777
Lecture notes on field theory in condensed matter physics Christopher Mudry Condensed matter theory group, Paul Scherrer Insti- tute, Switzerland E-mail address : [email protected]

Lecture notes on field theory in condensed matter physics

  • Upload
    buimien

  • View
    424

  • Download
    30

Embed Size (px)

Citation preview

  • Lecture notes on field theory in

    condensed matter physics

    Christopher Mudry

    Condensed matter theory group, Paul Scherrer Insti-tute, Switzerland

    E-mail address : [email protected]

  • PREFACE 1

    Preface

    Reading the books from Baym, [1] Messiah, [2] and Dirac [3], whilean undergraduate student at ETHZ, taught me how enjoyable and use-ful it is to learn quantum mechanics from different perspectives. Thisimpression was reinforced as I got exposed to statistical physics and tothe diversity of approaches to it found in the books from Becker, [4]Callen, [5] Huang, [6] and Feynman. [7]

    My initiation to quantum field theory was different. In those days,there seemed to be two separate communities doing many-body physics.I had taken a proseminar from Prof. Klaus Hepp on the theory ofrenormalization, who had told me that the only book on quantumfield theory relevant to his class was that of Itzykson and Zuber. [8]Although I already had taken a proseminar on Fermi liquid theoryand had started reading Kittels Quantum theory of solid, [9] I hadnot realized the close connection between the many-body physics ap-plied to high-energy physics, statistical physics, and condensed matterphysics. At the time, the few books on quantum field theory for high-energy physics were obsessed with propagators, Feynman diagrams,causality and positivity, and how to make sense of ultraviolet diver-gences. The venerable books on many-body physics in condensed mat-ter physics were deceptive. [10]-[13] Although they claimed to do many-body physics, single-particle physics was soon enough resuscitating asa mean-field approximation or with particles with diminutive names(quasiparticles). Moreover, these books were full of approximationswith mysterious acronyms such as the random phase approximation,involving some magical circular logic.

    This cultural divide manifest in the books published prior to the80s was however in the process of disappearing. As I was movingto UIUC to start my PhD, the standard model had established itselfas the fundamental theory for particle physics and a steady supply ofbooks devoted to it were being published by the late 80s. The renor-malization group had been applied to explain asymptotic freedom ofquantum chromodynamics (QCD), to solve the Kondo problem, andhad a profound impact on statistical physics. Lattice gauge theory, anapproach to solve QCD in the strong coupling problem, was turninginto a discipline of its own bridging relativistic quantum field theory tostatistical physics. Algebraic topology, an arcane discipline of mathe-matics to most physicists, had shown its value to classify defects in thevacua of quantum field theories and the order parameters of symmetrybroken phases in condensed matter physics or statistical physics. Al-gebraic topology could explain the quantization of the quantized Halleffect and the existence of collective excitations obeying exchange sta-tistics evading the spin statistics Theorem. Integrable models fromstatistical physics were used to explain the low-energy properties of

  • 2

    spin chains, impurity models, and quasi-one-dimensional metals. Thesecond remarkable discovery of the 80s in condensed matter physicswas of course that of high Tc superconductivity, a class of materialsthat defy a solution using perturbation theory to this date.

    The first books that I am aware of aiming at overcoming the culturaldifferences between high-energy physics, statistical physics, and con-densed matter physics of the 60s are those of Polyakov, [14] Parisi, [15]and Negele and Orland [16], whose authors had made seminal contri-butions to the revolution brought upon by the application of the renor-malization group to theoretical physics during the late 60s and 70s.Another generation of authors came along in the 90s with the intent toexplain how the machinery of quantum field theory should be appliedto condensed matter physics. [17]-[22]

    Since the turn of the 21st century, concepts and techniques havebeen shared from condensed matter theory to string theory. The breadthof topics in condensed matter physics makes it impossible to cover allapplications of quantum field theory to condensed matter physics ina single book. Correspondingly, the number of books applying quan-tum field theory to condensed matter physics is steadily increasing andgetting more specialized. A student has now the luxury of picking hisfavorite book and taking advantage of a variety of view points.

    This book is the result of teaching the class Quantum field theoryin condensed matter physics at ETHZ. My aim was to demystify someof the condensed matter jargon used in seminars in condensed matterphysics for a student at the level of a master degree in physics fromETHZ. I also wanted a student attending my class to obtain a hands-on experience of concepts such as spontaneous symmetry breaking,mean-field theory, random phase approximation, screening, quantumfluctuations, renormalization group flows, critical points, phase tran-sitions driven by topological defects, bosonization, etc. Many bookson quantum field theory devote space to the machinery of quantumfield theory before solving problems with it. I wanted my teaching todo the reverse, i.e., to develop the needed methodology one problemat a time. I also did not want quantum field theory to become theprimary interest. It had to remain a tool to explain as economically aspossible fundamental principles of condensed matter physics. I am ofthe opinion that the most efficient technique for this purpose is to sys-tematically use the path integral representation of quantum mechanics.Path integrals are thus pervasive in this book. However, I assume nomore prior knowledge than familiarity with quantum mechanics, at thelevel of Bayms book say.

  • PREFACE 3

    The book is organized in two parts. The first part deals with bosons,the second with fermions. In condensed matter physics, this organi-zation principle is not as obvious as would be implied by the stan-dard model of high-energy physics. The fundamental boson of con-densed matter physics is the photon. The fundamental fermions ofcondensed matter physics are the electron and the proton, the chargedconstituents of the atoms from the periodic table. On the relevant en-ergy and length scales of condensed matter physics, these elementaryconstituents interact through the rules of quantum electrodynamics ata non-vanishing density of fermionic matter in the ground state. Thisis the main difference with quantum field theory aiming at explain-ing high-energy scattering experiments, for which the ground state(the vacuum before and after scattering) has a vanishing density of(fermionic) matter. This difference is of a fundamental nature. Theatomic nucleus has a much larger mass than the electrons orbitingaround it. In a material, positive charge is localized in position space onthe sites of a crystal at low temperatures. As a result, the fermionic na-ture of the ionic constituents becomes irrelevant. What matters greatlyhowever is that the normal modes of this crystal are phonons, collec-tive excitations obeying Bose-Einstein statistics. The same can happenwith electrons. They can localize in position space, in which case thematerial is called an insulator. Some localized electrons can still inter-act through their internal spin-1/2 degree of freedom. It is often thecase that the collective excitations resulting from the interactions be-tween the spins of localized electrons are collective excitations obeyingBose-Einstein statistics. They are called magnons. Electrons need notbe localized in a material, which is then called a metal. In a metal,the mobile electrons exchange photons with each other, they interactthrough the Coulomb interaction, they interact with the localized pos-itive charge of the crystal, a one-body potential for the electrons, theyinteract with the phonons, and they might interact with some of theelectrons that are localized around the crystalline sites. Solving thismany-body problem from the Schrodinger equation is and will be im-possible. The Hilbert space is simply too large. Instead, effective the-ories motivated by phenomenology and simplicity have been the breadand butter of theoretical condensed matter physics. In these models,the elementary local constituents might be bosons, fermions, or morecomplicated objects of which the simplest examples are quantum spindegrees of freedom. The partition of this book into a part devoted tobosons and a part devoted to fermions refers to the situations whensome of the low-energy collective excitations can be shown to obeythe Bose-Einstein or Fermi-Dirac statistics, respectively. Even then,we shall show that bosons and fermions emerging from some interact-ing models on a one-dimensional lattice are interchangeable under therule of bosonization. The four chapters on bosons cover phonons (as

  • 4

    a way to introduce a quantum field theory), superfluidity, restorationof a continuous symmetry by fluctuations at the lower critical dimen-sion, and the Kosterlitz-Thouless phase transition, respectively. Thefive chapters on fermions cover non-interacting fermions, the randomphase approximation in the jellium model, superconductivity, dissipa-tive Josephson junction (an example of dissipative quantum mechan-ics), and bosonization, respectively. Each chapter ends with a sectionin which material is presented as a sequence of exercises. Each chap-ter also comes with an appendix. Some appendices provide distractingintermediary steps. Most appendices contain learning material.

    The books of Naoto Nagaosa in Ref. [20] and Mike Stone in Ref. [21]have been very influential when preparing my lectures. I am indebtedto these authors for these inspiring books.

    Gipf-Oberfrick, July 2013 Christopher Mudry

  • ACKNOWLEDGMENTS 5

    Acknowledgments

    I must start thanking Donald E. Knuth for developing TeX. I type-set my master thesis and had I needed to do the same for a book, Iwould have never written one.

    I am grateful to my home institution, the Paul Scherrer Institut(PSI), in the persons of Kurt Clausen who has been supportive of thisendeavor and Joel Mesot who has been a steady and reliable advocate ofthe condensed matter theory group. Since 1999 to this date, I benefitedat PSI from a great colleague, Rudolf Morf.

    I am indebted to my mentors Eduardo Fradkin (my PhD adviser),Xiao-Gang Wen (my host at MIT), and Bertrand Halperin (my hostat Harvard) for shaping my taste in physics. I am also indebted to myfriends and long term collaborators Claudio Chamon, Akira Furusaki,Piet Brouwer, and Shinsei Ryu who have been so influential on myunderstanding of physics.

    I have had the good fortune of directing the thesis of three talentedstudents: Andreas Schnyder, Sebastian Guerrero, and Titus Neupert.They were all teaching assistants of my class and made important con-tributions to the exercises. Titus had also the kindness and patiencefor converting my figures into artworks.

    In the last six months Maurizio Storni has helped me polish mylecture notes into this book. He even shares my compulsive obsessionwith using TeX for baroque notation! It has been my privilege tobenefit from his dedicated and critical reading. Maurizio has been thefairy-godmother of Cinderella for my lecture notes. I only hope there isno midnight deadline. Of course, as convention dictates, all remainingembarrassing mistakes are my responsibility.

