 1)|n〉 = a+(N + 1)|n〉 = (n + 1)a+|n〉 .

Problem 2.3Taking into account the results of the two previous problems, show that the spectrumof the number operator N = a+a, where a and a+ satisfy the commutation ruleaa+ − a+a = 1, is the set of integer numbers.

SolutionWe have seen that N has a non negative spectrum. This means that N possesses alowest eigenvalue n0 with |n0〉 its eigenstate. This eigenstate |n0〉 is such that

a|n0〉 = 0 .

In fact, on the basis of the results of the previous problem, a|n0〉 should be eigenstateof N with eigenvalue n0−1 but this is not possible because n0 is the lowest eigenvalueof N . Consequently the state a|n0〉 is not a good Fock state and we set it equal to 0.In addition, due to the fact that

N |n0〉 = n0|n0〉= a+a|n0〉 = a+ (

a|n0〉) = a+ (0) = 0

we find that n0 = 0. Thus, the state |0〉, called vacuum state, is the eigenstate of Nwith eigenvalue 0, i.e.

N |0〉 = 0|0〉 = 0 ,

but alsoa|0〉 = 0 .

Due to this equation, it follows that the eigenstates of N are only those generated byapplying m times the operator a+ on the vacuum state |0〉, namely

|m〉 = 1√m! (a

+)m |0〉 ,

Page 52: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

42 2 Second Quantization of Light

where the factorial is due to the normalization. Finally, we notice that it has beenshown in the previous problem that the state |m〉 has integer eigenvalue m.

Problem 2.4Consider the following quantumHamiltonian of the one-dimensional harmonic oscil-lator

H = p2

2m+ 1

2mω2 x2 .

By using the properties of the annihilation operator

a =√mω

2�

(x + i

mωp

),

determine the eigenfunction of the ground state of the system.

SolutionLet us observe the following property of the annihilation operator

〈x |a =√mω

2�

(x + �

∂x

)〈x | .

In addition, froma|0〉 = 0 ,

we find

〈x |a|0〉 =√mω

2�

(x + �

∂x

)〈x |0〉 = 0 .

Introducing the characteristic harmonic length

lH =√

mω,

the dimensionless coordinatex = x

lH,

and the dimensionless eigenfunction

ψn(x) = lH ψn(xlH ) ,

we obtain (x + ∂

∂ x

)ψ0(x) = 0 .

Page 53: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.5 Solved Problems 43

The solution of this first order differential equation is foundby separation of variables:

x d x = −dψ0

ψ0,

from which

ψ0(x) = 1

π1/4exp

(−x2

2

),

having imposed the normalization

∫dx |ψ0(x)|2 = 1 .

Problem 2.5Consider the quantum Hamiltonian of the two-dimensional harmonic oscillator

H = p21 + p222m

+ 1

2mω2(x21 + x22 ) .

By using the properties of the creation operators

a+k =

√mω

2�

(xk − i

mωpk

), k = 1, 2

determine the eigenfunctions of the quantum Hamiltonian with eigenvalue 6�ω.

SolutionThe eigenvalues of the two-dimensional harmonic oscillator are given by

En1n2 = �ω(n1 + n2 + 1) ,

where n1, n2 = 0, 1, 2, 3, . . . are the quantum numbers.The eigenstates |n1n2〉 corresponding to the eigenvalue 6�ω are:

|50〉 , |41〉 , |32〉 , |23〉 , |14〉 , |05〉 .

Thus the eigenfunctions to be determined are

φ50(x1, x2) , φ41(x1, x2) , φ32(x1, x2) , φ23(x1, x2) , φ14(x1, x2) , φ05(x1, x2) .

In general, the eigenfunctions φn1n2(x1, x2) = 〈x1x2|n1n2〉 can be factorized as fol-lows

φn1n2(x1, x2) = ψn1(x1) ψn2(x2)

Page 54: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

44 2 Second Quantization of Light

where ψn j (x j ) = 〈x j |n j 〉, j = 1, 2. It is now sufficient to calculate the followingeigenfunctions of the one-dimensional harmonic oscillator:

ψ0(x) , ψ1(x) , ψ2(x) , ψ3(x) , ψ4(x) , ψ5(x) .

We observe that the creation operator of the one-dimensional harmonic oscillatorsatisfies this property

〈x |a+ =√mω

2�

(x − �

∂x

)〈x | ,

and consequently

〈x |a+|n〉 =√mω

2�

(x − �

∂x

)〈x |n〉 =

√mω

2�

(x − �

∂x

)ψn(x) .

Reminding thata+|n〉 = √

n + 1|n + 1〉 ,

we get

〈x |n + 1〉 = 1√n + 1

〈x |a+|n〉 ,

and the iterative formula

ψn+1(x) = 1√n + 1

√mω

2�

(x − �

∂x

)ψn(x) .

Now we introduce the characteristic harmonic length

lH =√

mω,

the dimensionless coordinatex = x

lH,

and the dimensionless wavefunction

ψn(x) = lH ψn(xlH ) ,

finding

ψn(x) = 1√2n n!

(x − ∂

∂ x

)n

ψ0(x) .

Page 55: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.5 Solved Problems 45

The function ψ0(x) of the ground state is given by (see exercise 1.4)

ψ0(x) = 1

π1/4exp

(−x2

2

).

Let us now calculate the effect of the operator

(x − ∂

∂ x

)n

.

For n = 1: (x − ∂

∂ x

)exp

(−x2

2

)= 2x exp

(−x2

2

).

For n = 2:

(x − ∂

∂ x

)2

exp

(−x2

2

)= 2(2x2 − 1) exp

(−x2

2

).

For n = 3:

(x − ∂

∂ x

)3

exp

(−x2

2

)= (8x3 − 12x) exp

(−x2

2

).

For n = 4:

(x − ∂

∂ x

)4

exp

(−x2

2

)= (16x4 − 48x2 + 12) exp

(−x2

2

).

For n = 5:

(x − ∂

∂ x

)5

exp

(−x2

2

)= (32x5 − 160x3 + 120x) exp

(−x2

2

).

Finally, the eigenfunctions of H with eigenvalue 6�ω are linear combinations offunctions φn1n2(x1, x2), namely

�(x1, x2) =∑n1n2

cn1n2 φn1n2(x1, x2) δn1+n2,5 ,

where the coefficients cn1n2 are such that

∑n1n2

|cn1n2 |2 δn1+n2,5 = 1 .

Page 56: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

46 2 Second Quantization of Light

Problem 2.6Show that the coherent state |α〉, defined by the equation

a|α〉 = α|α〉 ,

can be written as

|α〉 =∞∑n=0

e−|α|2/2 αn

√n! |n〉 .

SolutionThe coherent state |α〉 can be expanded as

|α〉 =∞∑n=0

cn|n〉 .

Then we have

a|α〉 = a∞∑n=0

cn|n〉 =∞∑n=0

cna|n〉 =∞∑n=1

cn√n|n − 1〉 =

∞∑n=0

cn+1

√n + 1|n〉

α|α〉 = α

∞∑n=0

cn|n〉 =∞∑n=0

α cn|n〉 .

Since, by definitiona|α〉 = α|α〉 ,

it follows thatcn+1

√n + 1 = α cn

from whichcn+1 = α√

n + 1cn ,

namely

c1 = α√1c0 , c2 = α2

√2!c0 , c3 = α3

√3!c0 , . . .

and in general

cn = αn

√n!c0 .

Page 57: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.5 Solved Problems 47

Summarizing

|α〉 =∞∑n=0

c0αn

√n! |n〉 ,

where the parameter c0 is fixed by the normalization

1 = 〈α|α〉 = |c0|2∞∑n=0

α2n

n! = |c0|2 e|α|2 ,

and consequentlyc0 = e−|α|2/2 ,

up to a constant phase factor.

Problem 2.7Calculate the probability of finding the Fock state |n〉 in the vacuum state |0〉.SolutionThe probability p is given by

p = |〈0|n〉|2 = δ0,n ={1 if n = 00 if n �= 0

.

Problem 2.8Calculate the probability of finding the coherent state |α〉 in the vacuum state |0〉.SolutionThe probability is given by

p = |〈0|α〉|2 .

Because

〈0|α〉 = 〈0|∞∑n=0

e−|α|2/2 αn

√n! |n〉 =

∞∑n=0

e−|α|2/2 αn

√n! 〈0|n〉 = e−|α|2/2 1√

0! = e−|α|2/2 ,

we getp = e−|α|2 .

Problem 2.9Calculate the probability of finding the coherent state |α〉 in the Fock state |n〉.

Page 58: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

48 2 Second Quantization of Light

SolutionThe probability is given by

p = |〈n|α〉|2 .

Because

〈n|α〉 = 〈n|∞∑

m=0

e−|α|2/2 αm

√m! |m〉 =

∞∑m=0

e−|α|2/2 αm

√m! 〈n|m〉 = e−|α|2/2 αn

√n! ,

we get

p = e−|α|2 |α|2nn! ,

which is the familiar Poisson distribution.

Problem 2.10Calculate the probability of finding the coherent state |α〉 in the coherent state |β〉.SolutionThe probability is given by

p = |〈β|α〉|2 .

Because

〈β|α〉 = 〈β|∞∑

m=0

e−|α|2/2 αm

√m! |m〉 =

∞∑n=0

∞∑m=0

e−|β|2/2e−|α|2/2 (β∗)n√n!

αm

√m! 〈n|m〉

= e−(|α|2+|β|2)/2∞∑n=0

(β∗α)n

n! = e−(|α|2+|β|2)/2 eβ∗α = e−(|α|2+|β|2−2αβ∗)/2 ,

we getp = e−|α−β|2 .

This means that two generic coherent states |α〉 and |β〉 are never orthogonal to eachother.

Further Reading

For the second quantization of the electromagnetic field:F. Mandl and G. Shaw, Quantum Field Theory, Chap. 1, Sects. 1.1 and 1.2 (Wiley,New York, 1984).M.O. Scully and M.S. Zubairy, Quantum Optics, Chap. 1, Sects. 1.1 and 1.2 (Cam-bridge Univ. Press, Cambridge, 1997).U. Leonhardt, Measuring the Quantum State of Light, Chap. 2, Sects. 2.1, 2.2, and2.3 (Cambridge Univ. Press, Cambridge, 1997).

Page 59: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

2.5 Solved Problems 49

F.T. Arecchi, Measurement of the statistical distribution of Gaussian and lasersources, Phys. Rev. Lett. 15, 912 (1965).For the Casimir effect:R.W. Robinett, Quantum Mechanics: Classical Results, Modern Systems, and Visu-alized Examples, Chap. 19, Sect. 19.9 (Oxford Univ. Press, Oxford, 2006).S. K. Lamoreaux, Demonstration of the Casimir force in the 0.6 to 6 m range, Phys.Rev. Lett. 78, 5 (1997).G. Bressi, G. Carugno, R. Onofrio, and G. Ruoso, Measurement of the Casimir forcebetween parallel metallic surfaces, Phys. Rev. Lett. 88, 041804 (2002).For the quantum radiation field at finite temperature:K. Huang, Statistical Mechanics, Chap. 12, Sect. 12.1 (Wiley, New York, 1987).

Page 60: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 3Electromagnetic Transitions

In this chapter we investigate the crucial role of the quantum electromagnetic fieldon the spontaneous and stimulated transitions between two atomic quantum states.After reviewing some properties of classical electrodynamics, we analyze the quan-tum electrodynamics within the dipole approximation. We calculate the rate of spon-taneous emission, absorption, and stimulated emission and connect them with thetransition coefficients introduced by Einstein. Finally, we discuss the lifetime of anatomic state and the line width of an electromagnetic transition.

3.1 Classical Electrodynamics

Let us consider a classical system composed of N particles with masses mi , electriccharges qi , positions ri and linear momenta pi , under the presence of an electromag-netic field. The Hamiltonian of free matter is given by

Hf ree =N∑i=1

p2i2mi

, (3.1)

where p2i /(2mi ) = p2i /(2mi ) is the kinetic energy of i-th particle. The presence ofthe electromagnetic field is modelled by the Hamiltonian

H = Hshi f t + Hrad , (3.2)

where

Hshi f t =N∑i=1

(pi − qi Ai )2

2mi+ qi φi , (3.3)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_3

51

Page 61: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

52 3 Electromagnetic Transitions

with Ai = A(ri , t), φi = φ(ri , t), where φ(r, t) and A(r, t) are the electromagneticpotentials and

Hrad =∫

d3r(

ε0

2E2T (r, t) + 1

2μ0B2(r, t)

), (3.4)

is the radiation Hamiltonian. We observe that by using the Hamilton equations

ri = ∂Hshi f t

∂pi(3.5)

pi = −∂Hshi f t

∂ri(3.6)

on the shift Hamiltonian (3.3) it is straightforward to derive Newton equations withthe Lorentz force acting on the i-th particle

mi ri = qi

(−∇iφi − ∂Ai

∂t+ vi ∧ (∇i ∧ Ai )

)= qi (Ei + vi ∧ Bi ) (3.7)

where vi = ri and, as always, the electric field E is obtained as

E = −∇φ − ∂A∂t

, (3.8)

while the magnetic field B can be written as

B = ∇ ∧ A . (3.9)

In Eq. (3.4) it appears the transverse electric field ET which, in the Coulomb gauge∇ ·A = 0, is related to the total electric field E by the following decomposition intolongitudinal EL and transverse ET fields

E = EL + ET , (3.10)

such that

EL = −∇φ , ET = −∂A∂t

. (3.11)

Remarkably, the longitudinal electric field EL gives rise to the instantaneous elec-trostatic interaction between the charges. Indeed, Eq. (3.3) can be rewritten as

Hshi f t =N∑i=1

(pi − qi Ai )2

2mi+ HC , (3.12)

Page 62: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.1 Classical Electrodynamics 53

where

HC =N∑i=1

qi φi = 1

2

N∑i, j=1

qiq j

4πε0|ri − r j | , (3.13)

is theHamiltonian of the Coulomb interaction. TheHamiltonian Hshi f t can be furtherdecomposed as follows

Hshi f t = Hmatt + HI , (3.14)

whereHmatt = Hf ree + HC (3.15)

is the Hamiltonian of the self-interacting matter and

HI =N∑i=1

(−qimAi · pi + q2

i

2miA2

i

)(3.16)

is the interaction Hamiltonian between matter and radiation. Combining the aboveresults, we obtain the complete Hamiltonian

H = Hshi f t + Hrad = Hmatt + Hrad + HI . (3.17)

Notice, however, that this Hamiltonian does not take into account the possible spinof particles.

3.2 Quantum Electrodynamics in the DipoleApproximation

The quantization of electrodynamics is obtained promoting the classical Hamiltonianof the system to a quantum Hamiltonian. For simplicity we consider the hydrogenatom with Hamiltonian

Hmatt = p2

2m− e2

4πε0|r| , (3.18)

where p = −i�∇ is the linear momentum operator of the electron in the state |p〉,in the presence of the radiation field with Hamiltonian

Hrad =∑k

∑s

�ωk a+ks aks , (3.19)

where aks and aks are the annihilation and creation operators of the photon in thestate |ks〉. Here −e is the electric charge of the electron, e = 1.60 × 10−19 C is theelectric charge of the proton, and m = memp/(me +mp) � me is the reduced mass

Page 63: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

54 3 Electromagnetic Transitions

of the electron-proton system, with me = 9.11 × 10−31 kg the electron mass andmp = 1.67 × 10−27 kg the proton mass.

We take into account thematter-radiation interaction by using the so-called dipolarHamiltonian

HD = e

mA(0, t) · p . (3.20)

With respect to the complete interaction Hamiltonian HI , given in our case by Eq.(3.16) with N = 1, we neglect the term e2A2/(2m) of Eq. (3.16) which representsonly a tiny perturbation in the atomic system. In addition, we also neglect the spatialvariations in the vector potential operator

A(r, t) =∑k

∑s

√�

2ε0ωkV

[aks e

i(k·r−ωk t) + a+ks e

−i(k·r−ωk t)]

εks . (3.21)

This dipolar approximation, which corresponds to

eik·r = 1 + ik · r + 1

2(ik · r)2 + · · · � 1 , (3.22)

e−ik·r = 1 − ik · r + 1

2(ik · r)2 + · · · � 1 , (3.23)

is reliable if k · r � 1, namely if the electromagnetic radiation has a wavelengthλ = 2π/|k| very large compared to the linear dimension R of the atom. Indeed, theapproximation is fully justified in atomic physics where λ � 10−7 m and R � 10−10

m. Notice, however, that for the γ electromagnetic transitions of atomic nuclei thedipolar approximation is not good and it is more convenient to expand the vectorpotential operator A(r, t) into vector spherical harmonics, i.e. photons of definiteangular momentum.

The total Hamiltonian of our system is then given by

H = H0 + HD , (3.24)

whereH0 = Hmatt + Hrad (3.25)

is the unperturbed Hamiltonian, whose eigenstates are of the form

|a〉| . . . nks . . . 〉 = |a〉 ⊗ | . . . nks . . . 〉 (3.26)

where |a〉 is the eigenstate of Hmatt with eigenvalue Ea and | . . . nks . . . 〉 it theeigenstate of Hrad with eigenvalue

∑ks �ωknks , i.e.

Page 64: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.2 Quantum Electrodynamics in the Dipole Approximation 55

H0|a〉| . . . nkr . . . 〉 =(Hmatt + Hrad

)|a〉| . . . nks . . . 〉 =

⎛⎝Ea +

∑ks

�ωknks

⎞⎠ |a〉| . . . nks . . . 〉 .

(3.27)The time-dependent perturbing Hamiltonian is instead given by

HD(t) = e

m

∑k

∑s

√�

2ε0ωkV

[aks e

−iωk t + a+ks e

iωk t]

εks · p . (3.28)

Fermi golden rule: Given the initial |I 〉 and final |F〉 eigenstates of the unperturbedHamiltonian H0 under the presence to the time-dependent perturbing HamiltonianHD , the probability per unit time of the transition from |I 〉 to |F〉 is given by

WIF = 2π

�|〈F |HD(0)|I 〉|2 δ(EI − EF ) , (3.29)

with the constraint of energy conservation.

This is the so-called Fermi golden rule, derived in 1926 by Paul Dirac on the basisof the first order time-dependent perturbation theory, and named “golden rule” fewyears later by Enrico Fermi.

3.2.1 Spontaneous Emission

Let us now apply the Fermi golden rule to the very interesting case of the hydrogenatom in the state |b〉 and the radiation field in the vacuum state |0〉. We are thussupposing that the initial state is

|I 〉 = |b〉|0〉 . (3.30)

Notice that, because we are considering the hydrogen atom, one has

Hmatt |b〉 = Eb|b〉 , (3.31)

where

Eb = −mc2α2

2n2b� −13.6 eV

n2b(3.32)

is the well-known quantization formula of the nonrelativistic hydrogen atom withquantum number nb = 1, 2, 3, . . . and

Page 65: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

56 3 Electromagnetic Transitions

α = e2

(4πε0)�c� 1

137(3.33)

the fine-structure constant. In addition we suppose that the final state is

|F〉 = |a〉|ks〉 , (3.34)

i.e. the final atomic state is |a〉 and the final photon state is |ks〉 = |1ks〉 = a+ks |0〉.

Obviously, we haveHmatt |a〉 = Ea|a〉 , (3.35)

where

Ea = −mc2α2

2n2a� −13.6 eV

n2a(3.36)

with na = 1, 2, 3, . . .. In this process, where Eb > Ea , there is the spontaneousproduction of a photon from the vacuum: a phenomenon strictly related to the quan-tization of the electromagnetic field.

From Eqs. (3.28) and (3.29) one finds

Wspontba,ks = 2π

( e

m

)2(

2ε0ωkV

)|εks · 〈a|p|b〉|2 δ(Eb − Ea − �ωk) , (3.37)

becauseak′s ′ |I 〉 = ak′s ′ |b〉|0〉 = |b〉ak′s ′ |0〉 = 0 , (3.38)

whilea+k′s ′ |I 〉 = a+

k′s ′ |b〉|0〉 = |b〉a+k′s ′ |0〉 = |b〉|k′s ′〉 , (3.39)

and consequently

〈F |p ak′s ′ |I 〉 = 0 , 〈F |p a+k′s ′ |I 〉 = 〈a|p|b〉 δk′,k δs ′,s . (3.40)

Remember that p acts on a different Hilbert space with respect to aks and a+ks . On

the basis of Heisenberg equation of motion of the linear momentum operator p ofthe electron

pm

= drdt

= 1

i�[r, Hmatt ] , (3.41)

we get

〈a|p|b〉 = 〈a|m 1

i�[r, Hmatt ]|b〉 = m

i�〈a|rHmatt − Hmattr|b〉

= m

i�(Eb − Ea)〈b|r|a〉 = −imωba 〈a|r|b〉 , (3.42)

Page 66: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.2 Quantum Electrodynamics in the Dipole Approximation 57

where ωba = (Eb − Ea)/�, and consequently

Wspontba,ks = πω2

ba

V ε0ωk|εks · 〈a|e r|b〉|2 δ(�ωba − �ωk) . (3.43)

The delta function is eliminated by integrating over the final photon states

Wspontba =

∑k

∑s

Wspontba,ks = V

∫d3k

(2π)3

∑s=1,2

Wspontba,ks = V

8π3

∫dkk2

∫d�

∑s=1,2

Wspontba,ks ,

(3.44)where d� is the differential solid angle. Because εk1, εk2 and n = k/k form aorthonormal system of vectors, setting rab = 〈a|r|b〉 one finds

|rab|2 = |εk1·rab|2+|εk2·rab|2+|n·rab|2 =∑s=1,2

|εks ·rab|2+|rab|2 cos2 (θ) , (3.45)

where θ is the angle between rba and n. It follows immediately

∑s=1,2

|εks · rab|2 = |rab|2(1 − cos2 (θ)) = |rab|2 sin2(θ) , (3.46)

namely ∑s=1,2

|εks · 〈a|r|b〉|2 = |〈a|r|b〉|2 sin2(θ) . (3.47)

In addition, in spherical coordinates one can choose d� = sin (θ)dθdφ, with θ ∈[0,π] the zenith angle of colatitude and φ ∈ [0, 2π] the azimuth angle of longitude,and then ∫

d� sin2 (θ) =∫ 2π

0dφ

∫ π

0dθ sin3 (θ) = 8π

3. (3.48)

In this way from Eq. (3.44) we finally obtain

Wspontba = ω3

ba

3πε0�c3|〈a|d|b〉|2 , (3.49)

where the d = −e r is the classical electric dipole momentum of the hydrogen atom,i.e. the dipole of the electron-proton system where r is the position of the electron ofcharge −e < 0 with respect to the proton of charge e > 0, and 〈a|d|b〉 = −〈a|e r|b〉is the so-called dipole transition element.

Page 67: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

58 3 Electromagnetic Transitions

3.2.2 Absorption

We now consider the excitation from the atomic state |a〉 to the atomic state |b〉 dueto the absorption of one photon. Thus we suppose that the initial state is

|I 〉 = |a〉|nks〉 , (3.50)

while the final state is|F〉 = |b〉|nks − 1〉 , (3.51)

where Ea < Eb. From Eqs. (3.28) and (3.29) one finds

Wabsorpab,ks = 2π

( e

m

)2(

2ε0ωkV

)nks |εks · 〈b|p|a〉|2 δ(Ea + �ωk − Eb) , (3.52)

because

〈F |p ak′s ′ |I 〉 = √nks 〈b|p|a〉 δk′,k δs ′,s , 〈F |p a+

k′s ′ |I 〉 = 0 . (3.53)

Note that with respect to Eq. (3.37) in Eq. (3.52) there is also the multiplicative termnks . We can follow the procedure of the previous section to get

Wabsorpab,ks = πω2

ba

V ε0ωknks |εks · 〈b|e r|a〉|2 δ(�ωba − �ωk) . (3.54)

Again the delta function can be eliminated by integrating over the final photon statesbut here one must choose the functional dependence of nks . We simply set

nks = n(ωk) , (3.55)

and after integration over k and s, from Eq. (3.54) we get

Wabsorpab = ω3

ba

3πε0�c3|〈b|d|a〉|2 n(ωba) = Wspont

ba n(ωba) . (3.56)

Note that without the integration over the solid angle, which gives a factor 1/3,one obtains the absorption probability of one photon with a specific direction. Fora thermal distribution of photons, with ρ(ω) the energy density per unit of angularfrequency specified by the thermal-equilibrium Planck formula

ρ(ω) = �ω3

π2c3n(ω) , n(ω) = 1

e�ω/(kBT ) − 1, (3.57)

Page 68: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.2 Quantum Electrodynamics in the Dipole Approximation 59

where kB is the Boltzmann constant and T the absolute temperature, we can alsowrite

Wabsorpab = |〈b|d|a〉|2 1

3ε0�2ρ(ωba) = Wspont

ba

π2c3

�ω3ba

ρ(ωba) . (3.58)

3.2.3 Stimulated Emission

In conclusion we consider the stimulated emission of a photon from the atomic state|b〉 to the atomic state |a〉. Thus we suppose that the initial state is

|I 〉 = |b〉|nks〉 , (3.59)

while the final state is|F〉 = |a〉|nks + 1〉 , (3.60)

where Eb > Ea . From Eqs. (3.28) and (3.29) one finds

Wstimulba,ks = 2π

( e

m

)2(

2ε0ωkV

)(nks + 1) |εks · 〈a|p|b〉|2 δ(Eb − Ea − �ωk) ,

(3.61)because

aks |I 〉 = aks |b〉|nks〉 = |b〉aks |nks〉 = √nks |b〉|nks − 1〉 , (3.62)

while

a+ks |I 〉 = a+

ks |b〉|nks〉 = |b〉a+ks |nks〉 = √

nks + 1 |b〉|nks + 1〉 , (3.63)

and consequently

〈F |p aks |I 〉 = 0 , 〈F |p a+ks |I 〉 = √

nks + 1 〈a|p|b〉 . (3.64)

Note that with respect to Eq. (3.52) in Eq. (3.61) there is the factor nks + 1 insteadof nks . It is straightforward to follow the procedure of the previous section to obtain

Wstimulba = ω3

ba

3πε0�c3|〈a|d|b〉|2 (n(ωba) + 1) = Wabsorp

ab + Wspontba , (3.65)

which shows that the probabilityWstimulba of stimulated emission reduces to the spon-

taneous one Wspontba when n(ωba) = 0. It is then useful to introduce

W stimulba = Wstimul

ba − Wspontba (3.66)

Page 69: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

60 3 Electromagnetic Transitions

Fig. 3.1 Scheme of the 3 mechanisms of radiative transition

which is the effective stimulated emission, i.e. the stimulated emission without thecontribution due to the spontaneous emission. Clearly, for a very large number ofphotons (n(ωba) � 1) one gets W stimul

ba � Wstimulba . Moreover, for a thermal distrib-

ution of photons, with the energy density per unit of angular frequency ρ(ω), we canalso write

W stimulba = Wabsorp

ab = Wspontba

π2c3

�ω3ba

ρ(ωba) . (3.67)

Remarkably, the probability of stimulated emission is different from zero only if theemitted photon is in the same single-particle state |ks〉 of the stimulating ones (apartwhen the stimulating light is the vacuum |0〉). In the stimulated emission the emittedphoton is said to be “coherent” with the stimulating ones, having the same frequencyand the same direction (Fig. 3.1).

3.3 Selection Rules

It is clear that, within the dipolar approximation, no electromagnetic transition, eitherspontaneous or stimulated, will occur between the atomic states |a〉 and |b〉 unlessat least one component of the dipole transition matrix element 〈b|d|a〉 is nonzero. Itis possible to show that the matrix elements are zero for certain pairs of states. If so,the transition is not allowed (at least in the electric dipole approximation), and the

Page 70: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.3 Selection Rules 61

results can be summarized in terms of simple selection rules governing the allowedchanges in quantum numbers in transitions.

Since the electric dipole d = −er changes sign under parity (r → −r), thematrixelement 〈b|d|a〉 is zero if the states |a〉 and |b〉 have the same parity. Therefore

the pari t y o f the wave f unction must change in an electr ic dipole transi tion.

This means that if ψa(r) and ψb(r) are the eigenfunctions of the states |a〉 and |b〉and, for instance, both eigenfunctions are even, i.e.

ψa(−r) = ψa(r) and ψb(−r) = ψb(r) , (3.68)

then the dipole matrix element is such that

〈b|d|a〉 = −e∫

d3r ψ∗b(r) r ψa(r) = 0 . (3.69)

In fact, under the transformation r → −r one finds∫d3(−r) = ∫

d3r while〈b|d|a〉 = −〈b|d|a〉, and then it follows 〈b|d|a〉 = 0.

Let us recall that in our calculations a generic eigenstate |a〉 = |nlm〉 of the matterHamiltonian Hmatt is such that

ψa(r) = ψnlm(r) = 〈r|a〉 = 〈r|nlm〉 = Rnl(r)Ylm(θ,φ) , (3.70)

with Ynm(θ,φ) the spherical harmonic function. The spherical harmonics satisfy theorthonormalization condition

∫ π

0dθ sin (θ)

∫ 2π

0dφ Y ∗

l ′m ′(θ,φ) Ylm(θ,φ) = δl,l ′ δm,m ′ , (3.71)

where Y ∗l,m(θ,φ) = (−1)m Yl,−m(θ,φ). Moreover under parity transformation one

finds

ψnlm(−r) = Rnl(r)Ylm(π − θ,φ + π) = (−1)l Rnl(r)Ylm(θ,φ) = (−1)lψnlm(r) ,

(3.72)

thus the radial part Rnl(r) of the wavefunction is unchanged and the parity of thestate is fully determined from the angular part.

The generic dipole matrix element is given by

〈n′l ′m ′|d|nlm〉 = −e∫ ∞

0dr r3Rn′l ′(r)Rnl(r)

∫ π

0dθ sin (θ)

∫ 2π

0dφ Y ∗

l ′m ′(θ,φ)

× (cos (φ) sin (θ), sin (φ) sin (θ), cos (θ)) Ylm(θ,φ) , (3.73)

Page 71: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

62 3 Electromagnetic Transitions

with r = r (cos (φ) sin (θ), sin (φ) sin (θ), cos (θ)). Setting �l = l − l ′ and �m =m−m ′, by using the properties of the spherical harmonics it is possible to prove thatthe dipole matrix element 〈n′l ′m ′|d|nlm〉 is different from zero if only if �l = ±1and�m = 0,±1 (see, for instance, Problem 3.2). Thus, in the dipole approximation,the orbital angular momentum l of the atomic state and its third component m =−l,−l + 1, . . . , l − 1, l must satisfy the selection rules

�l = ±1 �m = 0,±1 . (3.74)

This means that in the electric dipole transitions the photon carries off (or brings in)one unit of angular momentum. It is important to stress that the above selection rulesare obtained within the dipole approximation. This means that they can be violatedby rare electromagnetic transitions involving higher multipolarities.

Concluding this section we observe that in the general case of N particles withpositions ri and electric charges qi , i = 1, 2, . . . , N , the electric dipole momentumis defined

d =N∑i=1

qi ri . (3.75)

Moreover, due to Eq. (3.42), the dipolar interaction Hamiltonian (3.20) can be effec-tively written as

HD = −d · E , (3.76)

which is the familiar expression of the coupling between the electric dipole d andthe electric field E.

3.4 Einstein Coefficients

In 1919 Albert Einstein observed that, given an ensemble of N atoms in two possibleatomic states |a〉 and |b〉, with Na(t) the number of atoms in the state |a〉 at time tand Nb(t) the number of atoms in the state |b〉 at time t , it must be

N = Na(t) + Nb(t) (3.77)

and consequentlydNa

dt= −dNb

dt. (3.78)

Einstein suggested that, if the atoms are exposed to an electromagnetic radiation ofenergy density per unit of frequency ρ(ω), the rate of change of atoms in the state|a〉 must be

dNa

dt= Aba Nb + Bba ρ(ωba) Nb − Bab ρ(ωba) Na . (3.79)

Page 72: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.4 Einstein Coefficients 63

where the parameters Aba , Bba , and Bab are known as Einstein coefficients. The firsttwo terms in this equation describe the increase of the number of atoms in |a〉 dueto spontaneous and stimulated transitions from |b〉, while the third term takes intoaccount the decrease of the number of atoms in |a〉 due to absorption with consequenttransition to |b〉.

Einstein was able to determine the coefficients Aba , Bba and Bab by supposing thatthe two rates in Eqs. (3.78) and (3.79) must be equal to zero at thermal equilibrium,i.e.

dNa

dt= −dNb

dt= 0 , (3.80)

In this way Einstein found

AbaNb

Na= ρ(ωba)

(Bab − Bba

Nb

Na

). (3.81)

Because the relative population of the atomic states |a〉 and |b〉 is given by a Boltz-mann factor

Nb

Na= e−βEb

e−βEa= e−β(Eb−Ea) = e−β�ωba , (3.82)

Einstein got

ρ(ωba) = Aba

Babeβ�ωba − Bba. (3.83)

At thermal equilibrium we know that

ρ(ωba) = �ω3ba

π2c31

eβ�ωba − 1. (3.84)

It follows that

Aba = Bba�ω3

ba

π2c3, Bab = Bba . (3.85)

Notice that in this way Einstein obtained the coefficient Aba of spontaneous decayby simply calculating the coefficient of stimulated decay Bba .

Actually, by using the results we have obtained in the previous section, it is clearthat

Aba = Wspontba = ω3

ba

3πε0�c3|〈a|d|b〉|2 , (3.86)

Bba = W stimulba

ρ(ωba)= π2c3

�ω3ba

Aba , (3.87)

Bab = Wabsorpab

ρ(ωba)= π2c3

�ω3ba

Aba . (3.88)

Page 73: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

64 3 Electromagnetic Transitions

It is important to stress that the laser device, invented in 1957 by Charles Townesand Arthur Schawlow at Bell Labs, is based on a generalization of Eqs. (3.78) and(3.79). In the laser device one has population inversion, which is achieved with anout-of-equilibrium pumping mechanism strongly dependent on the specific charac-teristics of the laser device.

3.4.1 Rate Equations for Two-Level and Three-Level Systems

Let us analyze under which conditions one can get population inversion. We haveseen that for a two-level system under the action of an electromagnetic pump withenergy density ρ(ωba) the rate equations of Einstein are given by

dNa

dt= Aba Nb + R (Nb − Na) , (3.89)

dNb

dt= −Aba Nb − R (Nb − Na) , (3.90)

where R = Bbaρ(ωba) is the pumping rate. We now suppose that the system is ina steady state, but not at thermal equilibrium, i.e. the energy density ρ(ωba) of theelectromagnetic pump is not given by the Planck distribution (3.84). At the steadystate we have

dNa

dt= dNb

dt= 0 , (3.91)

from which, taking into account that N = Na + Nb, we obtain

Na = NAba + R

Aba + 2R, (3.92)

Nb = NR

Aba + 2R, (3.93)

and the population imbalance ζ = (Na − Nb)/N reads

ζ = Aba

Aba + 2R. (3.94)

Equation (3.94) shows that in the absence of electromagnetic pump (R = 0) onehas ζ = 1, which simply means that all two-level atoms are in the lower state |a〉.Instead, if the electromagnetic pump is active (R �= 0) the population imbalanceζ decreases by increasing the pumping rate R, but ζ cannot change sign becauseζ → 0 for R → +∞. Thus, we conclude that the population inversion (ζ < 0), andconsequently laser light, is impossible in a strictly two-level system.

Let us now consider an ensemble of N atoms in three possible atomic states |a〉,|b〉 and |c〉, with Na(t) the number of atoms in the state |a〉 at time t , Nb(t) the

Page 74: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.4 Einstein Coefficients 65

number of atoms in the state |b〉 at time t , and Nc(t) the number of atoms in the state|b〉 at time t . In this ensemble of three-level atoms it must be

N = Na(t) + Nb(t) + Nc(t) . (3.95)

Moreover, we suppose that Ea < Eb < Ec, with Ea the energy of the state |a〉, Ebthe energy of the state |b〉, and Ec the energy of the state |c〉. In the presence of anelectromagnetic pump of energy density ρ(ωca), with ωca = (Ec − Ea)/�, betweenthe state |a〉 and |c〉 the rate equations of the three-level system are given by

dNa

dt= Aba Nb + Aca Nc + R (Nc − Na) + Bba ρba (Nb − Na) , (3.96)

dNb

dt= −Aba Nb + AcbNc − Bba ρba (Nb − Na) + Bcb ρcb (Nc − Nb) , (3.97)

dNc

dt= −Acb Nc − Aca Nc − R (Nc − Na) − Bcb ρcb (Nc − Nb) , (3.98)

where R = Bca ρ(ωca) is the pumping rate and the energy densities ρba and ρcbare induced by the electromagnetic transitions. These equations are simplified underthe following physical assumptions (Ruby laser): (i) there is no electromagnetictransition from |c〉 and |b〉, namely ρcb = 0; (ii) the time decay of the spontaneoustransition from |c〉 to |a〉 is very long, namely Aca � 0; (iii) the time decay of thespontaneous transition from |c〉 to |b〉 is very short and consequently Nc � Na . Inthis way the rate equations become

dNa

dt= Aba Nb − R Na + S (Nb − Na) , (3.99)

dNb

dt= −Aba Nb + AcbNc − S (Nb − Na) , (3.100)

dNc

dt= −Acb Nc + R Na , (3.101)

where S = Bba ρba is the stimulated rate. At stationarity one has

dNa

dt= dNb

dt= dNc

dt= 0 , (3.102)

from which we get

Na = NAba + S

Aba + 2S + R, (3.103)

Nb = NR + S

Aba + 2S + R, (3.104)

Nc = R

AcbNa , (3.105)

Page 75: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

66 3 Electromagnetic Transitions

and the population imbalance ζ = (Na − Nb)/N reads

ζ = Aba − R

Aba + 2S + R. (3.106)

This equation clearly shows that for R > Aba one gets ζ < 0, thus one obtains thedesired population imbalance.

3.5 Life-Time and Natural Line-Width

We have seen that the Einstein coefficient Aba = Wspontba gives the transition proba-

bility per unit of time from the atomic state |b〉 to the atomic state |a〉. This meansthat, according to Einstein, in the absence of an external electromagnetic radiationone has

dNb

dt= −Aba Nb (3.107)

with the unique solutionNb(t) = Nb(0) e

−Aba t . (3.108)

It is then quite natural to consider 1/Aba as the characteristic time of this spontaneoustransition.

More generally, the life-time τb of an atomic state |b〉 is defined as the reciprocalof the total spontaneous transition probability per unit time to all possible final atomicstates |a〉, namely

τb = 1∑a Aba

. (3.109)

Clearly, if |b〉 is the ground-state then Aba = 0 and τb = ∞.On the basis of the time-energy indetermination principle of Werner Heisenberg,

in the radiation energy spectrum the natural line-width �N due to the transition fromthe state |b〉 to the state |a〉 can be defined as

�N = �

(1

τb+ 1

τa

). (3.110)

Indeed it is possible to prove that in this transition the intensity of the emitted elec-tromagnetic radiation follows the Lorentzian peak

I (ε) = I0�2N/4

(ε − Eba)2 + �2N/4

, (3.111)

Page 76: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.5 Life-Time and Natural Line-Width 67

where ε = �ω is the energy of the emitted photon and Eba = Eb − Ea is the energydifference of the two atomic states. The Lorentzian peak is centered on ε = Eba and�N is its full width at half-maximum.

It is important to observe that the effective line-width � measured in the exper-iments is usually larger than �N because the radiating atoms move and collide. Infact, one can write

� = �N + �C + �D , (3.112)

where in addition to the natural width �N there are the so-called collisional broad-ening width �C and the Doppler broadening width �D .

3.5.1 Collisional Broadening

The collisional broadening is due to the collision among the radiating atoms of a gas.The collision reduces the effective life-time of an atomic state and the collisionalwidth can be then written as

�C = �

τcol, (3.113)

where τcol is the collisional time, i.e. the average time between two collision of atomsin the gas. According to the results of statistical mechanics, τcol is given by

τcol = 1

nσvmp, (3.114)

where n is the number density of atoms, σ is the cross-section, and vmp is the mostprobable speed of the particles in the gas.

In the quantum theory of scattering one usually assumes that an incoming particle,described by a plane wave along the z-axis, scatters with a target, modelled by anexternal potential V (r), such that the outcoming asymptotic wavefunction of theparticle reads

ψ(r) = eik·r + fk(θ)eikr

r, (3.115)

where fk(θ) is the scattering amplitude. The differential cross section is given bydσ/d� = | fk(θ)|2, which is the scattering probability per unit time of the particlewith mass m, wavevector k and energy �

2k2/(2m). If one considers only s-wavescattering the differential cross section does not depend on the angle θ, and the totalscattering cross section is just

σ(k) = 4π| fk |2 = 4π

k2sin2(δ0(k)) , (3.116)

Page 77: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

68 3 Electromagnetic Transitions

where δ0(k) is the so-called s-wave phase shift of the cross-section. The s-wavescattering length as , which characterizes the effective size of the target, is then definedas the following low-energy limit

as = − limk→0

1

ktan (δ0(k)) . (3.117)

In this way one clearly finds

limk→0

σ(k) = 4πa2s . (3.118)

Thus, for a dilute and cold gas of atoms the cross-section σ of low-energy atomscan be indeed approximated as

σ = 4πa2s , (3.119)

where the value of as can be indeed determined by solving the reduced single-particle Schrödinger equation with the actual inter-atomic potential V (r). Moreover,by considering the Maxwell-Boltzmann distribution of speeds in an ideal gas, themost probable speed vmp is given by

vmp =√2kBT

m, (3.120)

where T is the absolute temperature, kB is the Boltzmann constant andm is the massof each particle.

