Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets [5 ed.] 9814273562, 9789814273565

This is the fifth, expanded edition of the comprehensive textbook published in 1990 on the theory and applications of pa

216 86 234MB

English Pages 1579 [1626] Year 2009

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Contents
Preface
Preface to Fourth Edition
Preface to Third Edition
Preface to Second Edition
Preface to First Edition
List of Figures
List of Tables
1 Fundamentals
1.1 Classical Mechanics
1.2 Relati vistic Mechanics in Curved Spacetime
l.3 Quantum Mechanics
1.3.1 Bragg Reflections and Interference
1.3.2 Matter Waves
1.3.3 SchrOdinger Equation
1.3.4 Particle Current Conservation
1.4 Dirac's Bra-Ket Formalism.
1.4.1 Basis Transformations
1.4.2 Bracket Notation
1.4.3 Continuum Limit
1.4.4 Generalized Functions
1.4.5 SchrOdinger Equation in Dirac Notation
1.4.6 Momentum States
1.4.7 Incompleteness and Poisson's Summation Formula
1.5 Obscrvables
1.5.1 Uncertainty Relation.
1.5.2 Density Matrix and Wigner Function.
1.5.3 Generalization to Many Particles
1.6 Time Evolution Operator
l.7 Properties of the Time Evolution Operator
1.8 Heisenberg Picture of Quantum Mechanics
l.9 Interaction Picture and Perturbation Expansion
1.10 Time Evolution Amplitude
1.11 Fixed-Energy Amplitude
1.12 Free-Particle Amplitudes.
1.13 Quantum Mechanics of General Lagrangian Systems.
1.14 Particle on the Surface of a Sphere
1.15 Spinning Top
1.16 Scattering
1. 16.1 Scattering Matrix
1.16.2 Cross Section
1.16.3 Born Approximation
1. 16.4 Partial Wave Expansion and Eikonal Approximation
1. 16.5 Scattering Amplitude from Time Evolution Amplitude
1. 16.6 Lippmann-Schwinger Equation
1.17 Classical and Quantum Statistics
1. 17.1 Canonical Ensemble.
1. 17.2 Grand-Canonical Ensemble
1.18 Density of States and Ttacelog
Appendix l A Simple Time Evolution Operator.
Appendix I B Convergence of the Fresnel Integral
Appendix l C The Asymmetric Top
Notes and References.
2 Path Integrals - Elementary Properties and Simple Solutions
2.1 Path Integral Representation of Time Evolution Amplitudcs
2.1.1 Sliced Time Evolution Amplitude.
2.1.2 Zero-Hamiltonian Path Integral.
2.1.3 SchrOdinger Equation for Time Evolution Amplitude
2.1.4 Convergence of of the Time-Sliced Evolution Amplitude
2.1.5 Time Evolution Amplitude in Momentum Space.
2.1.6 Quantum-Mechanical Partition Function.
2.1.7 Feynman's Configurat ion Space Pat h Integral
2.2 Exact Solution for the Frec Particle
2.2. 1 Direct Solution
2.2.2 Fluctuations around the Classical Path
2.2.3 Fluctuation Factor
2.2.4 Finite Slicing Properties of Free-Particle Amplitude.
2.3 Exact Solution for Harmonic Oscillator
2.3. 1 Fluctuations around the Classical Path.
2.3.2 Fluctuation Factor
2.3.3 The i7]-Prescription and Maslov-Morse Index
2.3.4 Continuum Limit.
2.3.5 Useful Fluctuation Formulas.
2.3.6 Oscillator Amplitude on Finite Time Lattice
2.4 Gelfand-Yaglom Formula.
2.4. 1 Recursive Calculation of Fluctuation Detcrminant.
2.4.2 Examples
2.4.3 Calculation on Unsliced Time Axis
2.4.4 D'Alembert's Construction
2.4 .5 Another Simple Formula.