  • Contents

    Preface 1Acknowledgments 4

    Part 1. Bosons 1

    Chapter 1. The harmonic crystal 3Outline 31.1. Introduction 31.2. Classical one-dimensional crystal 31.3. Quantum one-dimensional crystal 101.4. Higher-dimensional generalizations 171.5. Problems 17

    Chapter 2. Bogoliubov theory of a dilute Bose gas 25Outline 252.1. Introduction 252.2. Second quantization for bosons 252.3. Bose-Einstein condensation and spontaneous symmetry

    breaking 292.4. Dilute Bose gas: Operator formalism at vanishing

    temperature 362.5. Dilute-Bose gas: Path-integral formalism at any

    temperature 432.6. Problems 57

    Chapter 3. Non-Linear-Sigma Models 65Outline 653.1. Introduction 653.2. Non-Linear-Sigma-Models (NLM) 653.3. Fixed point theories, engineering and scaling dimensions,

    irrelevant, marginal, and relevant interactions 783.4. General method of renormalization 893.5. Perturbative expansion of the two-point correlation

    function up to one loop for the two-dimensional O(N)NLM 90

    3.6. Callan-Symanzik equation obeyed by the spin-spincorrelator in the d = 2-dimensional O(N > 2) NLM 99

    3.7. Beta function in the d > 2-dimensional O(N > 2) NLM 109

    7

  • 8 CONTENTS

    3.8. Problems 119

    Chapter 4. Kosterlitz-Thouless transition 157Outline 1574.1. Introduction 1574.2. Classical two-dimensional XY model 1574.3. The Coulomb-gas representation of the classical 2dXY

    model 1684.4. Equivalence between the Coulomb gas and Sine-Gordon

    model 1694.5. Fugacity expansion of n-point functions in the Sine-Gordon

    model 1784.6. Kosterlitz renormalization-group equations 1844.7. Problems 193

    Part 2. Fermions 201

    Chapter 5. Non-interacting fermions 203Outline 2035.1. Introduction 2035.2. Second quantization for fermions 2035.3. The non-interacting jellium model 2075.4. Time-ordered Green functions 2265.5. Problems 240

    Chapter 6. Jellium model for electrons in a solid 251Outline 2516.1. Introduction 2516.2. Definition of the Coulomb gas in the Schrodinger picture 2516.3. Path-integral representation of the Coulomb gas 2556.4. The random-phase approximation 2586.5. Diagrammatic interpretation of the random-phase

    approximation 2636.6. Ground-state energy in the random-phase approximation 2666.7. Lindhard response function 2676.8. Random-phase approximation for a short-range interaction2826.9. Feedback effect on and by phonons 2846.10. Problems 286

    Chapter 7. Superconductivity in the mean-field and random-phase approximations 307

    Outline 3077.1. Pairing-order parameter 3077.2. Scaling of electronic interactions 3137.3. Time- and space-independent Landau-Ginzburg action 3217.4. Mean-field theory of superconductivity 3287.5. Nambu-Gorkov representation 332

  • CONTENTS 9

    7.6. Effective action for the pairing-order parameter 3347.7. Effective theory in the vicinity of T = 0 3357.8. Effective theory in the vicinity of T = Tc 3547.9. Problems 359

    Chapter 8. A single dissipative Josephson junction 367Outline 3678.1. Phenomenological model of a Josephson junction 3678.2. DC Josephson effect 3728.3. AC Josephson effect 3728.4. Dissipative Josephson junction 3738.5. Instantons in quantum mechanics 3848.6. The quantum-dissipative Josephson junction 4048.7. Duality in a dissipative Josephson junction 4098.8. Renormalization-group methods 4178.9. Conjectured phase diagram for a dissipative Josephson

    junction 4238.10. Problems 425

    Chapter 9. Abelian bosonization in two-dimensional space andtime 459

    Outline 4599.1. Introduction 4599.2. Abelian bosonization of the Thirring model 4619.3. Applications 4779.4. Problems 494

    Appendix A. The harmonic-oscillator algebra and its coherentstates 513

    A.1. The harmonic-oscillator algebra and its coherent states 513A.2. Path-integral representation of the anharmonic oscillator 517A.3. Higher dimensional generalizations 520

    Appendix B. Some Gaussian integrals 521B.1. Generating function 521B.2. Bose-Einstein distribution and the residue theorem 522

    Appendix C. Non-Linear-Sigma-Models (NLM) on Riemannianmanifolds 525

    C.1. Introduction 525C.2. A few preliminary definitions 525C.3. Definition of a NLM on a Riemannian manifold 528C.4. Classical equations of motion for NLM:

    Christoffel symbol and geodesics 529C.5. Riemann, Ricci, and scalar curvature tensors 531C.6. Normal coordinates and vielbeins for NLM 538C.7. How many couplings flow on a NLM? 557

  • 10 CONTENTS

    Appendix D. The Villain model 559

    Appendix E. Coherent states for fermions, Jordan-Wignerfermions, and linear-response theory 565

    E.1. Grassmann coherent states 565E.2. Path-integral representation for fermions 568E.3. Jordan-Wigner fermions 569E.4. The ground state energy and the single-particle

    time-ordered Green function 578E.5. Linear response 583

    Appendix F. Landau theory of Fermi liquids 599Introduction 599F.1. Adiabatic continuity 599F.2. Quasiparticles 601F.3. Topological stability of the Fermi surface 603F.4. Quasiparticles in the Landau theory of Fermi liquids as

    poles of the two-point Green function 612F.5. Breakdown of Landau Fermi liquid theory 612

    Appendix G. First-order phase transitions induced by thermalfluctuations 615

    Outline 615G.1. Landau-Ginzburg theory and the mean-field theory of

    continuous phase transitions 615G.2. Fluctuations induced by a local gauge symmetry 619G.3. Applications 626

    Appendix H. Useful identities 627H.1. Proof of Equation (8.75) 627

    Appendix I. Non-Abelian bosonization 635I.1. Introduction 635I.2. Minkowski versus Euclidean spaces 635I.3. Free massless Dirac fermions and the Wess-Zumino-Witten

    theory 637I.4. A quantum-mechanical example of a Wess-Zumino action 646I.5. Wess-Zumino action in (1 + 1)dimensional Minkowski

    space and time 650I.6. Equations of motion for the WZNW action 653I.7. One-loop RG flow for the WZNW theory 657I.8. The Polyakov-Wiegmann identity 659I.9. Integration of the anomaly in QCD2 660I.10. Bosonization of QCD2 for infinitely strong gauge coupling

    673

    Appendix. Bibliography 681

  • Part 1

    Bosons

  • CHAPTER 1

    The harmonic crystal

    Outline

    The classical equations of motion for a finite chain of atoms aresolved within the harmonic approximation. In the thermodynamiclimit, an approximate hydrodynamical description, i.e., a one-dimensionalclassical field theory, is obtained. Quantization of the finite harmonicchain is undertaken. In the thermodynamic limit, phonons in a one-dimensional lattice are approximated by a quantum hydrodynamicaltheory, i.e., a one-dimensional quantum field theory.

    1.1. Introduction

    To illustrate the transition from the one-body to the many-bodyphysics, the harmonic excitations of a crystal are derived classicallyand quantum mechanically. The thematic of crystallization, i.e., ofspontaneous-symmetry breaking of translation symmetry in positionspace, is addressed in section 1.5 from the point of view of an applica-tion of the Mermin-Wagner theorem.

    1.2. Classical one-dimensional crystal

    1.2.1. Discrete limit. For simplicity, we shall consider a one-dimensional world made of N point-like objects (atoms) of mass mand interacting through a potential V . We assume first that the po-tential V depends only on the coordinates n R, n = 1, , N , ofthe N atoms,

    V = V (1, , N). (1.1)Furthermore, we assume that V has a non-degenerate minimum at

    n = n a, n = 1, , N, (1.2)

    where a is the lattice spacing. For example,

    V (1, , N) =( a

    2

    )2

    N1n=1

    [1 cos

    (2

    a

    (n+1 n

    ))]

    +( a

    2

    )2m2

    Nn=1

    [1 cos

    (2

    an

    )]+ boundary terms.

    (1.3)

    3

  • 4 1. THE HARMONIC CRYSTAL

    The physical interpretation of the real-valued parameters and isobtained as follows. For small deviations n about minimum (1.2), itis natural to expand the potential energy according to

    V (1 + 1, , N + N ) =V (1, , N ) +N1n=1

    2

    (n+1 n

    )2+

    1

    2m2

    Nn=1

    (n)2 + + boundary terms.

    (1.4)

    The dimensionful constant is the elastic or spring constant. It mea-sures the strength of the linear restoring force between nearest-neighboratoms. The characteristic frequency measures the strength of an ex-ternal force that pins atoms to their equilibrium positions (1.2). Toput it differently, m2 is the curvature of the potential well that pinsan atom to its equilibrium position. Terms that have been neglected in are of several kinds. Only terms of quadratic order in the nearest-neighbor relative displacement n+1n have been accounted for, andall interactions beyond the nearest-neighbor range have been dropped.We have also omitted to spell out what the boundary terms are. Theyare specified once boundary conditions have been imposed. In the limitN , the choice of boundary conditions should be immaterial sincethe bulk potential energy should be of order L N a, whereas theenergy contribution arising from boundary terms should be of orderL0 = 1.

    To minimize boundary effects in a finite system, one often imposesperiodic boundary conditions

    n+N = n, n = 1, , N. (1.5)

    An open chain of atoms turns into a ring after imposing periodic bound-ary conditions. Furthermore, imposing periodic boundary conditionsendows the potential with new symmetries within the harmonic ap-proximation defined by 1

    Vhar(1+1, , N+N) :=Nn=1

    [

    2

    (n+1 n

    )2+

    1

    2m2 (n)

    2

    ].

    (1.7)First, changing labels according to

    n n+m, n = 1, , N, m Z, (1.8)

    1 Without loss of generality, we have set the classical minimum of the potentialenergy to zero,

    V (1, , N ) = 0. (1.6)

  • 1.2. CLASSICAL ONE-DIMENSIONAL CRYSTAL 5

    leaves Eq. (1.7) invariant. Second, translation invariance is recoveredin the absence of the pinning potential,

    Eq. (1.7) with = 0 =Vhar(1, , N) = Vhar(1 + x a, , N + x a)

    (1.9)

    for any real-valued x.The kinetic energy of an open chain of atoms is simply given by

    T (1 + 1, , N + N) =1

    2m

    Nn=1

    (dndt

    )2 1

    2m

    Nn=1

    (n

    )2.

    (1.10)As was the case for the potential energy, the choice of boundary con-ditions only affects the kinetic energy by terms of order L0. It is thusnatural to choose periodic boundary conditions if one is interested inextensive properties of the system.

    The classical Lagrangian L in the harmonic approximation and withperiodic boundary conditions is defined by subtracting from the kineticenergy (1.10) the potential energy (1.7),

    L :=Nn=1

    1

    2

    [m(

    n

    )2

    (n+1 n

    )2 m2 (n)2] . (1.11)The classical equations of motion follow from Euler-Lagrange equationsof motion

    d

    dt

    ( L

    n

    )=

    ( L

    n

    ), n = 1, , N. (1.12)

    They are

    m n = (n+1 + n1 2n

    )m2n, n = 1, , N,

    (1.13)with the complex-valued and traveling-wave solutions

    n(t) ei(knt), 2 = 2

    m(1 cos k) + 2. (1.14)

    Imposing periodic boundary conditions allows to identify the normalmodes. These are countably-many traveling waves with the frequencyto wave-number relation

    l =

    2

    m(1 cos kl) + 2, kl =

    2

    Nl, l = 1, , N. (1.15)

    The most general real-valued solution of Euler-Lagrange equations (1.13)obeying periodic boundary conditions is

    n(t) =Nl=1

    [Al e

    +i(klnlt) + Al ei(klnlt)

    ], n = 1, , N.

    (1.16)

  • 6 1. THE HARMONIC CRYSTAL

    Here, the complex-valued expansion coefficient Al remains arbitrary as

    long as initial conditions on n andn have not been specified.