3.5.2 Doppler Broadening

The Doppler broadening is due to the Doppler effect caused by the distribution ofvelocities of atoms. For non-relativistic velocities (vx � c) the Doppler shift infrequency is

ω = ω0

(1 − vx

c

), (3.121)

where ω is the observed angular frequency, ω0 is the rest angular frequency, vx isthe component of the atom speed along the axis between the observer and the atomand c is the speed of light. The Maxwell-Boltzmann distribution f (vx ) of speedsvx = −c(ω − ω0)/ω0 at temperature T , given by

f (vx ) dvx =(

m

2πkBT

)1/2

e−mv2x/(2kBT ) dvx , (3.122)

Page 78: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.5 Life-Time and Natural Line-Width 69

becomes

f (ω) dω =(

m

2πkBT

)1/2

e−mc2(ω−ω0)2/(2ω2

0kBT ) c

ω0dω (3.123)

in terms of the angular frequency ω. This is the distribution of frequencies seen bythe observer, and the full width at half-maximum of the Gaussian is taken as Dopplerwidth, namely

�D =√8 ln (2)kBT

mc2�ω0 . (3.124)

3.6 Minimal Coupling and Center of Mass

Up to now the interaction between atom and electromagnetic radiation has beeninvestigated by taking into account only the electronic degrees of freedom of theatoms. Here we consider also the effect of the atomic nucleus and for simplicity westudy the hydrogen atom, whose classical complete Hamiltonian reads

Hmatt = p2p2mp

+ p2e2me

− e2

4πε0|rp − re| , (3.125)

where rp and re are the positions of proton and electron, and pp and pe are theircorresponding linear momenta, with mp the proton mass and me the electron mass.

By imposing the minimal coupling to the electromagnetic vector potential A(r),in the Coulomb gauge we get

Hshi f t = (pp − eA(rp))2

2mp+ (pe + eA(re))2

2me− e2

4πε0|rp − re| . (3.126)

We now introduce center of mass position and relative position vectors:

rcm = mprp + mereM

, (3.127)

rrel = re − rp , (3.128)

where M = mp + me is the total mass and μ = mpme/M is the reduced mass, andalso the center of mass momentum and relative momentum vectors:

pcm = pp + pe , (3.129)

prel = mp

Mpe − me

Mpp . (3.130)

Page 79: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

70 3 Electromagnetic Transitions

First of all we notice that

Hmatt = p2cm2M

+ p2rel2μ

− e2

4πε0|rrel | , (3.131)

moreoverHshi f t = Hmatt + HI , (3.132)

where

HI = − e

Mpcm · A(rcm − me

Mrrel) + e

Mpcm · A(rcm + mp

Mrrel)

+ e

mpprel · A(rcm − me

Mrrel) + e

meprel · A(rcm + mp

Mrrel)

+ e2

2mpA(rcm − me

Mrrel)2 + e2

2meA(rcm + mp

Mrrel)2 . (3.133)

In this case, the dipole approximation means

A(rcm − me

Mrrel) � A(rcm + mp

Mrrel) � A(rcm) (3.134)

and the interaction Hamiltonian becomes

HI = e

μprel · A(rcm) + e2

2μA(rcm)2 . (3.135)

We stress that, within the dipole approximation, in the Hamiltonian HI the depen-dence on pcm disappears and that the only dependence on the center of mass isincluded in A(rcm). Indeed, this Hamiltonian is precisely the dipolar Hamiltonian ofEq. (3.20) by using the center of mass as origin of the coordinate system.

3.7 Solved Problems

Problem 3.1108 sodium atoms are excited in the first excited state of sodium by absorption oflight. Knowing that the excitation energy is 2.125 eV and the lifetime is 16 ns,calculate the maximum of the emitting power.

SolutionThe total absorbed energy is given by

E = Nε ,

Page 80: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.7 Solved Problems 71

where N is the number of atoms (and also the number of absorbed photons) while εis the transition energy

ε = 2.125 eV = 2.125 × 1.6 × 10−19 J = 3.4 × 10−19 J .

The absorbed energy is then

E = 108 × 3.4 × 10−19 J = 3.4 × 10−11 J .

The absorbed energy is equal to the energy emitted by spontaneous de-excitation.The emitting power P(t) decays exponentially with time t as

P(t) = P0e−t/τ ,

where P0 is the maximum of the emitting power and τ = 16 ns is the lifetime of theexcited state. It must be

E =∫ ∞

0P(t) dt = P0 τ ,

from which we get

P0 = E

τ= 3.4 × 10−11 J

16 × 10−9 s= 2.1 × 10−3 J

s= 2.1 mW .

Problem 3.2Derive the selection rule�m = 0 of the matrix element 〈n′l ′m ′|z|nlm〉 of the hydro-gen atoms by using the formula [L z, z] = 0.

SolutionRemember that for any eigenstate |nlm〉 of the electron in the hydrogen atom onehas

L z|nlm〉 = �m |nlm〉 .

Moreover[L z, z] = L z z − z L z .

We observe that

〈n′l ′m ′|[L z, z]|nlm〉 = 〈n′l ′m ′|L z z − z L z|nlm〉= 〈n′l ′m ′|L z z|nlm〉 − 〈n′l ′m ′|z L z|nlm〉= m ′

� 〈n′l ′m ′|z|nlm〉 − 〈n′l ′m ′|z|nlm〉 �m

= (m ′ − m)� 〈n′l ′m ′|z|nlm〉 .

Page 81: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

72 3 Electromagnetic Transitions

Because [L z, z] = 0 it follows that 〈n′l ′m ′|[L z, z]|nlm〉 = 0 and also

(m ′ − m)� 〈n′l ′m ′|z|nlm〉 = 0 .

From the last expression we get

〈n′l ′m ′|z|nlm〉 = 0

if m ′ �= m. This is exactly the selection rule �m = 0. We can then write

〈n′l ′m ′|z|nlm〉 = 〈n′l ′m|z|nlm〉 δm ′,m .

Problem 3.3Agas of hydrogen atoms is prepared in the state |n = 2, l = 1,m = 0〉. Calculate theelectric dipole moment associated to the transition to the ground-state |n = 1, l =0,m = 0〉.

SolutionThe generic eigenfunction of the electron in the hydrogen atom in spherical coordi-nates is given by

ψnlm(r, θ,φ) = Rnl(r) Ylm(θ,φ) ,

where Rnl(r) is the radial wavefunction, which depends on the quantum numbers n(principal) ed l (angular momentum), while Ylm(θ,φ) is the angular wavefunction,which depends on the quantum numbers l (angular momentum) and m (projectionof angular momentum).

The wavefunction of the the electron in the ground-state is

ψ100(r) = R10(r)Y00(θ,φ) ,

where

R10(r) = 2

r3/20

e−r/r0

Y00(θ,φ) = 1√4π

,

with r0 the Bohr radius. The wavefunction of the excited state is instead

ψ210(r) = R21(r)Y10(θ,φ) ,

where

R21(r) = 1√3(2r0)3/2

r

r0e−r/(2r0)

Page 82: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.7 Solved Problems 73

Y10(θ,φ) =√

3

4πcos θ .

The vector of the dipole transition is given by

rT = 〈ψ100|r|ψ210〉 =∫

d3r ψ∗100(r) r ψ210(r) .

Remembering that in spherical coordinates

r = (r cosφ sin θ , r sin φ sin θ , r cos θ )

with φ ∈ [0, 2π] and θ ∈ [0,π], and moreover

d3r = dr r2 dφ dθ sin θ ,

the 3 components of the transition vector rT = (xT , yT , zT ) read

xT =√3

∫ ∞

0drr2R10(r)r R21(r)

∫ 2π

0dφ cosφ

∫ π

0dθ sin2 θ cos θ ,

yT =√3

∫ ∞

0drr2R10(r)r R21(r)

∫ 2π

0dφ sin φ

∫ π

0dθ sin2 θ cos θ ,

zT =√3

∫ ∞

0drr2R10(r)r R21(r)

∫ 2π

0dφ

∫ π

0dθ sin θ cos2 θ .

One finds immediately thatxT = yT = 0 ,

because ∫ 2π

0dφ cosφ =

∫ 2π

0dφ sin φ = 0 ,

while ∫ π

0dθ sin θ cos2 θ =

∫ 1

−1d(cos θ) cos2 θ = 2

3.

As a consequence, we have

zT = 1

3√2r40

∫ ∞

0r4 e−3r/(2r0) dr = 25r0

36√2

∫ ∞

0η4e−η dη = 28r0

35√2

,

Page 83: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

74 3 Electromagnetic Transitions

because ∫ ∞

0ηne−η dη = �(n + 1) = n! .

Finally, the modulus of the electric dipole is

|d| = e|rT | = e zT = 28

35√2e r0 ,

where e is the electric charge of the electron. Performing the numerical calculationwe find

|d| = 1.6 × 10−19 C × 0.53 × 10−10 m × 0.74 = 6.3 × 10−30 C m .

Problem 3.4Calculate the rates of spontaneous emission Wspont to the ground-state in a gas ofhydrogen atoms which are initially:(a) in the state |n = 3, l = 2,m = 1〉;(b) in the state |n = 3, l = 1,m = 1〉.

SolutionThe ground-state corresponds to |n = 1, l = 0,m = 0〉.(a) In this case the transition probability is zero, within the dipolar approximation,because �l = 2 and the permitted transitions are only with �l = ±1 and �m =0,±1.(b) In this case the calculations must be done explicitly. The wavefunction of theelectron in the ground-state is

ψ100(r) = 1√π

1

r3/20

e−r/r0 ,

while the wavefunction of the excited state is

ψ311(r) = 2

33√

π

r

r5/20

(1 − r

6r0

)e−r/(3r0) eiφ sin θ ,

where r0 is the Bohr radius. The vector of the dipole transition is given by

rT = 〈ψ100|r|ψ311〉 =∫

d3r ψ∗100(r) r ψ311(r) .

Remembering that in spherical coordinates

r = (r cosφ sin θ , r sin φ sin θ , r cos θ ) ,

Page 84: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.7 Solved Problems 75

and alsod3r = dr r2 dφ dθ sin θ ,

the 3 components of the transition vector rT = (xT , yT , zT ) read

xT = − 2

33 π r40

∫ ∞

0dr r4

(1 − r

6r0

)e−4r/(3r0)

∫ 2π

0dφ cosφ eiφ

∫ π

0dθ sin3 θ ,

yT = − 2

33 π r40

∫ ∞

0dr r4

(1 − r

6r0

)e−4r/(3r0)

∫ 2π

0dφ sin φ eiφ

∫ π

0dθ sin3 θ ,

zT = − 2

33 π r40

∫ ∞

0dr r4

(1 − r

6r0

)e−4r/(3r0)

∫ 2π

0dφ eiφ

∫ π

0dθ cos θ sin2 θ .

One find immediately thatzT = 0 ,

because ∫ π

0dθ cos θ sin2 θ = 0 ,

while ∫ π

0dθ sin3 θ = 22

3.

In addition, we get

∫ 2π

0dφ cosφ eiφ =

∫ 2π

0dφ cos2 φ = π ,

and similarly ∫ 2π

0dφ sin φ eiφ = i

∫ 2π

0dφ sin2 φ = i π .

Instead, for the radial part

∫ ∞

0dr r4

(1 − r

6r0

)e−4r/(3r0) = 37

210r50 ,

and consequently

xT =(

− 2

33 π r40

) (37

210r50

) (π22

3

)= −33

27r0

Page 85: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

76 3 Electromagnetic Transitions

yT =(

− 2

33 π r40

) (37

210r50

) (i π

22

3

)= −i

33

27r0 .

The squared modulus of the transition vector is given by

|rT |2 = |xT |2 + |yT |2 = 236

214r20 = 0.088 r20 = 2.48 × 10−22 m2 ,

where r0 = 5.29 × 10−11 m.The rate of spontaneous transition is given by the formula

Wspont = 4

3c2α ω3

T |rT |2 ,

where c = 3 × 108 m/s, α = e2/(4πε0�c) � 1/137,

ωT = E3 − E1�

= 13.60 eV

(1 − 1

32

)= 13.60 × 1.60 × 10−19J

1.05 × 10−34 J s/rad

8

9= 1.84×1016 rad/s .

Finally, the rate of spontaneous transition reads

Wspont = 1.67 × 108 rad/s .

Problem 3.5Agas of hydrogen atoms is irradiated from all directions by lightwith spectral density

ρ(ω) = ρ0 exp

(− (ω − ω0)

2

1000 ω20

),

where ω0 is the frequency of the α-Lymann (1s–2p) transition.(a) Calculate the value of ρ0 which produces stimulated emission 2p → 1s with therate Wstimul = 12.5 × 108 s−1.(b) Determine the effective electric field Erms of the incoming light, defined as

1

2ε0E

2rms =

∫ ∞

0ρ(ω) dω .

Solution(a) From the Einstein relations we know that the rate of stimulated emissionWstimul

can be written asWstimul = π

3ε0�2|d|2ρ(ω0) ,

Page 86: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.7 Solved Problems 77

where d is the electric dipole momentum. In this transition we are considering

d = −e〈2p|r|1s〉 ,

where e is the electric charge of the electron. This dipole has been previously calcu-lated and its modulus is equal to

|d| = 6.3 × 10−30 C m .

Because in this case ρ(ω0) = ρ0, we obtain

ρ0 = 3ε0�2

π

Wstimul

|d|2 .

Finally, we getρ0 = 2.95 × 10−12 Js/m3 .

(b) The integral to be calculated is

E =∫ ∞

0ρ(ω) dω =

∫ ∞

0ρ0 exp

(− (ω − ω0)

2

1000 ω20

)dω .

With the position

t = ω − ω0√1000 ω0

the integral becomes

E = √1000ρ0 ω0

∫ ∞

− 1√1000

exp(−t2

)dt � 10

√10 ρ0 ω0

∫ ∞

0exp

(−t2)dt

= 10√10 ρ0 ω0

√π

2= 5

√10π ρ0 ω0 .

The frequency ω0 is

ω0 = −13.6 eV

(1

22− 1

12

)= 1.55 × 1016 s−1 ,

and consequentlyE = 1.28 × 106 J/m3 .

Page 87: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

78 3 Electromagnetic Transitions

The effective electric field reads

Erms =√2Eε0

= 5.38 × 108 V/m ,

where ε0 = 8.85 × 10−12 J/(mV2) is the dielectric constant in the vacuum.

Problem 3.6Calculate both natural and Doppler width of the α-Lymann line in a hydrogen gas atthe temperature 1000K, knowing that the lifetime of the excited state is 0.16×10−8 sand the wavelength of the transition is 1214 × 10−10 m.

SolutionThe α-Lymann line is associated to the 1s → 2p transition of the hydrogen atom.The natural width between the states |i〉 and | j〉 is given by

�νN = 1

(1

τi+ 1

τ j

)

where τi is the lifetime of the state |i〉 and τ j is the lifetime of the state | j〉. In ourproblem τ1s = ∞, because |1s〉 is the ground-state, while τ2p = 0.16 × 10−8 s forthe excited state |2p〉. It follows that

�νN = 1

(1

τ1s+ 1

τ1p

)= 1

6.28

(0 + 1

0.16 × 10−8

)s−1 = 9.9 × 107 Hz .

The Doppler width depends instead on the temperature T = 1000 ◦K and on thefrequency ν of the transition according to the formula

�νD = ν√8 ln(2)

√kBT

mHc2= ν

c

√8 ln(2)

√kBT

mH,

where kB = 1.3 × 10−23 J/K is the Boltzmann constant, mH = 1.6 × 10−27 kg isthe mass of an hydrogen atom, while c = 3 × 108 m/s is the speed of light in thevacuum. We know that

ν

c= 1

λ= 1

1.216 × 10−7 m= 8.22 × 106 m−1 ,

and then we get

�νD = 8.22 × 106 × 2.35 ×√1.3 × 10−23 × 103

1.6 × 10−27Hz = 5.6 × 1010 Hz .

Page 88: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

3.7 Solved Problems 79

Problem 3.7Calculate the pressure (collisional) width of the α-Lymann line for a gas of hydrogenatoms with density 1012 atoms/m3 and collisional cross-section 10−19 m2 at thetemperature 103 K.

SolutionThe collisional width is given by

�νC = 1

2πτcol,

where τcol is the collision time. This time depends on the cross-section σ = 10−19 m2

and on the gas density n = 1012 m−3 according to the formula

τcol = 1

n σ vmp,

where vmp = √2kBT/mH is the velocity corresponding the the maximum of the

Maxwell-Boltzmann distribution. We have then

�νC = n σ

√2kBT

mH.

Because T = 103 K, kB = 1.3 × 10−23 J/K and mH = 1.6 × 10−27 kg, we finallyobtain

�νC = 6.4 × 10−5 Hz .

Problem 3.8Helium atoms in a gas absorb light in the transition a → b, where |a〉 and |b〉 aretwo excited states. Knowing that the wavelength of the transition is 501.7 nm andthat the lifetimes are τb = 1.4 ns and τa = 1 ms, calculate: (a) the natural width;(b) the Doppler width. The gas is at the temperature 1000K.

SolutionThe natural width of the spectral line between the states |a〉 and |b〉 is given by

�νN = 1

(1

τa+ 1

τb

).

Consequently we obtain

�νN = 1

6.28

(1

10−3+ 1

1.4 × 10−9

)Hz = 1.14 × 108 Hz .

Page 89: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

80 3 Electromagnetic Transitions

The Doppler width is instead

�νD = ν

c

√8 ln(2)

√kBT

mHe,

where T = 103 K is the temperature of the gas, kB = 1.3×10−23 J/K is theBoltzmannconstant, mHe = 4 × 1.6 × 10−27 kg is the mass of a helium atom, while c =3 × 108 m/s is the velocity of light in the vacuum. We know that

ν

c= 1

λ= 1

501.7 × 10−9 m= 1.99 × 106 m−1 ,

and then we get�νD = 6.77 × 109 Hz .

Further Reading

For classical and quantum electrodynamics and radiative transitions:F. Mandl and G. Shaw, Quantum Field Theory, Chap. 1, Sects. 1.3 and 1.4 (Wiley,New York, 1984).For lifetime and line widths:B.H. Bransden and C.J. Joachain: Physics of Atoms and Molecules, Chap. 4,Sects. 4.6 and 4.7 (Prentice Hall, Upper Saddle River, 2003).

Page 90: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 4The Spin of the Electron

In this chapter we explain the origin of the intrinsic angular momentum, also knownas the spin, of the electron on the basis on the Dirac equation. After introducingthe Dirac equation for a relativistic massive and charged particle coupled to theelectromagnetic field, we study its non relativistic limit deriving the Pauli equation.This Pauli equation, which is nothing but the Schrödinger equationwith an additionalterm containing the spin operator, predicts very accurately the magnetic moment ofthe electron. Finally, we discuss the relativistic hydrogen atom and the fine-structurecorrections to the non relativistic one.

4.1 The Dirac Equation

The classical energy of a nonrelativistic free particle is given by

E = p2

2m, (4.1)

where p is the linear momentum and m the mass of the particle. The Schrödingerequation of the corresponding quantum particle with wavefunction ψ(r, t) is easilyobtained by imposing the quantization prescription

E → i�∂

∂t, p → −i�∇ . (4.2)

In this way one gets the time-dependent Schrödinger equation of the free particle,namely

i�∂

∂tψ(r, t) = − �

2

2m∇2ψ(r, t) , (4.3)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_4

81

Page 91: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

82 4 The Spin of the Electron

obtained for the first time in 1926 by Erwin Schrödinger. The classical energy of arelativistic free particle is instead given by

E =√p2c2 + m2c4 , (4.4)

where c is the speed of light in the vacuum. By applying directly the quantizationprescription (4.2) one finds

i�∂

∂tψ(r, t) =

√−�2c2∇2 + m2c4 ψ(r, t) . (4.5)

This equation is quite suggestive but the square-root operator on the right side is avery difficult mathematical object. For this reason in 1927 Oskar Klein and WalterGordon suggested to start with

E2 = p2c2 + m2c4 (4.6)

and then to apply the quantization prescription (4.2). In this way one obtains

− �2 ∂2

∂t2ψ(r, t) = (−�

2c2∇2 + m2c4)

ψ(r, t) (4.7)

the so-called Klein-Gordon equation, which can be re-written in the following form

(1

c2∂2

∂t2− ∇2 + m2c2

�2

)ψ(r, t) = 0 , (4.8)

i.e. a generalization of Maxwell’s wave equation for massive particles. This equationhas two problems: (i) it admits solutions with negative energy; (ii) the space integralover the entire space of the non negative probability density ρ(r, t) = |ψ(r, t)|2 ≥ 0is generally not time-independent, namely

d

dt

R3ρ(r, t) d3r �= 0 . (4.9)

Nowadaysweknow that to solve completely these twoproblems it is necessary to pro-mote ψ(r, t) to a quantum field operator. Within this second-quantization (quantumfield theory) approach the Klein-Gordon equation is now used to describe relativisticparticles with spin zero, like the pions or the Higgs boson.

In 1928 Paul Dirac proposed a different approach to the quantization of the rel-ativistic particle. To solve the problem of Eq. (4.9) he considered a wave equationwith only first derivatives with respect to time and space and introduced the classicalenergy

E = c α · p + β mc2 , (4.10)

Page 92: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.1 The Dirac Equation 83

such that squaring it one recovers Eq. (4.6). This condition is fulfilled only if α =(α1, α2, α3) and β satisfy the following algebra of matrices

α21 = α2

2 = α23 = β 2 = I , (4.11)

αi α j + α j αi = 0 , i �= j (4.12)

αi β + β αi = 0 , ∀i (4.13)

where 1 is the identity matrix and 0 is the zero matrix. The smallest dimension inwhich the so-called Dirac matrices αi and β can be realized is four. In particular, onecan write

αi =(02 σi

σi 02

), β =

(I2 0202 − I2

), (4.14)

where I2 is the 2 × 2 identity matrix, 02 is the 2 × 2 zero matrix, and

σ1 =(0 11 0

)σ2 =

(0 −ii 0

)σ3 =

(1 00 −1

)(4.15)

are the Pauli matrices. Equation (4.10) with the quantization prescription (4.2) gives

i�∂

∂t�(r, t) =

(−i�c α · ∇ + β mc2

)�(r, t) , (4.16)

which is the Dirac equation for a free particle. Notice that the wavefunction �(r, t)has four components in the abstract space of Dirac matrices, i.e. this spinor field canbe written

�(r, t) =

⎛⎜⎜⎝

ψ1(r, t)ψ2(r, t)ψ3(r, t)ψ4(r, t)

⎞⎟⎟⎠ . (4.17)

In explicit matrix form the Dirac equation is thus given by

i�∂

∂t

⎛⎜⎜⎝

ψ1(r, t)ψ2(r, t)ψ3(r, t)ψ4(r, t)

⎞⎟⎟⎠ = H

⎛⎜⎜⎝

ψ1(r, t)ψ2(r, t)ψ3(r, t)ψ4(r, t)

⎞⎟⎟⎠ (4.18)

Page 93: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

84 4 The Spin of the Electron

where

H =

⎛⎜⎜⎜⎝

mc2 0 −i�c ∂∂z −i�c( ∂

∂x − i ∂∂y )

0 mc2 −i�c( ∂∂x + i ∂

∂y ) i�c ∂∂z

−i�c ∂∂z −i�c( ∂

∂x − i ∂∂y ) −mc2 0

−i�c( ∂∂x + i ∂

∂y ) i�c ∂∂z 0 −mc2

⎞⎟⎟⎟⎠ .

(4.19)

It is easy to show that the Dirac equation satisfies the differential law of currentconservation. In fact, left-multiplying Eq. (4.16) by

�+(r, t) = (ψ∗1(r, t),ψ

∗2(r, t),ψ

∗3(r, t),ψ

∗4(r, t)

)(4.20)

we get

i��+ ∂�

∂t= −i�c�+α · ∇� + mc2�+β � . (4.21)

Considering the Hermitian conjugate of the Dirac equation (4.16) and right-multiplying it by �(r, t) we find instead

− i�∂�+

∂t� = i�c α · ∇�+� + mc2�+β � . (4.22)

Subtracting the last two equations we obtain the continuity equation

∂tρ(r, t) + ∇ · j(r, t) = 0 , (4.23)

where

ρ(r, t) = �+(r, t)�(r, t) =4∑

i=1

|ψi (r, t)|2 (4.24)

is the probability density, and j(r, t) is the probability current with three components

jk(r, t) = c�+(r, t)αk�(r, t) . (4.25)

Finally, we observe that from the continuity equation (4.23) one finds

d

dt

R3ρ(r, t) d3r = 0 , (4.26)

by using the divergence theorem and imposing a vanishing current density on theborder at infinity. Thus, contrary to the Klein-Gordon equation, the Dirac equationdoes not have the probability density problem.

Page 94: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.2 The Pauli Equation and the Spin 85

4.2 The Pauli Equation and the Spin

In this section we analyze the non-relativistic limit of the Dirac equation. Let ussuppose that the relativistic particle has the electric charge q. In presence of anelectromagnetic field, by using the Gauge-invariant substitution

i�∂

∂t→ i�

∂t− q φ(r, t) (4.27)

−i�∇ → −i�∇ − qA(r, t) (4.28)

in Eq. (4.16), we obtain

i�∂

∂t�(r, t) =

(c α · (p − qA(r, t)

) + β mc2 + q φ(r, t))

�(r, t) , (4.29)

where p = −i�∇, φ(r, t) is the scalar potential and A(r, t) the vector potential.To workout the non-relativistic limit of Eq. (4.29) it is useful to set

�(r, t) = e−imc2t/�

⎛⎜⎜⎝

ψ1(r, t)ψ2(r, t)χ1(r, t)χ2(r, t)

⎞⎟⎟⎠ = e−imc2t/�

(ψ(r, t)χ(r, t)

), (4.30)

where ψ(r, t) and χ(r, t) are two-component spinors, for which we obtain

i�∂

∂t

(ψχ

)=

(q φ c σ · (p − qA)

c σ · (p − qA) q φ − 2mc2

)(ψχ

)(4.31)

where σ = (σ1, σ2, σ3

). Remarkably, only in the lower equation of the system it

appears the mass term mc2, which is dominant in the non-relativistic limit. Indeed,under the approximation

(i� ∂

∂t − q φ + 2mc2)χ � 2mc2 χ, the previous equations

become

(i� ∂ψ

∂t0

)=

(q φ c σ · (p − qA)

c σ · (p − qA) −2mc2

)(ψχ

), (4.32)

from which

χ = σ · (p − qA)

2mcψ . (4.33)

Inserting this expression in the upper equation of the system (4.32) we find

i�∂

∂tψ =

([σ · (p − qA)

]22m

+ q φ

)ψ . (4.34)

Page 95: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

86 4 The Spin of the Electron

From the identity

[σ · (

p − qA)]2 = (p − qA)2 − i q (p ∧ A) · σ (4.35)

where p = −i�∇, and using the relation B = ∇ ∧ Awhich introduces the magneticfield we finally get

i�∂

∂tψ(r, t) =

((−i�∇ − qA(r, t))2

2m− q

mB(r, t) · S + q φ(r, t)

)ψ(r, t) ,

(4.36)that is the so-called Pauli equation with

S = �

2σ . (4.37)

the spin operator. This equation was introduced in 1927 (a year before the Diracequation) by Wolfgang Pauli as an extension of the Schrödinger equation with thephenomenological inclusion of the spin operator. If the magnetic field B is constant,the vector potential can be written as

A = 1

2B ∧ r (4.38)

and then

(p − qA

)2 = p2 − 2qA · p + q2A2 = p2 − qB · L + q2(B ∧ r)2 , (4.39)

with L = r ∧ p the orbital angular momentum operator. Thus, the Pauli equation fora particle of charge q in a constant magnetic field reads

i�∂

∂tψ(r, t) =

(−�

2∇2

2m− q

2mB ·

(L + 2S

)+ q2

2m(B ∧ r)2 + q φ(r, t)

)ψ(r, t) .

(4.40)In conclusion, we have shown that the spin S naturally emerges from the Dirac equa-tion. Moreover, the Dirac equation predicts very accurately the magnetic momentμS

of the electron (q = −e, m = me) which appears in the spin energy Es = −μS · Bwhere

μS = −geμB

�S (4.41)

with gyromagnetic ratio ge = 2 and Bohr magneton μB = e�/(2m) � 5.79 × 10−5

eV/T.

Page 96: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.3 Dirac Equation with a Central Potential 87

4.3 Dirac Equation with a Central Potential

We now consider the stationary Dirac equation with the confining spherically-symmetric potential V (r) = V (|r|), namely

(−i�c α · ∇ + β mc2 + V (r)

)�(r) = E �(r) . (4.42)

This equation is easily derived from Eq. (4.29) setting A = 0, qφ = V (r), and

�(r, t) = e−i Et/� �(r) . (4.43)

The relativistic Hamiltonian

H = −i�c α · ∇ + β mc2 + V (r) (4.44)

commutes with the total angular momentum operator

J = L + S = r ∧ p + �

2σ (4.45)

because the external potential is spherically symmetric. In fact, one can show that

[H , L] = −i�c α ∧ p = −[H , S] . (4.46)

Consequently one has[H , J] = 0 , (4.47)

and also[H , J 2] = 0 , [ J 2, Jx ] = [ J 2, Jy] = [ J 2, Jz] = 0 , (4.48)

where the three components Jx , Jy , Jz of the total angularmomentum J = ( Jx , Jy, Jz)satisfy the familiar commutation relations

[ Ji , J j ] = i� εi jk Jk (4.49)

with

εi jk =⎧⎨⎩

1 if (i, j, k) is (x, y, z) or (z, x, y) or (y, z, x)−1 if (i, j, k) is (x, z, y) or (z, y, x) or (y, x, z)0 if i = j or i = k or j = k

(4.50)

the Levi-Civita symbol (also called Ricci-Curbastro symbol). Note that these com-mutation relations can be symbolically synthesized as

J ∧ J = i� J . (4.51)

Page 97: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

88 4 The Spin of the Electron

Indicating the states which are simultaneous eigenstates of H , J 2 and Jz as |njm j 〉,one has

H |njm j 〉 = Enj |njm j 〉 , (4.52)

J 2|njm j 〉 = �2 j ( j + 1) |njm j 〉 , (4.53)

Jz|njm j 〉 = �m j |njm j 〉 , (4.54)

where j is the quantum number of the total angular momentum andm j = − j,− j +1,− j + 2, ..., j − 2, j − 1, j the quantum number of the third component of thetotal angular momentum.

In conclusion, we have found that, contrary to the total angular momentum J, theorbital angular momentum L and the spin S are not constants of motion of a particlein a central potential.

4.3.1 Relativistic Hydrogen Atom and Fine Splitting

Let us consider now the electron of the hydrogen atom. We set q = −e,m = me and

V (r) = − e2

4πε0|r| = − e2

4πε0 r. (4.55)

Then the eigenvalues Enj of H are given by

Enj = mc2√1 + α2(

n− j− 12 +

√( j+ 1

2 )2−α2)2

− mc2 , (4.56)

with α = e2/(4πε0�c) � 1/137 the fine-structure constant. We do not prove thisremarkable quantization formula, obtained independently in 1928 by Charles GaltonDarwin and Walter Gordon, but we stress that expanding it in powers of the fine-structure constant α to order α4 one gets

Enj = E (0)n

[1 + α2

n

(1

j + 12

− 3

4n

)], (4.57)

where

E (0)n = −1

2mc2

α2

n2= −13.6 eV

n2(4.58)

Page 98: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.3 Dirac Equation with a Central Potential 89

Fig. 4.1 Fine splitting forthe hydrogen atom. On theleft there are thenon-relativistic energy levelsobtained by solving theSchrödinger equation with aCoulomb potential(Coulomb). On the rightthere are the energy levelsobtained by taking intoaccount relativisticcorrections (Fine Structure)

is the familiar Bohr quantization formula of the non relativistic hydrogen atom. Theterm which corrects the Bohr formula, given by

�E = E (0)n

α2

n

(1

j + 12

− 3

4n

), (4.59)

is called fine splitting correction. This term removes the non relativistic degeneracyof energy levels, but not completely: double-degenerate levels remain with the samequantum numbers n and j but different orbital quantum number l = j ± 1/2.

We have seen that, strictly speaking, in the relativistic hydrogen atom nor theorbital angularmomentum L nor the spin S are constants ofmotion.As a consequencel, ml , s and ms are not good quantum numbers. Nevertheless, in practice, due to thesmallness of fine-splitting corrections, one often assumes without problems that bothL and S are approximately constants of motion (Fig. 4.1).

4.3.2 Relativistic Corrections to the Schrödinger Hamiltonian

It is important to stress that the relativistic Hamiltonian H of the Dirac equationin a spherically-symmetric potential, given by Eq. (4.44), can be expressed as thefamiliar non relativistic Schrödinger Hamiltonian

H0 = − �2

2m∇2 + V (r) (4.60)

Page 99: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

90 4 The Spin of the Electron

plus an infinite sum of relativistic quantum corrections. To this aim one can start fromthe Dirac equation, written in terms of bi-spinors, i.e. Eq. (4.31) with A(r, t) = 0and qφ(r, t) = V (r), which gives

E

(ψχ

)=

(V (r) c σ · pc σ · p V (r) − 2mc2

)(ψχ

)(4.61)

setting ψ(r, t) = ψ(r) e−i Et/� and χ(r, t) = χ(r) e−i Et/�. The lower equation ofthis system can be written as

χ = c σ · pE − V (r) + 2mc2

ψ . (4.62)

This is an exact equation. If E − V (r) � 2mc2 the equation becomes

χ = σ · p2mc

ψ , (4.63)

which is exactly the stationary version of Eq. (4.33) with A(r, t) = 0. We can dosomething better by expanding Eq. (4.62) with respect to the small term (E −V (r))/2mc2 obtaining

χ = σ · p2mc

(1 − E − V (r)

2mc2+ ...

)ψ . (4.64)

Inserting this expression in the upper equation of the system and neglecting the higherorder terms symbolized by the three dots, and after some tedious calculations, onefinds

E ψ = H ψ , (4.65)

whereH = H0 + H1 + H2 + H3 , (4.66)

with

H1 = − �4

8m3c2∇4 , (4.67)

H2 = 1

2m2c21

r

dV (r)

drL · S , (4.68)

H3 = �2

8m2c2∇2V (r) , (4.69)

with H1 the relativistic correction to the electron kinetic energy, H2 the spin-orbitcorrection, and H3 the Darwin correction.

Page 100: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.3 Dirac Equation with a Central Potential 91

If the external potential V(r) is that of the hydrogen atom, i.e. V (r) = −e2/(4πε0|r|), one finds immediately that H3 = (�2e2)/(8m2c2ε0)δ(r) because∇2(1/|r|) = −4πδ(r). In addition, by applying the first-order perturbation theoryto H with H0 unperturbed Hamiltonian, one gets exactly Eq. (4.57) of fine-structurecorrection. Physically one can say that the relativistic fine structure is due to thecoupling between the spin S and the orbital angular momentum L of the electron.Moreover, we observe that

L · S = 1

2

(J 2 − L2 − S2

)(4.70)

becauseJ 2 = (L + S)2 = L2 + S2 + 2 L · S , (4.71)

and this means that the Hamiltonian H of Eq. (4.66) commutes with L2 and S2 butH does not commute with L z and Sz .

Actually, also the nucleus (the proton in the case of the hydrogen atom) has itsspin I which couples to electronic spin to produce the so-called hyperfine structure.However, typically, hyperfine structure has energy shifts orders of magnitude smallerthan the fine structure.

4.4 Solved Problems

Problem 4.1One electron is set in a uniform magnetic field B = (0, 0, B0). Calculate the expec-tation value of the spin S along the x axis if at t = 0 the spin is along the z axis.

SolutionThe Hamiltonian operator of the spin is

H = −μS · B ,

where μ is the magnetic dipole moment of the electron, given by

μS = −geμB

�S = −1

2ge μB σ ,

with ge = 2.002319 � 2 the gyromagnetic ratio of the electron, μB = e�/(2m) =9.27 × 10−24 J/T theBohrmagneton and σ = (σ1, σ2, σ3) the vector of Paulimatri-ces:

σ1 =(0 11 0

)

Page 101: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

92 4 The Spin of the Electron

σ2 =(0 −ii 0

)

σ3 =(1 00 −1

).

The Hamiltonian operator can be written as

H = 1

2�ω0 σ3 = 1

2�ω0

(1 00 −1

),

where ω0 = geμB B0/� is the Larmor frequency of the system, with � = h/(2π) thereduced Planck constant. The initial state of the system is

|ψ(0)〉 =(10

)= | ↑ 〉 ,

while (01

)= | ↓ 〉

is the state along the third component of spin. The state at time t is then given by

|ψ(t)〉 = e−i H t/�|ψ(0)〉 = e−iω0σ3t/2| ↑ 〉 = e−iω0t/2| ↑ 〉 ,

becauseσ3| ↑ 〉 = | ↑ 〉

ande−iω0σ3t/2| ↑ 〉 = e−iω0t/2| ↑ 〉 .

The expectation value at time t of the spin component along the x axis is then

〈Sx (t)〉 = 〈ψ(t)|�2σ1|ψ(t)〉 = �

2〈↑ |eiω0t/2σ1e

−iω0t/2| ↑ 〉 = �

2〈↑ |σ1| ↑ 〉 .

Observing theσ1| ↑ 〉 = | ↓ 〉 ,

and also〈↑ | ↓ 〉 = 0 ,

we conclude that〈Sx (t)〉 = 0 .

Page 102: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.4 Solved Problems 93

This means that if initially the spin is in the same direction of the magnetic field itremains in that direction forever: it is a stationary state with a time-dependence onlyin the phase. Instead, the components of spin which are orthogonal to the magneticfield have always zero expectation value.

Problem 4.2One electron is set in a uniform magnetic field B = (0, 0, B0). Calculate the expec-tation value of the spin S along the x axis if at t = 0 the spin is along the x axis.

SolutionThe Hamiltonian operator of spin is given by

H = 1

2�ω0 σ3

where ω0 = geμB B0/� is the Larmor frequency of the system, with ge gyromagneticfactor and μB Bohr magneton. The initial state of the system is

|ψ(0)〉 = 1√2

(| ↑ 〉 + | ↓ 〉) ,

because

σ1|ψ(0)〉 = 1√2

(σ1| ↑ 〉 + σ1| ↓ 〉) = 1√

2(| ↓ 〉 + | ↑ 〉) = |ψ(0)〉 .

The state at time t is then

|ψ(t)〉 = e−i H t/�|ψ(0)〉 = e−iω0σ3t/2 1√2

(| ↑ 〉 + | ↓ 〉) = 1√2

(e−iω0t/2| ↑ 〉 + eiω0t/2| ↓ 〉

),

becauseσ3| ↑ 〉 = | ↑ 〉

σ3| ↓ 〉 = −| ↓ 〉

ande−iω0σ3t/2| ↑ 〉 = e−iω0t/2| ↑ 〉e−iω0σ3t/2| ↓ 〉 = eiω0t/2| ↓ 〉 .

The expectation value at time t of the spin component along x axis reads

〈Sx (t)〉 = 〈ψ(t)|�2

σ1|ψ(t)〉 = �

4

(〈↑ |eiω0t/2 + 〈↓ |e−iω0t/2

)σ1

(e−iω0t/2| ↑ 〉 + eiω0t/2| ↓ 〉

)

= �

4

(eiω0t 〈↑ | ↑ 〉 + e−iω0t 〈↓ | ↓ 〉

)= �

2cos (ω0t) .

Page 103: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

94 4 The Spin of the Electron

Problem 4.3One electronwith spin S, initially along z axis, is under the action of a time-dependentmagnetic field B = B1(cos (ωt)ex + sin (ωt)ey). Calculate the spin state at time t .

SolutionThe Hamiltonian operator of spin is given by

H(t) = 1

2�ω1

(cos (ωt) σ1 + sin (ωt) σ2

),

where �ω1 = geμB B1, with ge gyromagnetic factor of the electron and μB Bohrmagneton. In matrix form the Hamiltonian operator reads

H(t) = �

2

(0 ω1e−iωt

ω1eiωt 0

),

while the spin initial state is

|ψ(0)〉 = | ↑ 〉 =(10

).

The spin state at time t can be written instead as

|ψ(t)〉 = a(t)| ↑ 〉 + b(t)| ↓ 〉 = a(t)

(10

)+ b(t)

(01

),

where a(t) and b(t) are time-dependent functions to be determined with initial con-ditions

a(0) = 1 , b(0) = 0 .

The time-dependent Schrödinger equation

i�∂

∂t|ψ(t)〉 = H(t)|ψ(t)〉 ,

can be rewritten as

i�

(a(t)b(t)

)= �

2

(0 ω1e−iωt

ω1eiωt 0

) (a(t)b(t)

).