2.4.6 Generalization t o D Dimensions
2.5 Harmonic Oscillator with Time-Dependent Frequency
2.5.1 Coordinate Space.
2.5.2 Momentum Space
2.6 Free-Particle and Oscillator Wave Functions
2.7 General Time-Dependent Harmonic Action
2.8 Path Integrals and Quantum Statistics
2.9 Density Matrix
2.10 Quantum Statistics of the Harmonic Oscillator
2.11 Time-Dependent Harmonic Potential
2.12 Functional Measure in Fourier Space
2.13 Classical Limit.
2.14 Calculation Techniques on Sliced Time Axis via the Poisson Formula
2.15 Ficld·Theorctic Definition of Harmonic Path Integrals by Analytic Regularization
2.15.1 Zero--Temperature Evaluation of the Frequency Sum.
2. 15.2 Finite-Temperature Evaluation of the Frequency Sum.
2.15.3 Quantum.Mechanical Harmonic Oscillator
2. 15.4 Tracelog of the First·Order Differential Operator
2.15.5 Gradient Expansion of the One-Dimensional Tracclog
2.15.6 Duality Transformation and Low· Temperature Expansion
2.16 Finite-N Behavior of Thermodynamic Quantities
2.17 Time Evolution Amplitude of Freely Falling Particle.
2.18 Charged Particle in Magnetic Field
2.18.1 Action.
2.18.2 Gauge Properties.
2. 18.3 Time-Sliced Path Integration
2.18.4 Classical Action
2. 18.5 Translational Invariancc
2.19 Charged Particle in Magnetic Field plus Harmonic Potentia]
2.20 Gauge Invariance and Alternative Path Integral Representation
2.21 Velocity Path Integral.
2.22 Path Integral Representation of the Scattering Matrix
2.22.1 General Development
2.22.2 Improved Formulation
2.22.3 Eikonal Approximation to the Scattering Amplitude
2.23 Heisenberg Operator Approach to Time Evolution Amplitude.
2.23.1 Free Particle
2.23.2 Harmonic Oscillator
2.23.3 Charged Particle in Magnetic Field
Appendix 2A Baker-Campbell-Hausdorff Formula and Magnus Expansion
Appendix 28 Direct Calculation of the Time-Sliced Oscillator Amplitudc
Appendix 2C Derivation of Mehler Formula.