    To revert to the Hamilton-Jacobi formalism of classical mechanics,one introduces the canonical momentum n conjugate to n through

    n(t) :=L

    n

    = imNl=1

    l[Al e

    +i(klnlt) Al ei(klnlt)], n = 1, , N,

    (1.17)

    and construct the Hamiltonian

    H =Nn=1

    1

    2

    [(n)

    2

    m+

    (n+1 n

    )2+m2(n)

    2

    ](1.18)

    from the Lagrangian (1.11) through a Legendre transformation. Hamilton-Jacobi equations of motion are then

    n = +H

    n= {n,H}, n =

    H

    n= {n,H}, n = 1, , N,

    (1.19)where {, } stands for the Poisson brackets. 2

    In the long wave-number limit kl 1, the dispersion relation re-duces to

    2l =

    mk2l +

    2 +O(k4l ). (1.21)

    The pinning potential characterized by the curvature of the potentialwell has opened a gap in the spectrum of normal modes. No solutionsto Euler-Lagrange equations (1.13) can be found below the character-istic frequency . By switching off the pinning potential, = 0, thedispersion relation simplifies to

    2l =

    mk2l +O(k4l ). (1.22)

    The proportionality constant/m between frequency and wave num-

    ber is interpreted as the velocity of propagation of a sound wave in theone-dimensional harmonic chain in units for which the lattice spacinga has been set to unity.

    2 The Poisson bracket {f, g} of two functions f and g of the canonical variablesn and n is defined by

    {f, g} :=Nn=1

    (f

    n

    g

    n fn

    g

    n

    ). (1.20)

  • 1.2. CLASSICAL ONE-DIMENSIONAL CRYSTAL 7

    1.2.2. Thermodynamic limit. The thermodynamic limit N emerges naturally if one is interested in the response of a one-dimensional solid to external perturbations as can be induced, say,by compressions. If the characteristic wavelength of a perturbationapplied to a solid is much larger than the atomic separation, then the(elastic) response from this solid to this perturbation is dominated bynormal modes with arbitrarily small wave numbers k 0. If so, itis then much more economical not to account for the discrete natureof this solid as is done in the Lagrangian (1.11) when computing the(elastic) response. To this end, Eq. (1.11) is first rewritten as

    L =Nn=1

    a1

    2

    [m

    a

    (n

    )2 a

    (n+1 n

    a

    )2 m

    a2 (n)

    2

    ]

    =:Nn=1

    aLn.

    (1.23)

    We interpret

    :=m

    a, Y := a, :=

    n+1 na

    , and Ln, (1.24)

    as the mass per unit length, the Youngs modulus, 3 the elongation perunit length, and the local Lagrangian per unit length (the Lagrangiandensity) , respectively. Then, we write

    L =

    L0

    dx1

    2

    [

    (

    t

    )2 Y

    (

    x

    )2 22

    ]

    =:

    L0

    dxL,

    (1.26)

    whereby the following substitutions have been performed.[1] The discrete sum

    n has been replaced by the integral

    dx/a

    over the semi-open interval ]0, L].[2] The relative displacement n at time t has been replaced by the

    value of the real-valued function at the space-time coordinate (x, t)obeying periodic boundary conditions in position space,

    (x+ L, t) = (x, t), x ]0, L], t R. (1.27)

    3 For an elastic rode obeying Hookes law, the extension of the rode per unitlength is proportional to the exerted force F with the Youngs modulus Y as theproportionality constant,

    F = Y . (1.25)

  • 8 1. THE HARMONIC CRYSTAL

    [3] The time derivative of the relative displacement n at timet has been replaced by the value of the time derivative (t) at thespace-time coordinate (x, t).

    [4] The discrete difference n+1 n at time t has been replacedby the lattice constant times the value of the derivative (x) at thespace-time coordinate (x, t).

    [5] The integrand L in Eq. (1.26) is called the Lagrangian density.It is a real-valued function of space and time. From it, one obtains thecontinuum limit of Euler-Lagrange equations (1.12) according to

    tL(x, t)

    (t)(y, t)+ x

    L(x, t)(x)(y, t)

    =L(x, t)(y, t)

    . (1.28)

    Here, the symbol L(x, t) is to be interpreted as the infinitesimal func-tional change of L at the given space-time coordinates (x, t) inducedby the Taylor expansion

    L =L[+ , (x) + (x), (t) + (t)] L[, (x), (t)]

    =L

    +L

    (x)(x) +

    L(t)

    (t) + .

    (1.29)

    One must keep in mind that , (x), and (t) are independent vari-ables. Moreover, one must use the rule

    (x, t)

    (y, t)= (x y) =

    L0

    dx(x, t)

    (y, t)= 1, y ]0, L], (1.30)

    that extends the rule

    mn

    = m,n =Nm=1

    mn

    = 1, n = 1, , N, (1.31)

    to the continuum. Otherwise, all the usual rules of differentiation applyto /.

    [6] Equations of motion (1.13) become the one-dimensional soundwave equation

    (2t v22x + 2

    ) = 0, v :=

    Y

    , (1.32)

    after replacing the finite difference

    n+1 + n1 2n = +(n+1 n

    )(n n1

    )(1.33)

    by a2 times the value of the second-order space derivative (2x) at thespace-time coordinate (x, t).

    The Hamiltonian H in the continuum limit follows from Eq. (1.26)with the help of a (functional) Legendre transform or directly from the

  • 1.2. CLASSICAL ONE-DIMENSIONAL CRYSTAL 9

    continuum limit of Eq. (1.18),

    H =

    L0

    dx1

    2

    [2

    + Y

    (

    x

    )2+ 22

    ]

    =:

    L0

    dxH,

    (1.34a)

    where the field is the canonically conjugate to ,

    (x, t) :=

    L0

    dyL(y, t)

    (t)(x, t)= (t)(x, t). (1.34b)

    Probing the one-dimensional harmonic crystal on length scales muchlarger than the lattice spacing a blurs our vision to the point where thecrystal appears as an elastic continuum. Viewed without an atomicmicroscope, the relative displacements n, n = 1, , N , become afield (x, t) where x can be any real-valued number provided N is suf-ficiently large. 4

    The mathematics that justifies this blurring or coarse graining isthat, for functions f that vary slowly on the lattice scale,

    n

    f(n a)

    dx

    af(x). (1.35)

    In particular,

    f(m a) =n

    m,nf(n a) =n

    am,naf(n a) f(x) =

    dy (xy)f(y),

    (1.36)justifies the identification

    m,na (x y). (1.37)

    Equation (1.37) tells us that the divergent quantity (x = 0) in positionspace should be thought of as the inverse, 1/a, of the lattice spacing a.In turn, the number 1/a can be interpreted as the spacing of normalmodes in reciprocal space per unit volume 2/N in wave-number spaceby the following argument,

    1

    a=kl+1 kl

    a 1

    2/N, kl :=

    2

    Nl. (1.38)

    4 In mathematics, a (real-valued scalar) field is a mapping : Rd+1 R, (r, t) 7 (r, t). In physics, a field is often abbreviated by the value (r, t) ittakes at the point (r, t) in (d+ 1)-dimensional (position) space and time.

  • 10 1. THE HARMONIC CRYSTAL

    How does one go from a discrete Fourier sum to a Fourier integral?Start from an even number N of sites for which

    Nl=1

    eikl(mn) = Nm,n, kl :=2

    Nl. (1.39)

    Multiply both sides of this equation by the inverse of the system sizeL = N a,

    1

    L

    Nl=1

    eikl(mn) =m,na. (1.40)

    Since the right-hand side should be identified with (x y) in thethermodynamic limit N , the left-hand side should be identifiedwith

    1

    L

    Nl=1

    eikla

    (mn) a 2/a0

    dk

    2eik(xy)

    +

    dk

    2eik(xy), (1.41)

    wherebykla k, (m n) a x y. (1.42)

    To see this, recall first that the periodic boundary conditions tell us thatl = 1, , N could have equally well be chosen to run betweenN/2+1and +N/2. Hence, it is permissible to adopt the more symmetrical rule

    1

    L

    Nl=1

    f(kl) +/a/a

    dk

    2f(k) (1.43)

    to convert a finite summation over wave numbers into an integral overthe first Brillouin zone (reciprocal space) ]/a,+/a] as the thermo-dynamic limit N = L/a is taken. Now, if f(x) is a slowly varyingfunction on the lattice scale a, its Fourier transform f(k) will be essen-tially vanishing for |k| 1/a. In this case, the limits /a can safelybe replaced by the limits on the right-hand side of Eq. (1.43). Wethen arrive at the desired integral representation of the delta functionin position space,

    (x y) =+

    dk

    2eik(xy). (1.44)

    Observe that factors of 2 appear in an asymmetrical way in in-tegrals over position (x) and reciprocal (k) spaces. Although this ispurely a matter of convention when defining the Fourier transform,there is a physical reasoning behind this choice. Indeed, Eq. (1.43) im-plies that dk/(2) has the physical meaning of the number of normalmodes in reciprocal space with wave number between k and k + dk

  • 1.3. QUANTUM ONE-DIMENSIONAL CRYSTAL 11

    per unit volume L in position space. Correspondingly, the divergentquantity 2 (k = 0) in reciprocal space has the physical meaning ofbeing the divergent volume L of the system as is inferred from

    (x) =

    +

    dk

    2eikx 2(k) =

    +

    dx eikx. (1.45)

    1.3. Quantum one-dimensional crystal

    1.3.1. Reminiscences about the harmonic oscillator. We nowturn to the task of giving a quantum-mechanical description for a non-dissipative one-dimensional harmonic crystal. One possible route con-sists in the construction of a Hilbert space with the operators actingon it and whose expectation values can be related to measurable prop-erties of the crystal. 5 In this setting, the time evolution of physicalquantities can be calculated either in the Schrodinger or in the Heisen-berg picture. We will begin by reviewing these two approaches in thecontext of a single harmonic oscillator. The extension to the harmoniccrystal will then follow in a very natural way.

    The classical Hamiltonian that describes a single particle of unitmass m = 1 confined to a quadratic well with curvature 2 is

    H :=1

    2

    (p2 + 2x2

    ). (1.46)

    Hamilton-Jacobi equations of motion are

    dx(t)

    dt= {x,H} = +H

    p= p(t),

    dp(t)

    dt= {p,H} = H

    x= 2x(t).

    (1.47)

    Solutions to these classical equations of motion are

    x(t) = A cos(t) +B sin(t),

    p(t) = [A sin(t) +B cos(t)] .(1.48)

    The energy E of the particle is a constant of the motion that dependson the choice of initial conditions through the two real-valued constantsA and B,

    E =1

    2

    (A2 +B2

    )2. (1.49)

    In the Schrodinger picture of quantum mechanics, the position x ofthe particle and its canonical conjugate p become operators x and p that(i) act on the Hilbert space of twice-differentiable and square-integrablefunctions : R C and (ii) obey the canonical commutation relation

    [x, p] := x p p x = i~. (1.50)5 Another route to quantization is by means of the path-integral representation

    of quantum mechanics as is shown in appendix A.