Thus the complex functions a(t) and b(t) satisfy the following system of first-orderordinary differential equations:

2 i a(t) = ω1 e−iωt b(t) ,

2 i b(t) = ω1 eiωt a(t) .

Page 104: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.4 Solved Problems 95

The time-dependent phase factors can be removed introducing the new variables c(t)and d(t) such that

a(t) = e−iωt/2 c(t) ,

d(t) = eiωt/2 d(t) .

The differential system then becomes

2 i c(t) = −ω c(t) + ω1 d(t) ,

2 i d(t) = ω d(t) + ω1 c(t) .

This system can be solved by using the Laplace transform

F(s) =∫ ∞

0f (t) e−st dt .

Indeed, by applying the Laplace transform, the system becomes

2 i (s C(s) − c(0)) = −ωC(s) + ω1 D(s) .

2 i (s D(s) − d(0)) = ω D(s) + ω1 C(s) .

Taking into account the initial conditions c(0) = 1 and d(0) = 0, one can get D(s)from the second equation, namely

D(s) = ω1

2is − ωC(s) ,

and insert it into the first one. In this way

C(s) = s

ω2R + s2

+ iω

2

1

ω2R + s2

,

where ωR = 12

√ω21 + ω2 is the so-called Rabi frequency. By applying the Laplace

anti-transform we immediately obtain

c(t) = cos (ωRt) + iω

2ωRsin (ωRt) .

Coming back to the initial variables a(t) and b(t) we eventually find

|a(t)|2 = ω2

4ω2R

+ ω21

4ω2R

cos2 (ωRt) , |b(t)|2 = ω21

4ω2R

sin2 (ωRt) .

Page 105: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

96 4 The Spin of the Electron

which clearly satisfy the relation |a(t)|2 + |b(t)|2 = 1. Notice that for ω = 0 it fol-lows ωR = ω1/2 from which

|a(t)|2 = cos2 (ωRt) , |b(t)|2 = sin2 (ωRt) .

In this case there is complete spin-flip during the dynamics.

Problem 4.4Derive the non relativistic formula of Bohr for the spectrum of the hydrogen atomfrom the relativistic expression

Enj = mc2√1 + α2(

n− j− 12 +

√( j+ 1

2 )2−α2)2

− mc2

which is obtained from the Dirac equation, expanding it in powers of α � 1/137 atorder α2.

SolutionSetting x = α2 the relativistic spectrum can be written as

E(x) = mc2 f (x) − mc2 ,

where

f (x) =(1 + x

(A + √B − x)2

)−1/2

,

with A = n − j − 1/2 e B = ( j + 1/2)2. Let us now expand f (x) in MacLaurinseries at first order:

f (x) = f (0) + f ′(0)x ,

wheref (0) = 1

f ′(0) = −1

2

1

(A + √B)2

= −1

2

1

n2.

We get at first order in x

E(x) = mc2 − 1

2mc2

x

n2− mc2 = −1

2mc2

x

n2,

namely

En = −1

2mc2

α2

n2,

which is exactly the non relativistic Bohr formula.

Page 106: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

4.4 Solved Problems 97

Problem 4.5Calculate the fine splitting for the state |3p〉 of the hydrogen atom.

SolutionFrom the relativistic formula of the energy levels of the hydrogen atom

Enj = mc2√1 + α2(

n− j− 12 +

√( j+ 1

2 )2−α2)2

− mc2

expanding to order α4 we get

Enj = E (0)n

[1 + α2

n

(1

j + 12

− 3

4n

)]

where

E (0)n = −1

2mc2

α2

n2= −13.6 eV

n2.

The integer number j is the quantum number of the total angular momentum J =L + S, where j = 1/2 if l = 0 and j = l − 1/2, l + 1/2 if l �= 0. The state |3p〉means n = 3 and l = 1, consequently j = 1/2 or j = 3/2. The hyperfine correctionfor j = 1/2 reads

�E3, 12= E (0)

3

α2

3

(1 − 1

4

)= E (0)

3

α2

4= −13.6 eV

9

1

1372 · 4 = −2.01 × 10−5 eV .

The hyperfine correction for j = 3/2 is instead

�E3, 32= E(0)

3α2

3

(1

2− 1

4

)= E(0)

3α2

12= −13.6 eV

9

1

1372 · 12 = −6.71 × 10−6 eV .

Further Reading

For the Dirac equation and the Pauli equation:J.D. Bjorken and S.D. Drell, Relativistic Quantum Mechanics, Chap. 1, Sects. 1.1,1.2, 1.3, 1.4 (McGraw-Hill, New York, 1964).For the relativistic hydrogen atom and the fine splitting:V.B. Berestetskii, E.M. Lifshitz, L.P. Pitaevskii, Relativistic Quantum Theory, vol.4 of Course of Theoretical Physics, Chap. 4, Sects. 34, 35, 36 (Pergamon Press,Oxford, 1971).B.H. Bransden and C.J. Joachain: Physics of Atoms and Molecules, Chap. 5, Sect.5.1 (Prentice Hall, Upper Saddle River, 2003).

Page 107: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 5Energy Splitting and Shift Due to ExternalFields

In this chapter we study the Stark effect and the Zeeman effect on atoms, and inparticular on theHydrogen atom. In the Stark effect one finds that an external constantelectric field induces splitting of degenerate energy levels and energy shift of theground-state. Similarly, in the Zeeman effect an external constant magnetic fieldinduces splitting of degenerate energy levels, but the splitting properties stronglydepend on the intensity of the external magnetic field.

5.1 Stark Effect

Let us consider the Hydrogen atom under the action of a constant electric field E.We write the constant electric field as

E = E uz = (0, 0, E), (5.1)

choosing the z axis in the same direction of E. The Hamiltonian operator of thesystem is then given by

H = H0 + HI , (5.2)

where

H0 = p2

2m− e2

4πε0 r(5.3)

is the non-relativisticHamiltonian of the electron in the hydrogen atom (withm = me

the electron mass), while

HI = −e φ(r) = e E · r = −d · E = e E z (5.4)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_5

99

Page 108: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

100 5 Energy Splitting and Shift Due to External Fields

is the Hamiltonian of the interaction due to the electric scalar potential φ(r) = −E ·rsuch thatE = −∇φ, with−e electric charge of the electron and d = −er the electricdipole.

Let |nlml〉 be the eigenstates of H0, such that

H0|nlml〉 = E (0)n |nlml〉 (5.5)

with

E (0)n = −13.6

n2eV , (5.6)

the Bohr spectrum of the hydrogen atom, and moreover

L2|nlml〉 = �2 l(l + 1) |nlml〉 , L z|nlml〉 = � ml |nlml〉 . (5.7)

At the first order of degenerate perturbation theory the energy spectrum is given by

En = E (0)n + E (1)

n , (5.8)

where E (1)n is one of the eigenvalues of the submatrix H n

I , whose elements are

a(n)

l ′m ′l ,lml

= 〈nl ′m ′l |HI |nlml〉 = e E 〈nl ′m ′

l |z|nlml〉 . (5.9)

Thus, in general, there is a linear splitting of degenerate energy levels due to theexternal electric field E . This effect is named after Johannes Stark, who discovered itin 1913. Actually, it was discovered independently in the same year also by AntoninoLo Surdo.

It is important to stress that the ground-state |1s〉 = |n = 1, l = 0,ml = 0〉 of thehydrogen atom is not degenerate and for it 〈1s|z|1s〉 = 0. It follows that there is nolinear Stark effect for the ground-state. Thus, we need the second order perturbationtheory, namely

E1 = E (0)1 + E (1)

1 + E (2)1 , (5.10)

where

E (1)1 = 〈100|eEz|100〉 = 0 , (5.11)

E (2)1 =

∑nlml �=100

|〈nlml |eEz|100〉|2E (0)1 − E (0)

n

= e2E2∞∑n=2

|〈n10|z|100〉|2E (0)1 − E (0)

n

, (5.12)

where the last equality is due to the dipole selection rules. This formula shows thatthe electric field produces a quadratic shift in the energy of the ground state. Thisphenomenon is known as the quadratic Stark effect.

Page 109: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.1 Stark Effect 101

The polarizability αp of an atom is defined in terms of the energy-shift �E1 ofthe atomic ground state energy E1 induced by an external electric field E as follows:

�E1 = −1

2αp E2 . (5.13)

Hence, for the hydrogen atom we can write

αp = −2e2∞∑n=2

|〈n10|z|100〉|2E (0)1 − E (0)

n

= −9

4

e2r20E (0)1

, (5.14)

where the last equality is demonstrated in Problem 5.2, with r0 the Bohr radius.

5.2 Zeeman Effect

Let us consider the hydrogen atom under the action of a constant magnetic field B.According to the Pauli equation, the Hamiltonian operator of the system is given by

H =(p + eA

)22m

− e2

4πε0 r− µS · B (5.15)

where

µS = − e

mS (5.16)

is the spin dipole magnetic moment, with S the spin of the electron, and A is thevector potential, such that B = ∇ ∧A. Because the magnetic field B is constant, thevector potential can be written as

A = 1

2B ∧ r , (5.17)

and then

(p + eA

)2 = p2 + 2ep · A + e2A2 = p2 + 2eB · L + e2(B ∧ r)2 , (5.18)

with L = r ∧ p. In this way the Hamiltonian can be expressed as

H = H0 + HI , (5.19)

Page 110: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

102 5 Energy Splitting and Shift Due to External Fields

where

H0 = p2

2m− e2

4πε0 r(5.20)

is the non-relativistic Hamiltonian of the electron in the hydrogen atom, while

HI = −µ · B + e2

8m(B ∧ r)2 (5.21)

is the Hamiltonian of the magnetic interaction, with

µ = µL + µS = − e

2m(L + 2S) (5.22)

the total dipolemagneticmoment of the electron.Wenowwrite the constantmagneticfield as

B = B uz = (0, 0, B) , (5.23)

choosing the z axis in the same direction ofB. In this way the interactionHamiltonianbecomes

HI = eB

2m

(L z + 2Sz

)+ e2B2

8m(x2 + y2) . (5.24)

The first term, called paramagnetic term, grows linearly with the magnetic field Bwhile the second one, the diamagnetic term, grows quadratically. The paramagneticterm is of the order of μB B, where μB = e�/(2m) = 9.3 × 10−24 J/T = 5.29 ×10−5 eV/T is the Bohr magneton. Because the unperturbed energy of H0 is of theorder of 10 eV, the paramagnetic term can be considered a small perturbation.

5.2.1 Strong-Field Zeeman Effect

Usually the diamagnetic term is much smaller than the paramagnetic one, andbecomes observable only for B of the order of 106/n4 Tesla, i.e. mainly in theastrophysical context. Thus in laboratory the diamagnetic term is usually negligible(apart for very large values of the principal quantum number n) and the effectiveinteraction Hamiltonian reads

HI = eB

2m

(L z + 2Sz

). (5.25)

Page 111: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.2 Zeeman Effect 103

Thus (5.20) is the unperturbed Hamiltonian and (5.25) the perturbing Hamiltonian. Itis clear that this total Hamiltonian is diagonalwith respect to the eigenstates |nlmlms〉and one obtains immediately the following energy spectrum

En,ml ,ms = E (0)n + μB B (ml + 2ms) , (5.26)

where E (0)n is the unperturbed Bohr eigenspectrum and μB = e�/(2m) is the Bohr

magneton, with m the mass of the electron. Equation (5.26) describes the high-fieldZeeman effect, first observed in 1896 by Pieter Zeeman. The field B does not removethe degeneracy in l but it does remove the degeneracy in ml and ms . The selectionrules for dipolar transitions require �ms = 0 and �ml = 0,±1. Thus the spectralline corresponding to a transition n → n′ is split into 3 components, called Lorentztriplet (Fig. 5.1).

Fig. 5.1 Zeeman effect for the hydrogen atom. Due to a (strong) magnetic field the energy of thedegenerate excited state |2p〉 = |n = 2, l = 1〉 is split into 3 levels and the energy of the degenerateexcited state |3d〉 = |n = 3, l = 2〉 is split into 5 levels. But the spectral lines are always three(Lorentz triplet)

Page 112: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

104 5 Energy Splitting and Shift Due to External Fields

5.2.2 Weak-Field Zeeman Effect

In the hydrogen atom the strong-field Zeeman effect is observable if the magneticfield B is between about 1/n3 Tesla and about 106/n4 Tesla, with n the principalquantum number. In fact, as previously explained, for B larger than about 106/n4

Tesla the diamagnetic term is no more negligible. Instead, for B smaller than about1/n3 Tesla the splitting due to the magnetic field B becomes comparable with thesplitting due to relativistic fine-structure corrections. Thus, to study the effect ofa weak field B, i.e. the weak-field Zeeman effect, the unperturbed non-relativisticHamiltonian H0 given by the Eq. (5.20) is no more reliable. One must use insteadthe exact relativistic Hamiltonian or, at least, the non-relativistic one with relativisticcorrections, namely

H0 = H0,0 + H0,1 + H0,2 + H0,3 , (5.27)

where

H0,0 = − �2

2m∇2 − e2

4πε0 r(5.28)

H0,1 = − �4

8m3c2∇4 , (5.29)

H0,2 = 1

2m2c21

r

dV (r)

drL · S , (5.30)

H0,3 = �2

8m2c2∇2V (r) , (5.31)

with H0,0 the non relativistic Hamiltonian, H0,1 the relativistic correction to theelectronkinetic energy, H0,2 the spin-orbit correction, and H0,3 theDarwin correction.In any case, ml and ms are no more good quantum numbers because L z and Sz donot commute with the new H0.

The Hamiltonian (5.27) commutes instead with L2, S2, J 2 and Jz . Consequently,for this Hamiltonian the good quantum numbers are n, l, s, j andm j . Applying againthe first-order perturbation theory, where now (5.27) is the unperturbed Hamiltonianand (5.25) is the perturbing Hamiltonian, one obtains the following energy spectrum

En,l, j,m j = E (0)n, j + E (1)

n,l,s, j,m j, (5.32)

where E (0)n, j is the unperturbed relativistic spectrum and

E (1)n,l,s, j,m j

= eB

2m〈n, l, s, j,m j |L z + 2Sz|n, l, s, j,m j 〉 (5.33)

Page 113: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.2 Zeeman Effect 105

is the first-order correction, which is indeed not very easy to calculate. But we cando it. First we note that

〈n, l, s, j,m j |L z + 2Sz|n, l, s, j,m j 〉 = 〈n, l, s, j,m j | Jz + Sz|n, l, s, j,m j 〉= �m j + 〈n, l, s, j,m j |Sz|n, l, s, j,m j 〉 .

(5.34)

Then, on the basis of the Wigner-Eckart theorem,1 we have

�2 j ( j + 1)〈n, l, s, j,m j |Sz |n, l, s, j,m j 〉 = 〈n, l, s, j,m j |(S · J) Jz |n, l, s, j,m j 〉 ,

(5.35)

from which we obtain

�2 j ( j + 1)〈n, l, s, j,m j |Sz |n, l, s, j,m j 〉 = �m j 〈n, l, s, j,m j |S · J|n, l, s, j,m j 〉

= �m j 〈n, l, s, j,m j | 12(J2 + S2 − L2

)|n, l, s, j,m j 〉

= �m j1

2�2 ( j ( j + 1) + s(s + 1) − l(l + 1)) . (5.36)

In conclusion, for a weak magnetic field B the first-order correction is given by

E (1)n,l,s, j,m j

= μB B gl,s, j m j , (5.37)

where μB = e�/(2m) is the Bohr magneton, and

gl,s, j = 1 + j ( j + 1) + s(s + 1) − l(l + 1)

2 j ( j + 1)(5.38)

is the so-called Landé factor. Clearly, in the case of the electron s = 1/2 and theLandé factor becomes

gl, j = 1 + j ( j + 1) − l(l + 1) + 3/4

2 j ( j + 1). (5.39)

Strictly speaking, the energy splitting described by Eq. (5.37) is fully reliable onlyfor a weak magnetic field B in the range 0 Tesla ≤ B � 1/n3 Tesla, with n theprincipal quantum number. In fact, if the magnetic field B approaches 1/n3 Teslaone observes a complex pattern of splitting, which moves by increasing B from the

1The Wigner-Eckart theorem states that for any vector operator V = (V1, V2, V3) such that[ Ji , V j ] = i�εi jk Vk holds the identity

�2 j ( j + 1) 〈n, l, s, j,m j |V|n, l, s, j,m j 〉 = 〈n, l, s, j,m j |(V · J) J|n, l, s, j,m j 〉 .

In our case V = S and we have considered the z component only.

Page 114: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

106 5 Energy Splitting and Shift Due to External Fields

splitting described by Eq. (5.37) towards the splitting described by Eq. (5.26). Thistransition, observed in 1913 by Friedrich Paschen and Ernst Back, is now called thePaschen-Back effect.

5.3 Solved Problems

Problem 5.1Determine the Stark splitting for the state |2p〉 of the hydrogen atom in a electricfield of 107 V/m.

SolutionWe write the constant electric field as E = E uz , choosing the z axis in the samedirection of E. The Hamiltonian operator is given by

H = H0 + H1 ,

where

H0 = p2

2m− e2

4πε0 r

is the Hamiltonian of the unperturbed system, while

H1 = e E · r = e E z

is the Hamiltonian of the perturbation due to the electric field, with e electric chargeof the electron. Let |nlml〉 be the eigenstate of H0, such that

H0|nlml〉 = E (0)n |nlml〉

with

E (0)n = −13.6

n2eV ,

the Bohr spectrum of the hydrogen atom, and moreover

L2|nlml〉 = � l(l + 1) |nlml〉 , L z|nlml〉 = � ml |nlml〉 .

At the first order of perturbation theory the energetic spectrum is given by

En = E (0)n + E (1)

n ,

Page 115: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.3 Solved Problems 107

where E (1)n is one of the eigenvalues of the submatrix H n

1 , whose elements are

a(n)

l ′m ′l ,lml

= 〈nl ′m ′l |H1|nlml〉 = e E 〈nl ′m ′

l |z|nlml〉 .

If n = 2 then l = 0, 1 and ml = −l,−l + 1, ..., l − 1, l, i.e. ml = 0 if l = 0 andml = −1, 0, 1 if l = 1. It follows that the submatrix H (2)

1 is a 4 × 4 matrix, givenby

H (2)1 =

⎛⎜⎜⎜⎝

a(2)00,00 a(2)

00,10 a(2)00,11 a(2)

00,1−1

a(2)10,00 a(2)

10,10 a(2)10,11 a(2)

10,1−1

a(2)11,00 a(2)

11,10 a(2)11,11 a(2)

11,1−1

a(2)1−1,00 a

(2)1−1,10 a

(2)1−1,11 a

(2)1−1,1−1

⎞⎟⎟⎟⎠ .

Note that

〈nl ′m ′l |z|nlml〉 = 〈nl ′m ′

l |z|nlml〉 δm ′l ,ml δl+l ′,1 ,

i.e. the selection rules �l = ±1 e �ml = 0 hold. It follows that many elements ofthe submatrix H (2)

1 are zero:

H (2)1 =

⎛⎜⎜⎝

0 a(2)00,10 0 0

a(2)10,00 0 0 00 0 0 00 0 0 0

⎞⎟⎟⎠ ,

and also a(2)00,10 = a(2)

10,00. One finds immediately that 2 eigenvalues of H (2)1 are zero.

To determine the other 2 eigenvalues it is sufficient to study the 2 × 2 given by

H (2)1 =

(0 a(2)

00,10

a(2)00,10 0

),

whose eigenvalues are

λ1,2 = ±a(2)00,10 = ±e E 〈200|z|210〉 .

We must now calculate 〈200|z|210〉. By using the completeness relation

1 =∫

d3r|r〉〈r| ,

this matrix element can be written as

〈200|z|210〉 = 〈200|∫

d3r|r〉〈r|z|210〉 =∫

d3r〈200|r〉z〈r|210〉 =∫

d3rψ∗200(r)zψ

∗210(r) .

Page 116: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

108 5 Energy Splitting and Shift Due to External Fields

For the hydrogen atom one has

ψ200(r) = 1

2√2πr3/20

(1 − r

2r0)e−r/(2r0) ,

ψ210(r) = 1

4√2πr3/20

r

r0e−r/(2r0) cos θ ,

where r0 = 0.53 × 10−10 m is the Bohr radius. Remembering that in sphericalcoordinates z = r cos θ, we get

∫d3r ψ∗

200(r)zψ210(r) =∫ ∞

0drr2

∫ 2π

0dφ

∫ π

0dθ sin θ ψ∗

200(r) r cos θ ψ∗210(r) =

= 1

16πr40

∫ ∞

0dr e−r/r0 r4(1 − r

2r0)

∫ 2π

0dφ

∫ 1

−1d(cos θ) cos2 θ

= 1

12r40

∫ ∞

0dr e−r/r0 r4(1 − r

2r0) = 1

12r40(−36r50 ) = −3r0 ,

namely

〈200|z|210〉 = −3r0 .

Finally, the possible values of the perturbing energy E (1)2 read

E (1)2 =

⎧⎨⎩

3eEr00

−3eEr0.

It follows that the unperturbed energy E (0)2 splits into 3 levels, given by E2 =

E (0)2 + E (1)

2 , but the central one of them coincides with the unperturbed level. Inconclusion,the Stark splitting is

�E2 = ±3eEr0 = ±3e × 107V

m· 0.53 × 10−10 m = ±1.6 × 10−3 eV .

Problem 5.2Calculate the Stark shift of the state |1s〉 of the hydrogen atom in a electric field of107 V/m.

Page 117: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.3 Solved Problems 109

SolutionWe write the constant electric field as E = E ez , choosing the z axis in the samedirection of E. The Hamiltonian operator is given by

H = H0 + H1 ,

where

H0 = p2

2m− e2

4πε0r

is the Hamiltonian of the unperturbed system, while

H1 = e E · r = e E z

is the Hamiltonian of the perturbation due to the electric field, with e electric chargeof the electron. At first order of perturbation theory the energy of |1s〉 = |n = 1, l =0,ml = 0〉 is given by

E1 = E (0)1 + E (1)

1 ,

where

E (0)1 = −13.6 eV ,

while E (1)1 is

E (1)1 = 〈100|H1|100〉 = e E 〈100|z|100〉 = 0 .

This matrix element is zero due to the selection rules �l = ±1 e �ml = 0. Thus,we need the second order perturbation theory, namely

E1 = E (0)1 + E (1)

1 + E (2)1 ,

where

E (2)1 =

∑nlml �=100

|〈nlml |eEz|100〉|2E (0)1 − E (0)

n

=∞∑n=2

|〈n10|eEz|100〉|2E (0)1 − E (0)

n

,

by using again the selection rules. Remembering that

E (0)n = E (0)

1

n2,

Page 118: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

110 5 Energy Splitting and Shift Due to External Fields

we can write

E (2)1 = e2E2

E (0)1

∞∑n=2

n2

(n2 − 1)|〈n10|z|100〉|2 .

In addition, knowing that in spherical coordinates z = r cos θ, the matrix element〈n10|z|100〉 reads

〈n10|z|100〉 =∫

d3rψ∗n1o(r)zψ100(r) =

=∫ ∞

0drr3Rn1(r)R10(r)

∫ 2π

0dφ

∫dθ sin θ cos θY10(θ,φ)Y00(θ,φ) .

Because

Y00(θ,φ) = 1√4π

Y10(θ,φ) =√

3

4πcos θ ,

we get

〈n10|z|100〉 = 1√3

∫ ∞

0drr3Rn1(r)R10(r) = 1√

3r0 f (n) ,

where r0 = 0.53 × 10−10 m is the Bohr radius and f (n) is a function of the principalquantum number n. It is possible to show that

∞∑n=2

n2

n2 − 1f (n)2 = 27

8.

From this exact relation we obtain

E (2)1 = e2E2r20

3E (0)1

27

8= 9

8

e2E2r20E (0)1

.

Inserting the numerical values we finally get

E (2)1 = −2.3 × 10−15 eV .

Problem 5.3Determine the Zeeman splitting for the states |1s〉 e |2p〉 of the hydrogen atom in amagnetic field of 10T.

Page 119: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.3 Solved Problems 111

SolutionIn presence of a magnetic field with intensity B between 1/n3 Tesla and 106/n4

Tesla, one has the strong-field Zeeman effect and the energy spectrum is given by

En,m,ms = En + μB B(ml + 2ms) ,

where

En = −13.6 eV

n2

is the Bohr spectrum,ml is the quantum number of the third component of the orbitalangular momentum L and ms is the quantum number of the third component of thespin S. In addition μB = e�/(2m) = 9.3 × 10−24 J/T is the Bohr magneton. In ourproblem we are in the regime of validity of the strong-field Zeeman effect and

μB B = 9.3 × 10−24 J

T× 10 T = 9.3 × 10−23J = 9.3 × 10−23 × 1019

1.6eV = 5.8 × 10−4 eV .

For the state |1s〉 = |n = 1, l = 0,ml = 0〉 one has s = 12 and consequently

ms = − 12 ,

12 . It follows that

�E1,0,− 12

= −μB B = −5.8 × 10−4 eV

�E1,0, 12= μB B = 5.8 × 10−4 eV .

For the state |2p〉 = |n = 2, l = 1,ml = −1, 0, 1〉 one has s = 12 and then

ms = − 12 ,

12 . It follows that

�E2,−1,− 12

= −2μB B = −11.6 × 10−4 eV

�E2,0,− 12

= −μB B = −5.8 × 10−4 eV

�E2,1,− 12

= �E2,−1, 12= 0

�E2,0, 12= μB B = 5.8 × 10−4 eV

�E2,1, 12= 2μB B = 11.6 × 10−4 eV .

Problem 5.4Calculate the Zeeman splitting for the state |2p〉 of the hydrogen atom in a magneticfield of 5 Gauss.

Page 120: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

112 5 Energy Splitting and Shift Due to External Fields

SolutionThe magnetic field B = 5 Gauss = 5 × 10−4 T is quite weak (the magnetic fieldof the Earth is about 0.5 Gauss). With a field B between 0 and 10−2 T, the energyspectrum is well described by

En,l, j,m j = En + μB B gl, j m j ,

where μB = e�/(2m) = 9.3 × 10−24 J/T is the Bohr magneton, m j is the quantumnumber of the third component of the total angular momentum J = L + S, and gl, jis the Landé factor, given by

gl, j = 1 + j ( j + 1) − l(l + 1) + 3/4

2 j ( j + 1).

In this problem the magnetic field is very weak, and one has the weak-field Zeemaneffect. For the state |2p〉 = |n = 2, l = 1,ml = −1, 0, 1〉 one has s = 1

2 and thenms = − 1

2 ,12 . In addition, j = l ± 1

2 = 12 ,

32 from which m j = − 3

2 ,− 12 ,

12 ,

32 . In the

case j = 1/2 the Landé factor reads

g1, 12 = 1 + 3/4 − 2 + 3/4

3/2= 1 − 1/3 = 2

3.

In the case j = 3/2 the Landè factor is instead given by

g1, 32 = 1 + 15/4 − 2 + 3/4

15/2= 1 + 1

3= 4

3.

In our problem

μB B = 9.3 × 10−24 J

T× 5 × 10−4 T = 4.6 × 10−27J = 4.6 × 10−27 × 1019

1.6eV = 2.9 × 10−8 eV .

We can now calculate the Zeeman splittings. For j = 12 one has m j = − 1

2 ,12 and

then

�E1, 12 ,− 12

= −1

3μB B = −0.97 × 10−8 eV

�E1, 12 , 12= 1

3μB B = 0.97 × 10−8 eV .

For j = 32 one has instead m j = − 3

2 ,− 12 ,

12 ,

32 and then

�E1, 32 ,− 32

= −2μB B = −5.8 × 10−8 eV

�E1, 32 ,− 12

= −2

3μB B = −1.93 × 10−8 eV

Page 121: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

5.3 Solved Problems 113

�E1, 32 , 12= 2

3μB B = 1.93 × 10−8 eV

�E1, 32 , 32= 2μB B = 5.8 × 10−8 eV .

Further ReadingFor the Stark effect:R.W. Robinett, Quantum Mechanics: Classical Results, Modern Systems, and Visu-alized Examples, Chap. 19, Sect. 19.6 (Oxford Univ. Press, Oxford, 2006).For the Zeeman effect:B.H. Bransden and C.J. Joachain: Physics of Atoms and Molecules, Chap. 6,Sects. 6.1 and 6.2 (Prentice Hall, Upper Saddle River, 2003).

Page 122: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 6Many-Body Systems

In this chapter we want to analyze atoms with many electrons and, more gener-ally, systems with many interacting identical particles. We first consider the generalproperties of many identical particles with their bosonic or fermionic many-bodywavefunctions, and the connection between spin and statistics which explains thePauli principle and the main features of the periodic table of elements. We thendiscuss the Hartree-Fock approximation and also the density functional theory, his-torically introduced by Thomas and Fermi to model the electronic cloud of atomsand recently applied also to atomic Bose-Einstein condensates. Finally, we illus-trate the Born-Oppenheimer approximation, which is very useful to treat moleculescomposed by many electrons in interaction with several nuclei.

6.1 Identical Quantum Particles

First of all, we introduce the generalized coordinate x = (r,σ) of a particle whichtakes into account the spatial coordinate r but also the intrinsic spin σ pertaining tothe particle. For instance, a spin 1/2 particle has σ = −1/2, 1/2 =↓,↑. By usingthe Dirac notation the corresponding single-particle state is

|x〉 = |r σ〉 . (6.1)

We now consider N identical particles; for instance particles with the same mass andelectric charge. The many-body wavefunction of the system is given by

�(x1, x2, ..., xN ) = �(r1,σ1, r2,σ2, ..., rN ,σN ) , (6.2)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_6

115

Page 123: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

116 6 Many-Body Systems

According to quantum mechanics identical particles are indistinguishable. As a con-sequence, it must be

|�(x1, x2, ..., xi , ..., x j , ..., xN )|2 = |�(x1, x2, ..., x j , ..., xi , ..., xN )|2 , (6.3)

which means that the probability of finding the particles must be independent on theexchange of two generalized coordinates xi and x j . Obviously, for 2 particles thisimplies that

|�(x1, x2)|2 = |�(x2, x1)|2 . (6.4)

Experiments suggests that there are only two kind of identical particles which satisfyEq. (6.3): bosons and fermions. For N identical bosons one has

�(x1, x2, ..., xi , ..., x j , ..., xN ) = �(x1, x2, ..., x j , ..., xi , ..., xN ) , (6.5)

i.e. the many-body wavefunction is symmetric with respect to the exchange of twocoordinates xi and x j . Note that for 2 identical bosonic particles this implies

�(x1, x2) = �(x2, x1). (6.6)

For N identical fermions one has instead

�(x1, x2, ..., xi , ..., x j , ..., xN ) = −�(x1, x2, ..., x j , ..., xi , ..., xN ) , (6.7)

i.e. the many-body wavefunction is anti-symmetric with respect to the exchange oftwo coordinates xi and x j . Note that for 2 identical fermionic particles this implies

�(x1, x2) = −�(x2, x1) . (6.8)

An immediate consequence of the anti-symmetry of the fermionic many-body wave-function is the Pauli Principle: if xi = x j then the many-body wavefunction is zero.In other words: the probability of finding two fermionic particles with the samegeneralized coordinates is zero.

A remarkable experimental fact, which is often called spin-statistics theorembecause can be deduced from other postulates of relativistic quantum field theory, isthe following: identical particles with integer spin are bosonswhile identical particleswith semi-integer spin are fermions. For instance, photons are bosons with spin 1while electrons are fermions with spin 1/2. Notice that for a composed particle itis the total spin which determines the statistics. For example, the total spin (sum ofnuclear and electronic spins) of 4He atom is 0 and consequently this atom is a boson,while the total spin of 3He atom is 1/2 and consequently this atom is a fermion.

Page 124: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.2 Non-interacting Identical Particles 117

6.2 Non-interacting Identical Particles

The quantum Hamiltonian of N identical non-interacting particles is given by

H0 =N∑i=1

h(xi ) , (6.9)

where h(x) is the single-particle Hamiltonian. Usually the single-particle Hamil-tonian is given by

h(x) = − �2

2m∇2 +U (r) , (6.10)

withU (r) the external confining potential. In general the single-particle Hamiltonianh satisfies the eigenvalue equation

h(x) φn(x) = εn φn(x) , (6.11)

where εn are the single-particle eigenenergies and φn(x) the single-particle eigen-functions, with n = 1, 2, ....

The many-body wavefunction �(x1, x2, ..., xN ) of the system can be written interms of the single-particle wavefunctions φn(x) but one must take into account thespin-statistics of the identical particles. For N bosons the simplest many-body wavefunction reads

�(x1, x2, ..., xN ) = φ1(x1) φ1(x2) ... φ1(xN ) , (6.12)

which corresponds to the configuration where all the particles are in the lowest-energy single-particle state φ1(x). This is indeed a pure Bose-Einstein condensate.Note that for 2 bosons the previous expression becomes

�(x1, x2) = φ1(x1)φ1(x2) . (6.13)

Obviously there are infinite configuration which satisfy the bosonic symmetry of themany-body wavefunction. For example, with 2 bosons one can have

�(x1, x2) = φ4(x1)φ4(x2) , (6.14)

which means that the two bosons are both in the fourth eigenstate; another exampleis

�(x1, x2) = 1√2

(φ1(x1)φ2(x2) + φ1(x2)φ2(x1)) , (6.15)

where the factor 1/√2 has been included to maintain the same normalization of the

many-body wavefunction, and in this case the bosons are in the first two availablesingle-particles eigenstates (Fig. 6.1).

Page 125: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

118 6 Many-Body Systems

Fig. 6.1 Ground-state of a system of identical non-interacting bosons (a) and fermions (b) is aharmonic trap

For N fermions the simplest many-body wave function is instead very different,and it is given by

�(x1, x2, ..., xN ) = 1√N !

⎛⎜⎜⎝

φ1(x1) φ1(x2) ... φ1(xN )

φ2(x1) φ2(x2) ... φ2(xN )

... ... ... ...

φN (x1) φN (x2) ... φN (xN )

⎞⎟⎟⎠ (6.16)

that is the so-called Slater determinant of the N × N matrix obtained with the Nlowest-energy single particle wavefunctions ψn(x), with n = 1, 2, ..., N , calculatedin the N possible generalized coordinates xi , with i = 1, 2, ..., N . Note that for 2fermions the previous expression becomes

�(x1, x2) = 1√2

(φ1(x1)φ2(x2) − φ1(x2)φ2(x1)) . (6.17)

We stress that for non-interacting identical particles the Hamiltonian (6.9) is sep-arable and the total energy associated to the bosonic many-body wavefunction (6.12)is simply

E = N ε1 , (6.18)

while for the fermionic many-body wavefunction (6.16) the total energy (in theabsence of degenerate single-particle energy levels) reads

E = ε1 + ε2 + ... + εN , (6.19)

which is surely higher than the bosonic one. The highest occupied single-particleenergy level is called Fermi energy, and it indicated as εF ; in our case it is obviouslyεF = εN .

Page 126: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.2 Non-interacting Identical Particles 119

Notice the above definition of Fermi energy εF can be considered acceptablein many contexts and is a simple and effective definition, especially for the zero-temperature Fermi gas. Unfortunately it is imprecise in some contexts, e.g. finitesystems with a nonzero gap between the last full and first empty state. A betterdefinition, such as the limit for T → 0 of the chemical potential μ (see Sect. 7.5)would lead to some intermediate value between the energy of the highest occupiedstate and that of the lowest empty state.

6.2.1 Uniform Gas of Non-interacting Fermions

A quite important physical system is the uniform gas of non-interacting fermions. Itis indeed a good starting point for the description of all the real systems which havea finite interaction between fermions.

The non-interacting uniform Fermi gas is obtained setting to zero the confiningpotential, i.e.

U (r) = 0 , (6.20)

and imposing periodicity conditions on the single-particle wavefunctions, which areplane waves with a spinor

φ(x) = 1√V

eik·r χσ , (6.21)

where χσ is the spinor for spin-up and spin-down along a chosen z asis:

χ↑ =(10

), χ↓ =

(01

). (6.22)

At the boundaries of a cube having volume V and side L one has

eikx (x+L) = eikx x , eiky(y+L) = eiky y , eikz(z+L) = eikz z . (6.23)

It follows that the linear momentum k can only take on the values

kx = 2π

Lnx , ky = 2π

Lny , kz = 2π

Lnz , (6.24)

where nx , ny , nz are integer quantum numbers. The single-particle energies are givenby

εk = �2k2

2m= �

2

2m

4π2

L2(n2x + n2y + n2z ) . (6.25)

Page 127: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

120 6 Many-Body Systems

In the thermodynamic limit L → ∞, the allowed values are closely spaced and onecan use the continuum approximation

∑nx ,ny ,nz

→∫

dnx dny dnz , (6.26)

which implies∑k

→ L3

(2π)3

∫d3k = V

∫d3k

(2π)3. (6.27)

The total number N of fermionic particles is given by

N =∑

σ

∑k

�(εF − εk) , (6.28)

where theHeaviside step function�(x), such that�(x) = 0 for x < 0 and�(x) = 1for x > 0, takes into account the fact that fermions are occupied only up to theFermi energy εF , which is determined by fixing N . Notice that at finite temperatureT the total number N of ideal fermions is instead obtained from the Fermi-Diracdistribution, namely

N =∑

σ

∑k

1

eβ(εk−μ) + 1, (6.29)

where β = 1/(kBT ), with kB the Boltzmann constant, andμ is the chemical potentialof the system. In the limit β → +∞, i.e. for T → 0, the Fermi-Dirac distributionbecomes the Heaviside step function and μ is identified as the Fermi energy εF .

In the continuum limit and choosing spin 1/2 fermions one finds

N =∑

σ=↑,↓V

∫d3k

(2π)3�

(εF − �

2k2

2m

), (6.30)

from which one gets (the sum of spins gives simply a factor 2) the uniform density

ρ = N

V= 1

3π2

(2mεF

�2

)3/2

. (6.31)

The formula can be inverted giving the Fermi energy εF as a function of the densityρ, namely

εF = �2

2m

(3π2ρ

)2/3. (6.32)

Page 128: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.2 Non-interacting Identical Particles 121

In many applications the Fermi energy εF is written as

εF = �2k2F2m

, (6.33)

where kF is the so-called Fermi wave-number, given by

kF = (3π2ρ

)1/3. (6.34)

The total energy E of the uniform and non-interacting Fermi system is given by

E =∑

σ

∑k

εk �(εF − εk) , (6.35)

and using again the continuum limit with spin 1/2 fermions it becomes

E =∑

σ=↑,↓V

∫d3k

(2π)3

�2k2

2m�

(εF − �

2k2

2m

), (6.36)

from which one gets the energy density

E = E

V= 3

5ρ εF = 3

5

�2

2m

(3π2

)2/3ρ5/3 (6.37)

in terms of the Fermi energy εF and the uniform density ρ.

6.2.2 Atomic Shell Structure and the Periodic Table of theElements

The non-relativistic quantum Hamiltonian of Z identical non-interacting electronsin the neutral atom is given by

H0 =Z∑

i=1

h(ri ) , (6.38)

where h(r) is the single-particle Hamiltonian given by

h(r) = − �2

2m∇2 +U (r) , (6.39)

Page 129: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

122 6 Many-Body Systems

with

U (r) = − Ze2

4πε0 |r| (6.40)

the confining potential due to the attractive Coulomb interaction between the singleelectron and the atomic nucleus of positive charge Ze, with e > 0.

Because the confiningpotentialU (r) is spherically symmetric, i.e.U (r) = U (|r|),the single-particle Hamiltonian h satisfies the eigenvalue equation

h(r) φnlmlms (r) = εn(Z) φnlmlms (r) , (6.41)

where

εn(Z) = −13.6 eVZ2

n2(6.42)

are theBohr single-particle eigenenergies of the hydrogen-like atom, andφnlmlms (r) =Rnl(r)Ylml (θ,φ) are the single-particle eigenfunctions, which depends on the princi-pal quantum numbers n = 1, 2, ..., the angular quantum number l = 0, 1, ..., n − 1,the third-component angular quantum numberml = −l,−l + 1, ..., l − 1, l, and thethird-component spin quantum number ms = − 1

2 ,12 . Notice that here

φnlmlms (r) = φnlml (r,σ) (6.43)

with σ =↑ for ms = 12 and σ =↓ for ms = − 1

2 (Table6.1).

Table 6.1 Lightest atoms and their ground-state energy E on the basis of single-particle energies

Z Atom Symbol E

1 Hydrogen H ε1(1)

2 Helium He 2ε1(2)

3 Lihium Li 2ε1(3) + ε2(3)

4 Berylium Be 2ε1(4) + 2ε2(4)

5 Boron B 2ε1(5) + 3ε2(5)

6 Carbon C 2ε1(6) + 4ε2(6)

7 Nitrogen N 2ε1(7) + 5ε2(7)

8 Oxygen O 2ε1(8) + 6ε2(8)

9 Fluorine F 2ε1(9) + 7ε2(9)

10 Neon Ne 2ε1(10) + 8ε2(10)

11 Sodium Na 2ε1(11) + 8ε2(11) + ε3(11)

12 Magnesium Mg 2ε1(12) + 8ε2(12) + 2ε3(12)

13 Aluminium Al 2ε1(13) + 8ε2(13) + 3ε3(13)

14 Silicon Si 2ε1(14) + 8ε2(14) + 4ε3(14)

Page 130: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.2 Non-interacting Identical Particles 123

Due to the Pauli principle the ground-state energy E of this system of Z electronsstrongly depends of the degeneracy of single-particle energy levels. In the Table wereport the ground-state energy E of the lightest atoms on the basis of their single-particle energy levels εn(Z).