Notes and References
3 External Sources, Correlations, and Perturbation Theory
3.1 External Sources
3.2 Green Function of Harmonic Oscillator
3.2.1 Wronski Construction
3.2.2 Spectral Representation
3.3 Green Functions of First-Order Differential Equation
3.3.1 Time-Independent Frequency
3.3.2 Time-Dependent Frequency
3.4 Summing Spectral Representation of Green Function
3.5 Wronski Construction for Periodic and Antiperiodic Green F\mctions
3.6 Time Evolution Amplitude in Presence of Source Term
3.7 Time Evolution Amplitude at Fixed Path Average
3.8 External Source in Quantum-Statistical Path Integral
3.8.1 Continuation of Real-Time Result
3.8.2 Calculation at Imaginary Time
3.9 Lattice Green Function
3.10 Correlation Functions, Generating Functional, and Wick Expansion
3.10.1 Real-Time Correlation Functions.
3.11 Correlation Functions of Charged Particle in Magnetic Field
3.12 Correlation Functions in Canonical Path Integral.
3.12.1 Harmonic Correlation Functions
3.12.2 Relations between Various Amplitudes
3.12.3 Harmonic Generating Functionals.
3.13 Particle in Heat Bath
3.14 Heat Bath of Photons.
3.15 Harmonic Oscillator in Ohmic Heat Bath
3.16 Harmonic Oscillator in Photon Heat Bath
3.17 Perturbation Expansion of Anharmonic Systems
3.18 Rayleigh-SchrOdinger and Brillouin-Wigner Perturbation Expansion
3.19 Level-Shifts and Perturbed Wave Functions from SchrOdinger Equation
3.20 CrucuiatioIl of Perturbation Series via Fcynman Diagrams.
3.21 Perturbative Definition of Interacting Path Integrals.
3.22 Generating Functional of Connected Correlation Functions
3.22.1 Connectedness Structure of Correlation Functions.
3.22.2 Correlation Functions versus Connected Correlation Functions
3.22.3 Functional Generation of Vacuum Diagrams.
3.22.4 Correlation Functions from Vacuum Diagra.ms
3.22.5 Generating Functional for Vertex Functions. Effective Action
3.22.6 Ginzburg-Landau Approximation to Generating Functional
3.22.7 Composite Fields.
3.23 Path Integral Calculation of Effective Action by Loop Expansion
3.23.1 General Formalism
3.23.2 Mean-Field Approximation
3.23.3 Corrections from Quadratic Fluctuations.
3.23.4 Effective Action to Second Order in Ii
3.23.5 Finite-Temperature Two-Loop Effective Action
3.23.6 Background Field Method for Effective Action
3.24 Nambu-Goldstone Theorem
3.25 Effective Classical Potential
3.25.1 Effective Classical Boltzmann Factor
3.25.2 Effective Classical Hamiltonian
3.25.3 High- and Low-Temperature Behavior
3.25.4 Alternative Candidate for Effective Classical Potential
3.25.5 Harmonic Correlation FUnction without Zero Mode
3.25.6 Perturbation Expansion
3.25.7 Effective Potential and Magnetization Curves
3.25.8 First-Order Perturbative Result.
3.26 Perturbati vc Approach to Scattering Amplitude
3.26.1 Generating Functional
3.26.2 Application to Scattering Amplitude
3.26.3 First Correction to Eikonal Approximation
3.26.4 Rayleigh-SehrOdinger Expansion of Scattering Amplitude.
3.27 Functional Determinants from Green FUnctions
Appendix 3A Matrix Elements for General Potential.
Appendix 3B Energy Shifts for gx4 /4-Interaction.
Appendix 3C Recursion Relations for Perturbation Coefficients.
3C.l One-Dimensional Interaction X4
3C.2 General One-Dimensional Interaction.
3C.3 Cumulative Treatment of Interactions X4 and x3
3C.4 Ground-State Energy with External Current
3C.5 Recursion Relation for Effective Potential
3C.6 Interaction r4 in D-Dimensional Radial Oscillator.
3C.7 Interaction r2q in D Dimensions.
3C.8 Polynomial Interaction in D Dimensions
Appendix 3D Feynman Integrals for T =f 0
Notes and References
4 Semiclassical Time Evolution Amplitude
4. 1 Wentzel-Kramers-Brillouin (WKB) Approximation.