  • 12 1. THE HARMONIC CRYSTAL

    The time evolution (or dynamics in short) of the system is encoded bySchrodinger equation

    i~ t(x, t) = H (x, t), (1.51a)

    where the quantum Hamiltonian H is given by

    H =1

    2

    (p2 + 2x2

    ). (1.51b)

    The time evolution of the wave function (x, t) is unique once initialconditions (x, t = 0) are given.

    Solving the time-independent eigenvalue problem

    H n(x) = nn(x) (1.52a)

    is tantamount to solving the time-dependent Schrodinger equation throughthe Ansatz

    (x, t) =n

    cn n(x) eint/~. (1.52b)

    The expansion coefficients cn C are time independent and uniquelydetermined by the initial condition, say (x, t = 0).

    As is well known, the energy eigenvalues n are given by

    n =

    (n+

    1

    2

    )~, n = 0, 1, 2, . (1.53)

    The energy eigenfunctions n(x) are Hermite polynomials multiplyinga Gaussian,

    0(x) =( ~

    )1/4e

    12~ x

    2

    ,

    1(x) =

    [4

    (~

    )3]1/4xe

    12~ x

    2

    ,

    2(x) =(

    4~

    )1/4 (2

    ~x2 1

    )e

    12~ x

    2

    ,

    ...

    n(x) =

    [1

    2nn!

    (~

    )n]1/2 ( ~

    )1/4(~x d

    dx

    )ne

    12~ x

    2

    .

    (1.54)

    The Heisenberg picture of quantum mechanics is better suited thanthe Schrodinger picture to a generalization to quantum field theory. Inthe Heisenberg picture, and contrary to the Schrodinger picture, oper-ators are explicitly time dependent. For any operator O, the solutionto the equation of motion 6

    i~dO(t)

    dt= [O(t), H] (1.55a)

    6 The assumption that the system is non-dissipative has been used here in thatH does not depend explicitly on time, tH = 0.

  • 1.3. QUANTUM ONE-DIMENSIONAL CRYSTAL 13

    that replaces Schrodinger equation is

    O(t) = e+iHt~ O(t = 0) ei

    Ht~ . (1.55b)

    By definition, the algebra obeyed by operators in the Schrodinger pic-ture holds true in the Heisenberg picture provided operators are takenat equal time. For example,

    x(t) := e+iHt~ x(t = 0) ei

    Ht~ , p(t) := e+i

    Ht~ p(t = 0) ei

    Ht~ , (1.56)

    obey by construction the equal-time commutator

    [x(t), p(t)] = i~, t R. (1.57)Finding the commutator of x(t) and p(t) at unequal times t 6= t re-quires solving the dynamics of the system, i.e., Eq. (1.55a) with Osubstituted for x and p, respectively,

    dx(t)

    dt= +p(t),

    dp(t)

    dt= 2x(t). (1.58)

    In other words, the Heisenberg operators x(t) and p(t) satisfy the sameequations of motion as the classical variables they replace,

    d2x(t)

    dt2+ 2x(t) = 0, p(t) =

    dx(t)

    dt. (1.59)

    The solution (1.48) can thus be borrowed with the caveat that A and

    B should be replaced by time-independent operators A and B.At this stage, it is more productive to depart from following a strat-

    egy dictated by the real-valued classical solution (1.48). The key ob-servation is that the quantum Hamiltonian for the harmonic oscillatortakes the quadratic form 7

    H = ~[a(t) a(t) +

    1

    2

    ], (1.60a)

    if the pair of canonically conjugate Hermitean operators x(t) and p(t)is traded for the pair a(t) and a(t) of operators defined by

    x(t) =:

    ~

    2

    [a(t) + a(t)

    ], p(t) =:

    ~

    2

    [ia(t) + ia(t)

    ].

    (1.60b)Once the equal-time commutator [a(t), a(t)] is known, the Heisenbergequations of motion are easily derived from

    i~da(t)

    dt= [a(t), H], i~

    da(t)

    dt= [a(t), H]. (1.61)

    With the help of

    a(t) =

    2~

    [x(t) i p(t)

    ], a(t) =

    2~

    [x(t) + i

    p(t)

    ],

    (1.62a)

    7 We are anticipating that H does not depend explicitly on time.

  • 14 1. THE HARMONIC CRYSTAL

    one verifies that

    [x(t), p(t)] = i~, [x(t), x(t)] = [p(t), p(t)] = 0[a(t), a(t)] = 1, [a(t), a(t)] = [a(t), a(t)] = 0.

    (1.62b)

    The change of Hermitean operator-valued variables to non-Hermiteanoperator-valued variables is advantageous in that the equations of mo-tion for a(t) and a(t) decouple according to

    da(t)

    dt= +i a(t), a(t) = a(t = 0) e+it,

    da(t)

    dt= i a(t), a(t) = a(t = 0) eit,

    (1.63)

    respectively. Below, we will write a for a(t = 0) and similarly for a.

    The time evolution of x(t), p(t), and H is now explicitly given by

    x(t) =

    ~

    2

    (a eit + a e+it

    ), (1.64a)

    p(t) = i

    ~2

    (a eit a e+it

    ), (1.64b)

    H = ~(aa+

    1

    2

    ). (1.64c)

    As must be by the absence of dissipation, H is explicitly time indepen-dent, tH = 0.

    The Hilbert space can now be constructed explicitly with purelyalgebraic methods. The Hilbert space is defined by all possible linearcombinations of the eigenstates

    |n :=(a)n

    n!|0, H|n = n|n, n = 0, 1, 2, . (1.65)

    Here, the ground state or vacuum |0 is defined by the condition

    a|0 = 0. (1.66)

    One verifies that 0(x) in Eq. (1.54) uniquely (up to a phase) satisfiesEq. (1.66) by using the position-space representation of the operator a.

    1.3.2. Discrete limit. In the spirit of the Heisenberg picture forthe harmonic oscillator and guided by the Fourier expansions in Eqs. (1.16)

  • 1.3. QUANTUM ONE-DIMENSIONAL CRYSTAL 15

    and (1.17), we begin by defining the operators

    n(t) :=1N

    Nl=1

    ~

    2l

    [al e

    +i(klnlt) + al ei(klnlt)

    ], n = 1, , N,

    n(t) := i1N

    Nl=1

    ~l

    2

    [al e

    +i(klnlt) al ei(klnlt)

    ], n = 1, , N,

    (1.67a)

    where the frequency l and the integer label l are related by Eq. (1.15),i.e., (remember that we have chosen units in which the mass is givenby m = 1)

    l =

    2 (1 cos kl) + 2, kl :=2

    Nl, l = 1, , N, (1.67b)

    and the operator-valued expansion coefficients al and al obey the har-monic oscillator algebra

    [al, al ] = l,l , [al, al ] = [a

    l , al ] = 0, l, l

    = 1, , N. (1.67c)

    The normalization factor 1/N is needed to cancel the factor of N

    present in the Fourier series

    Nl=1

    eikl(mn) = Nm,n (1.68)

    that shows up when one verifies that the equal-time commutators

    [m(t), n(t)] = i~ m,n, [m(t), n(t)] = [m(t), n(t)] = 0, m, n = 1, , N,(1.69)

    hold for all times. We are now ready to define in a consistent waythe Hamiltonian H for the quantum one-dimensional harmonic crystal[compare with Eq. (1.18)]

    H :=Nn=1

    1

    2

    {[n(t)]

    2 + [n+1(t) n(t)

    ]2+ 2 [n(t)]

    2}. (1.70)

    With the help of the algebra (1.67c), one verifies that H is explicitlytime independent and given by

    H =Nl=1

    ~l(al al +

    1

    2

    ). (1.71)

    The next task is to construct the Hilbert space for the one-dimensionalquantum crystal by algebraic methods. Assume that there exists aunique (up to a phase) normalized state |0, the ground state or vac-uum, defined by

    0|0 = 1, al|0 = 0, l = 1, , N. (1.72)

  • 16 1. THE HARMONIC CRYSTAL

    If so, the state

    |n1, n2, , nN :=Nl=1

    1nl!

    (al

    )nl|0, n1, n2, , nN = 0, 1, 2, ,(1.73)

    is normalized to one and is an eigenstate of H with the energy eigen-value

    n1, ,nN :=Nl=1

    ~l(nl +

    1

    2

    ). (1.74)

    The ground-state energy is of order N and given by

    0, ,0 :=1

    2

    Nl=1

    ~l. (1.75)

    Excited states have at least one nl > 0. They are called phonons. Theeigenstate |n1, n2, , nN is said to have n1 phonons in the first mode,n2 phonons in the second mode, and so on. Phonons can be thoughtof as identical elementary particles since they possess a definite energyand momentum. Because the phonon occupation number

    nl = n1, , nl, , nN |al al|n1, , nl, , nN (1.76)

    is an arbitrary positive integer, phonons obey Bose-Einstein statistics.Upon switching on a suitable interaction [say by including cubic andquartic terms in the expansion (1.4)], phonons scatter off one other justas other -ons (mesons, photons, gluons, and so on) known to physicsdo. Although we are en route towards constructing the quantum field(x, t) out of n(t), we have encountered particles. The duality betweenquantum fields and particles is the essence of quantum field theory.

    The vector space spanned by the states labeled by the phonon oc-cupation numbers (n1, , nN) {0, 1, 2, }N in Eq. (1.73) is theHilbert space of the one-dimensional quantum crystal. The mathemat-ical structure of this Hilbert space is a symmetric tensor product of Ncopies of the Hilbert space for the harmonic oscillator. In physics, thissymmetric tensor product is called a Fock space when the emphasis isto think of the phonon as an elementary particle.

    1.3.3. Thermodynamic limit. Taking the thermodynamic limitN is a direct application to section 1.3.2 of the rules establishedin the context of the classical description of sections 1.2.2 Hence, with

  • 1.3. QUANTUM ONE-DIMENSIONAL CRYSTAL 17

    the identifications 8

    Nn=1

    a

    dx,

    1

    N a

    Nl=1

    dk

    2,

    l (k) =

    2v2

    a2[1 cos(k a)] + 2

    v2k2 + 2, if |k a| 1,

    kln kx,

    al 1N a

    a(k),

    n(t) a (x, t),

    n(t) a (x, t),

    (1.77)

    the canonically conjugate pairs of operators n(t) and n(t) are replacedby the quantum fields

    (x, t) :=

    dk

    2

    ~

    2(k)

    {a(k) e+i[kx(k)t] + a(k) ei[kx(k)t]

    },

    (x, t) := i

    dk

    2

    ~(k)

    2

    {a(k) e+i[kx(k)t] a(k) ei[kx(k)t]

    },

    (1.78)

    respectively. 9 Their equal-time commutators follow from the harmonicoscillator algebra

    [a(k), a(k)] = 2(k k), [a(k), a(k)] = [a(k), a(k)] = 0.(1.79)

    They are

    [(x, t), (y, t)] = i~ (xy), [(x, t), (y, t)] = [(x, t), (y, t)] = 0.(1.80)

    8 Limits of integrations in position and reciprocal spaces are left unspecifiedat this stage as we want to remain free to choose how the thermodynamic limitN is to be taken. For example, we could keep a finite in which case thethermodynamic limit N implies L . Alternatively, we could keep Lfinite in which case the thermodynamic limit N implies a 0.