The degeneracy of the single-particle energy level εn(Z) is clearly independenton Z and given by

deg(εn(Z)) =n−1∑l=0

2(2l + 1) = 2n2 , (6.44)

which is the maximum number of electrons with the same principal quantum numbern. The set of stateswith the same principal quantumnumber is called theoretical shell.The number of electrons in each theoretical shell are: 2, 8, 18, 32, 52. One expectsthat the more stable atoms are characterized by fully occupied theoretical shells.Actually, the experimental data, namely the periodic table of elements due to DmitriMendeleev, suggest that the true number of electrons in each experimental shell areinstead: 2, 8, 8, 18, 18, 32, because the noble atoms are characterized the followingatomic numbers:

2, 2 + 8 = 10, 2 + 8 + 8 = 18, 2 + 8 + 8 + 18 = 36,

2 + 8 + 8 + 18 + 18 = 54, 2 + 8 + 8 + 18 + 18 + 32 = 86 ,

corresponding to Helium (Z = 2), Neon (Z = 10), Argon (Z = 18), Krypton (Z =36), Xenon (Z = 54), and Radon (Z = 86). The experimental sequence is clearlysimilar but not equal to the theoretical one, due to repetitions of 8 and 18.

It is important to stress that the theoretical sequence is obtained under the verycrude assumption of non-interacting electrons. To improve the agreement betweentheory and experiment one must include the interaction between the electrons.

6.3 Interacting Identical Particles

The quantum Hamiltonian of N identical interacting particles is given by

H =N∑i=1

h(xi ) + 1

2

N∑i, j=1i = j

V (xi , x j ) = H0 + HI , (6.45)

where h is the single-particle Hamiltonian and V (xi , x j ) is the inter-particle potentialof the mutual interaction. In general, due to the inter-particle potential, the Hamil-tonian (6.45) is not separable and the many-body wavefunctions given by Eqs. (6.12)and (6.16) are not exact eigenfunctions of H .

Page 131: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

124 6 Many-Body Systems

6.3.1 Variational Principle

Many approaches to the determination of the ground-state of an interacting many-body system are based on the so-called variational principle, which is actually atheorem.

Theorem For any normalized many-body state |�〉, i.e. such that 〈�|�〉 = 1, whichbelongs to the Hilbert space on which acts the Hamiltonian H , one finds

〈�|H |�〉 ≥ Egs , (6.46)

where Egs is the ground-state energy of the system and the equality holds only if|�〉 = |�gs〉with |�gs〉 ground-state of the system, i.e. such that H |�gs〉 = Egs |�gs〉.

Proof The many-body Hamiltonian H satisfies the exact eigenvalue problem

H |�α〉 = Eα|�α〉 , (6.47)

where Eα are the ordered eigenvalues, i.e. such that E0 < E1 < E2 < ... withE0 = Egs , and |�α〉 the corresponding orthonormalized eigenstates, i.e. such that〈�α|�β〉 = δα,β with |�0〉 = |�gs〉. The generic many-body state |�〉 can be writtenas

|�〉 =∑

α

cα|�α〉 , (6.48)

where cα are the complex coefficients of the expansion such that

∑α

|cα|2 = 1 . (6.49)

Then one finds

〈�|H |�〉 =∑α,β

c∗α cβ 〈�α|H |�β〉 =

∑α,β

c∗α cβ Eβ 〈�α|�β〉 =

∑α,β

c∗α cβ Eβ δα,β

=∑

α

|cα|2 Eα ≥∑

α

|cα|2 E0 = E0 = Egs . (6.50)

Obviously, the equality holds only if c0 = 1 and, consequently, all the other coeffi-cients are zero.

In 1927 by Douglas Hartree and Vladimir Fock used the variational principle todevelop a powerful method for the study of interacting identical particles. We shallanalyze this variational method in the following subsections.

Page 132: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.3 Interacting Identical Particles 125

6.3.2 Hartree for Bosons

In the case of N identical interacting bosons the Hartree approximation is simplygiven by

�(x1, x2, ..., xN ) = φ(x1) φ(x2) ... φ(xN ) , (6.51)

where the single-particle wavefunction φ(x) is unknown and it must be determinedin a self-consistent way. Notice that, as previously stressed, this factorization impliesthat all particles belong to the same single-particle state, i.e. we are supposing thatthe interacting system is a pure Bose-Einstein condensate. This is a quite strongassumption, that is however reliable in the description of ultracold and dilute gasesmade of bosonic alkali-metal atoms (in 2001EricCornell, CarlWeiman, andWolfangKetterle got the Nobel Prize in Physics for their experiments with these quantumgases), andwhichmust be relaxed in the case of strongly-interacting bosonic systems(like superfluid 4He). In the variational spirit of the Hartree approach the unknownwavefunction φ(x) is determined by minimizing the expectation value of the totalHamiltonian, given by

〈�|H |�〉 =∫

dx1dx2...dxN�∗(x1, x2, ..., xN )H�(x1, x2, ..., xN ) , (6.52)

with respect to φ(x). In fact, by using Eq. (6.45) one finds immediately

〈�|H |�〉 = N∫

dx φ∗(x)h(x)φ(x) + 1

2N (N − 1)

∫dx dx ′ |φ(x)|2V (x, x ′)|φ(x ′)|2 ,

(6.53)which is a nonlinear energy functional of the single-particle wavefunction φ(x). Itis called single-orbital Hartree functional for bosons. In this functional the first termis related to the single-particle hamiltonian h(x) while the second term is related tothe inter-particle interaction potential V (x, x ′). Weminimize this functional with thefollowing constraint due to the normalization

∫dx |φ(x)|2 = 1 . (6.54)

We get immediately the so-called Hartree equation for bosons

[h(x) +Umf (x)

]φ(x) = ε φ(x) , (6.55)

where the mean-field potential Umf (x) reads

Umf (x) = (N − 1)∫

dx ′ V (x, x ′) |φ(x ′)|2 (6.56)

Page 133: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

126 6 Many-Body Systems

and ε is the Lagrange multiplier fixed by the normalization. It is important to observethat the mean-field potential Umf (x) depends on φ(x) and it must be obtained self-consistently. In other words, the Hartree equation of bosons is a integro-differentialnonlinear Schrödinger equation whose nonlinear term gives the mean-field potentialof the system.

In the case of spinless bosons, where |x〉 = |r〉, given the local bosonic density

ρ(r) = N |φ(r)|2 , (6.57)

under the assumption of a large number N of particles the Hartree variational energyreads

〈�|H |�〉 = N∫

d3r φ∗(r)

[− �

2

2m∇2 +U (r)

]φ(r) + 1

2

∫d3r d3r′ ρ(r)V (r − r′)ρ(r′)

(6.58)while the Hartree equation becomes

[− �

2

2m∇2 +U (r) +

∫d3r′ V (r − r′) ρ(r′)

]φ(r) = ε φ(r) . (6.59)

To conclude this subsection, we observe that in the case of a contact inter-particlepotential, i.e.

V (r − r′) = g δ(r − r′) , (6.60)

the previous Hartree variational energy becomes

〈�|H |�〉 = N∫

d3r φ∗(r)[− �

2

2m∇2 +U (r)

]φ(r) + g

2

∫d3r ρ(r)2 (6.61)

and the corresponding Hartee equation reads

[− �

2

2m∇2 +U (r) + g ρ(r)

]φ(r) = ε φ(r) , (6.62)

which is the so-called Gross-Pitaevskii equation, deduced in 1961 by Eugene Grossand Lev Pitaevskii.

6.3.3 Hartree-Fock for Fermions

In the case of N identical interacting fermions, the approximation developed byHartree and Fock is based on the Slater determinant we have seen previously, namely

Page 134: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.3 Interacting Identical Particles 127

�(x1, x2, ..., xN ) = 1√N !

⎛⎜⎜⎝

φ1(x1) φ1(x2) ... φ1(xN )

φ2(x1) φ2(x2) ... φ2(xN )

... ... ... ...

φN (x1) φN (x2) ... φN (xN )

⎞⎟⎟⎠ , (6.63)

where now the single-particle wavefunctions φn(x) are unknown and they are deter-mined with a variational procedure. In fact, in the Hartree-Fock approach theunknown wavefunctions φn(x) are obtained by minimizing the expectation valueof the total Hamiltonian, given by

〈�|H |�〉 =∫

dx1 dx2 ... dxN �∗(x1, x2, ..., xN )H�(x1, x2, ..., xN ) , (6.64)

with respect to the N single-particle wavefunctions φn(x). By using Eq. (6.45) andafter some tedious calculations one finds

〈�|H |�〉 =N∑

i=1

∫dx φ∗

i (x)h(x)φi (x) + 1

2

N∑

i, j=1i = j

[ ∫dx dx ′ |φi (x)|2V (x, x ′)|φ j (x

′)|2

−∫

dx dx ′ φ∗i (x)φ j (x)V (x, x ′)φ∗

j (x′)φi (x ′)

], (6.65)

which is a nonlinear energy functional of the N single-particle wavefunctions φi (x).In this functional the first term is related to the single-particleHamiltonian h(x)whilethe second and the third terms are related to the inter-particle interaction potentialV (x, x ′). The second term is called direct term of interaction and the third term iscalled exchange term of interaction. We minimize this functional with the followingconstraints due to the normalization

∫dx |φi (x)|2 = 1 , i = 1, 2, ..., N , (6.66)

where, in the case of spin 1/2 particles, one has

φi (x) = φi (r,σ) = φi (r,σ) χσ (6.67)

with χσ the two-component spinor, and the integration over x means

∫dx =

∫d3r

∑σ=↑,↓

, (6.68)

such that∫

dx |φi (x)|2 =∫

d3r∑

σ=↑,↓|φi (r,σ)|2 =

∫d3r

∑σ=↑,↓

|φi (r,σ)|2 , (6.69)

because χ∗σχσ = 1 and more generally χ∗

σχσ′ = δσ,σ′ .

Page 135: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

128 6 Many-Body Systems

After minimization of the energy functional we get the so-called Hartree-Fockequations [

h(x) + Um f (x)]φi (x) = εi φi (x) (6.70)

where εi are theLagrangemultipliers fixed by the normalization and Um f is a nonlocalmean-field operator. This nonlocal operator is given by

Um f (x) φi (x) = Ud(x) φi (x) −N∑j=1

U jix (x)φ j (x) , (6.71)

where the direct mean-field potential Ud(x) reads

Ud(x) =N∑j=1j =i

∫dx ′ V (x, x ′) |φ j (x

′)|2 , (6.72)

while the exchange mean-field potential U jix (x) is instead

U jix (x) =

∫dx ′ φ∗

j (x′) V (x, x ′) φi (x

′) . (6.73)

If one neglects the exchange term, as done by Hartree in his original derivation, theso-called Hartree equations

[h(x) +Ud(x)

]φi (x) = εi φi (x) , (6.74)

are immediately derived. It is clearly much simpler to solve the Hartree equationsinstead of the Hartree-Fock ones. For this reason, in many applications the latter areoften used. In the case of spin 1/2 fermions, given the local fermionic density

ρ(r) =∑

σ=↑,↓

N∑i=1

|φi (r,σ)|2 =∑

σ=↑,↓ρ(r,σ) , (6.75)

under the assumption of a large number N of particles the Hartree (direct) variationalenergy reads

ED =N∑i=1

∑σ=↑,↓

∫d3r φ∗

i (r,σ)

[− �

2

2m∇2 +U (r)

]φi (r,σ)

+ 1

2

∑σ,σ′=↑,↓

∫d3r d3r′ ρ(r,σ)V (r − r′)ρ(r′,σ′) (6.76)

Page 136: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.3 Interacting Identical Particles 129

and the corresponding Hartree equation becomes

⎡⎣− �

2

2m∇2 +U (r) +

∑σ′=↑,↓

∫d3r′ V (r − r′) ρ(r′,σ′)

⎤⎦φi (r,σ) = εi φi (r,σ) .

(6.77)The Hartree-Fock variational energy is slightly more complex because it includesalso the exchange energy, given by

EX = −1

2

N∑i, j=1i = j

∑σ=↑,↓

∫d3r d3r′ φ∗

i (r,σ)φi (r′,σ)V (r − r′)φ∗j (r

′,σ)φ j (r,σ) .

(6.78)Notice that in the exchange energy all the termswith opposite spins are zero due to thescalar product of spinors:χ∗

σχσ′ = δσ,σ′ . The existence of this exchange energy EX isa direct consequence of the anti-symmetry of themany-bodywave function, namely aconsequence of the fermionic nature of the particles we are considering. Historically,this term EX was obtained by Vladimir Fock to correct the first derivation of DouglasHartree who used a not anti-symmetrized many-body wavefunction.

To conclude this subsection, we observe that in the case of a contact inter-particlepotential, i.e.

V (r − r′) = g δ(r − r′) , (6.79)

the Hartee-Fock (direct plus exchange) variational energy reads

E =N∑

i=1

σ=↑,↓

∫d3r φ∗

i (r, σ)

[− �

2

2m∇2 +U (r)

]φi (r,σ) (6.80)

+ g

2

N∑

i, j=1i = j

σ,σ′=↑,↓

∫d3r

[|φi (r, σ)|2|φ j (r, σ

′)|2 − |φi (r, σ)|2|φ j (r,σ)|2 δσ,σ′]

.

It follows immediately that identical spin-polarized fermionswith contact interactionare effectively non-interacting because in this case the interaction terms of direct andexchange energy exactly compensate to zero.

6.3.4 Mean-Field Approximation

A common strategy to solve the interacting many-body problem is the so-calledmean-field approximation,which corresponds on finding a suitablemean-field poten-tial which makes the system quasi-separable. The idea is the following. The interac-tion Hamiltonian

Page 137: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

130 6 Many-Body Systems

HI = 1

2

N∑i, j=1i = j

V (xi , x j ) (6.81)

can be formally written as

HI = Hm f +(HI − Hm f

)= Hm f + Hres , (6.82)

where Hm f is an appropriate separable mean-field Hamiltonian, i.e. such that

Hm f =N∑i=1

Umf (xi ) , (6.83)

and Hres is the residual Hamiltonian. In this way the total Hamiltonian reads

H = H0 + Hm f + Hres = H ′ + Hres , (6.84)

where the effective Hamiltonian H ′ is given by

H ′ = H0 + Hm f =N∑i=1

[h(xi ) +Umf (xi )

]=

N∑i=1

[− �

2

2m∇2i +U (xi ) +Umf (xi )

].

(6.85)Neglecting the residual Hamiltonian Hres , the system is described by the effectiveHamiltonian H ′ and the N -body problem is reduced to the solution of the 1-bodyproblem [

h(x) +Umf (x)]φn(x) = εn φn(x) . (6.86)

Obviously, themaindifficulty of this approach is to determine themean-field potentialUmf (x) which minimizes the effect of the residual interaction Hres . The mean-fieldpotentialUmf (x) can be chosen on the basis of the properties of the specific physicalsystem or on the basis of known experimental data.

As shown in the previous subsections, Umf (x) can be also obtained by using theHartree-Fockmethod. In the case of theHartree-Fockmethod for bosons, introducingthe effective Hamiltonian

H ′ =N∑i=1

[h(xi ) +Umf (xi )

](6.87)

with Um f (xi )given byEq. (6.56), it is immediate to verify that the bosonicmany-bodywavefunction �(x1, x2, ..., xN ) of Eq. (6.51), with the single-particle wavefunctionφ(x) satisfying Eq. (6.55), is the lowest eigenfunction of H ′, and such that

Page 138: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.3 Interacting Identical Particles 131

H ′�(x1, x2, ..., xN ) = Nε �(x1, x2, ..., xN ) . (6.88)

On the other hand, �(x1, x2, ..., xN ) is not an eigenfunction of the total HamiltonianH , but it is an approximate variational wavefunction of the ground-state of the fullHamiltonian H .

In a similar way, in the case of the Hartree-Fock method for fermions, introducingagain the effective Hamiltonian

H ′ =N∑i=1

[h(xi ) + Um f (xi )

](6.89)

with Um f (xi ) given by Eq. (6.71), it is not difficult to verify that the fermionic many-body wavefunction �(x1, x2, ..., xN ) of Eq. (6.63), with the single-particle wave-function φ(x) satisfying Eq. (6.70), is the lowest eigenfunction of H ′, and such that

H ′�(x1, x2, ..., xN ) =N∑i=1

εi �(x1, x2, ..., xN ) . (6.90)

On the other hand, �(x1, x2, ..., xN ) is not an eigenfunction of the total HamiltonianH , but it is an approximate variational wavefunction of the ground-state of the fullHamiltonian H .

6.4 Density Functional Theory

In 1927 Llewellyn Thomas and Enrico Fermi independently proposed a statisticalapproach to the electronic structure of atoms with a large number N of electrons,which avoids tackling the solution of the many-body Schödinger equation by focus-ing on the electron local total number density ρ(r). The starting point is the (kinetic)energy density of a uniform ideal gas of electrons at zero temperature in the thermo-dynamic limit, given by

Eideal = 3

5

�2

2m(3π2)2/3ρ5/3 , (6.91)

where ρ is the uniform number density of electrons. Thomas and Fermi consideredN electrons in a atom with nuclear charge Ze, such that

V (r − r′) = e2

4πε0|r − r′| . (6.92)

the electron-electron Coulomb potential, while

Page 139: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

132 6 Many-Body Systems

U (r) = − Ze2

4πε0|r| (6.93)

is the nucleus-electron Coulomb potential with the nucleus centered at the orgin ofthe reference frame. Thomas and Fermi supposed that the ground-state energy of theelectrons can be captured by the functional

ET F [ρ] = TT F [ρ] + ED[ρ] +∫

d3r U (r) ρ(r) , (6.94)

where

TT F [ρ] =∫

d3r3

5

�2

2m(3π2)2/3ρ(r)5/3 (6.95)

is the Thomas-Fermi kinetic energy, i.e. the local approximation to the kinetic energyof the ideal Fermi gas,

ED[ρ] = 1

2

∫d3r d3r′ ρ(r) V (r − r′) ρ(r′) , (6.96)

is the (Hartree-like) direct energy of electron-electron interaction, and the third termin Eq. (6.94) is the energy of the nucleus-electron interaction.

It is important to stress that within this formalism the electron-nucleus potentialU (r) could be more general, taking into account the effect of many nucleons. Thefunctional is minimized under the constraints ρ(r) ≥ 0 and

N =∫

d3r ρ(r) (6.97)

to find the equilibrium condition

�2

2m(3π2)2/3ρ2/3 +U (r) +Umf (r) = μ , (6.98)

where

Umf (r) =∫

d3r′ V (r − r′) ρ(r′) (6.99)

is the Hartree-like mean-field potential acting on the electrons and μ is the Lagrangemultiplier fixed by the normalization. The previous equation can be written as

ρ(r) = (2m)3/2

3π2�3

(μ −U (r) −

∫d3r′ V (r − r′) ρ(r′)

)3/2

(6.100)

Equation (6.100) is an implicit integral equation for the local electronic density ρ(r)which can be solved numerically by using an iterative procedure.

Page 140: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.4 Density Functional Theory 133

Although this was an important first step, the Thomas-Fermi method is limitedbecause the resulting kinetic energy is only approximate, and also because themethoddoes not include the exchange energy of the electrons due to the Pauli principle.A bet-ter theoretical description is obtained with the Thomas-Fermi-Dirac-vonWeizsäckermodel, whose energy functional is given by

ET FDW [ρ] = ET F [ρ] + EX [ρ] + EW [ρ] , (6.101)

where ET F [ρ] is the Thomas-Fermi energy functional of Eq. (6.94),

EX [ρ] = −∫

d3r3

4

e2

4πε0(3

π)1/3ρ(r)4/3 (6.102)

is the functional correction introduced by Paul Dirac in 1930 to take into account theexchange energy which appears in the Hartree-Fock method, and

EW [ρ] =∫

d3r λ�2

2m(∇√

ρ(r))2 (6.103)

is the functional correction introduced by Carl Friedrich von Weizsäcker in 1953 toimprove the accuracy of the kinetic energy with λ an adjustable parameter. Noticethat a simple dimensional analysis shows that the Thomas-Fermi kinetic energydensity must scale as �

2ρ5/3/(2m), while the exchange energy density must scale ase2ρ4/3/(4πε0).

In 1964 the density functional approach was put on a firm theoretical footing byPierre Hohenberg and Walter Kohn. We follow their original reasoning which doesnot take into account explicitly the spin degrees. First of all they observed that, givena many-body wavefunction �(r1, r2, ..., rN ), the associated one-body density ρ(r)reads

ρ(r) = 〈�|N∑i=1

δ(r − ri )|�〉 = N∫

d3r2 ... d3rN |�(r, r2, ..., rN )|2 . (6.104)

Then they formulated two rigorous theorems for a system of identical particles(bosons or fermions) described by the Hamiltonian

H = T + U + V (6.105)

where

T =N∑i=1

− �2

2m∇2i (6.106)

is the many-body kinetic energy operator,

Page 141: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

134 6 Many-Body Systems

U =N∑i=1

U (ri ) (6.107)

is the many-body external potential operator, and

V = 1

2

N∑i, j=1i = j

V (ri , r j ) (6.108)

is the many-body interaction potential operator. We now state the two theorems ofHohenberg and Kohn, which are based on the variational principle discussed in theprevious section, but first of all we observe that for any �(r1, r2, ..., rN ) one has

〈�|H |�〉 = 〈�|T + V |�〉 + 〈�|U |�〉 = 〈�|T + V |�〉 +∫

U (r) ρ(r) d3r .

(6.109)This formula shows explicitly that the external energy 〈�|U |�〉 is a functional ofthe local density ρ(r). The Hohemberg-Kohn theorems ensure that also the internalenergy 〈�|F |�〉 = 〈�|T + V |�〉 is a functional of ρ(r).

Theorem 1 For a system of N identical interacting particles in an external potentialU (r) the density ρgs(r) of the non degenerate ground-state �gs(r1, r2, ..., rN ) isuniquely determined by the external potential. In other words, there is a one-to-onecorrespondence between U (r) and ρgs(r).

Proof The proof of this theorem proceeds by reductio ad absurdum. Let there be twodifferent external potentials, U1(r) and U2(r), that give rise to the same ground-state density ρgs(r). Due to the variational principle, the associated Hamiltoni-ans, H1 = F + U1 and H2 = F + U2 have different ground-state wavefunctions,�gs,1(r1, r2, ..., rN ) and �gs,2(r1, r2, ..., rN ), that each yield ρgs(r). Using the vari-ational principle

Egs,1 < 〈�gs,2|H1|�gs,2〉 = 〈�gs,2|H2|�gs,2〉 + 〈�gs,2|H1 − H2|�gs,2〉= Egs,2 +

∫ρgs(r)[U1(r) −U2(r)] d3r , (6.110)

where Egs,1 and Egs,2 are the ground-state energies of H1 and H2 respectively. Anequivalent expression for Eq. (6.110) holds when the subscripts are interchanged,namely

Egs,2 < 〈�gs,1|H2|�gs,1〉 = 〈�gs,1|H1|�gs,1〉 + 〈�gs,1|H2 − H1|�gs,1〉= Egs,1 +

∫ρgs(r)[U2(r) −U1(r)] d3r . (6.111)

Page 142: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.4 Density Functional Theory 135

Therefore adding the inequality (6.110) to the inequality (6.110) leads to the result:

Egs,1 + Egs,2 < Egs,2 + Egs,1 , (6.112)

which is a contradiction.

From this theorems one deduces immediately two corollaries.

Corollary 1 The many-body ground-state wavefunction �gs(r1, r2, ..., rN ) is abijective function of ρgs(r).

Proof As just shown in Theorem1, ρgs(r) determines U (r), and U (r) determinesH and therefore�gs(r1, r2, ..., rN ). This means that�gs(r1, r2, ..., rN ) is a bijectivefunction of ρgs(r). Symbolically we can write

�gs = G[ρgs] . (6.113)

Corollary 2 The internal energy 〈�gs |F |�gs〉 of the ground state is a universalfunctional F[ρ] of ρgs(r).

Proof Egs is a functional of �gs(r1, r2, ..., rN ). Symbolically we have

Egs = �[�gs] . (6.114)

But we have just seen that �gs(r1, r2, ..., rN ) is a bijective function of ρgs(r), inparticular

�gs = G[ρgs] . (6.115)

Consequently Egs is a functional of ρgs(r), namely

Egs = �[G[ρgs]] = E[ρgs] . (6.116)

Due to the simple structure of the external energy, the functional E[ρ] is clearly givenby

E[ρ] = F[ρ] +∫

d3r U (r) ρ(r) , (6.117)

where F[ρ] is a universal functional which describes the internal (kinetic and inter-action) energy.

Page 143: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

136 6 Many-Body Systems

Theorem 2 For a system of N identical interacting particles in an external potentialU (r) the density functional

E[ρ] = F[ρ] +∫

d3r U (r) ρ(r)

is such that E[ρ] ≥ E[ρgs] = Egs for any trial density ρ(r), and the equality holdsonly for ρ(r) = ρgs(r).

Proof Wehave seen that Egs is a functional of�gs(r1, r2, ..., rN ) but also a functionalof ρgs(r), i.e.

Egs = �[�gs] = �[G[ρgs]] = E[ρgs] . (6.118)

If �(r1, r2, ..., rN ) = �gs(r1, r2, ..., rN ) from the variational principle we have

�[�] = E[ρ] > Egs , (6.119)

where ρ(r) is a local density different from ρgs(r), due to Theorem1. It followsimmediately

E[ρ] > E[ρgs] , (6.120)

if ρ(r) = ρgs(r).

The functional F[ρ] is universal in the sense that it does not depend on the externalpotentialU (r) but only on the inter-particle potential V (r − r′), which is the familiarCoulomb potential in the case of electrons. Indeed, the description of electrons inatoms, molecules and solids is based on the choice of U (r) while F[ρ] is the same.

In the last forty years several approaches have been developed to find approximatebut reliable expressions for F[ρ] in the case of electrons. For instance, within theThomas-Fermi-Dirac-von Weizsäcker model we have seen that

F[ρ] = TT F [ρ] + ED[ρ] + EX [ρ] + EW [ρ] . (6.121)

Instead, for a Bose-Einstein condensate made of dilute and ultracold atoms, thebosonic Hatree-Fock method suggests

F[ρ] = EW [ρ] + g

2

∫d3r ρ(r)2 (6.122)

because V (r − r′) � g δ(r − r′), where g = 4π�2as/m, with m the atomic mass

and as the inter-atomic s-wave scattering length. Note that on the basis of empiricalobservations in the EW [ρ] term one usually takes λ = 1/6 for electrons and λ = 1for bosons.

Page 144: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.4 Density Functional Theory 137

Nowadays the most used density functional for electrons is the one proposedby Walter Kohn and Lu Jeu Sham in 1965. In the Kohn-Sham density functionalapproach the universal functional F[ρ] is given by

F[ρ] = TK S[ρ] + ED[ρ] + EXC [ρ] (6.123)

where

TK S[ρ] =N∑i=1

φ∗i (r)

(− �

2

2m∇2

)φi (r) (6.124)

is theKohn-Shamkinetic energy, where the orbitalsφi (r) determine the local density,namely

ρ(r) =N∑i=1

|φi (r)|2 , (6.125)

the direct (or Hartree-like) energy of interaction ED[ρ] has the familiar form ofEq. (6.96), and EXC [ρ] is the so-called exchange-correlation energy, which simplytakes into account the missing energy with respect to the exact result. The minimiza-tion of the Kohn-Sham density functional gives

[− �

2

2m∇2 +U (r) +

∫d3r′ V (r − r′) ρ(r′) + δEXC [ρ]

δρ(r)

]φi (r) = εi φi (r)

(6.126)which are the local Kohn-Sham equations for the orbitals φi (r), with εi the Lagrangemultipliers fixed by the normalization to one of the orbitals. Notice that the thirdterm on the left side of the previous equation is obtained by using

δEXC [ρ]δφ∗

i (r)= δEXC [ρ]

δρ(r)δρ(r)δφ∗

i (r)= δEXC [ρ]

δρ(r)φi (r) . (6.127)

In many applications the (usually unknown) exchange-correlation energy is writtenas

EXC [ρ] = EX [ρ] + EC [ρ] , (6.128)

where the correlation energy EC [ρ] is fitted fromMonte Carlo calculations. Contraryto the Hartree-Fock theory, in the Kohn-Sham approach the single-particle orbitalsφi (r) are not related to a many-body wavefunction but to the local density ρ(r): thisimplies that the orbitals φi (r) and energies εi of the Kohn-Sham density functionalhave no deep physical meaning. Nevertheless, from the numerical point of view, thesolution of the Kohn-Sham equations is much simpler than that of the Hartree-Fockones, since, contrary to the effective Hartree-Fock mean-field potential, the effectiveKohn-Sham mean-field potential is local.

Page 145: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

138 6 Many-Body Systems

It is important to stress that the density functional theory is an extremely usefulapproach for the description of atoms, molecules, and metals. Nowadays the successof density functional theory not only encompasses standard bulk materials but alsocomplex materials such as proteins, carbon nanotubes, and nuclear matter.

6.5 Molecules and the Born-Oppenheimer Approximation

Up to nowwe have considered only single atoms with many-electrons. In this sectionwe discuss systems composed by many atoms, and in particular molecules.

It is well known that a generic molecule is made of Nn atomic nuclei with electriccharges Zαe and masses Mα (α = 1, 2, ..., Nn) and Ne electrons with charges −eand masses m. Neglecting the finite structure of atomic nuclei, the Hamiltonian ofthe molecule can be written as

H = Hn + He + Vne , (6.129)

where

Hn =Nn∑

α=1

− �2

2Mα∇2

α + 1

2

Nn∑α,β=1α =β

ZαZβe2

4πε0

1

|Rα − Rβ | (6.130)

is the Hamiltonian of the atomic nuclei, with Rα the position of the α-th nucleus,

He =Ne∑i=1

− �2

2m∇2i + 1

2

Ne∑i, j=1i = j

e2

4πε0

1

|ri − r j | (6.131)

is the Hamiltonian of the electrons, with ri the position of the i-th electron, and

Vne = −Nn ,Ne∑α,i=1

Zαe2

4πε0

1

|Rα − ri | (6.132)

is the potential energy of the Coulomb interaction between atomic nuclei and elec-trons.

It is clear that the computation of the ground-state energy and the many-bodywavefunction of an average-size molecule is a formidable task. For instance, thebenzene molecule (C6H6) consists of 12 atomic nuclei and 42 electrons, and thismeans that its many-body wavefunction has (12 + 42) × 3 = 162 variables: the spa-tial coordinates of the electrons and the nuclei. The exact many-body Schrodingerequation for the ground-state is given by

H�(R, r) = E �(R, r) , (6.133)

Page 146: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.5 Molecules and the Born-Oppenheimer Approximation 139

where �(R, r) = �(R1, ...,RNn , r1, ..., rNe ) is the ground-state wavefunction, withR = (R1, ...,RNn ) and r = (r1, ..., rNe ) multi-vectors for nuclear and electroniccoordinates respectively.

In 1927 Max Born and Julius Robert Oppenheimer suggested a reliable approxi-mation to treat this problem. Their approach is based on the separation of the fast elec-tron dynamics from the slowmotion of the nuclei. In the so-calledBorn-Oppenheimerapproximation the many-body wave function of the molecule is factorized as

�(R, r) = �e(r; R) �n(R) , (6.134)

where �n(R) is the nuclear wavefunction and �e(r; R) is the electronic wavefunc-tion, which depends also on nuclear coordinates. Moreover, one assumes that theelectronic wavefunction �e(r; R) is a solution of the following eigenvalue equation

(He + Vne

)�e(r; R) = Ee(R)�e(r; R) . (6.135)

where the electronic eigenenergy Ee(R) depends on the nuclear coordinates becauseof the nucleon-electron potential Vne = Vne(R, r). This equation describes an elec-tronic eigenstate compatible with a given geometrical configuration R of the atomicnuclei. Physically, it corresponds on assuming that nuclear and electronic motionsare somehow decoupled. In this way one gets

H�e(r; R) �n(R) =(Hn + He + Vne

)�e(r; R) �n(R)

= Hn�e(r; R) �n(R) +(He + Vne

)�e(r; R) �n(R)

= Hn�e(r; R) �n(R) + Ee(R)�e(r; R) �n(R)

� �e(r; R)(Hn�n(R) + Ee(R)

)�n(R) (6.136)

where the last approximate equality � is sound because the effect of the kineticterm of the operator Hn on the electronic wavefunction �e(r; R) is small, beingthe nuclear masses Mα much larger than the electronic mass m. Finally, from theprevious equation and Eqs. (6.133) and (6.135) one finds

(Hn + Ee(R)

)�n(R) = E �n(R) , (6.137)

that is the equation of the adiabaticmotion of atomic nuclei in themean-field potentialEe(R) generated by the electrons.

Equations (6.135) and (6.137) are still quite complex, and they are usually solvedwithin some approximate quantummany-bodymethod, e.g. theHartree-Fockmethodor the density functional theory.

Page 147: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

140 6 Many-Body Systems

To conclude, we observe that in this section we have considered molecules, butthe Born-Oppenheimer approximation is crucial in any context where there is morethan a single atom around, which includes atomic gases, clusters, crystals, and manyother physical systems.

6.6 Solved Problems

Problem 6.1By using the Gaussian variational method calculate the approximate energy of theground-state of a one-dimensional quantum particle under the action of the quarticpotential

U (x) = A x4 .

SolutionThe stationary Schrödinger equation of the particle is given by

[− �

2

2m

∂2

∂x2+ A x4

]ψ(x) = ε ψ(x) .

This equation can be seen as the Euler-Lagrange equation obtained by minimizingthe energy functional

E =∫

dx ψ∗(x)

[− �

2

2m

∂2

∂x2+ A x4

]ψ(x) =

∫dx

[�2

2m

∣∣∣∂ψ(x)

∂x

∣∣∣2 + A x4|ψ(x)|2

],

with the normalization condition∫

dx |ψ(x)|2 = 1 .

According to the Gaussian variational method the wavefunction is supposed to begiven by

ψ(x) = e−x2/(2σ2)

π1/4σ1/2,

where σ is the variational parameter. Inserting this variational wavefunction in theenergy functional and integrating over x we get

E = �2

4mσ2+ 3

4A σ4 .

Page 148: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.6 Solved Problems 141

Minimizing this energy with respect to the variational parameter σ, i.e. setting dEdσ

=0, we find

σ =(

�2

6mA

)1/6

.

Substituting this value of σ in the energy E we finally obtain

E = 9

4

(�2

6m

)2/3

A1/3 ,

which is the approximate energy of the ground-state. This energy is surely larger orequal to the energy associated to the exact ground-state of the system.

Problem 6.2Derive the Gross-Pitaevskii equation for a system of N Bose-condensed particleswith contact interaction

V (r − r′) = g δ(r − r′) ,

and prove that the Lagrange multiplier ε of the Gross-Pitaevskii equation is thechemical potential μ of the system.

SolutionThe Hartree equation for N Bose-condensed particles is given by

[− �

2

2m∇2 +U (r) + (N − 1)

∫d3r′ |ψ(r′)|2V (r − r′)

]ψ(r) = ε ψ(r) ,

where ψ(r) is the wavefunction normalized to one. Inserting the contact potentialwe obtain [

− �2

2m∇2 +U (r) + (N − 1)g|ψ(r)|2

]ψ(r) = ε ψ(r) .

Under the reasonable hypotesis that N � 1 we get

[− �

2

2m∇2 +U (r) + Ng|ψ(r)|2

]ψ(r) = ε ψ(r) .

that is the Gross-Pitaevskii equation. This equation describes accurately a diluteBose-Einstein condensate where the true inter-particle potential can be approximatedwith the contact potential. In this case the parameter g can be written as

g = 4π�2as

m,

Page 149: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

142 6 Many-Body Systems

where as is the s-wave scattering length of the true inter-particle potential V (r − r).The energy functional associated to the Gross-Pitaevskii equation is given by

E = N∫

d3r{ψ∗(r)

[− �

2

2m∇2 +U (r)

]ψ(r) + 1

2Ng|ψ(r)|4

}.

The chemical potential is defined as

μ = ∂E

∂N.

On the basis of this definition we obtain

μ =∫

d3r{ψ∗(r)

[− �

2

2m∇2 +U (r)

]ψ(r) + Ng|ψ(r)|4

}.

On the other hand, if we insert ψ∗(r) on the left side of the Gross-Pitaevskii equationand we integrate over r we get

∫d3r

{ψ∗(r)

[− �

2

2m∇2 +U (r)

]ψ(r) + Ng|ψ(r)|4

}= ε ,

because ∫d3r|ψ(r)|2 = 1 .

By comparing the formulas of ε and μ we conclude that ε = μ. This result is alsoknown asKoopmans theorem. Indeed,more rigorously the chemical potential definedas

μ = EN − EN−1 ,

but also in this case one easily finds that ε = μ using the correct expression of theenergy

EN = N∫

d3r{ψ∗(r)

[− �

2

2m∇2 +U (r)

]ψ(r) + 1

2(N − 1)g|ψ(r)|4

},

where it appears (N − 1) instead of N in front of the interaction strength g.

Problem 6.3Write the energy functional associated to theGross-Pitaevskii equation as a functionalof the local density ρ(r) of the Bose-Einstein condensate.

Page 150: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

6.6 Solved Problems 143

SolutionWe have seen that the energy functional associated to the Gross-Pitaevskii equationof the Bose-Einstein condensate is given by

E = N∫

d3r{ψ∗(r)

[− �

2

2m∇2 +U (r)

]ψ(r) + 1

2Ng|ψ(r)|4

},

but it can also be written as

E = N∫

d3r{

�2

2m|∇ψ(r)|2 +U (r)|ψ(r)|2 + 1

2Ng|ψ(r)|4

},

if the wavefunction is zero at the surface of integration volume. We now introducethe local density of the Bose-Einstein condensate as

ρ(r) = Nψ(r)2 ,

supposing that the wavefunction is real. In this way the energy functional can beimmediately written as

E =∫

d3r{

�2

2m

(∇√

ρ(r))2 +U (r)ρ(r) + 1

2gρ(r)2

},

that is a fuctional of the local density ρ(r), as required.

Problem 6.4By using the Gaussian variational method on the Gross-Pitaevskii functional calcu-late the energy per particle of a Bose-Einstein condensate under harmonic confine-ment, given by

U (r) = 1

2mω2(x2 + y2 + z2) .

SolutionWe start from the Gaussian variational wave function

ψ(r) = 1

π3/4a3/2H σ3/2e−(x2+y2+z2)/(2a2Hσ2) ,

where aH = √�/(mω) is the characteristic length of the harmonic confinement and

σ is the dimensionless variational parameter. Inserting this wave function in theGross-Pitaevskii energy functional

E = N∫

d3r{ψ∗(r)

[− �

2

2m∇2 + 1

2mω2(x2 + y2 + z2)

]ψ(r) + 1

2Ng|ψ(r)|4

},

Page 151: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

144 6 Many-Body Systems

after integration we obtain the energy E as a function of σ, namely

E = N�ω

(3

4

1

σ2+ 3

4σ2 + γ

2

1

σ3

),

where γ = √2/πNas/aH is the dimensionless strength of the inter-particle interac-

tion with g = 4π�2as/m and as the s-wave scattering length. The best choice for σ

is obtained by extremizing the energy function:

dE

dσ= 0 ,

from which we obtainσ(σ4 − 1) = γ .

It is clear that σ growswith γ, if γ > 0. Instead, if γ < 0 there are two possible valuesof σ: one value corresponds to a minimum of the energy E (meta-stable solution)and the other corresponds to a maximum of the energy E (unstable solution). It isstraightforward to show that these two solutions exists only if γ > −4/55/4.

Further Reading

For the many-electron atom and the Hartree-Fock method:B.H. Bransden and C.J. Joachain, Physics of Atoms and Molecules, Chap. 8,Sects. 8.1, 8.2 and 8.4 (Prentice Hall, Upper Saddle River, 2003).For the density functional theory:B.H. Bransden and C.J. Joachain, Physics of Atoms and Molecules, Chap. 8,Sects. 8.3 and 8.6 (Prentice Hall, Upper Saddle River, 2003).E. Lipparini, Modern Many-Particle Physics: Atomic Gases, Quantum Dots andQuantumFluids, Chap. 4, Sects. 4.1, 4.2, and 4.3 (World Scientific, Singapore, 2003).

Page 152: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 7Second Quantization of Matter

In this chapter we discuss the second quantization of the non-relativistic matter field,that is the Schrödinger field. We show that the Schrödinger field can be expressedas a infinite sum of harmonic oscillators. These oscillators, which describe the pos-sible eigenfrequencies of the matter field, are quantized by introducing creation andannihilation operators acting on the Fock space of number representation. We showhow, depending on the commutation rule, these ladder operators can model accu-rately bosonic and fermionic particles which interact among themselves or with thequantum electromagnetic field. The second quantization (also called quantum fieldtheory) is the powerful tool to describe phenomena, both at zero and finite temper-ature, where the number of particles is not conserved or it is conserved only on theaverage.