4.2 Saddle Point Approximation
4.2. 1 Ordinary Integrals
4.2.2 Path Integrals
4.3 Van Vleck-Pauli-Morette Det erminant
4.4 Fundamental Composition Law for Semiclassical Time Evolution Amplitude
4 .5 Semiclassical Fix ed-Energy Amplitude
4.6 Se miclassical Amplitude in Momentum Space
4.7 Semiclassical Quantum-Mechanical Partition Function
4.8 Multi-Dimensional Systems
4.9 Quantum Corrections to Classical Density of States
4.9.1 One-Dimensional Case
4.9.2 Arbitrary Dimensions
4.9.3 Bilocal Density of States
4.9.4 Gradient Expansion of Tracelog of Hamiltonian Operator
4.9.5 Local De ns ity of States on Circle
4.9.6 Quantum Corrections to Bohr-Sommerfeld Approximation
4.10 Thomas-Fermi Model of Neutral Atoms
4.10.1 Semiclassical Limit
4.10.2 Self-Consistent Field Equation
4.10.3 Energy Functiona l of Thomas-Fermi Atom
4.10.4 Calcula tion of Energies
4.1 0.5 Virial Theorem
4.10.6 Exchange E nergy
4.10.7 Quantum Correction Near Origin
4.10.8 Systemat ic Quantum Corrections to Thomas-Fermi Energies
4.11 Classical Action of Coulomb System
4.12 Semiclassical Scattering
4.12.1 General Formulation
4.12.2 Semiclassical Cross Section of Mott Scattering
Appendix 4A Se miclassical Quantization for Pure Power Potentials
Appendix 4B Derivation of Se miclassical Time Evolution Amplitude
Notes and References
5 Variation al Perturbation T heory
5.1 Variational Approach to Effective Classical Partition Function
5.2 Local Harmonic Trial Partition Function
5.3 The Optimal Upper Bound
5.4 Accuracy of Variational Approximation
5.5 Weakly Bound Ground State Energy in Finite-Range Potential Well
5.6 Possible Direct Generalizations
5.7 Effective Classical Potential for Anharmonic Oscillator and Double-Well Potential
a) Case w2 > 0, Anharmonic Oscillator
b) Case w2 < 0: The Double-Well Potential
5.8 Particle Densities
5.9 Extension to D Dimensions
5.10 Application to Coulomb and Yukawa Potentials
5.11 Hydrogen Atom in Strong Magnetic Field
5.11.1 Weak-Field Behavior
5.11.2 Effective Classical Hamiltonian
5.12 Variational Approach to Excitation Energies
5.13 Systematic Improvement of Feynman-Kleinert Approximation. Variational Perturbation Theory
5.14 Applications of Variational Perturbation Expansion
5.14.1 Anharmonic Oscillator at T = 0
5.14.2 Anharmonic Oscillator for T > 0
5.15 Convergence of Variational Perturbation Expansion
5.16 Variational Perturbation Theory for Strong-Coupling Expansion
5.17 General Strong-Coupling Expansions
5.18 Variational Interpolation between Weak and StrongCoupling Expansions
5.19 Systematic Improvement of Excited Energies
5.20 Variational Treatment of Double-Well Potential
5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions
5.21.1 Evaluation of Path Integrals
5.21.2 Higher-Order Smearing Formula in D Dimensions
5.21.3 Isotropic Second-Order Approximation to Coulomb Problem
5.21.4 Anisotropic Second-Order Approximation to Coulomb Problem
5.21.5 Zero-Temperature Limit
5.22 Polarons
5.22.1 Partition Function
5.22.2 Harmonic Trial System
5.22.3 Effective Mass
5.22.4 Second-Order Correction
5.22.5 Polaron in Magnetic Field, Bipolarons, Small Polarons, Polaronic Excitons, and More
5.22.6 Variational Interpolation for Polaron Energy and Mass
5.23 Density Matrices
5.23.1 Harmonic Oscillator
5.23.2 Variational Perturbation Theory for Density Matrices
5.23.3 Smearing Formula for Density Matrices
5.23.4 First-Order Variational Approximation
5.23.5 Smearing Formula in Higher Spatial Dimensions
Appendix 5A Feynman Integrals for T=fo 0 without Zero Frequency
Appendix 5B Proof of Scaling Relation for Extrema of W N
Appendix 5C Second-Order Shift of Polaron Energy
Notes and References
6 Path Integrals with Topological Constraints
6. 1 Point Particle on Circle.