    9 The substitution rules al 1N a a(k), n(t) a (x, t), and n(t)

    a (x, t), are needed to cancel the volume factor N a inNl=1 N a

    dk2 .

  • 18 1. THE HARMONIC CRYSTAL

    The Hamiltonian is

    H =

    dx

    1

    2

    {[(x, t)]2 + v2 [x(x, t)]

    2 + 2 [(x, t)]2}

    =

    dk

    2

    1

    2~(k)

    [a(k)a(k) + a(k)a(k)

    ].

    (1.81)

    The excitation spectrum is obtained by making use of the commu-tator between a(k) and a(k). It is given by

    H E0 :=

    dk

    2~(k) a(k) a(k) (1.82)

    and is observed to vanish for the vacuum |0. The operation of sub-tracting from the Hamiltonian the ground state energy E0 is callednormal ordering. It amounts to placing all annihilation operators a(k)to the right of the creation operators a(k). The ground state energy

    E0 := 0|H|0

    =

    dk

    2

    1

    2~(k) 2(k = 0)

    = (Volume in position space)

    dk

    2

    1

    2~(k)

    =

    modes

    1

    2~modes

    (1.83)

    can be ill-defined for two distinct reasons. First, if N with aheld fixed, there exists an upper cut-off to the integral over reciprocalspace at the Brillouin zone boundaries /a and E0 is only infrareddivergent due to the fact that 2(k = 0) is the diverging volumeL = N a in position space. Second, even if L = N a is kept finitewhile both the infrared N and ultraviolet a 0 limits are taken,the absence of an upper cut-off in the k integral can cause the zero-point energy density E0/L to diverge as well. Divergences of E0 orE0/L are only of practical relevance if one can control experimentally(k) or the density of states

    modes and thereby measure changes in

    E0 or E0/L. For example, this can be achieved in a resonant cavitywhose size is variable. If so, changes of E0 with the cavity size can bemeasured. These changes in the zero point energy are known as theCasimir energy. Sensitivity to E0 with measurable consequences alsooccurs when, upon tuning of some internal parameters entering the mi-croscopic Hamiltonian, the vacuum state |0 becomes unstable, i.e., isnot the true ground state anymore. The system then undergoes a quan-tum phase transition. Finally, divergences of E0/L matter greatly if theenergy-momentum tensors of matter fields are dynamical variablesas is the case in cosmological models.

  • 1.5. PROBLEMS 19

    1.4. Higher-dimensional generalizations

    Generalizations to higher dimensions are straightforward. The co-ordinates x R1 and k R1 in position and reciprocal one-dimensionalspaces need only be replaced by the vectors r Rd and k Rd, inposition and reciprocal d-dimensional spaces, respectively.

    1.5. Problems

    1.5.1. Absence of crystalline order in one and two dimen-sions.

    Introduction. We are going to prove the Mermin-Wagner theoremfor the case of crystalline order in two (and one) dimensions of positionspace. [23] The Mermin-Wagner theorem states that classical particlesin a box, i.e., particles that are subject to hard-wall boundary con-ditions, cannot exhibit crystalline order in one and two dimensions,provided that the pair potential (r) through which they interact sat-isfies certain conditions [see Eq. (1.109)].

    Before we start with the derivation, let us set up some notation.Given the pair potential (r), the internal energy of a configurationof N particles with coordinates r1, , rN in d-dimensional positionspace is given by

    U(r1, , rN) =1

    2

    Ni6=j

    (ri rj). (1.84)

    Using this, we can define the (classical) ensemble average of a real-valued function f of the coordinates r1, , rN by

    f := 1Z

    B

    (Ni=1

    ddri

    )e U(r1, ,rN ) f(r1, , rN) (1.85a)

    and

    Z :=

    B

    (Ni=1

    ddri

    )e U(r1, ,rN ). (1.85b)

    Here, is the inverse temperature after the Boltzmann constant kBhas been set to unity and B denotes the box over which the integrationis taken.

  • 20 1. THE HARMONIC CRYSTAL

    Step 1: Proof of Bogoliubovs inequality. The proof of the Mermin-Wagner theorem will be crucially based on an inequality due to Bogoli-ubov, which for our purposes can be formulated as

    Ni=1

    (ri)

    2

    Ni=1

    (ri)(ri)2

    2

    Ni,j=1

    (ri rj)(ri) (rj)2 + N

    i=1

    |(ri)|2

    ,(1.86)

    for a real-valued function that is continuous and differentiable andvanishes on the boundary B of B, while is complex valued andsufficiently smooth. Our first task is to prove Eq. (1.86).

    Exercise 1.1: Convince yourself that the bilinear map

    , : L L R, (, ) 7 , := , (1.87)for two complex-valued functions and belonging to the set L ofcontinuous differentiable functions from B to R with the standard defi-nition of a product of two functions, is a scalar product. We then havethe Schwarz inequality

    |f1|2 |f2|2 |f1 f2|2 (1.88)

    for any pair of functions f1 and f2 from L at our disposal.Exercise 1.2: Use the Schwarz inequality (1.88) with the choice

    f1(r1, , rN) :=Ni=1

    (ri), (1.89a)

    f 2(r1, , rN) := 1

    e U(r1, ,rN )

    Ni=1

    i[(ri) e

    U(r1, ,rN )],(1.89b)

    to prove Eq. (1.86). Hint: Use partial integration.Step 2: Densities on the lattice. We now want to use the Bogoli-

    ubov inequality (1.86) to probe the tendency towards crystalline order(we specialize to d = 2 for simplicity, but without loss of generality).Suppose that the crystalline order has the Bravais lattice vectors a1and a2 and consists of N1 N2 sites so that

    B ={r Rd |r = x1 a1N1 + x2 a2N2, 0 x1, x2 < 1

    }. (1.90)

    The reciprocal lattice vectors K are given by

    K := n1 b1 + n2 b2, n1, n2 Z, bi aj = 2 ij, i, j = 1, 2,(1.91)

    and a general wave vector k is given by

    k :=n1N1b1 +

    n2N2b2, n1, n2 Z. (1.92)

  • 1.5. PROBLEMS 21

    To probe whether particles form a crystal, we have to compute theirdensity at the reciprocal lattice vectors. In position space, the densityof a configuration of N particles is

    (r) :=Ni=1

    (r ri). (1.93)

    Its Fourier component at momentum k is given by

    k :=

    B

    d2r eikr (r) =Ni=1

    eikri . (1.94)

    This allows us to sharpen a criterion for crystalline order as follows. Acrystal has formed, if

    limN1,N2

    1

    Nk = 0, if k is not a reciprocal lattice vector,

    limN1,N2

    1

    NK 6= 0, for at least one reciprocal lattice vector K.

    (1.95)

    The thermodynamic limit is taken in such a way that the filling of thesystem with particles n := N/(N1N2) is held constant.

    Exercise 2.1: Define the following momenta. Let k be an arbitrarywave vector from the first Brillouin zone as given by Eq. (1.92). Itscomponents with respect to the basis b1 and b2 of the reciprocal latticeare

    ki = ni bi/Ni (1.96)

    for i = 1, 2. Let K be a reciprocal lattice vector, for which Eq. (1.95)is claimed not to vanish.

    (a) Show that the functions defined by

    (r) := ei(k+K)r, (r) := sin(k1 r) sin(k2 r), (1.97)

    are such that vanishes on the boundary B of the box B.(b) Show that the Bogoliubov inequality (1.86) with these func-

    tions yields +k+K kK

    Nk,KDk,K

    , (1.98a)

    where the numerator is

    Nk,K :=(k +K)2

    16

    (K + K+2k K+2k1 K+2k2)2 , (1.98b)

  • 22 1. THE HARMONIC CRYSTAL

    while the denominator is ( is Laplace operator in two-dimensionalposition space)

    Dk,K :=1

    2

    Ni,j=1

    (ri rj)

    [sin(k1 ri) sin(k2 ri) sin(k1 rj) sin(k2 rj)

    ]2

    +1

    Ni=1

    |k2 sin(k1 ri) cos(k2 ri) + (1 2)|

    2.

    (1.98c)

    Exercise 2.2:

    (a) Show that there exists an A > 0 that depends only on b1 andb2 such that the estimates

    A (k1 + k2)2 k21 + k22, A (k1 + k2)

    2 (1 k1 + 2 k2)2 , (1.99)

    for any pair 1 and 2 of real numbers of magnitudes less orequal to one, |1| 1, |2| 1, hold. We are now going tomake use of these inequalities.

    (b) Establish upper bounds on the trigonometric functions in thedenominator (1.98c) to infer that

    1

    N

    +k+K kK

    1N2

    Nk,K

    Ak2

    (1 +

    N

    Ni,j=1

    (ri rj) (ri rj)2) .

    (1.100)

    It will be the quadratic k-dependence in the denominator on whichthe argument crucially relies (it would break down for a k-linear orconstant term). From here on, the task is to find suitable estimates forthe remaining factors.

    To refine the estimate of the denominator on the right-hand side ofEq. (1.100), we have to impose conditions on the asymptotic behaviorof the pair potential (r) at small and large r. To that end, we considera family of pair potentials labeled by a real number > 0

    (r) := (r) r2 |(r)|. (1.101)

    We define the free energy F to be the functional from the space of pairpotentials to the real-valued numbers that assigns to any pair potential the value

    F [] := lnB

    (Ni=1

    ddri

    )exp

    (

    2

    Ni6=j

    (ri rj)

    ). (1.102)

    Exercice 2.3: Show that

    F [0] F []N

    12N

    Ni,j=1

    (ri rj)2|(ri rj)|

    0. (1.103)

  • 1.5. PROBLEMS 23

    Hint: Use the representation

    F [0] F [] =

    0

    dF [ ]

    (1.104)

    and the fact that (prove!)

    F []

    = D , 2F []

    2=

    (D D)

    2, (1.105)

    where denotes the ensemble average using the potential and

    D(r1, , rN) :=1

    2

    Ni,j=1

    (ri rj)2|(ri rj)|. (1.106)

    The inequality that we are seeking applies to a restricted class oftwo-body potentials {}. This restriction comes about because we needto insure that the thermodynamic limit N1, N2 is well defined.More precisely, we need the existence of the free energy per particle

    f0 := limN1,N2

    F [0]

    N 0 for which

    f := limN1,N2

    F []

    N 0 that satisfies Eq. (1.108), we

    can then use the fraction (f0 f)/ to estimate the right-hand sideof Eq. (1.100) by writing

    1

    N

    +k+K kK

    1N2

    Nk,K

    Ak2[1 + 2

    (f0 f)

    ] . (1.110)Exercice 2.4: By assumption (1.95), the averages K+2k, K+2k1,

    and K+2k2 vanish in the thermodynamic limit if 2k, 2k1, and 2k2 arenot reciprocal lattice vectors, respectively. Starting from Eq. (1.110),show that

    1

    V N

    q

    g(|q|)q q

    1

    64A

    K20 g(|K|+ |K0|/2)1 + 2

    (f0 f)

    K2

    N21

    V

    |k| const |r|(2+

    ), (1.109)

    where const is a positive number and , are two positive numbers.