7.1 Schrödinger Field

In the second chapter we have seen that the light field is composed of an infinitenumber of quanta, the photons, as explained in 1927 by Paul Dirac. In the same yearEugene Wigner and Pascual Jordan proposed something similar for the matter. Theysuggested that in non-relativistic quantum mechanics the matter field is nothing butthe single-particle Schrödinger field ψ(r, t) of quantum mechanics, which satisfiesthe time-dependent Schrödinger equation

i�∂

∂tψ(r, t) =

[− �

2

2m∇2 +U (r)

]ψ(r, t) , (7.1)

where U (r) is the external potential acting on the quantum particle. Setting

ψ(r, t) = φα(r) e−iεαt/� (7.2)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_7

145

Page 153: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

146 7 Second Quantization of Matter

one finds [− �

2

2m∇2 +U (r)

]φα(r) = εαφα(r) (7.3)

that is the stationary Schrödinger equation for the eigenfunction φα(r) with eigen-value εα where α is the label which represents a set of quantum numbers. In this casethe Schrödinger field ψ(r, t) is a stationary state of the system, in fact

|ψ(r, t)|2 = |φα(r)|2 (7.4)

does not depend on time t .In general, the Schrödinger field ψ(r, t) is not a stationary state of the system

because the space dependence and time dependence cannot expressed as in Eq. (7.2),but, as suggested by Wigner and Jordan, one can surely expand the field as

ψ(r, t) =∑

α

cα(t) φα(r) (7.5)

where ∫d3r φ∗

α(r)φβ(r) = δαβ . (7.6)

In this way, from the time-dependent Schrödinger equation we get

i� cα(t) = εα cα(t) , (7.7)

whose general solution is obviously

cα(t) = cα(0) e−iεαt/� . (7.8)

In this case the Schrödinger field ψ(r, t) is not a stationary state of the system, infact

|ψ(r, t)|2 =∑α,β

c∗α(0)cβ(0) e−i(εα−εβ)t/� φ∗

α(r)φβ(r) (7.9)

depends on time t , while obviously its integral

∫d3r|ψ(r, t)|2 =

∑α

|cα(0)|2 (7.10)

does not.The constant of motion associated to the Schrödinger field ψ(r, t) is the average

total energy of the system, given by

Page 154: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.1 Schrödinger Field 147

H =∫

d3r ψ∗(r, t)[− �

2

2m∇2 +U (r)

]ψ(r, t) . (7.11)

Inserting the expansion (7.5) into the total energy we find

H =∑

α

εα

2

(c∗αcα + cαc

∗α

). (7.12)

This energy is obviously independent on time: the time dependence of the complexamplitudes c∗

α(t) and cα(t) cancels due to Eq. (7.8).Instead of using the complex amplitudes c∗

α(t) and cα(t) one can introduce thereal variables

qα(t) =√2�

ωα

1

2

(cα(t) + c∗

α(t))

(7.13)

pα(t) = √2�ωα

1

2i

(cα(t) − c∗

α(t))

(7.14)

such that the energy of the matter field reads

H =∑

α

(p2α2

+ 1

2ω2

α q2α

), (7.15)

whereωα = εα

�(7.16)

is the eigenfrequency associated to the eigenenergy εα.This energy resembles that of infinitely many harmonic oscillators with unitary

mass and frequency ωα. It is written in terms of an infinite set of real harmonicoscillators: one oscillator for each mode characterized by quantum numbers α andfrequency ωα. It is clear the analogy between the energy of the Schrödinger field andthe energy of the classical electromagnetic field.

7.2 Second Quantization of the Schrödinger Field

The canonical quantization of the classical Hamiltonian (7.15) is obtained by pro-moting the real coordinates qα and the real momenta pα to operators:

qα → qα , (7.17)

pα → pα , (7.18)

Page 155: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

148 7 Second Quantization of Matter

satisfying the commutation relations

[qα, pβ] = i� δαβ , (7.19)

The quantum Hamiltonian is thus given by

H =∑

α

(p2α2

+ 1

2ω2

α q2α

). (7.20)

The formal difference between Eqs. (7.15) and (7.20) is simply the presence of the“hat symbol” in the canonical variables.

We now introduce annihilation and creation operators

cα =√

ωα

2�

(qα + i

ωαpα

), (7.21)

c+α =

√ωα

2�

(qα − i

ωαpα

), (7.22)

which satisfy the commutation relations

[cα, c+β ] = δα,β , [cα, cβ] = [c+

α , c+β ] = 0 , (7.23)

and the quantum Hamiltonian (7.20) becomes

H =∑

α

εα

(c+α cα + 1

2

). (7.24)

Obviously this quantumHamiltonian can be directly obtained from the classical one,given by Eq. (7.12), by promoting the complex amplitudes cα and c∗

α to operators:

cα → cα , (7.25)

c∗α → c+

α , (7.26)

satisfying the commutation relations (7.23).The operators cα and c+

α act in the Fock space of the “particles” of the Schrödingerfield. A generic state of this Fock space is given by

| . . . nα . . . nβ . . . nγ . . . 〉 , (7.27)

meaning that there are nα particles in the single-particle state |α〉, nβ particles inthe single-particle state |β〉, nγ particles in the single-particle state |γ〉, et cetera.The operators cα and c+

α are called annihilation and creation operators because theyrespectively destroy and create one particle in the single-particle state |α〉, namely

Page 156: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.2 Second Quantization of the Schrödinger Field 149

cα| . . . nα . . . 〉 = √nα | . . . nα − 1 . . . 〉 , (7.28)

c+α | . . . nα . . . 〉 = √

nα + 1 | . . . nα + 1 . . . 〉 . (7.29)

Note that these properties follow directly from the commutation relations (7.23). Thevacuum state, where there are no particles, can be written as

|0〉 = | . . . 0 . . . 0 . . . 0 . . . 〉 , (7.30)

and

cα|0〉 = 0 , (7.31)

c+α |0〉 = |1α〉 = |α〉 , (7.32)

where |α〉 is such that〈r|α〉 = φα(r) . (7.33)

From Eqs. (7.28) and (7.29) it follows immediately that

Nα = c+α cα (7.34)

is the number operator which counts the number of particles in the single-particlestate |α〉, i.e.

Nα| . . . nα . . . 〉 = nα | . . . nα . . . 〉 . (7.35)

Notice that the quantum Hamiltonian of the matter field can be written as

H =∑

α

εα Nα , (7.36)

after removing the puzzling zero-point energy.This is the second-quantizationHamil-tonian of non-interacting matter.

Introducing the time-evolution unitary operator

U (t) = e−i H t/� , (7.37)

one finds that obviously

U+(t) cα(0) U (t) = cα(0) e−iεαt/� . (7.38)

In fact, one can write

Page 157: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

150 7 Second Quantization of Matter

i�d

dt

(U+(t) cα(0) U (t)

)= i�

d

dt

(ei H t/� cα(0) e−i H t/�

)

= i�

(i

�H cα(t) − i

�cα(t) H

)

= cα(t) H − H cα(t)

= [cα(t), H ]= [cα(t),

∑β

εβ c+β (0) cβ(0)]

=∑

β

εβ[cα(t), c+β (t) cβ(t)]

=∑

β

εβ[cα(t), c+β (t)] cβ(t)

=∑

β

εβ δα,β cβ(t)

= εα cα(t) . (7.39)

Thus, one obtains

i�d

dt

(U+(t) cα(0) U (t)

)= εα cα(t) , (7.40)

with solutionU+(t) cα(0) U (t) = cα(0) eiεαt/� . (7.41)

This is clearly the operatorial version of the classical solution (7.8), where

cα(t) = U+(t) cα(0) U (t) (7.42)

is the familiar expression of the time-evolution of the operator.

7.2.1 Bosonic and Fermionic Matter Field

The annihilation and creation operators cα and c+α which satisfy the commutation

rules (7.23) are called bosonic operators and the corresponding quantum field oper-ator

ψ(r, t) =∑

α

cα(t) φα(r) (7.43)

is the bosonic field operator. Indeed the commutation rules (7.23) imply Eqs. (7.28)and (7.29) and, as expected for bosons, there is no restriction on the number ofparticles nα which can occupy in the single-particle state |α〉.

Page 158: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.2 Second Quantization of the Schrödinger Field 151

To obtain fermionic properties it is sufficient to impose anti-commutation rulesfor the operators cα and c+

α , i.e.

{cα, c+β } = δαβ , {cα, cβ} = {c+

α , c+β } = 0 , (7.44)

where { A, B} = A B + B A are the anti-commutation brackets. An important conse-quence of anti-commutation is that

(c+α )2 = 0 . (7.45)

Notice that Eqs. (7.28) and (7.29) formally work also for fermions but with theconstraints that nα = 0 or nα = 1 for any single-particle state |α〉. In fact, forfermions

N 2α = (

c+α cα

)2 = c+α cαc

+α cα (7.46)

= c+α

(1 − c+

α cα

)cα = c+

α cα (7.47)

= Nα , (7.48)

which implies that the eigenvalues of Nα can be only 0 and 1. For a fermionictwo-particle state we have

|1α 1β〉 = c+α c

+β |0〉 = −c+

β c+α |0〉 = −|1β 1α〉 , (7.49)

i.e. the state is anti-symmetric under interchange of particle labels, and moreover

|2α〉 = (c+α )2|0〉 = 0 , (7.50)

expressing the Pauli principle that two particles cannot be created in the same single-particle state. For fermions the field operator (7.43) is called fermionic field operator.

A remarkable property of the field operator ψ+(r, t), which works for bosons andfermions, is the following:

ψ+(r, t)|0〉 = |r, t〉 (7.51)

that is the operator ψ+(r, t) creates a particle in the state |r, t〉 from the vacuum state|0〉. In fact,

ψ+(r, t)|0〉 =∑

α

c+α (t) φ∗

α(r)|0〉 =∑

α

c+α e

iεαt/�〈α|r〉|0〉

=∑

α

eiεαt/�|α〉〈α|r〉 = |r, t〉 , (7.52)

Page 159: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

152 7 Second Quantization of Matter

because

|r, t〉 = ei H t/�|r〉 =∑

α

|α〉〈α|ei H t/�∑

β

|β〉〈β|r〉 =∑

α

eiεαt/�|α〉〈α|r〉 (7.53)

with c+α = c+

α (0), 〈α|β〉 = δαβ , and∑

α |α〉〈α| = 1.It is straightforward to show that the bosonic field operator satisfies the following

equal-time commutation rules

[ψ(r, t), ψ+(r′, t)] = δ(r − r′) , (7.54)

while for the fermionic field operator one gets

{ψ(r, t), ψ+(r′, t)} = δ(r − r′) . (7.55)

Let us prove Eq. (7.54). By using the expansion of the field operators one finds

[ψ(r, t), ψ+(r′, t)] =∑α,β

φα(r)φ∗β(r′) [cα, c+

β ]

=∑α,β

φα(r)φ∗β(r′) δα,β =

∑α

φα(r)φ∗α(r′)

=∑

α

〈r|α〉 〈α|r′〉 = 〈r|∑

α

|α〉 〈α| r′〉

= 〈r|r′〉 = δ(r − r′) . (7.56)

The proof for fermions is practically the same, with anti-commutators instead ofcommutators.

Finally, we observe that the many-body quantum Hamiltonian (7.36) can be writ-ten, both for bosons and fermions, in the elegant form

H =∫

d3r ψ+(r, t)[− �

2

2m∇2 +U (r)

]ψ(r, t) . (7.57)

This quantum Hamiltonian can be directly obtained from the classical one, given byEq. (7.11), by promoting the complex classical field ψ(r, t) and ψ(r, t) to quantumfield operators:

ψ(r, t) → ψ(r, t) , (7.58)

ψ∗(r, t) → ψ+(r, t) , (7.59)

satisfying the commutation relations (7.54) of bosons or the anti-commutation rela-tions (7.55) of fermions.

Page 160: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.3 Connection Between First and Second Quantization 153

7.3 Connection Between First and Second Quantization

In this section we shall analyze relationships between the formalism of second quan-tization, where the number of particles is apriori not fixed, and the formalism of firstquantization, where the number of particles is instead fixed.

First we observe that, within the formalism of second quantization of field oper-ators, the time-independent one-body density operator is defined as

ρ(r) = ψ+(r) ψ(r) , (7.60)

and it is such that

N =∫

d3r ρ(r) (7.61)

is the total number operator. By using the expansion

ψ(r) =∑

α

cα φα(r) (7.62)

one finds immediatelyρ(r) =

∑α,β

c+α cβ φ∗

α(r)φβ(r) (7.63)

and alsoN =

∑α

c+α cα =

∑α

Nα , (7.64)

because we consider an orthonormal basis of single-particle wavefunctions φα(r).A remarkable connection between second quantization and first quantization is

made explicit by the formula

ρ(r) |r1r2 . . . rN 〉 =N∑i=1

δ(r − ri ) |r1r2 . . . rN 〉 , (7.65)

where ρ(r) = ∑Ni=1 δ(r − ri ) is the one-body density function introduced in the

previous chapter. This formula can be proved, for bosons, as follows:

ρ(r) |r1r2 . . . rN 〉 = ψ+(r) ψ(r) |r1r2 . . . rN 〉= ψ+(r) ψ(r) ψ+(r1) ψ+(r2) . . . ψ+(rN ) |0〉= ψ+(r)

(δ(r − r1) + ψ+(r1) ψ(r)

)ψ+(r2) . . . ψ+(rN ) |0〉

= δ(r − r1) |r1r2 . . . rN 〉+ ψ+(r) ψ+(r1) ψ(r) ψ+(r2) . . . ψ+(rN ) |0〉= δ(r − r1) |r1r2 . . . rN 〉

Page 161: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

154 7 Second Quantization of Matter

+ ψ+(r) ψ+(r1)(

δ(r − r2) + ψ+(r2) ψ(r))

ψ+(r3) . . . ψ+(rN ) |0〉= δ(r − r1) |r1r2 . . . rN 〉 + δ(r − r2) |r1r2 . . . rN 〉+ ψ+(r) ψ+(r2) ψ(r) ψ+(r3) . . . ψ+(rN ) |0〉= . . .

=N∑

i=1

δ(r − ri ) |r1r2 . . . rN 〉

+ ψ+(r) ψ+(r2) ψ+(r3) . . . ψ+(rN ) ψ(r) |0〉

=N∑

i=1

δ(r − ri ) |r1r2 . . . rN 〉 , (7.66)

taking into account the commutation relations of the field operators ψ(r) and ψ+(r).For fermions the proof proceeds in a very similar way.

We can further investigate the connection between second quantization and firstquantization by analyzing the Hamiltonian operator. In first quantization, the non-relativistic quantum Hamiltonian of N interacting identical particles in the externalpotential U (r) is given by

H (N ) =N∑i=1

[− �

2

2m∇2i +U (ri )

]+1

2

∑i, j=1i = j

V (ri−r j ) =N∑i=1

hi+1

2

N∑i, j=1i = j

Vi j , (7.67)

where V (r − r′) is the inter-particle potential. In second quantization, the quantumfield operator can be written as

ψ(r) =∑

α

cα φα(r) (7.68)

where the φα(r) = 〈r|α〉 are the eigenfunctions of h such that h|α〉 = εα|α〉, andcα and c+

α are the annihilation and creation operators of the single-particle state |α〉.We now introduce the quantum many-body Hamiltonian

H =∑

α

εα c+α cα +

∑αβγδ

Vαβγδ c+α c

+β cδ cγ , (7.69)

where

Vαβδγ =∫

d3r d3r′ φ∗α(r) φ∗

β(r′) V (r − r′) φδ(r′) φγ(r) . (7.70)

This Hamiltonian can be also written as

Page 162: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.3 Connection Between First and Second Quantization 155

H =∫

d3r ψ+(r)[− �

2

2m∇2 +U (r)

]ψ(r)

+ 1

2

∫d3r d3r′ ψ+(r) ψ+(r′) V (r − r′) ψ(r′) ψ(r) . (7.71)

We are now ready to show the meaningful connection between the second-quantization Hamiltonian H and the first-quantization Hamiltonian H (N ), whichis given by the formula

H |r1r2 . . . rN 〉 = H (N )|r1r2 . . . rN 〉 . (7.72)

This formula can be proved following the same path of Eq. (7.66). In fact, one findsthat

ψ+(r) h(r) ψ(r) |r1r2 . . . rN 〉 =N∑i=1

h(ri )δ(r − ri ) |r1r2 . . . rN 〉 (7.73)

and also

ψ+(r) ψ+(r′) V (r, r′) ψ(r′) ψ(r) |r1r2 . . . rN 〉 =N∑

i, j=1i = j

V (ri , ri )δ(r− ri ) δ(r′ − r j ) |r1r2 . . . rN 〉 .

(7.74)From these two expressions Eq. (7.72) follows immediately, after space integration.Similarly to Eqs. (7.65), (7.72) displays a deep connection between second and firstquantization.

Up to now in this section we have considered time-independent quantum fieldoperators. The time-dependent equation of motion for the field operator ψ(r, t) iseasily obtained from the Hamiltonian (7.71) by using the familiar Heisenberg equa-tion

i�∂

∂tψ = [ψ, H ] , (7.75)

which gives

i�∂

∂tψ =

[− �

2

2m∇2 +U (r)

]ψ(r)+

∫d3r′ ψ+(r′) V (r− r′) ψ(r′) ψ(r) . (7.76)

This exact equation for the field operator ψ(r, t) is second-quantized version of theHartree-like time-dependent nonlinear Schrödinger equation of a complexwavefunc-tion (classical field) ψ(r, t). Note that, in practice, one can formally use the simplerformula

i�∂

∂tψ = δ H

δψ+ (7.77)

Page 163: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

156 7 Second Quantization of Matter

to determine the Heisenberg equation of motion, where

δ H

δψ+ =(

δH

δψ∗

)

ψ=ψ,ψ∗=ψ+. (7.78)

Obviously, the equation of motion for the time-dependent field operator ψ+(r, t) isobtained in the same way.

Similarly, the time-dependent equation of motion of the field operator cα(t) isobtained from the Hamiltonian (7.70) by using the Heisenberg equation

i�d

dtcα = [cα, H ] , (7.79)

which gives

i�d

dtcα = cα +

∑βγδ

Vαβγδ c+β cδ cγ . (7.80)

Again, one can formally can the simpler formula

i�d

dtcα = ∂ H

∂c+α

(7.81)

to determine the Heisenberg equation of motion of cα, where

∂ H

∂a+α

=(

∂H

∂a∗α

)

aα=aα,a∗α=a+

α

. (7.82)

7.4 Coherent States for Bosonic and Fermionic MatterFields

Nowadays it is possible to produce dilute and ultracold bosonic gases made of alkali-metal atoms in a confining magnetic or optical potential U (r). For instance, onemillion of 87Rb atoms at the temperature of 100 nK in the lowest single-particle state.The system can be so dilute that the inter-atomic interaction can be neglected. Underthese conditions one has a pure Bose-Einstein condensate made of non-interactingbosons described by the Hamiltonian

H =∫

d3r ψ+(r)[− �

2

2m∇2 +U (r)

]ψ(r) =

∑α

εα c+α cα . (7.83)

Page 164: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.4 Coherent States for Bosonic and Fermionic Matter Fields 157

If there are exactly N bosons in the lowest single-particle state of energy ε0, theground-state of the system is a Fock state given by

|FN 〉 = 1√N !

(c+0

)N |0〉 = |N , 0, 0, . . . 〉 . (7.84)

It is then straightforward to show that

〈FN |ψ(r)|FN 〉 = 0 , (7.85)

while the expectation value of ψ+(r)ψ(r) is given by

〈FN |ψ+(r)ψ(r)|FN 〉 = N |ψ0(r)|2 . (7.86)

These results are the exact analog of what we have seen for the light field in Chap.2.In the spirit of quantum optics one can suppose that the number of massive bosonsin the matter field is not fixed, in other words that the system is not in a Fock state.We then introduce the coherent state |c0〉, such that

c0|c0〉 = c0|c0〉 , (7.87)

with〈c0|c0〉 = 1 . (7.88)

The coherent state |c0〉 is thus the eigenstate of the annihilation operator c0 withcomplex eigenvalue c0 = |c0|eiθ0 . |c0〉 does not have a fixed number of bosons, i.e.it is not an eigenstate of the number operator N nor of N0, and it is not difficult toshow that |c0〉 can be expanded in terms of number (Fock) states |N , 0, 0, . . . 〉 asfollows

|c0〉 = e−|c0|2/2∞∑N=0

cN0√N ! |N , 0, 0, . . . 〉 . (7.89)

From Eq. (7.87) one immediately finds

N = 〈c0|N |c0〉 = |c0|2 , (7.90)

and it is natural to setc0 =

√N eiθ0 , (7.91)

where N is the average number of bosons in the coherent state, while θ0 is the phaseof the coherent state. The expectation value of the matter field ψ(r) in the coherentstate |c0〉 reads

〈c0|ψ(r)|c0〉 =√N eiθ0 φ0(r) , (7.92)

Page 165: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

158 7 Second Quantization of Matter

with ω0 = ε0/�, while the expectation value of ψ+(r)ψ(r) is given by

〈c0|ψ+(r)ψ(r)|c0〉 = N |φ0(r)|2 . (7.93)

Up to now we have considered coherent states only for bosonic fields: the elec-tromagnetic field and the bosonic matter field. Are there coherent states also for thefermionic field? The answer is indeed positive.

For simplicity, let us consider a single mode of the fermionic field described bythe anti-commuting fermionic operators c and c+, such that

cc+ + c+c = 1 , c2 = (c+)2 = 0 . (7.94)

The fermionic coherent state |c〉 is defined as the eigenstate of the fermionic annihi-lation operator c, namely

c|c〉 = c|c〉 , (7.95)

where c is the corresponding eigenvalue. It is immediate to verify that, for mathe-matical consistency, this eigenvalue c must satisfy the following relationships

cc + cc = 1 , c2 = c2 = 0 , (7.96)

where c is such that〈c|c+ = c〈c| . (7.97)

Obviously c and c are not complex numbers. They are instead Grassmann numbers,namely elements of the Grassmann linear algebra {1, c, c, cc} characterized by theindependent basis elements 1, c, c, with 1 the identity (neutral) element. The mostgeneral function on this Grassmann algebra is given by

f (c, c) = f11 + f12 c + f21 c + f22 cc , (7.98)

where f11, f12, f21, f22 are complex numbers. In fact, the function f (c, c) does nothave higher powers of c, c and cc because they are identically zero. The differentiationwith respect to the Grassmann variable c is defined as

∂cf (c, c) = f12 − f22 c , (7.99)

where the minus sign occurs because one need to permute c and c before differenti-ation. This is called left differentiation. Clearly, one has also

∂cf (c, c) = f21 + f22 c (7.100)

Page 166: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.4 Coherent States for Bosonic and Fermionic Matter Fields 159

and∂2

∂c∂cf (c, c) = − ∂2

∂c∂cf (c, c) = − f22 . (7.101)

The integration over Grassmann numbers is a bit more tricky. It is defined as equiv-alent to differentiation, i.e.

∫dc = ∂

∂c,

∫dc = ∂

∂c,

∫dc dc = ∂2

∂c∂c,

∫dc dc = ∂2

∂c∂c. (7.102)

These integrals are not integrals in the Lebesgue sense. They are called integralssimply because they have some properties of Lebesgue integrals: the linearity andthe fundamental property of ordinary integrals over functions vanishing at infinitythat the integral of an exact differential form is zero. Notice that the mathematicaltheory of this kind of integral (now called Berezin integral) with anticommutingGrassmann variables was invented and developed by Felix Berezin in 1966. TheBerezin integral is mainly used for the path integral formulation of quantum fieldtheory. A discussion of the path integral formulation of quantum field theory, alsocalled functional integration, is outside the scope of this book.

To conclude this section we stress that, in full generality, the classical analog ofthe bosonic quantum field operator

ψ(r) =∑j

φ j (r) c j (7.103)

is the complex classical field

ψ(r) =∑j

φα(r) c j (7.104)

such thatψ(r)|ψ〉 = ψ(r)|ψ〉 , (7.105)

where|ψ〉 =

∏j

|c j 〉 (7.106)

is the bosonic coherent state of the system, |c j 〉 is the coherent state of the bosonicoperator c j , and c j is its complex eigenvalue. Similarly, one can introduce the pseudo-classical Grassmann analog of the fermionic field operator by using fermionic coher-ent states.

Page 167: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

160 7 Second Quantization of Matter

7.5 Quantum Matter Field at Finite Temperature

Let us consider the non-interacting matter field in thermal equilibrium with a bath atthe temperature T . The relevant quantity to calculate all thermodynamical propertiesof the system is the grand-canonical partition function Z , given by

Z = Tr [e−β(H−μN )] (7.107)

where β = 1/(kBT ) with kB the Boltzmann constant,

H =∑

α

εα Nα , (7.108)

is the quantum Hamiltonian,N =

∑α

Nα (7.109)

is total number operator, and μ is the chemical potential, fixed by the conservationof the average particle number. This implies that

Z =∑

{nα}〈 . . . nα . . . |e−β(H−μN )| . . . nα . . . 〉 =

{nα}〈 . . . nα . . . |e−β

∑α(εα−μ)Nα | . . . nα . . . 〉

=∑

{nα}e−β

∑α(εα−μ)nα =

{nα}

∏α

e−β(εα−μ)nα =∏α

∑nα

e−β(εα−μ)nα

=∏α

∞∑

n=0

e−β(εα−μ) n =∏α

1

1 − e−β(εα−μ)for bosons (7.110)

=∏α

1∑

n=0

e−β(εα−μ) n =∏α

(1 + e−β(εα−μ)

)for fermions (7.111)

Quantum statistical mechanics dictates that the thermal average of any operator A isobtained as

〈 A〉T = 1

Z Tr [ A e−β(H−μN )] . (7.112)

Let us suppose that A = H , it is then quite easy to show that

〈H ′〉T = 1

Z Tr [(H−μN ) e−β(H−μN )] = − ∂

∂βln

(Tr [e−β(H−μN ]

)= − ∂

∂βln(Z) .

(7.113)By using Eq. (7.110) or Eq. (7.111) we immediately obtain

ln(Z) = ∓∑

α

ln(1 ∓ e−β(εα−μ)

), (7.114)

Page 168: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.5 Quantum Matter Field at Finite Temperature 161

where − is for bosons and + for fermions, and finally from Eq. (7.113) we get

〈H〉T =∑

α

εα 〈Nα〉T , (7.115)

with

〈N 〉T =∑

α

1

eβ(εα−μ) ∓ 1. (7.116)

Notice that the zero-temperature limit, i.e. β → ∞, for fermions gives

〈H〉0 =∑

α

εα 〈Nα〉0 (7.117)

with〈N 〉0 =

∑α

�(μ − εα) , (7.118)

where the chemical potential μ at zero temperature is nothing but the Fermi energyεF , i.e. εF = μ(T = 0).

7.6 Matter-Radiation Interaction

InChap.3wehave discussed the quantumelectrodynamics by usingfirst quantizationfor thematter and secondquantization of the electromagnetic radiation.Herewewritedown the fully second-quantized Hamiltonian. It is given by

H = Hmatt + Hrad + HI (7.119)

where

Hmatt =∫

d3r ψ+(r, t)[− �

2

2m∇2 +U (r)

]ψ(r, t) , (7.120)

Hrad =∫

d3r(ε0

2

(∂A(r, t)∂t

)2 + 1

2μ0

(∇ ∧ A(r, t))2)

, (7.121)

HI =∫

d3r ψ+(r, t)[− e

mA(r, t) · (−i�∇) + e2

2mA(r, t)2

]ψ(r, t) .(7.122)

with ψ(r, t) the scalar field operator of electrons in the external potentialU (r), whichincludes the instantaneous electrostatic Coulomb potential, and A(r, t) the vectorfield operator of the electromagnetic radiation. Notice that the coupling HamiltonianHI describes the very specific situation of the charged matter (usually the electron

Page 169: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

162 7 Second Quantization of Matter

gas) coupled to radiation,which is an important but by nomeans unique case ofmatterinteracting with electromagnetic radiation using second quantization methods. Oneoften uses similar techniques to describe e.g. neutral atomic gases as we shall showin the last section.

Within the dipole approximation, where one neglects the quadratic term of thevector potential and moreover one assumes that the spatial behavior of the vectorpotential field changes more slowly than the matter field, the interaction Hamiltonianbecomes

HI HD =∫

d3r ψ+(r, t)[− e

mA(0, t) · (−i�∇)

]ψ(r, t) . (7.123)

By performing the following expansion for the matter field

ψ(r, t) =∑

α

cα e−iεαt/� φα(r) , (7.124)

where the φα(r) = 〈r|α〉 are the eigenfunctions of h = − �2

2m∇2 + U (r), and thefollowing expansion for the radiation field

A(r, t) =∑ks

√�

2ε0ωkV

[aks e

i(k·r−ωk t) + a+ks e

−i(k·r−ωk t)]

εks , (7.125)

we finally obtain

Hmatt =∑

α

εα c+α cα , (7.126)

Hrad =∑ks

�ωk a+ks aks , (7.127)

HD = �

∑αβks

gαβks c+β cα

(aks e

−iωk t + a+ks e

iωk t)ei(εβ−εα)t/� (7.128)

where

gαβks = − e

m

√1

2ε0�ωkVεks ·

∫d3r φ∗

β(r)(−i�∇)φα(r) . (7.129)

Notice that∫

d3r φ∗β(r)(−i�∇)φα(r) = 〈β|p|α〉 = i m ωβα〈β|r|α〉 , (7.130)

Page 170: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.6 Matter-Radiation Interaction 163

where ωβα = (εβ − εα)/� and consequently we can write

HD = �

∑αβks

gαβks c+β cα

(aks e

−iωk t + a+ks e

iωk t)eiωβαt (7.131)

with

gαβks = −i

√1

2ε0�ωkVωβα εks · 〈β|er|α〉 , (7.132)

and 〈β|er|α〉 the electric dipole moment of the electromagnetic transition betweenthe electronic states |β〉 and |α〉.

7.6.1 Cavity Quantum Electrodynamics

The development of lasers has produced a new field of research in quantum optics:cavity quantumelectrodynamics (CQED). This is the study of the interaction betweenlaser light confined in a reflective cavity and atoms or other particles which are insidethe cavity.

The simplest system of CQED is a laser light interacting with a two-level atom.This system can be described by the Hamiltonian (7.119) with Eqs. (7.126), (7.127),(7.128) but considering only one mode of the electromagnetic field and only twomodes of the matter field, namely

H = �ω a+a + ε1 c+1 c1 + ε2 c

+2 c2 + �g

[c+2 c1

(a e−i(ω−ω21)t + a+ ei(ω+ω21)t

)

+ c+1 c2

(a e−i(ω+ω21)t + a+ ei(ω−ω21)t

) ], (7.133)

where for simplicity we set ω = ωk , ω21 = (ε2 − ε1)/�, a = aks , a+ = a+ks and

g = g12 = g21. If the applied electromagnetic radiation is near resonance withan atomic transition, i.e. ω ω21 or equivalently (ω − ω21) 0, the previousHamiltonian can be approximated as

H = �ω a+a + ε1 c+1 c1 + ε2 c

+2 c2 + �g

[c+2 c1

(a + a+ ei(ω+ω21)t

)

+ c+1 c2

(a e−i(ω+ω21)t + a+) ]

. (7.134)

Moreover, within the so-called rotating-wave approximation, one can neglect theremaining time-dependent terms which oscillate rapidly, obtaining

H = �ω a+a + ε1 c+1 c1 + ε2 c

+2 c2 + �g

[c+2 c1 a + c+

1 c2 a+]

. (7.135)

Page 171: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

164 7 Second Quantization of Matter

This is the Jaynes-Cummings Hamiltonian, originally proposed in 1963 by EdwinJaynes and Fred Cummings. This model Hamiltonian can be considered thedrosophila of quantum optics: it contains enough physics to describe many phe-nomena in CQED and atom optics.

Let us analyze some properties of the Jaynes-Cummings Hamiltonian. The pho-tonic Fock state |n〉 is eigenstate of the bosonic number operator a+a. The photoncan be in any state |n〉 or or in a superposition of them, namely

|photon〉 =∞∑n=0

αn|n〉 , (7.136)

where αn are complex coefficients such that∑∞

n=0 |αn|2 = 1. In addition, the elec-tronic Fock state |g〉 = |11〉 is eigenstate of the fermionic number operator c+

1 c1,while the electronic Fock state |e〉 = |12〉 is eigenstate of the fermionic numberoperator c+

1 c1. In this two-level atom the electron can be in the state |g〉 or in thestate |e〉, or in a superposition of both, namely

|electron〉 = α|g〉 + β|e〉 , (7.137)

where α and β are two complex coefficients such that |α|2 +|β|2 = 1 and obviously

c+1 c1 + c+

2 c2 = 1 . (7.138)

It is then useful to introduce the following pseudo-spin operators

S+ = c+2 c1 , (7.139)

S− = c+1 c2 , (7.140)

Sz = 1

2

(c+2 c2 − c+

1 c1)

, (7.141)

which have these remarkable properties

S+|g〉 = |e〉 , S+|e〉 = 0 , (7.142)

S−|g〉 = 0 , S−|e〉 = |g〉 , (7.143)

Sz|g〉 = −1

2|g〉 , Sz|e〉 = 1

2|e〉 . (7.144)

By using these pseudo-spin operators the Jaynes-Cummings Hamiltonian ofEq. (7.135) becomes

H = �ω a+a + �ω21 Sz + �g(S+ a + S− a+

)+ ε , (7.145)

Page 172: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.6 Matter-Radiation Interaction 165

where ε = (ε1 + ε2)/2 is a constant energy shift. Due to the deep analogy betweentwo-level electronic system and two-spin states it is quite natural to set

| ↓ 〉 = |g〉 = |11〉 =(01

), (7.146)

| ↑ 〉 = |e〉 = |12〉 =(10

). (7.147)

A generic Fock state of the system is then given by

|n,↓ 〉 = |n〉 ⊗ | ↓ 〉 (7.148)

or|n,↑ 〉 = |n〉 ⊗ | ↑ 〉 . (7.149)

We rewrite theHamiltonian (7.145) neglecting the unrelevant energy shift ε as follows

H = H0 + HI (7.150)

where

H0 = �ω a+a + �ω21 Sz (7.151)

HI = �g(S+ a + S− a+

). (7.152)

Except for the state |0,↓ 〉, any other eigenstate of H0 belongs to one of the infiniteset of two-dimensional subspaces {|n − 1,↑ 〉, |n,↓ 〉}. Thus, for a fixed and finitevalue of n, the infinite matrix representing the Jaynes-Cummings Hamiltonian H onthe basis of eigenstates of H0 splis into an infinite set of 2 × 2 matrices

Hn =(

�ω (n − 1) + �ω21/2 �g√n

�g√n �ω n − �ω21/2

), (7.153)

because

S+a|n − 1,↑ 〉 = √n − 1 S+|n − 1,↑ 〉 = 0 , (7.154)

S+a|n,↓ 〉 = √n S+|n − 1,↓ 〉 = √

n |n − 1,↑ 〉 , (7.155)

S−a+|n − 1,↑ 〉 = √n S−|n,↑ 〉 = √

n |n,↓ 〉 , (7.156)

S−a+|n,↓ 〉 = √n + 1 S−|n + 1,↓ 〉 = 0 . (7.157)

The matrix Hn can be easily diagonalized yeldings the following eigenvalues

E (±)n = (n − 1

2)�ω ± 1

2� (7.158)

Page 173: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

166 7 Second Quantization of Matter

where� =

√δ2 + 4�2g2n , (7.159)

is the energy splitting withδ = �(ω21 − ω) (7.160)

the detuning between the atomic (electronic) transition frequency ω21 and the photonfrequency ω. The corresponding eigenstates are instead given by

|u(+)n 〉 = cos (θ)|n,↓ 〉 − sin (θ)|n − 1,↑ 〉 (7.161)

|u(−)n 〉 = sin (θ)|n,↓ 〉 + cos (θ)|n − 1,↑ 〉 , (7.162)

where

θ = arctan

√1 − δ/�

1 + δ/�. (7.163)

The states |u(±)n 〉, known as “dressed states”, were introduced by Claude Cohen-

Tannoudji and Serge Haroche in 1968–1969. These two scientists got the NobelPrize in Physics (Cohen-Tannoudji in 1997 and Haroche in 2012) for their studieson the manipulation of atoms with lasers.

7.7 Bosons in a Double-Well Potential

In this last section we consider a quite interesting physical problem: interactingbosons in a quasi-1D double-well potential. The starting point is the quantum-field-theory Hamiltonian

H =∫

d3r ψ+(r)[− �

2

2m∇2 +U (r)

]ψ(r)

+ 1

2

∫d3r d3r′ ψ+(r) ψ+(r′) V (r − r′) ψ(r′) ψ(r), (7.164)

where the external trapping potential is given by

U (r) = VDW (x) + 1

2mω2

⊥(y2 + z2) , (7.165)

that is a generic double-well potential VDW (x) in the x axial direction and a har-monic potential in the transverse (y, z) plane. We assume that the system of bosons,described by the field operator ψ(r), is dilute and approximate the inter-particlepotential with a contact Fermi pseudo-potential, namely

Page 174: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.7 Bosons in a Double-Well Potential 167

V (r − r′) = g δ(r − r′) , (7.166)

with g the strength of the interaction. If the frequency ω⊥ is sufficiently large, thesystem is quasi-1D and the bosonic field operator can be written as

ψ(r) = φ(x)e−(y2+z2)/(2l2⊥)

π1/2l⊥. (7.167)

We are thus supposing that in the transverse (y, z) plane the system is Bose-Einsteincondensed into the transverse single-particle ground-state, which is aGaussianwave-function of width

l⊥ =√

mω⊥, (7.168)

that is the characteristic length of the harmonic confinement. Inserting Eq. (7.167)into the Hamiltonian (7.164) and integrating over y and z variabiles we obtain theeffective 1D Hamiltonian

H =∫

dx φ+(x)

[− �

2

2m

d2

dx2+ VDW (x) + �ω⊥

]φ(x)

+ g1D

2

∫dx φ+(x) φ+(x) φ(x) φ(x) , (7.169)

whereg1D = g

2πl2⊥(7.170)

is the effective 1D interaction strength. Notice that in the one-body part of the Hamil-tonian it appears the single-particle ground-state energy �ω⊥ of the harmonic con-finement. We suppose that the barrier of the double-well potential VDW (x), with itsmaximum located at x = 0, is quite high such that there several doublets of quasi-degenerate single-particle energy levels. Moreover, we suppose that only the lowestdoublet (i.e. the single-particle ground-state and the single-particle first excited state)is occupied by bosons. Under these assumptions we can write the bosonic field oper-ator as

φ(x) = aL φL(x) + aR φR(x) (7.171)

that is the so-called two-mode approximation, where φL(x) and φR(x) are single-particle wavefunctions localized respectively on the left well and on the right well ofthe double-well potential. These wavefunctions (which can be taken real) are linearcombinations of the even (and postitive) wavefunction φ0(x) of the ground state andthe odd (and positive for x < 0) wavefunction φ1(x) of the first excited state, namely

Page 175: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

168 7 Second Quantization of Matter

φL(x) = 1√2

(φ0(x) + φ1(x)) , (7.172)

φR(x) = 1√2

(φ0(x) − φ1(x)) . (7.173)

Clearly the operator a j annihilates a boson in the j th site (well) while the operatora+j creates a boson in the j-th site ( j = L , R). Inserting the two-mode approximation

of the bosonic field operator in the effective 1D Hamiltonian, we get the followingtwo-site Hamiltonian

H = εL NL + εR NR − JLRa+L aR − JRL a

+R aL + UL

2NL(NL − 1)+ UR

2NR(NR − 1) ,

(7.174)where N j = a+

j a j is the number operator of the j-th site,

ε j =∫

dx φ j (x)

[− �

2

2m

d2

dx2+ VDW (x) + �ω⊥

]φ j (x) (7.175)

is the kinetic plus potential energy on the site j ,

Ji j =∫

dx φi (x)

[− �

2

2m

d2

dx2+ VDW (x) + �ω⊥

]φ j (x) , (7.176)

is the hopping energy (tunneling energy) between the site i and the site j , and

Uj = g1D

∫dx φ j (x)

4 (7.177)

is the interaction energy on the site j . The Hamiltonian (7.174) is the two-site Bose-HubbardHamiltonian, named after JohnHubbard introduced a similarmodel in 1963to describe fermions (electrons) on a periodic lattice. If the double-well potentialVDW (x) is fully symmetric then εL = εR = ε, JLR = JRL = J , UL = UR = U , andthe Bose-Hubbard Hamiltonian becomes

H = ε(NL + NR

)− J

(a+L aR + a+

R aL) + U

2

[NL(NL − 1) + NR(NR − 1)

].