  • 24 1. THE HARMONIC CRYSTAL

    where K0 is the reciprocal lattice vector with smallest magnitude andthe positive function g : R R+, k g(k) > 0 is a Gaussian centeredat the origin.

    The strategy to complete the proof will now be as follows. By in-spection of the right-hand side of Eq. (1.111), we anticipate that thefactor that is underbraced is non-vanishing but finite in the thermo-dynamic limit. In contrast, the sum over 1/k2, once turned into anintegral, diverges logarithmically near the origin. (What happens ifd = 1 or d > 2?) If the left-hand side of Eq. (1.111) turns out tohave a finite upper bound in the thermodynamic limit, the logarithmicdivergence forces

    K2

    N2N1,N2 0. (1.112)

    This is not compatible with our criterion for crystalline order.To make this line of arguments work, it thus remains to show that

    the left-hand side of Eq. (1.111) is bounded from above in the thermo-dynamic limit. To that end, we define the function

    (r) :=

    d2q

    (2)2g(|q|) eiqr. (1.113)

    Exercice 2.5: Show that, in the thermodynamic limit,

    (0) + 2

    (F [] F [ ]

    N

    ) (0) + 1

    N

    Ni6=j

    (ri rj)

    =

    1

    V N

    q

    g(|q|)+q q

    ,

    (1.114)

    where F [] is the free energy for the system with the pair potential as defined in Eq. (1.102).

    The left-hand side of Eq. (1.114) contains the difference in freeenergy per particle for the pair potential and . We require(and used above already) that F []/N is finite in the thermodynamiclimit. If a Gaussian is added to the pair potential, this behavior isunaffected, as the additive contribution (r) is well behaved both forsmall and large r. It follows that also F [ ]/N is finite in thethermodynamic limit.

    The estimate (1.114) thus allows us to recast the inequality in (1.111)for sufficiently large N in the form

    c >K

    2

    N21

    V

    |k|

  • 1.5. PROBLEMS 25

    where c < is a constant independent of N . For large but finite N ,the k-sum diverges as

    1

    V

    |k|

  • CHAPTER 2

    Bogoliubov theory of a dilute Bose gas

    Outline

    Second quantization for bosons is reviewed. Bose-Einstein con-densation for non-interacting bosons is interpreted as an example ofspontaneous-symmetry breaking. The spectrum of a dilute Bose gaswith hardcore repulsion is calculated within Bogoliubov mean-fieldtheory using the operator formalism. It is shown that a Goldstonemode, an acoustic phonon, emerges in association with spontaneous-symmetry breaking. Landau criterion for superfluidity is presented.Bose-Einstein condensation as well as superfluidity at non-vanishingtemperatures are treated using the path integral formalism.

    2.1. Introduction

    This chapter is devoted to the study of a dilute Bose gas with arepulsive contact interaction. We shall see that the phenomenon of su-perfluidity takes place at sufficiently low temperatures. Superfluidityis an example of the spontaneous breaking of a continuous symme-try. The continuous symmetry is the global U(1) gauge symmetry thatis responsible for conservation of total particle number. We shall alsocarefully distinguish Bose-Einstein condensation from superfluidity. In-teractions are necessary for superfluidity to take place. Interactions arenot needed for Bose-Einstein condensation.

    We begin this chapter with the formalism of second quantization forbosons. We then interpret Bose-Einstein condensation at zero temper-ature as an example of the spontaneous breaking of a continuous sym-metry through an explicit construction of a ground state that breaksthe global U(1) gauge symmetry. The emphasis is here on how theglobal U(1) gauge symmetry organizes the Hilbert space spanned byeigenstates of the Hamiltonian. In this construction, the thermody-namic limit plays an essential role.

    Next, we treat a repulsive contact interaction through a mean-fieldapproximation first proposed by Bogoliubov.

    We revisit this approximation using path-integral techniques toshow that it is nothing but a saddle-point approximation. We alsopresent two effective field theories with different physical contents. Thefirst one deals with single-particle excitations. The second one dealswith collective excitations.

    27

  • 28 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    2.2. Second quantization for bosons

    The terminology second quantization is rather unfortunate inthat it might be perceived as implying concepts more difficult to graspthan the passage from classical to quantum mechanics. Quite to thecontrary the relation between second and first quantization 1 isnothing but a matter of convenience. Going from first to second quan-tization is like going from a real-space representation of Schrodingerequation to a momentum-space representation when the Hamiltonianhas translation symmetry.

    Second quantization is a formalism that aims at describing a sys-tem made of identical particles, bosons or fermions, in which cre-ation and annihilation of particles is easily and naturally accountedfor. Hence, the quantum particle number need not be sharp in thisrepresentation, very much in the same way as position is not a sharpquantum number for a momentum eigenstate. Another analogy for therelationship between first quantization, in which the quantum parti-cle number is a sharp quantum number, and second quantization, inwhich it need not be, is that between the canonical and grand-canonicalensembles of statistical mechanics. In the canonical ensemble, parti-cle number is given. In the grand-canonical ensemble, particle numberfluctuates statistically as it has been traded for a fixed chemical poten-tial.

    The formalism of second quantization can already be introducedat the level of a single harmonic oscillator, but it is for interactingmany-body systems that it becomes very powerful. It is neverthelessinstructive to develop the formalism already at the level of a single-particle Hamiltonian since, to a large extent, many-body physics isglorified perturbative physics about some non-interacting limit.

    We shall now generalize the construction of a second-quantized for-malism in terms of creation and annihilation operators for the one-dimensional harmonic oscillator that we presented in chapter 1. Weshall thus consider a finite volume V of d-dimensional space on whichthe single-particle Hilbert space H(1) of square-integrable and twice-differentiable functions is defined. In turn, the single-particle Hamil-tonian is represented by (~ = 1 and is Laplaces operator in d-dimensional space)

    H = 2m

    + U(r), (2.1a)

    1 By first quantization is meant Schrodinger equation.

  • 2.2. SECOND QUANTIZATION FOR BOSONS 29

    and possesses the complete, orthogonal, and normalized basis of eigen-functions

    H n(r) = n n(r),

    V

    ddr m(r)n(r) = m,n,n

    n(r)n(r) = (rr).

    (2.1b)The index n belongs to a countable set after appropriate boundaryconditions, say periodic, have been imposed at the boundaries of thefinite volume V . We assume that the single-particle potential U(r)is bounded from below, i.e., there exists a single-particle and non-degenerate ground-state energy, say 0. Hence the energy eigenvalueindex runs over the positive integers, n = 0, 1, 2, . The time evolu-tion of any solution of Schrodinger equation

    it(r, t) = H (r, t), (r, t = 0) given, (2.2a)

    can be written as

    (r, t) =n

    An n(r) eint, An =

    V

    ddr n(r) (r, t = 0).

    (2.2b)The formalism of second quantization starts with the following two

    postulates.

    (1) There exists a set of pairs of adjoint operators an (creationoperator) and an (annihilation operator) labeled by the energyeigenvalue index n and obeying the bosonic algebra 2

    [am, an] = m,n, [am, an] = [a

    m, a

    n] = 0, m, n = 0, 1, 2, .

    (2.3)(2) There exists a non-degenerate vacuum state |0 that is anni-

    hilated by all annihilation operators,

    an|0 = 0, n = 0, 1, 2, . (2.4)

    With these postulates in hand, we define the Heisenberg representationfor the operator-valued field (in short, quantum field),

    (r, t) :=n

    an n(r) e

    +int (2.5a)

    together with its adjoint

    (r, t) :=n

    an n(r) eint. (2.5b)

    2 The conventions for the commutator and anticommutator of any two objectsA and B are [A,B] := AB BA and {A,B} := AB +BA, respectively.

  • 30 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    The bosonic algebra (2.3) endows the quantum fields (r, t) and (r, t)with the equal-time algebra 3

    [(r, t), (r, t)] = (rr), [(r, t), (r, t)] = [(r, t), (r, t)] = 0.(2.9)

    The quantum fields (r, t) and (r, t) act on the big many-particlespace

    F :=N=0

    N

    symH(1)

    . (2.10a)Here, each

    NsymH(1) is spanned by states of the form

    |m0, ,mi1,mi,mi+1, :=i

    (ai

    )mimi!|0, mi = 0, 1, 2, ,

    (2.10b)with the condition on the non-negative integers mi that

    i

    mi = N. (2.10c)

    The algebra obeyed by the as and their adjoints ensures thatN

    symH(1)

    is theN -th symmetric power ofH(1), i.e., that the state |m0, ,mi1,mi,mi+1, made of N identical particles of which mi have energy i is left un-changed by any permutation of theN particles. Hence, the big many-particle Hilbert space (2.10a) is the sum over the subspaces spanned bywave functions for N identical particles that are symmetric under anypermutation of the particles labels. This big many-particle Hilbertspace is called the bosonic Fock space in physics.

    The rule to change the representation of operators from the Schrodingerpicture to the second quantized language is best illustrated by the fol-lowing examples.

    3 Alternatively, if we start from the classical Lagrangian density

    L := (it)(r, t)1

    2m||2(r, t) ||2(r, t)U(r), (2.6)

    we can elevate the field (r, t) and its momentum conjugate

    (r, t) :=L

    (t)(r, t)= i(r, t) (2.7)

    to the status of quantum fields (r, t) and (r, t) = i(r, t) obeying the equal-timebosonic algebra

    [(r, t), (r, t)] = i(r r), [(r, t), (r, t)] = [(r, t), (r, t)] = 0. (2.8)

  • 2.2. SECOND QUANTIZATION FOR BOSONS 31

    Example 1: The second-quantized representation H of the single-particle Hamiltonian (2.1a) is

    H :=

    V

    ddr (r, t)H (r, t)

    =n

    nan an.

    (2.11)

    As it should be, it is explicitly time independent.Example 2: The second-quantized total particle-number operator

    Q is

    Q :=

    V

    ddr (r, t) 11 (r, t)

    =n

    an an.

    (2.12)

    It is explicitly time independent as follows from the continuity equation

    0 = (t)(r, t) + ( J)(r, t),(r, t) := |(r, t)|2,

    J(r, t) :=1

    2mi[(r, t) () (r, t) () (r, t)(r, t)] ,

    (2.13a)

    obeyed by Schrodinger equation (2.2a). The number operator Q is theinfinitesimal generator of global gauge transformations by which allstates in the bosonic Fock space are multiplied by the same operator-valued phase factor. Thus, for any q R, a global gauge transformationon the Fock space is implemented by the operation

    |m0, ,mi1,mi,mi+1, e+iq Q |m0, ,mi1,mi,mi+1, (2.14)

    on states, or, equivalently, 4

    an e+iq Q an eiq Q = eiq an, (2.16a)and

    an e+iq Q an eiq Q = e+iq an, (2.16b)for all pairs of creation and annihilation operators, respectively. Equa-tion (2.16b) teaches us that any creation operator carries the particlenumber +1. Equation (2.16a) teaches us that any annihilation operatorcarries the particle number 1.

    4 we made use of

    [aa, a] = aaaaaa = aaaaaa+aaaaaa = a[a, a]+[a, a]a = a, (2.15a)and, similarly,

    [aa, a] = +a. (2.15b)

  • 32 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    Example 3: The second-quantized local particle-number densityoperator and the particle-number current density operator J are

    (r, t) := (r, t) 11 (r, t), (2.17a)

    and

    J(r, t) :=1

    2mi

    [(r, t) () (r, t)

    (

    )(r, t)(r, t)

    ], (2.17b)

    respectively. The continuity equation

    0 = (t)(r, t) +( J

    )(r, t) (2.17c)

    that follows from evaluating the commutator between and H isobeyed as an operator equation.

    The operators H, Q, , and J all act on the Fock space F . Theyare thus distinct from their single-particle counterparts H, Q, , and Jwhose actions are restricted to the Hilbert space H(1). By construction,the action of H, Q, and J on the subspace

    1symH(1) of F coincides

    with the action of H, Q, , and J on H(1), respectively.

    2.3. Bose-Einstein condensation and spontaneous symmetrybreaking

    Given a many-body system made of identical bosons, say atomscarrying an integer-valued total angular momentum, how does one con-struct the ground state? The simplest answer to this question occurswhen bosons are non-interacting. In this case, the ground state is sim-ply obtained by putting all bosons in the lowest energy single-particlestate. If the number of bosons is taken to be N , then the ground stateis |N, 0, with energy N0. This straightforward observation under-lies the phenomenon of Bose-Einstein condensation. A non-vanishingfraction of bosons occupies the single-particle energy level 0 below theBose-Einstein transition temperature TBE in the thermodynamic limitof infinite volume V but non-vanishing particle density.

    From a conceptual point of view, it is more fruitful to associateBose-Einstein condensation with the phenomenon of the spontaneousbreaking of a continuous symmetry than with macroscopic occupationof a single-particle level. The continuous symmetry in question is thefreedom in the choice of the global phase of the many-particle wavefunctions. This symmetry is responsible for total particle-number con-servation. In mathematical terms, the vanishing commutator

    [H, Q] = 0 (2.18)

    between the total number operator Q and the single-particle Hamil-tonian H implies a global U(1) gauge symmetry.

    The concept of spontaneous symmetry breaking is subtle. For onething it can never take place when the normalized ground state |0 of

  • 2.3. BOSE-EINSTEIN CONDENSATION AND SPONTANEOUS SYMMETRY BREAKING33

    the many-particle Hamiltonian (possibly interacting) is non-degenerate,i.e., unique up to a phase factor. Indeed, the transformation law ofthe ground state |0 under any symmetry of the Hamiltonian mustthen be multiplication by a phase factor. Correspondingly, the groundstate |0 must transform according to the trivial representation of thesymmetry group, i.e., |0 transforms as a singlet. In this case thereis no room for the phenomenon of spontaneous symmetry breaking bywhich the ground state transforms non-trivially under some symmetrygroup of the Hamiltonian.

    Now, the Perron-Frobenius theorem for finite dimensional matri-ces with positive entries, see Refs. [24] and [25], or its extension, seeRef. [26], to single-particle Hamiltonians of the form (2.1a) guaran-tees that the ground state is non-degenerate for a non-interacting N -body Hamiltonian defined on the Hilbert space

    NsymH(1). When the

    ground state of an interacting Hamiltonian defined on the Hilbert spaceNsymH(1) is non-degenerate, then spontaneous symmetry breaking is

    ruled out for this interacting Hamiltonian.Before evading this no-go theorem by taking advantage of the

    thermodynamic limit of infinite volume V but non-vanishing particledensity, we want to investigate more closely the consequences of hav-ing a non-degenerate ground state. We consider the cases of bothnon-interacting many-body Hamiltonians such as H in Eq. (2.11) and

    interacting many-body Hamiltonians 5 that commute with Q. TheHilbert space will be the bosonic Fock space F in Eq. (2.10a) on whichthe quantum field operator (r, t) in Eq. (2.5b) is defined. We shallsee that the expectation value of (r, t) in the ground state |0 ofthe many-body system can be used as a signature of the spontaneousbreaking of the U(1) symmetry. More generally, we shall interpret thequantum statistical average of (r, t) as a temperature dependent orderparameter.

    As follows from Eq. (2.16a), the quantum field (r, t) transformsaccording to

    e+iq Q (r, t) eiq Q = eiq (r, t), r, t, (2.19)under any global gauge transformation labeled by the real-valued num-ber q. The quantum field (r, t) carries U(1) charge 1 as it lowersthe bosonic occupation numbers

    imi by one on any state (2.10b) of

    the bosonic Fock space F . By hypothesis, the ground state |0 of His non-degenerate. Thus, it transforms like a singlet under U(1),

    Q0 R, eiq Q |0 = eiq Q0 |0, 0| e+iq Q = 0| e+iq Q0 .(2.20)

    What then follows for the expectation value 0|(r, t)|0?5 Interactions are easily introduced through polynomials in creation and anni-

    hilation operators of degree larger than 2.

  • 34 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    It must vanish. Indeed,

    0|(e+iq Q (r, t)eiq Q

    )|0 = eiq 0|(r, t)|0, r, t, (2.21)

    by Eq. (2.19) and(0|e+iq Q

    )(r, t)

    (eiq Q|0

    )= 0|(r, t)|0, r, t, (2.22)

    by Eq. (2.20) hold simultaneously for any q R. The vanishing of0|(r, t)|0, in view of the fact that (r, t) carries U(1) charge 1and thus transforms non-trivially under U(1), can be traced to theassumption that the ground state |0 is unique, i.e., that |0 is aneigenstate of Q. In more intuitive terms, the action of (r, t) on an

    eigenstate of Q such as |0 is to lower the total number of particlesby one, thereby producing a state orthogonal to |0. Conversely, anon-vanishing expectation value of (r, t) in some state | F is onlypossible if | F is not an eigenstate of Q. 6

    Evading the no-go theorem for spontaneous symmetry breakingthus requires quantum degeneracy of the ground state with orthogonalground states that are related by the action of the U(1) symmetrygroup. In turn, this can be achieved by constructing a ground state| F that is an eigenstate of (r, t) and thus cannot be an eigenstateof Q.

    A prerequisite to evade the no-go theorem for spontaneous sym-metry breaking is that the thermodynamic limit of infinite volume Vbut non-vanishing particle density be taken. This idealized mathe-matical limit is often an excellent approximation in condensed-matterphysics or in cold-atom physics. When the thermodynamic limitN, V with N/V held fixed is well defined, there is no difference betweenapproaching this limit by working at fixed volume and at fixed par-ticle number with the Hilbert space

    NsymH(1) or approaching the

    thermodynamic limit by working at fixed external pressure and at fixedchemical potential with the Fock space F =

    N=0

    NsymH(1). The

    first approach to the thermodynamic limit defines the so-called canon-ical ensemble of quantum statistical mechanics. The second approachto the thermodynamic limit defines the so-called grand-canonical en-semble of quantum statistical mechanics. The thermodynamic limit isalso needed to recover spontaneous symmetry breaking even when theHilbert space of finitely-many degrees of freedom is endowed with thestructure of a Fock space. 7

    6 It is impossible for (r, t) to acquire an expectation value onN

    symH(1).7 This occurs when the bosons of the many-body system are collective excita-

    tions, say phonons in a solid, spin waves in an antiferromagnet, or excitons in asemiconductor, i.e., when the finitely-many degrees of freedom are ions, spins, orband electrons, respectively.

  • 2.3. BOSE-EINSTEIN CONDENSATION AND SPONTANEOUS SYMMETRY BREAKING35

    To underscore the role played by the thermodynamic limit to evadethe no-go theorem for spontaneous symmetry breaking, we now re-strict ourself to the many-body and non-interacting Hamiltonian

    H := H Q (2.23a)

    with

    H = 2m

    (2.23b)

    in Eq. (2.1a) so that translation invariance holds at the single-particlelevel. The real-valued parameter is called the chemical potential.Since H commutes with Q by hypothesis, an eigenstate of H is also aneigenstate of H and conversely. Eigenenergies of H and H may differ,however. For example, the single-particle eigenfunctions n(r) of H in

    Eq. (2.1a) are also single-particle eigenfunctions of H on H(1) but withthe rigidly shifted spectrum of energy eigenvalues n. Furthermore,the dimensionalities of the eigenspaces of H can change dramaticallyby the addition of Q. To see this, observe that the choice = 0insures that the single-particle ground-state energy of H vanishes and

    that the corresponding normalized eigenfunction (r) = 1/V . 8 This

    choice also guarantees that all states

    |m0, 0, =(a0)

    m0m0!|0, m0 = 0, 1, 2, , (2.25)

    are orthogonal eigenstates of H in F with the same vanishing energy. 9

    The choice = 0 guarantees that H has countably-many orthogonalground states provided the volume V is finite.

    Any linear combination of states of the form (2.25) is a ground state

    of H with = 0. Of all these possible linear combinations, consider

    the continuous family of normalized 10 ground states labeled by the

    8 A time-dependent gauge transformation plays the same role as the chemicalpotential if one chooses to work in the canonical instead of the grand-canonicalstatistical ensemble. For example, setting 0 to 0 in the single-particle Hilbertspace H(1) is achieved with the help of the time-dependent gauge transformation

    (r, t) ei0t(r, t) (2.24)

    on the single-particle Schrodinger equation (2.2a).9 The same states are also eigenstates of H in F but with distinct energy

    eigenvalues m00.10 Observe that the operator

    D(1, 2, ) :=n

    e(nan

    nan), 1, 2, C, (2.26)

    is unitary.

  • 36 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    complex-valued parameter ,

    |gs := eV2||2

    m0=0

    (V )m0

    m0!|m0, 0,

    = eV2||2 e

    V a0|0

    a0|0 = 0 = eV

    2||2 e+

    V a0 e

    V a0|0

    = eV ( a0a0)|0

    =: D(V , 0, )|0.

    (2.27)

    To reach the penultimate line, we made use of [[A,B], A] = [[A,B], B] =

    0 = eA eB = e[A,B]/2 eA+B. Here, the unitary operator D(V , 0, )

    rotates the vacuum into the bosonic coherent state (see appendix A)

    |V , 0, cs := e

    V a0 |0, (2.28)

    up to the proportionality constant exp(V2||2). Bosonic coherent

    states form an overcomplete set of the Fock space (see appendix A).The overlap between any two coherent states is always non-vanishing(see appendix A),

    cs0, 1, |0, 1, cs =n

    enn , n, n C,

    cs0, 1, | := 0|n

    enan , n C,

    |0, 1, cs :=n

    en an|0, n C.