(7.178)The Heisenberg equation of motion of the operator a j is given by

i�d

dta j = [a j , H ] = ∂ H

∂a+j

, (7.179)

from which we obtain

i�d

dta j = ε a j − J ai +U (N j − 1

2)a j , (7.180)

Page 176: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.7 Bosons in a Double-Well Potential 169

where j = L , R and i = R, L . By averaging this equation with the coherent state|αLαR〉 = |αL〉 ⊗ |αR〉, such that a j (t)|α j 〉 = α j (t)|α j 〉, we find

i�d

dtα j = ε α j − Jαi +U (|α j |2 − 1

2)α j , (7.181)

whereα j (t) = N j (t) e

iθ j (t) , (7.182)

with N j (t) the average number of bosons in the site j at time t and θ j (t) the corre-sponding phase angle at the same time t . Working with a fixed number of bosons,i.e.

N = NL(t) + NR(t) , (7.183)

and introducing population imbalance

z(t) = NL(t) − NR(t)

N(7.184)

and phase differenceθ(t) = θR(t) − θL(t) , (7.185)

the time-dependent equations for αL(t) and αR(t) can be re-written as follows

dz

dt= −2J

√1 − z2 sin (θ) , (7.186)

dt= 2J

z√1 − z2

cos (θ) + U

�z . (7.187)

These are the so-called Josephson equations of theBose-Einstein condensate. Indeed,under the condition of small population imbalance (|z| � 1), small on-site interactionenergy (|U |z/� � 1) and small phase difference (θ � 1) one finds

dz

dt= −2J

�sin (θ) , (7.188)

dt= 2J

�z , (7.189)

which are the equations introduced in 1962 by Brian Josephson to describe thesuperconducting electric current (made of quasi-bosonic Cooper pairs of electrons)between two superconductors separated by a thin insulating barrier (Nobel Prize in1973). Remarkably, as shown in 2007 at Heidelberg by the experimental group ofMarkus Oberthaler, Eqs. (7.186) and (7.187) describe accurately also the dynamicsof a Bose-Einstein condensate made of alkali-metal atoms confined in the quasi-1Ddouble-well potential by counter-propagating laser beams.

Page 177: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

170 7 Second Quantization of Matter

7.7.1 Analytical Results with N = 1 and N = 2

In this subsection we discuss the two-site Bose-Hubbard Hamiltonian

H = −J(a+L aR + a+

R aL) + U

2

[NL(NL − 1) + UR

2NR(NR − 1)

], (7.190)

in the case of N = 1 and N = 2 bosons. Note that, without loss of generality, wehave removed the constant energy εN .

The ground-state |GS〉 of the system with N bosons can be written as

|GS〉 =N∑i=0

ci |i, N − i〉 , (7.191)

where |i, N − i〉 = |i〉 ⊗ |N − i〉 is the state with i bosons in the left well and N − ibosons in the right well and ci is its probability amplitude such that

N∑i=0

|ci |2 = 1 . (7.192)

In the case of N = 1 bosons the ground-state is simply

|GS〉 = c0|0, 1〉 + c1|1, 0〉 , (7.193)

where|c0|2 + |c1|2 = 1 . (7.194)

Moreover, the on-site energy is identically zero and consequently it is immediate tofind

c0 = c1 = 1√2

(7.195)

for any value of J and U .In the case of N = 2 bosons the ground-state is instead given by

|GS〉 = c0|0, 2〉 + c1|1, 1〉 + c2|2, 0〉 , (7.196)

where|c0|2 + |c1|2 + |c2|2 = 1 . (7.197)

By direct diagonalization of the corresponding Hamiltonian matrix, one finds

Page 178: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.7 Bosons in a Double-Well Potential 171

c0 = c2 = 2√16 + ζ2 + ζ

√ζ2 + 16

(7.198)

and

c1 = c0ζ + √

ζ2 + 16

2√2

, (7.199)

where

ζ = U

J(7.200)

is the effective dimensionless interaction strength which controls the nature of theground state |GS〉. Thus, the ground-state is strongly dependent on the ratio betweenthe on-site energyU and the tunneling (hopping) energy J . In particular, one deducesthat

|GS〉 =

⎧⎪⎨⎪⎩

|1, 1〉 for ζ → +∞12

(|0, 2〉 + √

2|1, 1〉 + |2, 0〉)

for ζ → 01√2(|0, 2〉 + |2, 0〉) for ζ → −∞

. (7.201)

7.8 Solved Problems

Problem 7.1Consider the operator N = f + f , where f and f + satisfy the anti-commutation rulef f + + f + f = 1. Show that if |n〉 is an eigenstate of N with eigenvalue n thenf +|n〉 is eigenstates of N with eigenvalue 1 − n.

SolutionWe have

N f |n〉 = ( f + f ) f |n〉 .

The anti-commutation relation between f and f + can be written as

f f + = 1 − f + f .

This implies that

N f +|n〉 = ( f + f ) f +|n〉 = f +( f f +)|n〉 = f +(1− f + f )|n〉 = f +(1− N )|n〉 = (1−n) f +|n〉 .

Notice that in Sect. 7.2 we have shown that the operator N of fermions has onlyeigenvalues 0 and 1. This means that f +|0〉 = |1〉 and f +|1〉 = 0.

Page 179: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

172 7 Second Quantization of Matter

Problem 7.2Show that the following equation of motion

i�∂

∂tψ(r, t) =

[− �

2

2m∇2 +U (r)

]ψ(r, t) + g |ψ(r, t)|2ψ(r, t)

of the classical Schrödinger field ψ(r, t), that is the time-dependent Gross-Pitaevskiiequation, can be deduced by extremizing the action

S =∫

dtd3r L ,

where

L = ψ∗(r, t)[i�

∂t+ �

2

2m∇2 −U (r)

]ψ(r, t) − 1

2g |ψ(r, t)|4

is the Lagrangian density of the system.

SolutionFirst of all, we observe that

∇ · (ψ∗∇ψ

) = ψ∗∇2ψ + ∇ψ∗ · ∇ψ .

We can then write∫

Vd3r ψ∗∇2ψ =

Vd3r ∇ · (

ψ∗∇ψ) −

Vd3r ∇ψ∗ · ∇ψ

=∫

Sd2r

(ψ∗∇ψ

) · n −∫

Vd3r ∇ψ∗ · ∇ψ ,

where S is the surface of the domain of volume V . In the limit of a very large V wecan suppose that the Schrödinger field and its derivatives are zero on the surface S,and consequently ∫

Vd3r ψ∗∇2ψ = −

Vd3r ∇ψ∗ · ∇ψ .

This means that the Lagrangian density can be rewritten as

L = i�ψ∗ ∂ψ

∂t− �

2

2m|∇ψ|2 −U (r)|ψ|2 − 1

2g|ψ|4 ,

where |∇ψ|2 = ∇ψ∗ · ∇ψ. In this way the Lagrangian density is a function of theSchrödinger field and only its first derivatives, namely

L = L(ψ,ψ∗,∂ψ

∂t,∇ψ,∇ψ) .

Page 180: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.8 Solved Problems 173

The Schrödinger field ψ(r, t) = ψR(r, t) + iψI (r, t) is complex, depending on thetwo real fields ψR(r, t) and ψI (r, t) which can be varied independently, but in thisway the complex conjugate field ψ∗(r, t) = ψR(r, t)− iψI (r, t) is fully determined.Alternatively, one can consider ψ(r, t) and ψ∗(r, t) as independent fields: this is ourchoice and the action S is then a functional of ψ(r, t) and ψ(r, t). The first variationof the action S gives

δS =∫

dtd3r(

δS

δψδψ + δS

δψδψ∗

),

where

δS

δψ= ∂L

∂ψ− ∂

∂t

∂L∂ ∂ψ

∂t

− ∇ · ∂L∂∇ψ

,

δS

δψ∗ = ∂L∂ψ∗ − ∂

∂t

∂L∂ ∂ψ∗

∂t

− ∇ · ∂L∂∇ψ∗ .

Imposing thatδS = 0

we obtain two Euler-Lagrange equations

∂L∂ψ

− ∂

∂t

∂L∂ ∂ψ

∂t

− ∇ · ∂L∂∇ψ

= 0 ,

∂L∂ψ∗ − ∂

∂t

∂L∂ ∂ψ∗

∂t

− ∇ · ∂L∂∇ψ∗ = 0 .

By using our Lagrangian density the first Euler-Lagrange equation gives

i�∂

∂tψ =

[− �

2

2m∇2 +U (r)

]ψ + g |ψ|2ψ ,

while the second Euler-Lagrange equation gives

−i�∂

∂tψ∗ =

[− �

2

2m∇2 +U (r)

]ψ∗ + g |ψ|2ψ∗ ,

that is the complex conjugate of the previous one.To conclude,we observe that if the Lagrangian density depends also on higher spa-

tial derivatives the Euler-Lagrange equations are modified accordingly. For instance,in the presence of terms like ∇2ψ or ∇2ψ∗ one gets

Page 181: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

174 7 Second Quantization of Matter

∂L∂ψ

− ∂

∂t

∂L∂ ∂ψ

∂t

− ∇ · ∂L∂∇ψ

+ ∇2 ∂L∂∇2ψ

= 0 ,

∂L∂ψ∗ − ∂

∂t

∂L∂ ∂ψ∗

∂t

− ∇ · ∂L∂∇ψ∗ + ∇2 ∂L

∂∇2ψ∗ = 0 .

Problem 7.3Derive the Gross-Pitaevskii equation

i�∂

∂tψ0(r, t) =

[− �

2

2m∇2 +U (r)

]ψ0(r, t) + g N |ψ0(r, t)|2ψ0(r, t)

of the classical field ψ0(r, t) from the Heisenberg equation of motion

i�∂

∂tψ(r, t) =

[− �

2

2m∇2 +U (r)

]ψ(r, t) + g ψ+(r, t)ψ(r, t)ψ(r, t)

of the bosonic quantum field operator ψ(r, t).Suggestion: expand the quantum field operator as

ψ(r, t) =∑

α

cα φα(r, t)

and use the coherent state |c0〉 of c0.SolutionFirst of all we notice that

〈c0|i� ∂

∂tψ(r, t)|c0〉 = i�

∂t〈c0|ψ(r, t)|c0〉

= i�∂

∂t〈c0|

∑α

cα φα(r, t)|c0〉

= i�∂

∂t

∑α

φα(r, t)〈c0|cα|c0〉

= i�∂

∂t

∑α

φα(r, t)c0 δα,c0

= c0 i�∂

∂tφ0(r, t) .

Similarly

Page 182: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

7.8 Solved Problems 175

〈c0|[− �

2

2m∇2 +U (r)

]ψ(r, t)|c0〉 =

[− �

2

2m∇2 +U (r)

]〈c0|ψ(r, t)|c0〉

=[− �

2

2m∇2 +U (r)

]〈c0|

∑α

cα φα(r, t)|c0〉

=[− �

2

2m∇2 +U (r)

] ∑α

φα(r, t)〈c0|cα|c0〉

=[− �

2

2m∇2 +U (r)

] ∑α

φα(r, t)c0 δα,0

= c0

[− �

2

2m∇2 +U (r)

]φ0(r, t) .

Moreover

〈c0|ψ+(r, t)ψ(r, t)ψ(r, t)|c0〉 = 〈c0|∑α

c+α φ∗

α(r, t)∑

β

cβ φβ(r, t)∑γ

cγ φγ(r, t)|c0〉

=∑

αβγ

φ∗α(r, t)φβ(r, t)φγ(r, t)〈c0|c+

α cβ cγ |c0〉

=∑

αβγ

φ∗α(r, t)φβ(r, t)φγ(r, t)|c0|2c0δα,0δβ,0δγ,0

= |c0|2c0|φ0(r, t)|2φ0(r, t) .

Finally, after setting N = |c0|2, we conclude that

〈c0|i� ∂

∂tψ(r, t)|c0〉 = 〈c0|

[− �

2

2m∇2 +U (r)

]ψ(r, t) + g ψ+(r, t)ψ(r, t)ψ(r, t)|c0〉

is exactly

i�∂

∂tψ0(r, t) =

[− �

2

2m∇2 +U (r)

]ψ0(r, t) + g N |ψ0(r, t)|2ψ0(r, t) .

Problem 7.4Calculate the Brezin integral ∫

dc dc e−cAc ,

where c and c are Grassmann variables, and A is a complex number.

SolutionFirst of all we notice that, due to the fact thatwe areworkingwithGrasmann numbers,one finds

Page 183: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

176 7 Second Quantization of Matter

e−cAc = 1 − cAc ,

because all the other terms of the Taylor-MacLaurin expansion are identically zero.Consequently ∫

dc dc e−cAc =∫

dc dc (1 − cAc) .

Finally, on the basis of the definition of the Brezin integral where the integration isdefined as equivalent to differentiation, we get

∫dc dc (1 − cAc) = ∂2

∂c∂c(1 − cAc) = A .

Further Reading

For the second quantization of particles and non-relativistic quantum field theory:N. Nagaosa, Quantum Field Theory in CondensedMatter Physics, Chap. 1, Sects. 1.1and 1.2 (Springer, Berlin, 1999).A.L. Fetter and J.D. Walecka, Quantum Theory of Many-Particle Systems, Chap.1,Sects. 1.1 and 1.2 (Dover Publications, New York, 2003).H.T.C. Stoof, K.B. Gubbels, and D.B.M. Dickerscheid, Ultracold Quantum Fields,Chap. 6, Sects. 6.1, 6.2, 6.3, and 6.4 (Springer, Berlin, 2009).A. Atland and B. Simons, Condensed Matter Field Theory, Chap. 2, Sect. 2.1 (Cam-bridge Univ. Press, Cambridge, 2006).For the quantum properties of bosons in a double-well potential:G. Mazzarella, L. Salasnich, A. Parola, and F. Toigo, Coherence and entanglementin the ground-state of a bosonic Josephson junction:from macroscopic Schrdingercats to separable Fock states, Phys. Rev. A 83, 053607 (2011).

Page 184: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 8Functional Integration for the Bosonic Field

In this chapter we consider the quantum statistical mechanics of identical bosons byintroducing the formalism of functional integration for the bosonic field. We thenderive the equation of state of non-interacting and weakly-interacting bosons takinginto account beyond-mean-field regularized Gaussian fluctuations. We also analyzevarious properties of the mean-field time-dependent Gross-Pitaevskii equation. Inparticular we show that this equation is equivalent to the zero-temperature equa-tions of superfluid hydrodynamics proposed in 1949 by Landau. Finally, we discusstopological and solitonic solutions of the Gross-Pitaevskii equation.

8.1 Partition Function of Interacting Bosons

For a system of non-relativistic interacting bosons the partition function is given by

Z = Tr [e−β(H−μN )] , (8.1)

where the second-quantized Hamiltonian is

H =∫

d3r ψ+(r)(

− �2

2m∇2 +U (r)

)ψ(r)

+ 1

2

∫d3r d3r′ ψ+(r) ψ+(r′) V (r, r′) ψ(r′) ψ(r) , (8.2)

with U (r) the external potential acting on spinless bosons described by the fieldoperator ψ(r), V (r, r′) the inter-particle interaction, and the number operator reads

N =∫

d3r ψ+(r)ψ(r) . (8.3)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_8

177

Page 185: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

178 8 Functional Integration for the Bosonic Field

In Eq. (8.1), as usual, β = 1/(kBT ) with kB the Boltzmann constant and T thetemperature, while μ is the chemical potential.

In the previous chapter we have seen that, by expanding the field operator ψ(r)as

ψ(r) =∑

α

cα φα(r) , (8.4)

where φα(r) are the orthonormal eigenfunctions of the non-interacting problem withsingle-particle energies εα, one finds

H =∑

α

εα c+α cα +

∑αβγδ

Vαβγδ c+α c

+β cδ cγ , (8.5)

where

Vαβδγ =∫

d3r d3r′ φ∗α(r) φ∗

β(r′) V (r, r′) φδ(r′) φγ(r) , (8.6)

and alsoN =

∑α

c+α cα . (8.7)

Moreover, we have seen that the generalized coherent state

|ψ〉 =∏α

|cα〉 = e−∑α |cα|2/2 e∑

α cα c+α |0〉 , (8.8)

is such thatψ(r) |ψ〉 = ψ(r) |ψ〉 , (8.9)

whereψ(r) =

∑α

cα φα(r) (8.10)

is the “classical” Schrödinger field with cα the complex eigenvalues of cα. Clearly,it follows that

cα|ψ〉 = cα|ψ〉 for any α . (8.11)

Note that |ψ〉 is the tensor product of all single-mode coherent states |cα〉. Due totheir normalization we get

〈ψ|ψ〉 = 1 . (8.12)

The overlap between two generalized coherent states |ψ〉 and |ψ〉 is instead given by

〈ψ|ψ〉 = e− 12

∑α[c∗

α(cα−cα)−(c∗α−c∗

α)cα] . (8.13)

The single-mode coherent state |cα〉 satisfies the remarkable completeness relation

Page 186: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.1 Partition Function of Interacting Bosons 179

∫dc∗

αdcα

2πi|cα〉〈cα| = 1 . (8.14)

It is possible to prove that the set of single-mode coherent states {|cα〉} obtainedby varying α is over-complete and the generalized coherent state |ψ〉 satisfies thecompleteness relation ∫

d[ψ∗,ψ] |ψ〉〈ψ| = 1 , (8.15)

where the integration measure is defined by

∫d[ψ∗,ψ] =

∫ ∏α

dc∗αdcα

2πi(8.16)

with ψ∗(r) the complex conjugate of ψ(r) such that

〈ψ| ψ+(r) = ψ∗(r) 〈ψ| . (8.17)

Notice that cα and ψ(r) represent the same complex field in reciprocal spaces, andthe integration measure can also be written as

∫d[ψ∗,ψ] =

∫ ∏r

dψ∗(r)dψ(r)2πi

. (8.18)

Moreover, Eq. (8.13) can also be written as

〈ψ|ψ〉 = e− 12

∫d3r[ψ∗(r)(ψ(r)−ψ(r))−(ψ∗(r)−ψ∗(r))ψ(r)] . (8.19)

The key point is that the partition function Z can be expressed as a functionalintegral taking into account that the trace of an operator can be performed by usingany set of basis states, and in particular the basis of generalized coherent states. Thus,we write

Z = Tr [e−β(H−μN )] =∫

d[ψ∗,ψ] 〈ψ|e−β(H−μN )|ψ〉 , (8.20)

where |ψ〉 is the generalized coherent state.

8.1.1 Semiclassical Approximation and Imaginary Time

The main problem in the functional representation of the partition function Z is thecalculation of

〈ψ|e−β(H−μN )|ψ〉 . (8.21)

Page 187: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

180 8 Functional Integration for the Bosonic Field

The simple semiclassical approximation

〈ψ|e−β(H−μN )|ψ〉 � e−β〈ψ|(H−μN )|ψ〉 (8.22)

gives

Z =∫

d[ψ∗,ψ] e−β(E[ψ∗,ψ]−μN [ψ∗,ψ]) , (8.23)

where

E[ψ∗,ψ] =∫

d3r ψ∗(r)(

− �2

2m∇2 +U (r)

)ψ(r)

+ 1

2

∫d3r d3r′ |ψ(r)|2 V (r, r′) |ψ(r′)|2 (8.24)

and

N [ψ∗,ψ] =∫

d3rψ∗(r)ψ(r) . (8.25)

The partition function given by Eq. (8.23) is called semiclassical because it is usuallyreliable only at high temperatures, where quantum statistics plays a marginal role.

To better treat the expectation value of Eq. (8.21) we first write |ψ0〉 instead of|ψ〉 and also (with M → ∞)

�τ = �β

M. (8.26)

As we shall see immediately, this number �τ plays the role of the step interval ofan imaginary time τ . In fact, we obtain

〈ψ0|e−β(H−μN )|ψ0〉 = 〈ψ0|e−M�τ (H−μN )/�|ψ0〉= 〈ψ0|

(e−�τ (H−μN )/�

)M |ψ0〉 (8.27)

= 〈ψ0|e−�τ (H−μN )/� . . . e−�τ (H−μN )/�|ψ0〉 .

Then we insert M − 1 completeness relations for the general coherent state |ψ j 〉( j = 1, ..., M − 1) ∫

d[ψ∗j ,ψ j ] |ψ j 〉〈ψ j | = 1 (8.28)

to get

〈ψ0|e−β(H−μN )|ψ0〉 =∫ ⎛⎝

M−1∏j=1

d[ψ∗j ,ψ j ]

⎞⎠

M∏j=1

〈ψ j |e−�τ (H−μN )/�|ψ j−1〉 (8.29)

Page 188: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.1 Partition Function of Interacting Bosons 181

imposing that |ψ0〉 = |ψM〉. In the limit M → ∞ one has �τ → 0 and also

〈ψ j |e−�τ (H−μN )/�|ψ j−1〉 = 〈ψ j |ψ j−1〉 e−�τ (E[ψ∗j ,ψ j ]−μN [ψ∗

j ,ψ j ])/� . (8.30)

Moreover, from Eq. (8.19) the overlap between two general coherent states reads

〈ψ j |ψ j−1〉 = e− �τ2

∫d3r(ψ∗

j (r)[ψ j (r)−ψ j−1(r)]

�τ −ψ j (r)[ψ∗

j (r)−ψ∗j−1(r)]

�τ

). (8.31)

Finally, setting τ = j�τ , ψ(r, τ ) = ψ j (r) and

∂τψ(r, τ ) = [ψ j (r) − ψ j−1(r)]

�τ, (8.32)

we obtain the partition function

Z =∫

D[ψ,ψ∗] exp

{− S[ψ,ψ∗]

}, (8.33)

where

S[ψ,ψ∗] =∫

�β

0dτ

∫d3rL(ψ∗,ψ) (8.34)

is the Euclidean (i.e. with imaginary time) action of the Euclidean Lagrangian density

L(ψ∗,ψ) = �

2

(ψ∗(r, τ )

∂τψ(r, τ ) − ψ(r, τ )

∂τψ∗(r, τ )

)

+ ψ∗(r, τ )(

− �2

2m∇2 +U (r) − μ

)ψ(r, τ ) (8.35)

+ 1

2

∫d3r′ |ψ(r′, τ )|2 V (r, r′) |ψ(r, τ )|2

with time-periodic boundary conditions ψ(r, 0) = ψ(r, �β) and

∫D[ψ,ψ∗] =

∫ ∏(r,τ )

dψ∗(r, τ ) dψ(r, τ )

2πi. (8.36)

We stress again that a semiclassical treatment of the expectation value (8.21)gives a semiclassical partition function (8.23), which is written as a functional inte-gration of the time-independent energy functional. Instead, the exact treatment of Eq.(8.21) produces the exact partition function (8.33), which is expressed as a functionalintegration of the imaginary-time action functional.

Page 189: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

182 8 Functional Integration for the Bosonic Field

8.2 Broken Symmetry and Order Parameter

We consider a D-dimensional (D = 1, 2, 3) Bose gas either noninteracting or witha contact interaction, namely

V (r − r′) = g δ(r − r′) , (8.37)

where δ(x) is the Dirac delta function and g the strength of the interaction. We adoptthe path integral formalism, where the atomic bosons are described by the complexfield ψ(r, τ ).

The Euclidean Lagrangian density of the free system in a D-dimensional box ofvolume LD and with chemical potential μ is then given by

L = ψ∗[�

∂τ− �

2

2m∇2 − μ

]ψ + 1

2g |ψ|4 . (8.38)

Notice that this Lagrangian density is equivalent to the one of Eq. (8.35) becausethey give the same Euler-Lagrange equations.

We have seen that the partition function Z of the system at temperature T canthen be written as

Z =∫

D[ψ,ψ∗] exp

{− S[ψ,ψ∗]

}, (8.39)

where

S[ψ,ψ∗] =∫

�β

0dτ

LD

dDr L(ψ,ψ∗) (8.40)

is the Euclidean action and β ≡ 1/(kBT ) with kB being Boltzmann’s constant.The grand potential � of the system, which is a function of μ, T and g, is then

obtained as

� = − 1

βlnZ . (8.41)

We work in the superfluid phase where the global U(1) gauge symmetry of thesystem is spontaneously broken. For this reason we set

ψ(r, τ ) = ψ0 + η(r, τ ) , (8.42)

where η(r, τ ) is the complex field of bosonic fluctuations around the order parameterψ0 (condensate in 3D or quasi-condensate in 1D and 2D) of the system. We supposethat ψ0 is constant in time, uniform in space and real.

Page 190: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.2 Broken Symmetry and Order Parameter 183

8.2.1 Ideal Bose Gas: Critical Temperatureand Condensate Fraction

First we analyze the case with g = 0, where exact analytical results can be obtainedin any spatial dimension D. In this case the Lagrangian density reads

L = −μψ20 + η∗(r, τ )

[�

∂τ− �

2

2m∇2 − μ

]η(r, τ ) . (8.43)

Introducing the Fourier representation of the fluctuation field

η(r, τ ) = 1√�βLD

∑Q

ηQ ei(q·r−ωτ ) , (8.44)

where Q = (q, iωn) is the D+1 vector denoting the momentaq and frequencies ωn =2πn/(β�) (which are usually called bosonic Matsubara frequencies), the Euclideanaction of the ideal Bose gas can be written in a diagonal form as

S[ψ,ψ∗] = −μψ20 �β LD +

∑Q

�λQ η∗QηQ (8.45)

with

λQ = β (−i�ωn + �2q2

2m− μ) . (8.46)

Here ηQ = ηq,iωn = ηq,n is the Fourier transform of η(r, τ ) and the quantization offrequencies ωn is a consequence of the periodicity η(r, τ + �β) = η(r, τ ).

Taking into account that

Z =∫

D[η, η∗] exp

⎧⎨⎩−

∑Q

λQη∗QηQ

⎫⎬⎭ = 1∏

Q λQ(8.47)

and +∞∑n=−∞

ln (−i n + a) = 1

2

+∞∑n=−∞

ln (n2 + a2) , (8.48)

one finds the grand potential

� = −μψ20L

D + 1

β

∑Q

ln (λQ)

= −μψ20L

D + 1

∑q

+∞∑n=−∞

ln [β2(�2ω2n + ξ2

q)] , (8.49)

Page 191: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

184 8 Functional Integration for the Bosonic Field

where ξq is the shifted free-particle spectrum, i.e.

ξq = �2q2

2m− μ . (8.50)

The sum over bosonic Matsubara frequencies gives

1

+∞∑n=−∞

ln [β2(�2ω2n + ξ2

q)] = ξq

2+ 1

βln (1 − e−β ξq ) . (8.51)

In fact, the derivative with respect to ξq of Eq. (8.51) reads

1

β

+∞∑n=−∞

ξq

ω2n + ξ2

q

= 1

2+ 1

eβξq − 1= 1

2coth (

βξq

2) , (8.52)

and this formula is a straightforward consequence of the exact result

+∞∑n=−∞

1

n2 + a2= π

acoth (π a) (8.53)

remembering that the bosonic Matsubara frequencies are given by ωn = 2πn/β.The grand potential finally reads

� = �0 + �(0) + �(T ) , (8.54)

where�0 = −μψ2

0LD (8.55)

is the grand potential of the order parameter,

�(0) = 1

2

∑q

ξq (8.56)

is the zero-point energy of bosonic single-particle excitations, i.e. the zero-temperature contribution of quantum fluctuations, and

�(T ) = 1

β

∑q

ln(1 − e−βξq

)(8.57)

takes into account thermal fluctuations.

Page 192: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.2 Broken Symmetry and Order Parameter 185

In the continuum limit, where∑

q → LD∫dDq/(2π)D , the zero-point energy

�(0)

LD= 1

2

SD(2π)D

∫ +∞

0dq qD−1

(�

2q2

2m− μ

)(8.58)

of the ideal Bose gas is clearly ultraviolet divergent at any integer dimension D,i.e. for D = 1, 2, 3. Here SD = 2πD/2/�(D/2) is the solid angle in D dimensionswith �(x) the Euler gamma function. We shall show that this divergent zero-pointenergy of the ideal Bose gas is completely eliminated by dimensional regularization.Consequently, the exact grand potential of the ideal Bose gas is given by

LD= −μ ψ2

0 + 1

βLD

∑q

ln(1 − e−βξq

). (8.59)

We notice that ψ0 is not a free parameter but must be determined by minimizing�0, namely (

∂�0

∂ψ0

)

μ,T,LD

= 0 , (8.60)

from which one finds that

ψ0 ={

0 if μ �= 0any value if μ = 0

(8.61)

The number density n = N/LD is obtained from the thermodynamic relation

n = − 1

LD

(∂�

∂μ

)

T,LD ,ψ0

, (8.62)

which gives:

n = ψ20 + 1

LD

∑q

1

eβξq − 1. (8.63)

In the continuum limit∑

q → LD∫dDq/(2π)D , by setting μ = 0 (condensed

phase) from Eq. (8.63) one gets

n = n0 +∫

dDq(2π)D

1

e�2q2

2mkB T − 1, (8.64)

which gives the Bose-Einstein condensate density n0 = ψ20 as a function of the

temperature T .The critical temperature Tc of Bose-Einstein condensation is obtained setting

ψ0 = 0 in the previous equation. In this way one finds

Page 193: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

186 8 Functional Integration for the Bosonic Field

kB Tc =⎧⎨⎩

12πζ(3/2)2/3

�2

m n2/3 for D = 30 for D = 2

no solution for D = 1(8.65)

where ζ(x) is the Riemann zeta function. This result was extended to interactingsystems by David Mermin and Herbert Wagner in 1966. The so-called Mermin-Wagner theorem states that there is no Bose-Einstein condensation at finite tempera-ture in homogeneous systems with sufficiently short-range interactions in dimensionsD ≤ 2.

In the three-dimensional case (D = 3) from the equation

n = n0 +∫

d3q(2π)3

1

e�2q2

2mkB T − 1, (8.66)

we find

n = n0 + ζ(3/2)

(mkBT

2π�2

)3/2

. (8.67)

It follows that

n0

n= 1 − ζ(3/2)

( mkB2π�2

)3/2T 3/2

n= 1 − ζ(3/2)

( mkB2π�2

)3/2T 3/2

ζ(3/2)( mkB

2π�2

)3/2T 3/2c

. (8.68)

Thus, the condensate fraction reads

n0

n= 1 −

(T

Tc

)3/2

. (8.69)

The critical temperature of Eq. (8.65) and this formula for the condensate fraction ofnon-interacting massive bosons were obtained in 1925 by Albert Einstein extendingprevious results derived by Satyendra Nath Bose for a gas a photons (massless bosons)(Fig. 8.1).

8.2.2 Interacting Bose Gas: Bogoliubov Spectrum

Let us now consider a system of bosons with a repulsive contact interaction, i.e. withg > 0 in

L = ψ∗[�

∂τ− �

2

2m∇2 − μ

]ψ + 1

2g |ψ|4 . (8.70)

Page 194: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.2 Broken Symmetry and Order Parameter 187

-2 -1.5 -1 -0.5 0 0.5 1 1.5 2ψ

0

-1

0

1

2

3

Ω0 /

LD

μ < 0

μ > 0

Fig. 8.1 Mean-field grand potential �0 as a function of the real order parameter ψ0 for an interactingBose gas, see Eq. (8.72)

In this case one finds immediately the partition function of the uniform and constantorder parameter ψ0 as

Z0 = exp

{− S0

}= exp {−β �0} , (8.71)

where S0 = S[ψ0] and the grand potential �0 reads

�0

LD= −μψ2

0 + 1

2g ψ4

0 . (8.72)

Again, the constant, uniform and real order parameter ψ0 is obtained by minimiz-ing �0 as (

∂�0

∂ψ0

)

μ,T,LD

= 0 , (8.73)

from which one finds the relation between order parameter and chemical potential

μ = g ψ20 . (8.74)

showing that in the superfluid broken phase the chemical potential is positive and

ψ0 =√

μ

g, (8.75)

Page 195: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

188 8 Functional Integration for the Bosonic Field

Inserting this relation into Eq. (8.72) we find

�0

LD= − μ2

2 g. (8.76)

Clearly, this equation of state is lacking important informations encoded in quantumand thermal fluctuations.

Thus we consider again

ψ(r, τ ) = ψ0 + η(r, τ ) (8.77)

and expand the action S[ψ,ψ∗] around ψ0 up to quadratic (Gaussian) order in η(r, τ )

and η(r, τ ). We find

Z = Z0

∫D[η, η∗] exp

{− Sg[η, η∗]

}, (8.78)

where

Sg[η, η∗] = 1

2

∑Q

(η∗Q, η−Q) MQ

(ηQ

η∗−Q

)(8.79)

is the Gaussian action of fluctuations in reciprocal space with Q = (q, iωn) the D+1vector denoting the momenta q and bosonic Matsubara frequencies ωn = 2πn/(β�),and

MQ = β

(−i�ωn + �

2q2

2m − μ + 2gψ20 −gψ2

0

−gψ20 i�ωn + �

2q2

2m − μ + 2gψ20

)(8.80)

is the inverse fluctuation propagator.Integrating over the bosonic fields η(Q) and η(Q) in Eq. (8.78) one finds the

Gaussian grand potential

�g = 1

∑Q

ln Det(MQ)

= 1

∑q

+∞∑n=−∞

ln [β2(�2ω2n + E2

q)] , (8.81)

where Eq is given by

Eq =√(

�2q2

2m− μ + 2gψ2

0

)2

− g2ψ40 . (8.82)

Page 196: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.2 Broken Symmetry and Order Parameter 189

0 0.5 1 1.5 2 2.5q

0

1

2

3

4

Eq

Bogoliubov spectrum

Phonon spectrum

Fig. 8.2 Bogoliubov spectrum, given by Eq. (8.83), and its low-momentum phonon spectrumEq = cB �q, where cB = √

μ/m is the sound velocity. Energy Eq in units of μ and momentum q

in units of√mμ/�2

By using ψ0 = √μ/g the spectrum becomes

Eq =√

�2q2

2m

(�2q2

2m+ 2μ

)� cB �q for small q , (8.83)

which is the familiar Bogoliubov spectrum, with cB = √μ/m (Fig. 8.2).

Again, the sum over bosonic Matsubara frequencies gives

1

+∞∑n=−∞

ln [β2(�2ω2n + E2

q)] = Eq

2+ 1

βln (1 − e−βEq ) . (8.84)

The total grand potential may then be written as

� = �0 + �(0)g + �(T )

g , (8.85)

where �0 is given by Eq. (8.76).

�(0)g = 1

2

∑q

Eq (8.86)

Page 197: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

190 8 Functional Integration for the Bosonic Field

is the zero-point energy of bosonic collective excitations, i.e. the zero-temperaturecontribution of quantum Gaussian fluctuations, while

�(T )g = 1

β

∑q

ln(1 − e−βEq

)(8.87)

takes into account thermal Gaussian fluctuations.

8.2.3 Dimensional Regularization of Gaussian Fluctuations

We notice that the continuum limit of the zero-point energy for the Gaussian fluctu-ation of the interacting Bose gas

�(0)g

LD= 1

2

SD(2π)D

∫ +∞

0dq qD−1

√�2q2

2m

(�2q2

2m+ 2μ

)(8.88)

is ultraviolet divergent at any integer dimension D.We may rewrite Eq. (8.88) for the zero point energy of a repulsive Bose gas in

dimension D as:

�(0)g

LD= SD(2μ)

D2 +1

4(2π)D

(2m

�2

) D2

B(D + 1

2,−D + 2

2

), (8.89)

where B(x, y) is the Euler Beta function

B(x, y) =∫ +∞

0dt

t x−1

(1 + t)x+y, Re(x), Re(y) > 0 (8.90)

which may be continued to complex values of x and y as

B(x, y) = �(x) �(y)

�(x + y). (8.91)

We rewrite Eq. (8.88) using Eq. (8.91) to get

�(0)g

LD= SD(2μ)

D2 +1

4(2π)D

(2m

�2

) D2 �( D+1

2 ) �(− D+22 )

�(− 12 )

. (8.92)

This expression is now finite when D = 1 or D = 3, while it is still divergent ifD = 2 since �(p) diverges for integers p ≤ 0. In particular, setting D = 3 in Eq.(8.92) we get

Page 198: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.2 Broken Symmetry and Order Parameter 191

�(0)g

L3= 8

15π2(m

�2)3/2μ5/2 . (8.93)

In conclusion, the total grand potential of the three-dimensional Bose gas is thengiven by

L3= − μ2

2 g+ 8

15π2(m

�2)3/2μ5/2 + 1

βL3

∑q

ln(1 − e−βEq

). (8.94)

This result was obtained for the first time in 1957 by Tsung-Dao Lee, Kerson Huang,and Chen Ning Yang without invoking dimensional regularization.

8.3 Gross-Pitaevskii Equation

Let us consider again a system of bosons with a repulsive contact interaction, i.e. letus set g > 0 in

L = ψ∗[�

∂τ− �

2

2m∇2 − μ +U (r)

]ψ + 1

2g |ψ|4 , (8.95)

where U (r) is the external trapping potential.In this inhomogeneous case the partition function of the mean-field (saddle-point)

order parameter ψ0(r, τ ) reads

Z0 = exp

{− S0

}= exp {−β �0} , (8.96)

where

S0 = S[ψ0(r, τ ),ψ∗0(r, τ )] =

∫dt∫

d3r L(ψ0,ψ∗0) (8.97)

and ψ0(r, τ ) is the field ψ(r, τ ) which minimizes S[ψ(r, τ )], namely such that

δS[ψ(r, τ ),ψ∗(r, τ )] = 0 . (8.98)

The minimization of the action S gives the Euler-Lagrange equation

∂L∂ψ∗ − ∂

∂τ

∂L∂( ∂

∂τψ∗)

− ∇ ∂L∂(∇ψ∗)

+ ∇2 ∂L∂(∇2ψ∗)

= 0 (8.99)

from which we get

Page 199: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

192 8 Functional Integration for the Bosonic Field

�∂

∂τψ0(r, τ ) =

[− �

2

2m∇2 − μ +U (r) + g |ψ0(r, τ )|2

]ψ0(r, τ ) . (8.100)

This is the Gross-Pitaevskii equation with imaginary time τ . Moreover, at the mean-field level, the number of bosons reads

N = −∂�0

∂μ=∫

d3r |ψ0(r, τ )|2 . (8.101)

Going back to real time t = iτ we obtain

i�∂

∂tψ0(r, t) =

[− �

2

2m∇2 − μ +U (r) + g |ψ0(r, t)|2

]ψ0(r, t) , (8.102)

that is the real-time (time-dependent) Gross-Pitaevskii equation, with the normaliza-tion condition

N =∫

d3r |ψ0(r, t)|2 . (8.103)

Immediately one deduces the corresponding stationary Gross-Pitaevskii equation

[− �

2

2m∇2 +U (r) + g |ψ0(r)|2

]ψ0(r) = μψ0(r) . (8.104)

8.3.1 Superfluid Hydrodynamics

Quite remarkably the zero-temperature time-dependent Gross-Pitaevskii equation(8.102) can be rewritten as the equations of superfluid hydrodynamics. In fact, setting

ψ0(r, t) = n(r, t)1/2 eiθ(r,t) , (8.105)

and inserting this formula into Eq. (8.102) one finds

∂tn + ∇ · (n v) = 0 , (8.106)

m∂

∂tv + ∇

[1

2mv2 +U (r) + gn − �

2

2m

∇2√n√n

]= 0 , (8.107)

where

v(r, t) = �

m∇θ(r, t) (8.108)

Page 200: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.3 Gross-Pitaevskii Equation 193

is the local velocity field, that is (by definition) irrotational, i.e. such that

∇ ∧ v = 0 . (8.109)

Equations (8.106) and (8.107) are, respectively, the equation of continuity and theequation of conservation of linear momentum for a irrotational and inviscid fluid. Thezero-temperature equation of state of this superfluid, i.e. the local chemical potentialas a function of the local density and its derivatives, can be written as

μ(n,∇2n) = gn − �2

2m

∇2√n√n

, (8.110)

where the second term, which is usually called quantum pressure, becomes negligiblein the high-density regime. In 1941 Lev Landau, on the basis of previous importanttheoretical hints of Fritz London and Laszlo Tisza, developed the two-fluid theoryof superfluid 4He. At zero temperature this two-fluid theory gives Eqs. (8.106) and(8.107).

We observe that quantum effects are encoded not only in this equation of statebut also into the properties of the local field v(r, t): it is proportional to the gradientof a scalar field, θ(r, t), that is the angle of the phase of the single-valued complexwavefunction ψ0(r, t). Consequently, v(r, t) satisfies the equation

Cv · dr = 2π

mk (8.111)

for any closed contour C, with k an integer number. In other words, the circulationis quantized in units of �/m, and this property is strictly related to the existence ofquantized vortices.

For the sake of completeness we observe that, in general, for a rotational andviscous fluid with quite generic zero-temperature equation of state μ(n,∇n,∇2n, ...)the equations of hydrodynamics are given by

∂tn + ∇ · (n v) = 0 (8.112)

m∂

∂tv − η∇2v + ∇

[1

2mv2 +U (r) + μ(n, ∇n, ∇2n, ...)

]= mv ∧ (∇ ∧ v) (8.113)

where η is the viscosity and a rotational term appears on the right side of Eq. (8.113).These equations are called zero-temperature Navier-Stokes equations. Any genericfluid (in the collisional regime) satisfies these equations. Also a superfluid is describedby Eqs. (8.112) and (8.113), but only if η = 0 and under the additional constraintsgiven by Eqs. (8.109) and (8.111).

Page 201: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

194 8 Functional Integration for the Bosonic Field

Let us consider Eqs. (8.106) and (8.107) assuming that U (r) = 0. We set

n(r, t) = neq + δn(r, t) , (8.114)

v(r, t) = 0 + δv(r, t) , (8.115)

where δn(r, t) and δv(r, t) represent small variations with respect to the uniform andconstant stationary configuration neq . In this way, neglecting quadratic terms in thevariations (linearization) from Eqs. (8.106) and (8.107) we get the linear equationsof motion

∂tδn + neq∇ · δv = 0 , (8.116)

∂tδv + c2

s

neq∇δn − �

2

4m2neq∇(∇2δn) = 0 , (8.117)

where cs is the sound velocity of the bosonic superfluid, given by

mc2s = g neq . (8.118)

The linear equations of motion can be arranged in the form of the following waveequation [ ∂2

∂t2− c2

s∇2 + �2

4m2∇4]δn(r, t) = 0 . (8.119)

This wave equation admits monochromatic plane-wave solutions, where the fre-quency ω and the wave vector q are related by the dispersion formula ω = ω(q)

given by

� ω(q) =√

�2q2

2m

(�2q2

2m+ 2mc2

s

). (8.120)

We have thus obtained, in a different way, the Bogoliubov spectrum of elementaryexcitations.

8.3.2 Bright and Dark Solitons

The real-time Gross-Pitaevskii equation (8.102) can be seen as the Euler-Lagrangeequation associated to the minimization of the following (real-time) action functional

S =∫

dt d3r{ψ∗

0(r, t)[i�

∂t+ �

2

2m∇2 −U (r)

]ψ0(r, t) − 1

2g |ψ0(r, t)|4

}.

(8.121)Let us consider a very strong harmonic confinement of frequency ω⊥ along x and ybut no confinement along z, namely

Page 202: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.3 Gross-Pitaevskii Equation 195

U (r) = 1

2mω2

⊥(x2 + y2) . (8.122)

We adopt the variational ansatz

ψ0(r, t) = f (z, t)1

π1/2a2⊥

exp

(x2 + y2

2a2⊥

), (8.123)

where f (z, t) is the axial wave function and a⊥ = √�/(mω⊥) is the character-

istic length of the transverse harmonic confinement. By inserting Eq. (8.123) intoEq. (8.121) and integrating along x and y, the resulting effective action functionaldepends on the field f (z, t). One easily finds that the Euler-Lagrange equation ofthe axial wavefunction f (z, t) reads

i�∂

∂tf (z, t) =

[− �

2

2m

∂2

∂z2+ g1D| f (z, t)|2

]f (z, t) , (8.124)

whereg1D = g

2πa2⊥

(8.125)

is the effective one-dimensional interaction strength and the additive constant �ω⊥has been omitted because it does not affect the dynamics.

Equation (8.124) is a one-dimensional cubic nonlinear Schrödinger equation. In1972 Vladimir Zakharov and Aleksei Shabat found that this equation admits solitonic(i.e. shape invariant) analytical solutions. In particular, for g1D < 0 (self-focusingnonlinearity) one finds the localized bright soliton solution

f (z, t) =√m|g1D|

8�2Sech

[m|g1D|

4�2(z − vt)

]eiv(z−vt)ei(mv2/2−μ)t/� , (8.126)

where Sech[x] is the hyperbolic secant, v is the arbitrary velocity of propagationof the shape-invariant bright soliton and μ = −mg2

1D/(16�2). Instead, for g1D > 0

(self-defocusing nonlinearity) one finds the localized but not normalized dark solitonsolution

f (z, t) =√

g1Dφ2∞ − mv2

g1D

(Tanh

[(z − vt)

√g1Dφ2∞ − v2

]+ i

√mv2

g1Dφ2∞ − mv2

)e−iμt/� ,

(8.127)where Tanh[x] is the hyperbolic tangent, φ∞ > 0 is the constant value of the fieldamplitude | f (z, t)| at infinity (z → ±∞), and v is the velocity of propagation of thedark soliton. The parameter μ = g1Dφ2∞ is obtained by the asymptotic behavior ofthe field f (z, t). For 0 < |v| < cs , where cs = √gφ2∞ is the axial sound velocity, theminimum value of the field f (z, t) is greater than zero and the wave is called graysoliton. If v = 0 (stationary dark soliton) the minimum of the field f (z, t) is zero

Page 203: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

196 8 Functional Integration for the Bosonic Field

and the wave is called black soliton. The velocity cs is the maximal velocity of thedark soliton, namely the velocity of a wave of infinitesimal amplitude, i.e. a soundwave, propagating in a medium of density ρ∞ = φ2∞.

8.4 Solved Problems

Problem 8.1Consider a Bose-Einstein condensate made of N ultracold atoms with mass m andscattering lengthas , under external harmonic confinement of frequency ω. Neglectingthe kinetic term in the Gross-Pitaevskii equation derive the density profile n(r) andthe chemical potential μ of the system.

SolutionThe stationary Gross-Pitaevskii equation of the Bose-Einstein condensate is givenby [

− �2

2m∇2 +U (r) + g |ψ0(r)|2

]ψ0(r) = μψ0(r) ,

where

U (r) = 1

2mω2(x2 + y2 + z2)

and

g = 4π�2

mas .

Neglecting the kinetic term (second derivatives) in the Gross-Pitaevskii equation oneimmediately finds the density profile

n(r) = |ψ0(r)|2 = 1

g(μ −U (r)) � (μ −U (r))

with �(x) the Heaviside step function. Imposing the condition

N =∫

d3r n(r)

which fixes the chemical potential μ, after integration one gets

μ = 1

2�ω

(15NasaH

)2/5

with aH = √�/(mω) the characteristic length of the isotropic harmonic potential.

Page 204: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.4 Solved Problems 197

Problem 8.2Consider a bosonic gas of N non-interacting spinless particles under harmonic con-finement given by the external trapping potential

U (r) = 1

2mω2(x2 + y2 + z2) .

Calculate the critical temperature Tc of Bose-Einstein condensation as a functionof the number N of bosons and the condensate fraction N0/N as a function of thetemperature T .

SolutionOne must generalize Eq. (8.63) taking into account the external potentialU (r) actingon bosons. This can be done introducing the shifted single-particle energy

ξq(r) = �2q2

2m+U (r) − μ .

In this way one can write the local number density as

n(r) = n0(r) +∫

d3q(2π)3

1

eξq (r)kB T − 1

,

where n0(r) is the local condensate density. Below Tc one can set μ = 0. Moreover,at Tc by definition n0(r) = 0 and consequently

n(r) =∫

d3q(2π)3

1

e�2q2

2m +U (r)kB Tc − 1

.

After integration over momenta one obtains

n(r) = 1

λ3Tc

g3/2(e− U (r)

kB Tc ) ,

where

λTc =(

2π�2

mkBTc

)1/2

is the de Broglie wavelength at the critical temperature Tc and gn(x) is the Bosefunction, defined as

gn(x) = 1

�(n)

∫ +∞

0dy

x yn−1 e−y

1 − x e−y=

+∞∑i=1

xi

in,

Page 205: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

198 8 Functional Integration for the Bosonic Field

with �(n) the Euler gamma function, and the last equality holds only if |x | < 1.Moreover, integrating over coordinates one gets

N =∫

d3r n(r) = g3(1)

(�ω

kBTc

)3

,

from which

Tc = �ω

g3(1) kBN 1/3 .

For T ≤ Tc one can write

N = N0 +∫

d3rd3q(2π)3

1

e�2q2

2m +U (r)kB T − 1

and consequently

N0

N= 1 − 1

N

∫d3rd3q(2π)3

1

e�2q2

2m +U (r)kB T − 1

.

After integration over coordinates and momenta, and taking into account the formulaof Tc, one finally gets

N0

N= 1 −

(T

Tc

)3

.

Problem 8.3Calculate, at the Gaussian level, the zero-temperature grand potential � of a 1Duniform Bose gas with repulsive 1D strengthg1D by using dimensional regularization.

SolutionAt the Gaussian level the zero-temperature grand potential can be written as

� = �0 + �(0)g ,

where

�0 = − μ2

2g1D

is the mean-field grand potential with μ the chemical potential. Within the dimen-sional regularization, in a D-dimensional space, the beyond-mean-field contribution�(0)

g is instead given by Eq. (8.89). Setting D = 1 one immediately finds

Page 206: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.4 Solved Problems 199

�(0)g = − 2

(m�2

)1/2μ1/2 .

Problem 8.4Given the 1D stationary Gross-Pitaevskii equation

[− �

2

2m

d2

dz2+ g1D| f (z)|2

]f (z) = μ f (z) ,

with g1D the attractive (negative) strength, prove by construction that it admits thestationary solution

f (z) =√m|g1D|

8�2Sech

[m|g1D|

4�2z

]

with

μ = −m g21D

16 �2.

SolutionLet us assume that f (z) is real. Then the 1D stationary Gross-Pitaevskii equationcan be rewritten as

f ′′(z) = −∂W ( f )

∂ f,

where

W ( f ) = 1

2

m|g1D|�2

f 4 + mμ

�2f 2 .

Thus, f (z) can be seen as the “coordinate” for a fictitious particle at “time” z. Theconstant of motion of the problem reads

K = 1

2f ′(z)2 + W ( f ) ,

from which one findsd f

dz= √K − W ( f ) .

Imposing that f (z) → 0 as |z| → ∞ one gets K = 0 and consequently

d f√−W ( f )= dz ,

Page 207: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

200 8 Functional Integration for the Bosonic Field

or explicityd f√

− 12m|g1D |

�2 f 4 + m|μ|�2 f 2

= dz ,

with μ < 0. Inserting the integrals one obtains

∫ f (z)

f (0)

d f√− 1

2m|g1D |

�2 f 4 + m|μ|�2 f 2

= z

Setting f ′(0) = 0, from the definition of K and using K = 0 one finds W ( f (0)) = 0and therefore

f (0) =√

2|μ||g1D| .

After integration

1√m|μ|

ArcSech

[√|g1D|2|μ| f (z)

]= z

from which

f (z) =√

2|μ||g1D| Sech

[√m|μ|�2

z

].

Imposing the normalization condition

∫dz f (z)2 = 1 ,

one obtains

μ = −m g21D

16 �2.

Further Reading

For the functional integration of the non-relativistic bosonic field:N. Nagaosa, Quantum Field Theory in Condensed Matter Physics, Chap. 1, Sects. 1.1and 1.2 (Springer, Berlin, 1999).H.T.C. Stoof, K.B. Gubbels, and D.B.M. Dickerscheid, Ultracold Quantum Fields,Chap. 6, Sects. 6.1, 6.2, 6.3, and 6.4 (Springer, Berlin, 2009).A. Altland and B. Simons, Condensed Matter Field Theory, Chap. 2, Sect. 2.1 (Cam-bridge Univ. Press, Cambridge, 2006).

Page 208: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

8.4 Solved Problems 201

For a review of methods of regularization in the case of interacting bosons:L. Salasnich and F. Toigo, Zero-point energy of ultracold atoms, Phys. Rep. 640, 1(2016).For bright and dark solitons in the nonlinear Schrödinger equation:A. Scott (ed.), Encyclopedia of Nonlinear Science (Taylor and Francis, New York,2005).

Page 209: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Chapter 9Functional Integrationfor the Fermionic Field

In this chapter we consider the quantum statistical mechanics of identical fermions byintroducing the formalism of functional integration for the fermionic field. We thenderive the equation of state of non-interacting and weakly-interacting repulsive fermi-ons. We also discuss the paring of electrons within the mean-field Bardeen-Cooper-Schrieffer (BCS) theory of low-temperature superconductivity and the Ginzburg-Landau theory, which is quite successful near the critical temperature. The BCStheory is then extended to investigate the BCS-BEC crossover of ultracold atomsfrom the weak-coupling BCS regime of Cooper pairs to the strong-coupling regimewhere composite bosons form giving rise to Bose-Einstein condensation (BEC) ofmolecules.

9.1 Partition Function of Interacting Fermions

For a system of non-relativistic interacting two-spin-component fermions the parti-tion function is given by

Z = Tr [e−β(H−μ↑ N↑−μ↓ N↓)] , (9.1)

where the quantum-field-theory Hamiltonian is

H =∑

σ=↑,↓

∫d3r ψ+

σ (r)(

− �2

2m∇2 +Uσ(r)

)ψσ(r)

+ 1

2

∑σ,σ′=↑,↓

∫d3r d3r′ ψ+

σ (r) ψ+σ′(r′) Vσ,σ′(r, r′) ψσ′(r′) ψσ(r) , (9.2)

with Uσ(r) the external potential acting on fermions with spin σ, Vσ,σ′(r, r′) theinter-particle interaction, while the number operator for fermions with spin σ reads

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1_9

203

Page 210: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

204 9 Functional Integration for the Fermionic Field

Nσ =∫

d3r ψ+σ (r)ψσ(r) . (9.3)

As usual, β = 1/(kBT ) with kB the Boltzmann constant and T the temperature, whileμσ is the chemical potential for the fermions with spin σ.

The Hamiltonian (9.2) generalizes the one, valid for spinless fermions, describedin Chap. 7. Assuming that (i) the external confinement does not depend on spin σ, (ii)there is a contact interaction of strength g between fermions, (iii) there is chemicalequilibrium, i.e. μ↑ = μ↓ = μ, we can write

Z = Tr [e−β(H−μN )] (9.4)

withN = N↑ + N↓ (9.5)

and

H =∑

σ=↑,↓

∫d3r ψ+

σ (r)(

− �2

2m∇2 +U (r)

)ψσ(r) + g

2

∫d3r ψ+

↑ (r) ψ+↓ (r) ψ↓(r) ψ↑(r) .

(9.6)Notice that, as shown in Chap. 6, in the presence of a contact inter-particle potential,there is an interaction only between fermions with opposite spins.

In the previous chapter we have seen that the partition function of identical bosonscan be expressed as a functional integral, taking into account that the trace of anoperator can be performed by using any complete set of basis states, and in particularthe basis of generalized coherent states of bosons. The same reasoning can be appliedto the partition function of identical fermions by using the basis of generalizedcoherent states of fermions. As shown in Chap. 7, in the case of fermionic coherentstates one must use complex Grassmann variables and complex Grassmann fields.Thus, we can write

ψσ(r) |ψ〉 = ψσ(r) |ψ〉 , (9.7)

where ψσ(r) is the complex Grassmann field with spin σ and |ψ〉 is the generalizedfermionic coherent state. The complex conjugate of the Grassmann field ψσ(r) isdenoted as ψσ(r) and it satisfies the equation

〈ψ|ψ+σ (r) = 〈ψ|ψσ(r) . (9.8)

By expanding the fermionic field operator ψσ(r) as

ψσ(r) =∑

α

cα,σ φα(r) , (9.9)

where φα(r) are orthonormal eigenfunctions of single-particle states |α〉 and cα,σ isthe fermionic operator which annihilates a fermion with spin σ in the state |α〉, we

Page 211: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.1 Partition Function of Interacting Fermions 205

clearly haveψσ(r) =

∑α

cα,σ φα(r) (9.10)

withcα,σ|ψ〉 = cα,σ|ψ〉 for any α and σ =↑,↓ (9.11)

and cα,σ the complex Grassmann variables for the single-particle state |α〉 and spinσ.

Making a long story short, in full analogy with the bosonic case, the partitionfunction of identical fermions described by the many-body Hamiltonian (9.6) can bewritten as

Z =∫

D[ψσ, ψσ] exp

{−1

�S[ψσ, ψσ]

}, (9.12)

where

S[ψσ, ψσ] =∫

�β

0dτ

∫d3r L(ψs, ψs) (9.13)

is the Euclidean action functional and

L =∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 +U (r) − μ

]ψσ + g ψ↑ ψ↓ ψ↓ ψ↑ (9.14)

is the Euclidean Lagrangian density, where ψσ(r, τ ) and ψσ(r, τ ) are the Grassmannfields which depend also on the imaginary time τ . Notice that one picks up a minussign each time one permutes two Gaussian fields. In the fermionic functional integral(9.12) the integration measure reads

∫D[ψσ, ψσ] =

∫ ∏σ=↑,↓

∏(r,τ )

dψσ(r, τ ) dψσ(r, τ ) , (9.15)

and the boundary conditions for the imaginary time τ are anti-periodic, i.e. ψ(r, 0) =−ψ(r, �β).

9.2 Ideal Fermi Gas

We consider a D-dimensional (D = 1, 2, 3) Fermi gas of noninteracting fermions ina D-dimensional box of volume LD . Under these conditions exact analytical resultscan be obtained in any spatial dimension D. In this case the Euclidean Lagrangiandensity is given by

Page 212: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

206 9 Functional Integration for the Fermionic Field

L =∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 − μ

]ψσ , (9.16)

and the Euclidean action can be written in a diagonal form as

S[ψ,ψ∗] =∑

σ=↑,↓

∑Q

�λQ ψQ,σψQ,σ (9.17)

where Q = (q, iωn) is the D + 1 vector denoting the momenta q and frequenciesωn = 2π(n + 1/2)/(β�) (which are usually called fermionic Matsubara frequen-cies), and

λQ = β (−i�ωn + �2q2

2m− μ) . (9.18)

Here ψQ,σ = ψq,iωn ,σ = ψq,n,σ is the Fourier transform of ψσ(r, τ ) and the quanti-zation of frequencies ωn is a consequence of the anti-periodicity ψσ(r, τ + �β) =−ψσ(r, τ ).

Taking into account that

Z =∫

D[ψσ, ψσ] exp

⎧⎨⎩−

∑σ=↑,↓

∑Q

λQ,σψQ,σψQ,σ

⎫⎬⎭ =

∏σ=↑,↓

∏Q

λQ , (9.19)

the grand potential reads

� = − 1

βlnZ = − 2

β

∑Q

ln (λQ) = − 1

β

∑q

+∞∑n=−∞

ln [β2(�2ω2n + ξ2

q)] , (9.20)

summing over the two spin states and with ξq the shifted free-particle spectrum, i.e.

ξq = �2q2

2m− μ . (9.21)

The Matsubara frequencies of fermions are given by ωn = (2n + 1)π/β due to theanti-symmetry of the Grassmann fields. In this way one gets the exact result

1

β

+∞∑n=−∞

ln [β2(�2ω2n + ξ2

q)] = ξq + 2

βln (1 + e−β ξq ) . (9.22)

Thus, the grand potential finally reads

� = �(0) + �(T ) , (9.23)

Page 213: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.2 Ideal Fermi Gas 207

where�(0) = −

∑q

ξq (9.24)

is the zero-point energy of fermionic single-particle excitations, i.e. the zero-temperature contribution of quantum fluctuations, and

�(T ) = − 2

β

∑q

ln(1 + e−βξq

)(9.25)

takes into account thermal fluctuations.In the continuum limit, where

∑q → LD

∫dDq/(2π)D , the zero-point energy

�(0)

LD= − SD

(2π)D

∫ +∞

0dq qD−1

(�

2q2

2m− μ

)(9.26)

of the ideal Fermi gas is clearly ultraviolet divergent at any integer dimension D, i.e.for D = 1, 2, 3. Here SD = 2πD/2/�(D/2) is the solid angle in D dimensions with�(x) the Euler gamma function. This divergent zero-point energy of the ideal Fermigas is completely eliminated by dimensional regularization. Consequently, the exactgrand potential of the ideal Fermi gas is given by

LD= − 2

β

∫dDq

(2π)Dln(1 + e−βξq

), (9.27)

and the corresponding number density reads

n = 1

LD

∂�

∂μ= 2

∫dDq

(2π)D

1

eβξq + 1. (9.28)

At zero temperature this formula becomes

n = 1

LD

∂�

∂μ= 2

∫dDq

(2π)D�(−ξq) , (9.29)

where �(x) is the Heaviside step function. After integration one obtains

μ =

⎧⎪⎨⎪⎩

�2

2m (3π2n)2/3 for D = 3�

2

2m 2πn for D = 2�

2

2mπ2

4 n2 for D = 1

(9.30)

where the chemical potential μ can be identified with the Fermi energy εF .

Page 214: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

208 9 Functional Integration for the Fermionic Field

9.3 Repulsive Fermions: Hartree-Fock Approximation

Let us now consider identical fermions with a repulsive contact interaction. Thissystem is described by the Lagrangian density (9.14) with g > 0. In addition, wesuppose that the system has the same number of spin up and spin down fermions,i.e. n↑ = n↓ = n/2, with n the total number density of fermions.

The problem cannot be solved exactly and we adopt a mean-field Hartee-Fockapproximation for the quartic term, namely

g ψ↑ ψ↓ ψ↓ ψ↑ � gn

2ψ↑ψ↑ + g

n

2ψ↓ψ↓ − g

n2

4. (9.31)

In this way the Lagrangian density (9.14) becomes quadratic

L =∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 − μ + 1

2gn

]ψσ − g

n2

4, (9.32)

By repeating the procedure of the previous section it is straightforward to showthat the mean-field grand potential reads

LD= − 2

β

∫dDq

(2π)Dln(1 + e−βEq

)− gn2

4, (9.33)

where the single-particle Hartree-Fock energy Eq given by

Eq = �2q2

2m− μ + 1

2gn . (9.34)

The number density n must satisfy the equation

n = 1

LD

∂�

∂μ= 2

∫dDq

(2π)D

1

eβEq + 1, (9.35)

and, at zero-temperature, after integration one obtains

μ =

⎧⎪⎨⎪⎩

�2

2m (3π2n)2/3 + 12gn for D = 3

�2

2m 2πn + 12gn for D = 2

�2

2mπ2

4 n2 + 1

2gn for D = 1

(9.36)

In the three-dimensional case this result was obtained for the first time in 1957 byKerson Huang and Chen Ning Yang.

Page 215: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 209

9.4 Attractive Fermions: BCS Approximation

Let us now consider identical fermions with an attractive contact interaction. Thissystem is described by the Lagrangian density (9.14) with g < 0. In addition, wesuppose that the system has the same number of spin up and spin down fermions,i.e. n↑ = n↓ = n/2, with n the total number density of fermions.

As discussed in the previous section, the problem cannot be solved exactly due tothe presence of the quartic term in the Lagrangian density. Also in this case we adopta mean-field treatment which is however quite different with respect to the Hartree-Fock one of repulsive fermions. We suppose that, due to the attractive interaction,fermions with opposite spin can create a sort of bosonic particle with spin zero.

In particular we use the following mean-field Bardeen-Cooper-Schriffer (BCS)approximation

− g0 ψ↑ ψ↓ ψ↓ ψ↑ � −�∗ψ↓ψ↑ − �ψ↑ψ↓ + |�|2g0

, (9.37)

where g0 = −g > 0 and � is a complex order parameter usually called pairingenergy or pairing energy gap. For attractive fermions this BCS approximation givesmuch better agreement with exact calculations (and also with experimental data) thanthe Hartree-Fock approximation. In this way the Lagrangian density (9.14) becomesquadratic and given by

L = ψσ

[�∂τ − �

2

2m∇2 − μ

]ψσ − �∗ ψ↓ ψ↑ − �ψ↑ ψ↓ + |�|2

g0. (9.38)

Given the BCS Lagrangian density (9.38), the corresponding BCS Euclidean actioncan be written in a diagonal form as

S[ψ,ψ∗] =∑

σ=↑,↓

∑Q

�λQ ψQ,σψQ,σ + �2

g0LD

�β , (9.39)

whereλQ = β (−i�ωn + Eq) (9.40)

with

Eq =√(

�2q2

2m− μ

)2

+ |�|2 (9.41)

the single-particle fermionic excitations.The mean-field grand potential can then be written as

� = �0 + �(0) + �(T ) (9.42)

Page 216: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

210 9 Functional Integration for the Fermionic Field

including the grand potential of the order parameter �

�0 = �2

g0LD , (9.43)

the zero-point energy of fermionic single-particle excitations

�(0) = −∑q

Eq , (9.44)

and the finite-temperature grand potential of the fermionic single-particle excitations

�(T ) = 2

β

∑q

ln (1 + e−β Eq ) . (9.45)

The energy gap � which appears in the fermionic energy spectrum (9.41) isobtained by imposing that

∂�

∂�= 0 , (9.46)

from which one gets

1

g0= 1

LD

∑q

tanh(

βEq

2

)

2Eq(9.47)

that is the so-called gap equation. The number density n is instead obtained from

n = − 1

LD

∂�

∂μ, (9.48)

which gives the so-called number equation

n = 1

LD

∑q

(1 −

�2q2−μ2m

Eq

)tanh

(βEq

2

). (9.49)

In three-dimensions (D = 3) the gap equation (9.47) is ultraviolet divergent. Itbecomes convergent by introducing an ultraviolet cutoff � such that

1

g0= 1

L3

∑|q|<�

tanh(

βEq

2

)

2Eq. (9.50)

For electrons in superconducting metals the cutoff energy �2�2/(2m) can be iden-

tified with the Debye energy �ωD , where ωD is the Debye frequency, that is the

Page 217: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 211

maximum frequency of vibration for the atoms that make up the ionic crystal of themetal. The cutoff � is then quite natural because it accounts for the fact that phononsof ionic vibrations, which mediate the attraction between condition electrons, can atmost transfer an energy �ωD .

In the thermodynamic limit, where the volume L3 goes to infinity,∑

q can bereplaced by L3

∫d3q/(2π)3 = L3

∫N (ξ)dξ with N (ξ) = ∫

d3q/(2π)3 δ(ξ − ξq),ξq = εq − μ, and εq = �

2q2/(2m). In metals the condition �ωD � μ is always sat-isfied, consequently we can use the approximation

∫N (ξ)dξ � N (0)

∫dξ, where

N (0) =∫

d3q(2π)3

δ(μ − εq) (9.51)

is the density of states at the Fermi surface. In this way the gap equation (9.50)becomes

1

g0N (0)=∫

�ωD

0

tanh(β2

√ξ2 + �2)√

ξ2 + �2dξ . (9.52)

At zero-temperature, from Eq. (9.52) we find immediately a simple formula for thezero temperature energy gap �(0):

1

g0N (0)= ln

⎛⎝ �ωD

�(0)+√

1 + �2ω2D

�(0)2

⎞⎠ . (9.53)

Under the condition �(0) � �ωD , from Eq. (9.53) we find the weak-coupling BCSresult

�(0) = 2�ωD exp(− 1

g0N (0)) (9.54)

for the zero-temperature energy gap. Clearly the energy gap � decreases by increas-ing the temperature T and it becomes zero at the critical temperature Tc of thesuperconducting transition. Setting � = 0 in Eq. (9.52) one finds

kBTc = 1.13 �ωD exp(− 1

g0N (0)) . (9.55)

that is the BCS critical temperature obtained by John Bardeen, Leon Cooper and JohnRobert Schrieffer in 1957. Remarkably, taking into account Eqs. (9.54) and (9.55), theBCS theory predicts that the zero-temperature energy gap �(0) of superconductorsin the weak-coupling regime is related to the critical temperature Tc by the simpleequation

�(0) = 1.764 kBTc . (9.56)

Page 218: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

212 9 Functional Integration for the Fermionic Field

9.4.1 Condensate Fraction of Cooper Pairs

In the mean-field BCS approximation we have introduced the energy gap �. Ingeneral, this pairing order parameter is defined as

�(r, τ ) = g0 〈ψ↓(r, τ )ψ↑(r, τ )〉 . (9.57)

We have seen that a nonzero value of this quantity implies the formation of Cooperpairs (two fermions with opposite spins) which give rise to a quasi-bosonic particlesof spin zero. As suggested by Chen Ning Yang in 1962, the condensate wavefunctionof these Cooper pairs is given by

�(r, r′, τ ) = 〈ψ↓(r, τ ) ψ↑(r′, τ )〉 . (9.58)

This two-particle wave function is strictly related to the largest eigenvalue N0 oftwo-body density matrix of the system. N0 gives the number of Fermi pairs (Cooperpairs) in the lowest state, i.e. the condensate number of Fermi pairs, and it can bewritten as

N0 =∫

|�(r, r′, τ )|2 d3r d3r′ . (9.59)

A finite value for the condensate fraction f = N0/(N/2) in the thermodynamic limitN → ∞ implies the so-called off-diagonal-long-range-order.

By using the formalism of functional integration, the condensate wavefunction ofCooper pairs can be written as

�(r, r′, τ ) = 1

Z

∫D[ψσ, ψσ] ψ↓(r, τ ) ψ↑(r′, τ ) exp

{−1

�S[ψσ, ψσ]

}. (9.60)

Inserting the BCS Euclidean action (9.39), after some straightforward but lengthy cal-culations one finds that the condensate number of Cooper pairs satisfies this expres-sion

N0 =∑q

�2

4E2q

tanh2(βEq

2) . (9.61)

For the electrons in a superconducting metal, taking into account Eq. (9.51) one theneasily finds

n0 = 1

2N (0)�2

∫�ωD

0

tanh2(β2

√ξ2 + �2)

ξ2 + �2dξ (9.62)

where n0 = N0/L3 is the density of electrons in the condensate. At zero temperatureEq. (9.62) gives

n0(0) = 1

2N (0)�(0) arctan(

�ωD

�(0)) . (9.63)

Page 219: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 213

This expression shows that the condensate density n(0) can be expressed in terms ofdensity of states N (0), energy gap �(0) and Debye energy �ωD . Consequently, thezero-temperature condensate fraction f (0) = n0(0)/(n/2) is given by

f (0) = 1

2

N (0)

n�(0) arctan(

�ωD

�(0)) , (9.64)

where n is the density of conduction electrons. Moreover, under the condition�(0) � �ωD , the condensate density can be written as

n0(0) = π

4N (0)�(0) . (9.65)

We stress that in many real superconductors the simple mean-field BCS approxi-mation is not accurate, and one has to take into account the retarded electron-electroninteraction via phonons and also the Coulomb repulsion. The results obtained aboveby using the mean-field BCS theory are reliable in the weak-coupling regime, i.e. for�(0) � �ωD , where gN (0) ≤ 0.3. Therefore we will continue our analysis of thecondensate fraction only for a class of superconductors which satisfy this condition.

Within the free-electrons Sommerfeld approximation, the energy spectrum εqof conduction electrons has a simple quadratic behavior εq = �

2q2/(2m∗) but m∗,the effective mass of electrons, does not coincide with the physical mass m. Thefree particle density of states N f ree(0) is related to the total density of conductionelectrons by n = 4N f ree(0)μ/3 and the zero-temperature condensate fraction readsf (0) = 3π�(0)/(8μ). To get a better estimate, we correct the free electron value ofN (0) by an effective mass as obtained from specific heat measurements, i.e. we usethe expression N (0) = (m∗/m)N f ree(0).

In the first two columns of Table 9.1 we show the zero-temperature condensatedensity n0 and condensate fraction f (0) of simple metals obtained from Eqs. (9.63)and (9.64) by using experimental data of �(0) and ωD . In the third column we report

Table 9.1 BCS predictions for weak-coupling superconducting metals: n0 is the zero-temperaturecondensate density, f (0) = n0(0)/(n/2) is the zero-temperature condensate fraction, gN (0) isthe electron-phonon strength, and Tc is the critical temperature. Experimental data taken fromN.W. Ashcroft and N.D. Mermin, Solid State Physics (Holt-Sounders, Philadelphia, 1976) and J.P.Carbotte, Rev. Mod. Phys. 62, 1027 (1990)

n0(0) [10−33 m−3] f (0) [10−5] gN (0) Tc [K]

Cd 4.18 0.9 0.179 0.51

Zn 7.72 1.2 0.172 0.79

Al 22.4 2.5 0.168 1.15

Tl 31.0 5.9 0.263 2.43

In 62.1 10.8 0.267 3.46

Sn 62.5 8.4 0.254 3.68

Page 220: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

214 9 Functional Integration for the Fermionic Field

the electron-phonon strength gN (0) calculated with Eq. (9.53) knowing f (0). Thetable shows that indeed these simple metals are all in the weak-coupling regime. Forcompleteness, in the fourth column we insert the theoretical determination (see Eq.(9.55)) of the critical temperature Tc, which is very reliable for these simple metals,when compared with the experimental data.

Let us now consider the behavior of the condensate density n0 at finite temperatureT . Under the condition �ωD � kBTc, from Eq. (9.52) one finds that near Tc the energygap goes to zero according to the power law

�(T ) = 3.06 kBTc

(1 − T

Tc

)1/2

. (9.66)

Instead for the condensate densityn0(T ), from Eq. (9.62) and the previous expression,we find near Tc

n0(T ) = 0.43 N (0)�(T )2

kBTc= 4.03 N (0)kBTc

(1 − T

Tc

). (9.67)

Taking into account Eq. (9.64), the BCS theory predicts that the zero-temperaturecondensate density in superconductors in the weak-coupling regime, can be writtenas

n0(0) = 1.39 N (0)kBTc . (9.68)

9.4.2 Ginzburg-Landau Theory

In 1950, seven years before the BCS theory, Lev Landau and Vitaly Ginzburg pro-posed a phenomenological approach to describe the superconducting phase transi-tion. The main idea is that, close to the critical temperature, the grand potential of asuperconducting material can be written as

� = �n + �s , (9.69)

where �n is the contribution due to the normal component and �s is the contributiondue to the emergence of a superconducting complex order parameter �(r) belowa critical temperature. Within the Ginzburg-Landau approach, for a D-dimensionalsystem of area LD , the super component �s is given by

�s =∫

LD

dDr{a(T ) |�(r)|2 + b

2|�(r)|4 + γ |∇�(r)|2

}, (9.70)

wherea(T ) = α kB (T − Tc) (9.71)

Page 221: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 215

is a parameter which depends on the temperature T (kB is the Boltzmann constant)and becomes zero at the mean-field critical temperature Tc, while b > 0 and γ > 0are temperature-independent phenomenological parameters.

The partition function Z of the system is given by

Z = e−β�n Zs (9.72)

where

Zs =∫

d[�∗, �] e−β�s [�,�∗] , (9.73)

with β = 1/(kBT ), and the thermal average of an observable A that is a functionalof �(r) and �∗(r) reads

〈A〉 = 1

Zs

∫d[�,�∗] A[�,�∗] e−β�s [�,�∗] . (9.74)

Notice that in these functional integrals the complex field �(r) does not depend ontime. As shown in the previous chapter, they are fully reliable at high temperatureswhere quantum fluctuations are negligible.

Assuming a real and uniform order parameter, i.e.

�(r) = �0 , (9.75)

the energy functional (9.70) with Eqs. (9.71) and (9.75) becomes

�s0

LD= a(T )�2

0 + b

2�4

0 . (9.76)

Minimizing �s0 with respect to �0 one immediately finds

a(T )�0 + b�30 = 0 , (9.77)

and consequently

�0 ={

0 for T ≥ Tc√− a(T )

b for T < Tc. (9.78)

Thus, the uniform order parameter �0 becomes different from zero only below themean-field critical temperature Tc, where, by definition, Tc is such that

a(Tc) = 0 . (9.79)

In 1959 Lev Gor’kov showed that the phenomenological Ginzburg-Landau theorycan be deduced from the microscopic BCS theory. In particular, the order parameter�0 can be identified with the BCS energy gap �, while the coefficients α, b and γ of

Page 222: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

216 9 Functional Integration for the Fermionic Field

the Ginzburg-Landau energy functional (9.70) are directly related to the parametersof the BCS Lagrangian density (9.38), taking into account Eqs. (9.54), (9.55), (9.71),and (9.78). We notice, however, that the Ginzburg-Landau theory is somehow betterthan the the mean-field BCS theory because, in general, the Ginzburg-Landau orderparameter �(r) is not uniform, while the mean-field BEC energy gap � is assumedto be uniform.

Let us now consider the effects of a space-dependent Ginzburg-Landau orderparameter �(r). Extremizing the functional (9.70) with respect to ψ∗(r) one gets theEuler-Lagrange equation

a(T )ψ + b |�|2� − γ ∇2� = 0 . (9.80)

We write the space-dependent order parameter �(r) in the following way

�(r) = 0 + η(r) , (9.81)

where η(r) represents a fluctuation with respect to the real and uniform configuration0 with the condition

〈η〉 = 〈η∗〉 = 0 , (9.82)

where 〈·〉 is the thermal average. Inserting Eq. (9.81) into Eq. (9.80), we find

a(T )0 + b30 + a(T ) η + 2b2

0η + b20η

∗ + 2b0|η|2+ b0η

2 + b |η|2η − γ∇2η = 0 , (9.83)

and after thermal averaging we obtain

[a(T ) + 2b 〈|η|2〉 + b 〈η2〉] 0 + b3

0 + b 〈|η|2η〉 = 0 . (9.84)

Clearly, only removing all thermal averages one recovers Eq. (3.4) and finds that 0

is equal to �0 and given by Eq. (9.78).We take into account thermal fluctuations of the order parameter by keeping 〈|η|2〉

but neglecting the anomalous averages 〈η2〉 and 〈|η|2η〉. In this way we get

(a(T ) + 2b 〈|η|2〉) 0 + b3

0 = 0 , (9.85)

and consequently

0 ={

0 for T ≥ Tc,B√− a(T )+2b〈|η|2〉

b for T < Tc,B. (9.86)

In this case the uniform order parameter 0 becomes different from zero only belowthe beyond-mean-field critical temperature Tc,B given by the equation

a(Tc,B) + 2b 〈|η|2〉c,B = 0 . (9.87)

Page 223: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 217

It is important to stress that, to treat self-consistently the fluctuating field η(r), ourapproximation requires |η|2η � 2〈|η|2〉η and η2 � 0 in Eq. (9.83), which becomes

a(T ) η + 2b φ20η + 2b 〈|η|2〉η + b φ0η

∗ − γ∇2η = 0 , (9.88)

taking into account Eq. (9.85).Let us work with T ≥ Tc,B where φ0 = 0. Then Eq. (9.88) reduces to

a(T ) η + 2b 〈|η|2〉η − γ∇2η = 0 , (9.89)

which is the Euler-Lagrange equation of the Gaussian energy functional

�sg =∫

LD

dDr{(a(T ) + 2b〈|η|2〉)|η(r)|2 + γ |∇η(r)|2

}. (9.90)

Expanding the field η(r) in a plane-wave basis, i.e.

η(r) =∑k

ψkeik·r

LD/2, (9.91)

it is straightforward to find

�sg =∑k

(a(T ) + 2b〈|η|2〉 + γk2) |ψk|2 . (9.92)

The superconductive partition function Zs of the system is then given by

Zs =∫

d[η∗, η] e−β�sg [η,η∗] , (9.93)

where β = 1/(kBT ) and the thermal average

〈|η|2〉 = 1

Zs

∫d[η∗, η] |η(r)|2 e−β�sg[η,η∗] (9.94)

reads

〈|η|2〉 = 1

LD

∑k

1

β (a(T ) + 2b〈|η|2〉 + γk2). (9.95)

At T = Tc,B , where Eq. (9.87) holds, we get

〈|η|2〉c = 1

LD

∑k

kBTc,Bγk2

. (9.96)

Page 224: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

218 9 Functional Integration for the Fermionic Field

In the continuum of momentak, where∑

k = LD∫dDk/(2π)D , this thermal average

(9.96) goes to infinity, however, usually there are physical cutoffs which constrainthe values of k.

In the two-dimensional case (D = 2), where fluctuations of the mean-field criticaltemperature are expected to be relevant, from Eq. (9.96) we find

〈|η|2〉c = kBTc,B2πγ

ln

(�

k0

), (9.97)

where k0 is an infrared cutoff and � an ultraviolet cutoff. Then Eq. (9.87) gives

Tc − Tc,BTc,B

= 4 Gi ln

(�

k0

)(9.98)

with

Gi = b

4παγ(9.99)

the so-called Ginzburg-Levanyuk number. In addition, one can write � = 1/(2ξ)with ξ = (γ/(αkBTc,B))1/2 the coherence length and k0 = (kBTc,BαGi/γ)1/2. Itfollows that �/k0 = 1/(4 Gi)1/2 and consequently

Tc − Tc,BTc,B

= 2 Gi ln

(1

4 Gi

), (9.100)

that is a simple analytical formula for the shift of the critical temperature due tothermal fluctuations of the Ginzburg-Landau order parameter.

9.4.3 BCS-BEC Crossover

Scattering theory plays an essential role in the description of a three-dimensionalattractive Fermi gas of atoms which undergoes BCS-BEC crossover, i.e. the crossoverof a fermionic superfluid from weakly-bound BCS-like Cooper pairs to the Bose-Einstein condensation (BEC) of strongly-bound molecules.

In 1947 Bogoliubov found a regularization procedure relating the physical s-wavescattering length aF of the actual interatomic potential to the strength g of the modelcontact interaction:

m

4π�2aF= 1

g+ 1

L3

∑q

m

�2q2. (9.101)

This equation may be easily deduced from the Lippman-Schwinger equation for theT -matrix T (E) of scattering theory. In the low energy limit, the Lippman-Schwinger

Page 225: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 219

equation renormalizes the coupling constant of a contact interaction V (r) = g δ(r)as

1

gr (E)≡ 1

T (E + i0+)= 1

g+ 1

LD

∑q

m

�2q2 − E − i0+ , (9.102)

where T (E) = Tk,k′(E) ≡ 〈k|T (E)|k′〉 when k = k ′ and E = �2k2/m.