    (2.29)

    The same is true of the overlaps (see appendix A)

    gs|0 = eV2||2 ,

    gs|gs = eV||2

    2 .(2.30)

    The rational for having scaled the arguments of the unitary opera-tor D(

    V , 0, ) by the square root of the volume V of the system

    in Eq. (2.27) is to guarantee that all the rotated vacua in Eq. (2.27)become orthogonal in the thermodynamic limit. The thermodynamiclimit is thus essential in providing an escape to the absence of spon-taneous symmetry breaking in systems of finite sizes. In the ther-modynamic limit, we need not distinguish H defined on

    NsymH(1)

    from H defined on F . It is only in the thermodynamic limit thatthe ground-state manifold = C of H, = 0, in Eq. (2.27) becomesthe ground-state manifold = C of H. Where does this degeneracyof H comes from? When V and N are finite and H is restricted toN

    symH(1) the ground-state energy is N0. The ground-state energy

  • 2.3. BOSE-EINSTEIN CONDENSATION AND SPONTANEOUS SYMMETRY BREAKING37

    of H inN1

    symH(1) differs from that inN

    symH(1) by a term of orderN0 namely 0. In the Fock space F , the energy difference per particlebetween N, 0, |H|N, 0, and N N, 0, |H|N N, 0, scales like 1/N as the thermodynamic limit N , N/N 0, andN/V non-vanishing is taken. Hence, more and more states have anenergy of order N0 above the ground-state energy N0 as the systemsize is increased. The surprising result is that it is not a mere countableinfinity of states that become degenerate with the ground state in thethermodynamic limit but an uncountable infinity.

    It remains to verify that each ground state |gs in Eq. (2.27) is aneigenstate of the quantum fields (r, t), 11 but is not an eigenstate of

    Q,

    (r, t) |gs = |gs,

    eiQ |gs = |eigs.(2.31)

    The U(1) multiplet structure of the manifold of ground states = C inEq. (2.27) is displayed by Eq. (2.31). Circles in the complex plane C correspond to U(1) multiplets. Normalization of the single-particleeigenfunction 0(r) = 1/

    V and the property that coherent states are

    eigenstates of annihilation operators guaranty that the quantum field(r, t) acquires the expectation value C with the particle density||2 in the ground-state manifold (2.27),

    gs|(r, t)|gs = ,

    gs|(r, t)(r, t)|gs = ||2.(2.32)

    In an interacting system the non-interacting trick relying on finetuning of the chemical potential 0 to construct explicitly themany-body ground state breaks down. The chemical potential is choseninstead by demanding that the particle density,

    0|(r, t)(r, t)|0 =N

    V, (2.33)

    at zero temperature, 12 be held fixed to the value N/V as the thermo-dynamic limit is taken. At non-vanishing temperature the right-handside is unchanged whereas the left-hand side becomes a statistical aver-age in the grand-canonical ensemble. A degenerate manifold of groundstates satisfying Eqs. (2.32) is not anymore parametrized by C butby arg() [0, 2[, since the modulus ||2 = N/V is now given. The

    11 Remember that the single-particle ground-state wave function 0(r) is the

    constant 1/V . Make then use of the expansion (2.5b) applied to (2.27) whereby

    1Va0|V cs = 1V (

    V )|

    V cs must be used.

    12As before, |0 denotes the many-body ground state which, in practice, can-not be constructed exactly when interactions are present. We are implicitly as-suming translation invariance. This is the reason why the right-hand side does notdepend on r.

  • 38 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    U(1) symmetry group parametrized by exp(i Q), [0, 2[ is saidto act transitively on the ground-state manifold. Construction of theground-state manifold relies on approximate schemes such as mean-field theory. These approximations are non-perturbative in the sensethat they yield variational wave functions that cannot be derived fromthe non-interacting limit to any finite order of the perturbation theoryin the interaction strength.

    Spontaneous symmetry breaking is said to occur when the groundstate |0 of a many-body system is no longer a singlet under the ac-tion of a symmetry group of the system. A quantity like 0|(r, t)|0that must vanish when the ground state is a singlet, but becomes non-vanishing in a phase with spontaneous symmetry breaking is calledan order parameter. An order parameter is a probe to detect sponta-neous symmetry breaking. In condensed-matter physics, some orderparameters can be directly observed in static measurements. For ex-ample, elastic-neutron scattering can show Bragg peaks correspondingto crystalline or magnetic order. An order parameter can also be indi-rectly observed in a dynamical measurement. For instance, inelastic-neutron scattering can show a gapless branch of excitations, Goldstonemodes, corresponding to phonons or spin waves. Some consequences ofsymmetries such as selections rules and degeneracies of the excitationspectrum no longer hold in their simplest forms when the phenomenonof spontaneous symmetry breaking occurs. The mass distributions ofmesons, hadrons, photon, W and Z bosons are interpreted as a man-ifestation of spontaneous symmetry breaking leading to the standardmodel of strong, weak, and electromagnetic interactions.

    How does one go about detecting spontaneous symmetry breakingin the canonical ensemble? This question is of relevance to numericalsimulations where the dimensionality of the Hilbert space is necessarilyfinite. A probe for spontaneous symmetry breaking is off-diagonal long-range order. Let |N be the ground state of the many-body systemin the Hilbert space

    NsymH(1). We denote with (r) the quantum

    field (r, t = 0) in the Schrodinger picture. Here, the Schrodinger pic-ture can be implemented numerically through exact diagonalization ofmatrices say. We assume translation invariance, i.e., the single-particlepotential U(r) = 0 in Eq. (2.1a). Define the one-particle density matrixby

    R(r, r) :=1

    VN |(r) (r)|N. (2.34)

    By translation invariance

    R(r, r) = R(r r), (2.35)

  • 2.3. BOSE-EINSTEIN CONDENSATION AND SPONTANEOUS SYMMETRY BREAKING39

    which we use to deduce the dependence on r r. First, we insert intoEq. (2.34) the Fourier expansions (2.5)

    R(r, r) =

    ddk

    (2)d

    ddk

    (2)dei(krk

    r) N |akak|N. (2.36)

    Second, we take advantage of R(r, r) = R(r r) to do

    R(r, r) =1

    V

    ddyR(r + y, r + y). (2.37)

    Third, we combine Eqs. (2.36) and (2.37) into

    R(r, r) =1

    V

    ddy

    ddk

    (2)d

    ddk

    (2)dei[k(r+y)k

    (r+y)] N |akak|N

    =1

    V

    ddk

    (2)d

    ddk

    (2)d(2)d(k k)ei(krkr) N |a

    kak|N.

    (2.38)

    Finally, the integration over the momentum k yields

    R(r, r) =1

    V

    ddk

    (2)deik(rr

    ) N |akak|N

    =:

    ddk

    (2)deik(rr

    ) nk.

    (2.39)

    The ground-state expectation value nk is the number of particles perunit volume with momentum k. When r r = 0, the one-particledensity matrix R(r r) is just the total number of particles per unitvolume n0 = N/V . Bose-Einstein condensation means that

    nk = n0(2)d(k) + f(k), (2.40a)

    with f(k) some smooth function that satisfiesddk

    (2)df(k) = 0. (2.40b)

    In position space, Bose-Einstein condensation thus amounts to

    R(r, r) = n0+F (rr), F (r) :=

    ddk

    (2)deikrf(k), lim

    |r|F (r) = 0.

    (2.41)The non-vanishing of lim|r|R(r,0) is another signature of sponta-neous symmetry breaking associated to Bose-Einstein condensation.

    We conclude this section with some field-theoretical terminology.States | for which

    lim|r1r2|

    |O1(r1)O2(r2)| = |O1(r1)||O2(r2)| (2.42)

    holds for any pair of operators O1(r) and O2(r) defined on the Fockspace F are said to satisfy the cluster decomposition property or to beclustering. The ground state |N in Eq. (2.41) does not satisfy the

  • 40 2. BOGOLIUBOV THEORY OF A DILUTE BOSE GAS

    clustering property. 13 The manifold of states | Cgs in Eq. (2.27)does satisfy the clustering property by Eq. (2.32).

    2.4. Dilute Bose gas: Operator formalism at vanishingtemperature

    2.4.1. Operator formalism. Bogoliubov introduced in 1947 aninteracting model for superfluid 4He. [27] This model turns out notto be a very good one for superfluid 4He in that the assumption ofpairwise interactions made by Bogoliubov fails. However, this modelhas been conceptually very important. Moreover, this is a realisticmodel in the field of cold atoms that came into maturity in 1995 withthe experimental realization of Bose-Einstein condensation. [28]

    The model for weakly interacting bosons proposed by Bogoliubov, adilute Bose gas in short, is defined by the second-quantized Hamiltonian

    H, =

    V

    ddr

    [(r, t)

    (

    2m

    )(r, t) +

    2

    (

    )2(r, t)

    ].

    (2.43a)The chemical potential determines the number N() of particles inthe interacting ground state |gs from

    N() =

    gs

    V

    ddr(

    )(r, t)

    gs. (2.43b)

    Conversely, fixing the total particle number toN determines (N). Theinteraction is a two-body, short-range, and repulsive density-densityinteraction. In the limit in which the range of this interaction is muchsmaller than the average particle separation, this interaction is wellapproximated by a delta function repulsion (this is the justification forthe adjective dilute),

    H :=

    2

    V

    ddr

    V

    ddr (r, t) (rr) (r, t), (r, t) :=(

    )(r, t).

    (2.43c)The real-valued parameter 0 measures the strength of the repulsiveinteraction and carries the units of (energyvolume). Bosons are saidto have a hardcore.

    When periodic boundary conditions are imposed in the volume V ,it is natural to expand the pair of canonical conjugate quantum fields

    13 Choose O1 = and O2 = . The left-hand side of Eq. (2.42) is non-

    vanishing. On the other hand, since the ground state has a well-defined number Nof particle, the right-hand side must vanish.

  • 2.4. DILUTE BOSE GAS: OPERATOR FORMALISM AT VANISHING TEMPERATURE41

    (r, t) and i(r, t) in the basis of plane waves,

    (r, t) =1V

    k

    ak ei(krkt), (r, t) =

    1V

    k

    ak e+i(krkt).

    (2.44a)Here, the summation over reciprocal space is infinite but countable,

    k =2

    Ll, l Zd, Ld V, (2.44b)

    and we have introduced the single-particle dispersion

    k =k2

    2m. (2.44c)

    We observe that the single-particle plane wave with the lowest en-ergy is

    0(r) =1V, 0 = 0. (2.45)

    The representation of the Hamiltonian in terms of creation andannihilation operators ak and ak, respectively, is

    H, =k

    (k +

    2(r = 0)

    )akak+

    2V

    k1,k2,k3,k4

    k1+k2,k3+k4 ak1ak2 ak3 ak4 .

    (2.46)Normal ordering has resulted in the (divergent) shift in the chemicalpote