At very low energy one may relate the s-wave scattering length aF , a physicalmeasurable quantity, to T (E) according to

T (E + i0+) =

⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩

4π�2

maF

1+iaF√

mE�2

for D = 3

− 2π�2

m1

ln (aB√

mE�2 e

γ/2)+iπ/2for D = 2

− 2�2

m

i√

mE�2

1+iaF√

mE�2

for D = 1

(9.103)

with γ = 0.577... the Euler-Mascheroni constant. In three dimensions (D = 3) one cansafely set E = 0, so that T (0 + i0+) = 4π�

2aF/m, and Eq. (9.101) is immediatelyrecovered. In 2D and 1D one must use Eq. (9.102) with a finite value of E , while in1D one gets T (E + i0+) = −2�

2/(maF ) under the condition aF√mE/�2 � 1.

As recalled above, the bare interaction strength g appearing in the Lagrangiandensity (9.38) is related to the physical s-wave scattering length aF of fermions by

m

4π�2aF= 1

g+ 1

L3

∑|q|<�

m

�2q2, (9.104)

where, as usual, the ultraviolet cutoff � is introduced to avoid the divergence of thesecond term on the right side. The low energy scattering length is negative for anattractive potential if no bound states are present, while it becomes positive whenthe interaction is so attractive as to admit a bound state. Equation (9.104) allows forthe change of sign of aF as the strength of the attractive potential becomes moreand more negative. In fact, in the continuum limit

∑k → L3

∫d3k/(2π)3, after

integration over momenta, it reads

m

4π�2aF= 1

g+ m

2π2�2� . (9.105)

Therefore, in the weak-coupling BCS limit, where g → 0−, the first term on the r.h.s.of Eq. (9.105) dominates and aF = mg/(4π�

2) → 0−, while in the strong-couplingBEC limit, where g → −∞, the second term on the r.h.s. of Eq. (9.105) dominatesand aF = π/(2�) → 0+ when � is sent to infinity.

In addition, from Eq. (9.105) one finds that a bound state, that is a pole in theT-matrix of Eq. (9.102), is possible only if g < −2π2

�2/(m�). See Fig. 9.1. In the

presence of a bound state of energy −εB , for the T-matrix one has

Page 226: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

220 9 Functional Integration for the Fermionic Field

-2 -1.5 -1 -0.5 0g

-20

-15

-10

-5

0

5

10

15

20

a F

BCS regime

BEC regime

Unitarity

Fig. 9.1 Fermionic scattering length aF as a function of the bare interaction strength g, for a finiteultraviolet cutoff �, see Eq. (9.105). In the plot aF is in units of m/(4π�

2) and m�/(2π2�

2) = 1

1

T (−εB)= 0 , (9.106)

and consequently from Eq. (9.101) it follows

− 1

g= 1

L3

∑|k|<�

m

�2k2 + mεB. (9.107)

After integration over momenta one obtains

− 1

g= m

2π2�2

⎛⎝� −

√mεB

�2arctan

⎛⎝ �√

mεB�2

⎞⎠⎞⎠ , (9.108)

and comparing Eq. (9.105) with Eq. (9.108) one finds

aF =√

�2

mεB

π

2 arctan(�

√�2

mεB

) . (9.109)

In the limit � → +∞ from this interesting formula one gets

Page 227: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.4 Attractive Fermions: BCS Approximation 221

εB = �2

ma2F

, (9.110)

that is the familiar relation between the bound-state energy and the 3D s-wave scat-tering length aF .

The full BCS-BEC crossover in 3D, which has been experimentally observed in2004 with ultracold and dilute alkali-metal atoms, can be theoretically investigatedby inserting Eq. (9.104) into Eq. (9.50). In this way, in the limit � → +∞, one gets

m

4π�2aF= 1

L3

∑q

⎛⎝ tanh

(βEq

2

)

2Eq− m

�2q2

⎞⎠ (9.111)

that is a renormalized gap equation valid both for aF < 0 and aF > 0. Solving thisrenormalized gap equation, coupled to the number equation (9.49), one finds simpleanalytical results for the chemical potential μ, the energy gap �, and the condensatefraction n0/n (see Eq. 9.61) in limiting cases. In particular, introducing the adimen-sional strength y = 1/(kFaF ) with kF = (3π2n)1/3, three relevant limiting cases are:(i) the BCS regime, where y � −1, μ/� � 1 and the size of weakly bound Cooperpairs exceeds the typical interparticle spacing k−1

F ; (ii) the unitarity limit, with y = 0(thus aF = ±∞); (iii) the BEC regime, where y � 1, μ/� � −1 and the fermionscondense as a gas of tight dimers. The main results are the following:

(i) BCS regime: μ approaches the non-interacting Fermi energy εF = �2k2

F/(2m).One finds that the condensate density is given by

n0 � mkF8π�2

� = 3π

2e2n exp

2y)

, (9.112)

with an exponentially small energy gap � = 8e−2 εF exp (y π/2).(ii) Unitarity limit: one finds μ � 0.59εF and � � 0.69εF , and the condensate

fraction is2 N0

N� 0.6994 . (9.113)

(iii) BEC regime: the condensate fraction approaches the ideal Bose gas value

2 N0

N� 1 , (9.114)

with all pairs (molecular dimers) moving into the condensate.

Page 228: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

222 9 Functional Integration for the Fermionic Field

9.5 Solved Problems

Problem 9.1Given two quantities A and B with averages 〈A〉 and 〈B〉, prove the mean-field result

A B � 〈A〉 B + A 〈B〉 − 〈A〉 〈B〉

neglecting quadratic fluctuations.

Solution

First we set

A = 〈A〉 + δA

B = 〈B〉 + δB

where

δA = A − 〈A〉δB = B − 〈B〉

are the fluctuations of A and B. Then we can write

A B =(〈A〉 + δA

) (〈B〉 + δB

)= 〈A〉 〈B〉 + 〈A〉 δB + 〈B〉 δA + δA δB .

Neglecting the term δA δB and taking into account the defintion of fluctuations, theprevious equation becomes

A B � 〈A〉 〈B〉 + 〈A〉 δB + 〈B〉 δA = 〈A〉 〈B〉 + 〈A〉 (B − 〈B〉) + 〈B〉 (A − 〈A〉)= 〈A〉 〈B〉 + 〈A〉 B − 〈A〉 〈B〉 + 〈B〉 A − 〈B〉 〈A〉= 〈A〉 B + A 〈B〉 − 〈A〉 〈B〉 ,

by using the fact that averages commutes with themselves and also with quantities.

Problem 9.2Consider a two-component three-dimensional (D = 3) Fermi gas with repulsivecontact interaction of strength g. Within the mean-field Hartree-Fock approximation,find the critical strength gc above which the zero-temperature uniform unpolarizedconfiguration becomes unstable (3D Stoner instability).

Solution

Assuming that μ↑ �= μ↓ and n↑ �= n↓, the mean-field Hartee-Fock approximationfor the quartic term gives

Page 229: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.5 Solved Problems 223

g ψ↑ ψ↓ ψ↓ ψ↑ � g n↓ψ↑ψ↑ + g n↑ψ↓ψ↓ − gn↑n↓ .

In this way the effective Lagrangian density is given by

L =∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 − μσ + gnσ

]ψσ − gn↑n↓ ,

and the mean-field grand potential reads

L3= − 1

β

∑σ=↑,↓

∫d3q

(2π)Dln(

1 + e−βE (σ)q

)− gn↑n↓ ,

where the single-particle Hartree-Fock energy E (σ)q given by

E (σ)q = �

2q2

2m− μσ + gnσ′

and σ′ �= σ. At zero-temperature, after integration over momenta one finds the fol-lowing internal energy density

E

L3= �

L3+ μ↑n↑ + μ↓n↓ = 3

5

�2

2m(6π2)2/3

(n5/3

↑ + n5/3↓)

+ gn↑n↓ ,

and chemical potentials

μσ = �2

2m(6π2)2/3n2/3

σ + gnσ′

with σ �= σ′. Given the Jacobian

J =(

∂μ↑∂n↑

∂μ↑∂n↓

∂μ↓∂n↑

∂μ↓∂n↓

)=(

23

�2

2m (6π2)2/3n−1/3↑ g

g 23

�2

2m (6π2)2/3n−1/3↓

)

the critical strength gc of the unpolarized configuration is obtained setting n↑ = n↓ =n/2 and the determinant of J equal to zero. It follows

gc = 2

3

�2

2m(6π2)2/3(

n

2)−1/3 .

Problem 9.3Consider a two-component two-dimensional (D = 2) Fermi gas with repulsive con-tact interaction of strength g. Within the mean-field Hartree-Fock approximation,find the critical strength gc above which the zero-temperature uniform unpolarizedconfiguration becomes unstable (2D Stoner instability).

Page 230: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

224 9 Functional Integration for the Fermionic Field

Solution

The procedure is quite similar to the one of the previous problem. At zero-temperature, after integration over momenta one finds the following internal energydensity

E

L2= �

L2+ μ↑n↑ + μ↓n↓ = π

�2

m

(n2

↑ + n2↓)+ gn↑n↓ ,

and chemical potentials

μσ = 2π�

2

mnσ + gnσ′

with σ �= σ′. Also in this two-dimensional case, given the Jacobian

J =(

∂μ↑∂n↑

∂μ↑∂n↓

∂μ↓∂n↑

∂μ↓∂n↓

)=(

2π �2

m gg 2π �

2

m

)

the critical strength gc of the unpolarized configuration is obtained setting n↑ = n↓ =n/2 and the determinant of J equal to zero. It follows

gc = 2π�

2

m.

Problem 9.4Consider a two-component one-dimensional (D = 1) Fermi gas with repulsive con-tact interaction of strength g. Within the mean-field Hartree-Fock approximation,find the critical strength gc above which the zero-temperature uniform unpolarizedconfiguration becomes unstable (1D Stoner instability).

Solution

The procedure is quite similar to the ones of the two previous problems. At zero-temperature, after integration over momenta one finds the following internal energydensity

E

L= �

L+ μ↑n↑ + μ↓n↓ = π2 �

2

6m

(n3

↑ + n3↓)+ gn↑n↓ ,

and chemical potentials

μσ = π2 �2

2mn2

σ + gnσ′

with σ �= σ′. Also in this one-dimensional case, given the Jacobian

J =(

∂μ↑∂n↑

∂μ↑∂n↓

∂μ↓∂n↑

∂μ↓∂n↓

)=(

π2 �2

m n↑ gg π2 �

2

m n↓

)

Page 231: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.5 Solved Problems 225

the critical strength gc of the unpolarized configuration is obtained setting n↑ = n↓ =n/2 and the determinant of J equal to zero. It follows

gc = π2 �2

m

n

2.

Problem 9.5Prove that the fermionic partition function

Z =∫

D[ψσ, ψσ] exp

{−1

�S[ψσ, ψσ]

},

where

S[ψσ, ψσ] =∫

�β

0dτ

∫d3r

⎧⎨⎩∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 − μ

]ψσ − g0 ψ↑ ψ↓ ψ↓ ψ↑

⎫⎬⎭

with g0 > 0, can be rewritten (Hubbard-Stratonovich transformation) as

Z = N∫

D[ψσ, ψσ]D[�,�∗] exp

{−1

�Se[ψσ, ψσ,�,�∗]

},

where

Se[ψσ, ψσ,�,�∗] =∫

�β

0dτ

∫d3r

{ ∑

σ=↑,↓ψσ

[�∂τ − �

2

2m∇2 − μ

]ψσ − �∗ ψ↓ ψ↑

−�ψ↑ ψ↓ + |�|2g0

},

�(r, τ ) is a complex scalar field, and N is a normalization constant that disappearswhen one calculates thermal (and quantum) averages.

Solution

First of all we remark that in the BCS Lagrangian of Eq. (9.38) the complex orderparameter � is constant and uniform, while here �(r, τ ) is complex field of bothspace and time. Moreover, simply by inspection, it is easy to realize that one mustindeed prove the following equation

exp

{−1

∫�β

0dτ

∫d3r

(−g0 ψ↑ ψ↓ ψ↓ ψ↑)}

= N∫

D[�,�∗] exp

{−1

∫�β

0dτ

∫d3r

(−�∗ ψ↓ ψ↑ − �ψ↑ ψ↓ + |�|2

g0

)}.

Page 232: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

226 9 Functional Integration for the Fermionic Field

Now we prove the simple relation

eg0 AA = N∫

dzdz∗

2πie(z∗A+z A− |z|2

g0).

In fact, we observe that

g0| zg0

− A|2 = |z|2g0

− z∗A − z A + g0 AA

and consequently

∫dzdz∗

2πie(z∗A+z A− |z|2

g0) = eg0 AA

∫dzdz∗

2πie−g0| z

g0−A|2 = eg0 AAg2

0

∫dζdζ∗

2πie−g0|ζ|2 = eg0 AA g0 ,

setting ζ = z/g0 − A. It follows that, in this case, the normalization constant isN = 1/g0.

More generally, considering complex vectors z = (z1, ..., zN ) and A = (A1, ...,

AN ) it is not difficult to prove that

eg0A·A = N∫

dz · dz∗

2πie(z∗·A+z·A− |z|2

g0)

with N = 1/gN0 . Finally, it is straightforward to extend this approach to complex

fields obtaining what is required.

Further Reading

For the functional integration of the non-relativistic bosonic and fermionic fields:N. Nagaosa, Quantum Field Theory in Condensed Matter Physics, Chap. 1, Sects.1.1 and 1.2 (Springer, Berlin, 1999).H.T.C. Stoof, K.B. Gubbels, and D.B.M. Dickerscheid, Ultracold Quantum Fields,Chap. 6, Sects. 6.1, 6.2, 6.3, and 6.4 (Springer, Berlin, 2009).A. Altland and B. Simons, Condensed Matter Field Theory, Chap. 2, Sect. 2.1 (Cam-bridge Univ. Press, Cambridge, 2006).For the dimensional regularization of Gaussian fluctuations in the case of interactingbosons and fermions:L. Salasnich and F. Toigo, Zero-point energy of ultracold atoms, Phys. Rep. 640, 1(2016).For the BCS theory in metals with various experimental data:N.W. Ashcroft and N.D. Mermin, Solid State Physics (Holt-Sounders, Philadelphia,1976).J.P. Carbotte, Rev. Mod. Phys. 62, 1027 (1990).For the condensate fraction of superfluid fermions:L. Salasnich, N. Manini, and A. Parola, Condensate Fraction of a Fermi Gas in theBCS-BEC Crossover, Phys. Rev. A 72, 023621 (2005).

Page 233: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

9.5 Solved Problems 227

For a detailed account of the Ginzburg-Landau theory and its developments:A. Larkin and A. Varlamov, Theory of Fluctuations in Superconductors (OxfordUniv. Press, Oxford, 2009).For a comprehensive review on the state of the art in the study of the BCS-BECcrossover:W. Zwerger (Ed.), The BCS-BEC Crossover and the Unitary Fermi Gas, LectureNotes in Physics (Springer, Berlin, 2012).

Page 234: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix ADirac Delta Function

In 1880 the self-taught electrical scientist Oliver Heaviside introduced the followingfunction

�(x) ={1 for x > 00 for x < 0

(A.1)

which is now called Heaviside step function. This is a discontinous function, witha discontinuity of first kind (jump) at x = 0, which is often used in the context ofthe analysis of electric signals. Moreover, it is important to stress that the Havisidestep function appears also in the context of quantum statistical physics. In fact, theFermi-Dirac function (or Fermi-Dirac distribution)

Fβ(x) = 1

eβ x + 1, (A.2)

proposed in 1926 by Enrico Fermi and Paul Dirac to describe the quantum statisticaldistribution of electrons in metals, where β = 1/(kBT ) is the inverse of the absolutetemperature T (with kB the Boltzmann constant) and x = ε − μ is the energy ε ofthe electron with respect to the chemical potential μ, becomes the function �(−x)in the limit of very small temperature T , namely

limβ→+∞

Fβ(x) = �(−x) ={0 for x > 01 for x < 0

. (A.3)

Inspired by the work of Heaviside, with the purpose of describing an extremelylocalized charge density, in 1930 Paul Dirac investigated the following “function”

δ(x) ={+∞ for x = 0

0 for x �= 0(A.4)

imposing that ∫ +∞

−∞δ(x) dx = 1 . (A.5)

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1

229

Page 235: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

230 Appendix A: Dirac Delta Function

Unfortunately, this property of δ(x) is not compatible with the definition (A.4).In fact, from Eq. (A.4) it follows that the integral must be equal to zero. In otherwords, it does not exist a function δ(x) which satisfies both Eqs. (A.4) and (A.5).Dirac suggested that a way to circumvent this problem is to interpret the integral ofEq. (A.5) as ∫ +∞

−∞δ(x) dx = lim

ε→0+

∫ +∞

−∞δε(x) dx , (A.6)

where δε(x) is a generic function of both x and ε such that

limε→0+

δε(x) ={+∞ for x = 0

0 for x �= 0, (A.7)

∫ +∞

−∞δε(x) dx = 1 . (A.8)

Thus, theDirac delta function δ(x) is a “generalized function” (but, strictly-speaking,not a function) which satisfy Eqs. (A.4) and (A.5) with the caveat that the integral inEq. (A.5) must be interpreted according to Eq. (A.6) where the functions δε(x) satisfyEqs. (A.7) and (A.8).

There are infinite functions δε(x)which satisfy Eqs. (A.7) and (A.8). Among themthere is, for instance, the following Gaussian

δε(x) = 1

ε√

πe−x2/ε2 , (A.9)

which clearly satisfies Eq. (A.7) and whose integral is equal to 1 for any value of ε.Another example is the function

δε(x) ={

1εfor |x | ≤ ε/2

0 for |x | > ε/2, (A.10)

which again satisfies Eq. (A.7) and whose integral is equal to 1 for any value of ε.In the following we shall use Eq. (A.10) to study the properties of the Dirac deltafunction.

According to the approach ofDirac, the integral involving δ(x)must be interpretedas the limit of the corresponding integral involving δε(x), namely

∫ +∞

−∞δ(x) f (x) dx = lim

ε→0+

∫ +∞

−∞δε(x) f (x) dx , (A.11)

for any function f (x). It is then easy to prove that

∫ +∞

−∞δ(x) f (x) dx = f (0) . (A.12)

Page 236: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix A: Dirac Delta Function 231

by using Eq. (A.10) and the mean value theorem. Similarly one finds

∫ +∞

−∞δ(x − c) f (x) dx = f (c) . (A.13)

Several other properties of the Dirac delta function δ(x) follow from its definition.In particular

δ(−x) = δ(x) , (A.14)

δ(a x) = 1

|a| δ(x) with a �= 0 , (A.15)

δ( f (x)) =∑i

1

| f ′(xi )| δ(x − xi ) with f (xi ) = 0 . (A.16)

Up to nowwe have considered theDirac delta function δ(x)with only one variablex . It is not difficult to define a Dirac delta function δ(D)(r) in the case of a D-dimensional domainRD , where r = (x1, x2, ..., xD) ∈ R

D is aD-dimensional vector:

δ(D)(r) ={+∞ for r = 0

0 for r �= 0(A.17)

and ∫RD

δ(D)(r) dDr = 1 . (A.18)

Notice that sometimes δ(D)(r) is written using the simpler notation δ(r). Clearly,also in this case one must interpret the integral of Eq. (A.18) as

∫RD

δ(D)(r) dDr = limε→0+

∫RD

δ(D)ε (r) dDr , (A.19)

where δ(D)ε (r) is a generic function of both r and ε such that

limε→0+

δ(D)ε (r) =

{+∞ for r = 00 for r �= 0

, (A.20)

limε→0+

∫δ(D)ε (r) dDr = 1 . (A.21)

Several properties of δ(x) remain valid also for δ(D)(r). Nevertheless, some prop-erties of δ(D)(r) depend on the space dimension D. For instance, one can prove theremarkable formula

δ(D)(r) ={

12π ∇2 (ln |r|) for D = 2

− 1D(D−2)VD

∇2(

1|r|D−2

)for D ≥ 3

, (A.22)

Page 237: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

232 Appendix A: Dirac Delta Function

where ∇2 = ∂2

∂x21+ ∂2

∂x22+ · · · + ∂2

∂x2Dand VD = πD/2/�(1 + D/2) is the volume of

a D-dimensional ipersphere of unitary radius, with �(x) the Euler Gamma function.In the case D = 3 the previous formula becomes

δ(3)(r) = − 1

4π∇2

(1

|r|)

, (A.23)

which can be used to transform the Gauss law of electromagnetism from its integralform to its differential form.

Page 238: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix BFourier Transform

It was known from the times of Archimedes that, in some cases, the infinite sumof decreasing numbers can produce a finite result. But it was only in 1593 that themathematician Francois Viete gave the first example of a function, f (x) = 1/(1 −x), written as the infinite sum of power functions. This function is nothing else thanthe geometric series, given by

1

1 − x=

∞∑n=0

xn , for |x | < 1 . (B.1)

In 1714 Brook Taylor suggested that any real function f (x) which is infinitelydifferentiable in x0 and sufficiently regular can be written as a series of powers, i.e.

f (x) =∞∑n=0

cn (x − x0)n , (B.2)

where the coefficients cn are given by

cn = 1

n! f(n)(x0) , (B.3)

with f (n)(x) the n-th derivative of the function f (x). The series (B.2) is now calledTaylor series and becomes the so-called Maclaurin series if x0 = 0. Clearly, thegeometric series (B.1) is nothing else than the Maclaurin series, where cn = 1. Weobserve that it is quite easy to prove the Taylor series: it is sufficient to suppose thatEq. (B.2) is valid and then to derive the coefficients cn by calculating the derivativesof f (x) at x = x0; in this way one gets Eq. (B.3).

In 1807 Jean Baptiste Joseph Fourier, who was interested on wave propagationand periodic phenomena, found that any sufficiently regular real function functionf (x) which is periodic, i.e. such that

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1

233

Page 239: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

234 Appendix B: Fourier Transform

f (x + L) = f (x) , (B.4)

where L is the periodicity, can be written as the infinite sum of sinusoidal functions,namely

f (x) = a02

+∞∑n=1

[an cos

(n2π

Lx

)+ bn sin

(n2π

Lx

)], (B.5)

where

an = 2

L

∫ L/2

−L/2f (y) cos

(n2π

Ly

)dy , (B.6)

bn = 2

L

∫ L/2

−L/2f (y) sin

(n2π

Ly

)dy . (B.7)

It is quite easy to prove also the series (B.5), which is now called Fourier series. Infact, it is sufficient to suppose that Eq. (B.5) is valid and then to derive the coeffi-cients an and bn by multiplying both side of Eq. (B.5) by cos

(n 2π

L x)and cos

(n 2π

L x)

respectively and integrating over one period L; in this way one gets Eqs. (B.6) and(B.7).

It is important to stress that, in general, the real variable x of the function f (x)can represent a space coordinate but also a time coordinate. In the former case Lgives the spatial periodicity and 2π/L is the wavenumber, while in the latter case Lis the time periodicity and 2π/L the angular frequency.

Taking into account the Euler formula

ein2πL x = cos

(n2π

Lx

)+ i sin

(n2π

Lx

)(B.8)

with i = √−1 the imaginary unit, Fourier observed that his series (B.5) can bere-written in the very elegant form

f (x) =+∞∑

n=−∞fn e

in 2πL x , (B.9)

where

fn = 1

L

∫ L/2

−L/2f (y) e−in 2π

L y dy (B.10)

are complex coefficients, with f0 = a0/2, fn = (an − ibn)/2 if n > 0 and fn =(a−n + ib−n)/2 if n < 0, thus f ∗

n = f−n .The complex representation (B.9) suggests that the function f (x) can be periodic

but complex, i.e. such that f : R → C. Moreover, one can consider the limit L →+∞ of infinite periodicity, i.e. a function which is not periodic. In this limit Eq. (B.9)

Page 240: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix B: Fourier Transform 235

becomes the so-called Fourier integral (or Fourier anti-transform)

f (x) = 1

∫ +∞

−∞f (k) eikx dk (B.11)

with

f (k) =∫ ∞

−∞f (y) e−iky dy (B.12)

the Fourier transform of f (x). To prove Eqs. (B.11) and (B.12) we write Eq. (B.9)taking into account Eq. (B.10) and we find

f (x) =+∞∑

n=−∞

(1

L

∫ L/2

−L/2f (y) e−in 2π

L y dy

)ein

2πL x . (B.13)

Setting

kn = n2π

Land �k = kn+1 − kn = 2π

L(B.14)

the previous expression of f (x) becomes

f (x) = 1

+∞∑n=−∞

(∫ L/2

−L/2f (y) e−ikn y dy

)eikn x �k . (B.15)

In the limit L → +∞ one has �k → dk, kn → k and consequently

f (x) = 1

∫ +∞

−∞

(∫ +∞

−∞f (y) e−iky dy

)eikx dk , (B.16)

which gives exactly Eqs. (B.11) and (B.12). Note, however, that one gets the sameresult (B.16) if the Fourier integral and its Fourier transform are defined multiplyingthem respectively with a generic constant and its inverse. Thus, we have found thatany sufficiently regular complex function f (x) of real variable x which is globallyintegrable, i.e. such that ∫ +∞

−∞| f (x)| dx < +∞ , (B.17)

can be considered as the (infinite) superposition of complex monocromatic waveseikx . The amplitude f (k) of the monocromatic wave eikx is the Fourier trans-form of f (x).

The Fourier transform f (k) of a function f (x) is sometimes denoted asF[ f (x)](k), namely

f (k) = F[ f (x)](k) =∫ ∞

−∞f (x) e−ikx dx . (B.18)

Page 241: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

236 Appendix B: Fourier Transform

Table B.1 Fouriertransforms F[ f (x)](k) ofsimple functions f (x), whereδ(x) is the Dirac deltafunction, sgn(x) is the signfunction, and �(x) is theHeaviside step function

f (x) F[ f (x)](k)0 0

1 2πδ(k)

δ(x) 1

�(x) 1ik + π δ(k)

eik0x 2π δ(k − k0)

e−x2/(2a2) a√2πe−a2k2/2

e−a|x | 2aa2+k2

sgn(x) 2ik

sin (k0x)πi [δ(k − k0) − δ(k + k0)]

cos (k0x) π [δ(k − k0) + δ(k + k0)]

The Fourier transformF[ f (x)](k) has many interesting properties. For instance, dueto the linearity of the integral the Fourier transform is clearly a linear map:

F[a f (x) + b g(x)](k) = aF[ f (x)](k) + bF[g(x)](k) . (B.19)

Moreover, one finds immediately that

F[ f (x − a)](k) = e−ika F[ f (x)](k) , (B.20)

F[eik0x f (x)](k) = F[ f (x)](k − k0) . (B.21)

F[x f (x)](k) = i f ′(k) , (B.22)

F[ f (n)(x)](k) = (ik)n f (k) , (B.23)

where f (n)(x) is the n-th derivative of f (x) with respect to x (TableB.1).In the Table we report the Fourier transforms F[ f (x)](k) of some elementary

functions f (x), including the Dirac delta function δ(x) and the Heaviside step func-tion �(x). We insert also the sign function sgn(x) defined as: sgn(x) = 1 for x > 0and sgn(x) = −1 for x < 0. The table of Fourier transforms clearly shows that theFourier transform localizes functions which is delocalized, while it delocalizes func-tions which are localized. In fact, the Fourier transform of a constant is a Diracdelta function while the Fourier transform of a Dirac delta function is a constant. Ingeneral, it holds the following uncertainty theorem

�x �k ≥ 1

2, (B.24)

where

�x2 =∫ ∞

−∞x2 | f (x)|2 dx −

(∫ ∞

−∞x | f (x)|2 dx

)2

(B.25)

Page 242: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix B: Fourier Transform 237

and

�k2 =∫ ∞

−∞k2 | f (k)|2 dk −

(∫ ∞

−∞k | f (k)|2 dk

)2

(B.26)

are the spreads of the wavepackets respectively in the space x and in the dual spacek. This theorem is nothing else than the uncertainty principle of quantum mechanicsformulated by Werner Heisenberg in 1927, where x is the position and k is thewavenumber. Another interesting and intuitive relationship is the Parseval identity,given by ∫ +∞

−∞| f (x)|2dx =

∫ +∞

−∞| f (k)|2dk . (B.27)

It is important to stress that the power series, the Fourier series, and the Fourierintegral are special cases of the quite general expansion

f (x) =∑∫

fα φα(x) dα (B.28)

of a generic function f (x) in terms of a set φα(x) of basis functions spanned by theparameterα, which can be a discrete or a continuous variable. A large part of modernmathematical analysis is devoted to the study of Eq. (B.28) and its generalization.

The Fourier transform is often used in electronics. In that field of research thesignal of amplitude f depends on time t , i.e. f = f (t). In this case the dual variableof time t is the frequency ω and the fourier integral is usually written as

f (t) = 1

∫ +∞

−∞f (ω) e−iωt dk (B.29)

with

f (ω) = F[ f (t)](ω) =∫ ∞

−∞f (t) eiωt dt (B.30)

the Fourier transform of f (t). Clearly, the function f (t) can be seen as the Fourieranti-transform of f (ω), in symbols

f (t) = F−1[ f (ω)](t) = F−1[F[ f (t)](ω)](t) , (B.31)

which obviously means that the composition F−1 ◦ F gives the identity.More generally, if the signal f depends on the 3 spatial coordinates r = (x, y, z)

and time t , i.e. f = f (r, t), one can introduce Fourier transforms from r to k, fromt to ω, or both. In this latter case one obviously obtains

f (r, t) = 1

(2π)4

∫R4

f (k,ω) ei(k·x−ωt) d3k dω (B.32)

Page 243: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

238 Appendix B: Fourier Transform

with

f (k,ω) = F[ f (r, t)](k,ω) =∫R4

f (r, t) e−i(k·r−ωt) d3r dt . (B.33)

Also in this general case the function f (r, t) can be seen as the Fourier anti-transformof f (k,ω), in symbols

f (r, t) = F−1[ f (k,ω)](r, t) = F−1[F[ f (r, t)](k,ω)](r, t) . (B.34)

Page 244: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix CLaplace Transform

The Laplace transform is an integral transformation, similar but not equal to theFourier transform, introduced in 1737 by Leonard Euler and independently in 1782by Pierre-Simon de Laplace. Nowaday the Laplace transform is mainly used to solvenon-homogeneous ordinary differential equations with constant coefficients.

Given a sufficiently regular function f (t) of time t , the Laplace transform of f (t)is the function F(s) such that

F(s) =∫ +∞

0f (t) e−st dt , (C.1)

where s is a complex number. Usually the integral converges if the real part Re(s) ofthe complex number s is greater than critical real number xc, which is called abscissaof convergence and unfortunately depends on f (t). The Laplace transform F(s) ofa function f (t) is sometimes denoted as L[ f (t)](s), namely

F(s) = L[ f (t)](s) =∫ ∞

0f (t) e−st dt . (C.2)

For the sake of completeness and clarity, we write also the Fourier transform f (ω),denoted as F[ f (t)](ω), of the same function

f (ω) = F[ f (t)](ω) =∫ ∞

−∞f (t) eiωt dt . (C.3)

First of all we notice that the Laplace transform depends on the behavior of f (t) fornon negative values of t , while the Fourier transform depends also on the behaviorof f (t) for negative values of t . This is however not a big problem, because we canset f (t) = 0 for t < 0 (or equivalently we can multiply f (t) by the Heaviside stepfunction �(t)), and then the Fourier transform becomes

© Springer International Publishing AG 2017L. Salasnich, Quantum Physics of Light and Matter, UNITEXT for Physics,DOI 10.1007/978-3-319-52998-1

239

Page 245: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

240 Appendix C: Laplace Transform

Table C.1 Laplacetransforms L[ f (t)](s) ofsimple functions f (t), whereδ(t) is the Dirac deltafunction and �(t) is theHeaviside step function, andτ > 0 and n positive integer

f (t) L[ f (t)](s)0 0

1 1s

δ(t − τ ) e−τs

�(t − τ ) e−τs

s

tn n!sn+1

e−at 1s+a

e−a|t | 2aa2−s2

sin (at) as2+a2

cos (at) ss2+a2

f (ω) = F[ f (t)](ω) =∫ ∞

0f (t) eiωt dt . (C.4)

Moreover, it is important to observe that, comparing Eq. (C.2) with Eq. (C.4), if bothF and L of f (t) exist, we obtain

F(s) = L[ f (t)](s) = F[ f (t)](is) = f (is) , (C.5)

or equivalently

f (ω) = F[ f (t)](ω) = L[ f (t)](−iω) = F(−iω) . (C.6)

Remember thatω is a real variable while s is a complex variable. Thus we have foundthat for a generic function f (t), such that f (t) = 0 for t < 0, the Laplace transformF(s) and the Fourier transform f (ω) are simply related to each other.

In the following Table we report the Laplace transforms L[ f (t)](s) of someelementary functions f (t), including the Dirac delta function δ(t) and the Heavisidestep function �(t), forgetting about the possible problems of regularity (TableC.1).

We now show that writing f (t) as the Fourier anti-transform of f (ω) one candeduce the formula of the Laplace anti-transform of F(s). In fact, one has

f (t) = F−1[ f (ω)](t) = 1

∫ ∞

−∞f (ω) e−iωt dω . (C.7)

Because f (ω) = F(−iω) one finds also

f (t) = 1

∫ ∞

−∞F(−iω) e−iωt dω . (C.8)

Page 246: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix C: Laplace Transform 241

Using s = −iω as integration variable, this integral representation becomes

f (t) = 1

2π i

∫ +i∞

−i∞F(s) est ds , (C.9)

where the integration is now a contour integral along any path γ in the complex planeof the variable s, which starts at s = −i∞ and ends at s = i∞. What we have foundis exactly the Laplace anti-transform of F(s), in symbols

f (t) = L−1[F(s)](t) = 1

2π i

∫ +i∞

−i∞F(s) est ds . (C.10)

Remember that this function is such that f (t) = 0 for t < 0. The fact that the functionf (t) is the Laplace anti-transform of F(s) can be symbolized by

f (t) = L−1[F(s)](t) = L−1[L[ f (t)](s)](t) , (C.11)

which means that the composition L−1 ◦ L gives the identity.The Laplace transform L[ f (t)](s) has many interesting properties. For instance,

due to the linearity of the integral the Laplace transform is clearly a linear map:

L[a f (t) + b g(t)](s) = a L[ f (t)](s) + bL[g(t)](s) . (C.12)

Moreover, one finds immediately that

L[ f (t − a)�(t − a)](s) = e−as L[ f (t)](s) , (C.13)

L[eat f (t)](s) = L[ f (t)](s − a) . (C.14)

For the solution of non-homogeneous ordinary differential equations with constantcoefficients, the most important property of the Laplace transform is the following

L[ f (n)(t)](s) = sn L[ f (t)](s) − sn−1 f (0) − sn−2 f (1)(0) − · · · − s f (n−2) − f (n−1)(0)(C.15)

where f (n)(t) is the n-th derivative of f (t) with respect to t . For instance, in thesimple cases n = 1 and n = 2 one has

L[ f ′(t)](s) = s F(s) − f (0) , (C.16)

L[ f ′′(t)](s) = s2 F(s) − s f ′(0) − f (0) , (C.17)

by using F(s) = L[ f (t)](s). The proof of Eq. (C.15) is straightforward performingintegration by parts. Let us prove, for instance, Eq. (C.16):

Page 247: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

242 Appendix C: Laplace Transform

L[ f ′(t)](s) =∫ +∞

0f ′(t) e−st dt = [

f (t) e−st]+∞0 −

∫ +∞

0f (t)

d

dt

(e−st

)dt

= − f (0) + s∫ +∞

0f (t) e−st dt = − f (0) + s L[ f (t)](s) (C.18)

= − f (0) + s F(s) .

Wenowgive a simple example of theLaplacemethod to solve ordinary differentialequations by considering the differential problem

f ′(t) + f (t) = 1 with f (0) = 2 . (C.19)

We apply the Laplace transform to both sides of the differential problem

L[ f ′(t) + f (t)](s) = L[1](s) (C.20)

obtaining

s F(s) − 2 + F(s) = 1

s. (C.21)

This is now an algebric problem with solution

F(s) = 1

s(s + 1)+ 2

s + 1= 1

s− 1

s + 1+ 2

s + 1= 1

s+ 1

s + 1. (C.22)

Finally, we apply the Laplace anti-transform

f (t) = L−1

[1

s+ 1

s + 1

](t) = L−1

[1

s

](t) + L−1

[1

s + 1

](t) . (C.23)

By using our Table of Laplace transforms we find immediately the solution

f (t) = 1 + e−t . (C.24)

The Laplace transform can be used to solve also integral equations. In fact, onefinds

L[∫ t

0f (y) dy

](s) = 1

sF(s) , (C.25)

and more generally

L[∫ t

−∞f (y) g(t − y) dy

](s) = I0

s+ F(s)G(s) , (C.26)

Page 248: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

Appendix C: Laplace Transform 243

where

I0 =∫ 0

−∞f (y) g(t − y) dy . (C.27)

Notice that the integral which appears in Eq. (C.25) is, by definition, the convolution( f ∗ g)(t) of two functions, i.e.

( f ∗ g)(t) =∫ t

0f (y) g(t − y) dy . (C.28)

Bibliography

1. F.T. Arecchi, Measurement of the statistical distribution of Gaussian and laser sources. Phys.Rev. Lett. 15, 912 (1965)

2. A. Altland, B. Simons, Condensed Matter Field Theory (Cambridge University Press, Cam-bridge, 2006)

3. V.B. Berestetskii, E.M. Lifshitz, L.P. Pitaevskii, Relativistic Quantum Theory, vol. 4 of Courseof Theoretical Physics (Pergamon Press, Oxford, 1971)

4. J.D. Bjorken, S.D. Drell, Relativistic Quantum Mechanics (McGraw-Hill, New York, 1964)5. B.H. Bransden, C.J. Joachain, Physics of Atoms and Molecules (Prentice Hall, Upper Saddle

River, 2003)6. G. Bressi, G. Carugno, R. Onofrio, G. Ruoso, Measurement of the Casimir force between

parallel metallic surfaces. Phys. Rev. Lett. 88, 041804 (2002)7. J.P. Carbotte, Rev. Mod. Phys. 62, 1027 (1990)8. C. Cohen-Tannoudji, B. Dui, F. Laloe, Quantum Machanics (Wiley, New York, 1991)9. P.A.M. Dirac, The Principles of Quantum Mechanics (Oxford University Press, Oxford, 1982)10. A.L. Fetter, J.D. Walecka, Quantum Theory of Many-Particle Systems (Dover Publications,

New York, 2003)11. N.W. Ashcroft, N.D. Mermin, Solid State Physics (Holt-Sounders, Philadelphia, 1976)12. K. Huang, Statistical Mechanics (Wiley, New York, 1987)13. S.K. Lamoreaux, Demonstration of the Casimir force in the 0.6 to 6 m range. Phys. Rev. Lett.

78, 5 (1997)14. A. Larkin, A. Varlamov, Theory of Fluctuations in Superconductors (Oxford University Press,

Oxford, 2009)15. M. Le Bellac, A Short Introduction to Quantum Information and Quantum Computation (Cam-

bridge University Press, Cambridge, 2006)16. U. Leonhardt,Measuring theQuantumState of Light (CambridgeUniversity Press, Cambridge,

1997)17. E. Lipparini, Modern Many-Particle Physics: Atomic Gases, Quantum Dots and Quantum

Fluids (World Scientific, Singapore, 2003)18. F. Mandl, G. Shaw, Quantum Field Theory (Wiley, New York, 1984)19. G. Mazzarella, L. Salasnich, A. Parola, F. Toigo, Coherence and entanglement in the ground-

state of a bosonic Josephson junction:from macroscopic Schrdinger cats to separable Fockstates. Phys. Rev. A 83, 053607 (2011)

20. N. Nagaosa, Quantum Field Theory in Condensed Matter Physics (Springer, Berlin, 1999)21. W. Rindler, Introduction to Special Relativity (Oxford University Press, Oxford, 1991)22. R.W. Robinett, Quantum Mechanics: Classical Results, Modern Systems, and Visualized Ex-

amples (Oxford University Press, Oxford, 2006)23. L. Salasnich, N. Manini, A. Parola, Condensate fraction of a Fermi gas in the BCS-BEC

crossover. Phys. Rev. A 72, 023621 (2005)

Page 249: Quantum physics of light and matter : photons, atoms, and strongly correlated systems

244 Appendix C: Laplace Transform

24. L. Salasnich, F. Toigo, Zero-point energy of ultracold atoms. Phys. Rep. 640, 1 (2016)25. A. Scott (ed.), Encyclopedia of Nonlinear Science (Taylor and Francis, New York, 2005)26. M.O. Scully, M.S. Zubairy, Quantum Optics (Cambridge University Press, Cambridge, 1997)27. H.T.C. Stoof,K.B.Gubbels,D.B.M.Dickerscheid,UltracoldQuantumFields (Springer,Berlin,

2009)28. W. Zwerger (ed.), The BCS-BEC Crossover and the Unitary Fermi Gas, Lecture Notes in

Physics (Springer, Berlin, 2012)