6.2 Infinite Wall
6.3 Point Particle in Box
6.4 Strong-Coupling Theory for Particle in Box.
6.4. 1 Partition Function
6.4.2 Perturbation Expansion
6.4.3 Variational Strong-Coupling Approximations
6.4.4 Special Properties of Expansion
6.4.5 ExponentiaJly Fast Convergence
Notes and References
7 Many Particle Orbits - Statistics and Second Quantization
7.1 Ensembles of Bose and Fermi Particle Orbits
7.2 Bose-Einstein Condensation
7.2. 1 Free Bose Gas
7.2.2 Bose Gas in Finite Box
7.2.3 Effect of Interactions
7.2.4 Bose-Einstein Condensation in Harmonic 'frap
7.2.5 Thermodynamic Functions
7.2.6 Critical Temperature
7.2.7 More General Anisotropic Trap
7.2.8 Rotating Bose-Einstein Gas
7.2.9 Finite-Size Corrections
7.2.10 Entropy and Specific Heat
7.2. 11 Interactions in Harmonic Trap
7.3 Gas of Free Fermions
7.4 Statistics Interaction
7.5 Fractional Statistics
7.6 Second-Quantized Bose Fields
7.7 Fluctuating Bose Fields.
7.8 Coherent States
7.9 Second-Quantized Fermi Fields
7.10 Fluctuating Fermi Fields
7.10.1 Grassmann Variables
7.10.2 Fermionic Functional Determinant
7.10.3 Coherent States for Fermions
7.11 Hilbert Space of Quantized Grassmann Variable
7.11.1 Single Real Grassmann Variable
7.11.2 Quantizing Harmonic Oscillator with Grassmann Variables
7.11.3 Spin System with Grassmann Variables
7.12 External Sources in a", a -Path Integral
7.13 Generalization to Pair Terms.
7.14 Spatial Degrees of Freedom.
7.14.1 Grand-Canonical Ensemble of Particle Orbits from Free Fluctuating Field
7.14.2 First versus Second Quantization
7.14.3 Interacting Fields.
7.14.4 Effective Classical Field Theory.
7.15 Bosonization
7.15.1 Collective Field.
7.15.2 Bosonized versus Original Theory
Appendix 7A Treatment of Singularities in Zeta-Function
7 A.I Finite Box
7 A.2 Harmonic Trap
Appendix 7B Experimental versus Theoretical Would-be Critical Temperature
Notes and References
8 Path Integrals in Polar and Spherical Coordinates
8.1 Angular Deromposition in Two Dimensions.
8.2 Trouble with Feynman's Path Integral Formula in Radial Coordinates
8.3 Cautionary Remarks
8.4 Time Slicing Corrections
8.5 Angular Decomposition in Three and More Dimensions
8.5.1 Three Dimensions
8.5.2 D Dimensions
8.6 Radial Path Integral for Harmonic Oscillator and Free Particle
8.7 Particle ncar the Surface of a Sphere in D Dimensions.
8.8 Angular Barriers ncar the Surface of a Spherc
8.8.1 Angular Barriers in Three Dimensions
8.8.2 Angular Barriers in Four Dimensions
8.9 Motion on a Sphere in D Dimensions
8.10 Path Integrals on Group Spaces
8.11 Path Integral of Spinning Top
8.12 Path Integral of Spinning Particle
8.13 Berry Phase.
8.14 Spin Precession
Notes and References
9 Wave Functions
9.1 Free Particle in D Dimensions
9.2 Harmonic Oscillator in D Dimensions
9.3 Free Particle from w --+ 0 -Limit of Oscillator.
9.4 Charged Particle in Uniform Magnetic Field
9.5 Dirac a-Function Potential
Notes and References
10 Spaces with Curvature and Torsion
10.1 Einstein's Equivalence Principle
10.2 Classical Motion of Mass Point in General Mctric·Affi ne Space
10.2. 1 Equations of Motion
10. 2.2 Nonholonomic Mapping to Spaces with Torsion
10.2.3 New Equivalence Principle.
10.2.4 Classical Action Principle for Spaces with Curvature and Torsion
10.3 Path Integral in Metric·Affine Space
10.3.1 Nonholonomic Transformation of Action
10.3.2 Measure of Path Integration.
10.4 Completing the Solution of Path Integral on Surface of Sphere
10.5 External Potent ials and Vector Potentials
10.6 Perturbative Calculat ion of Path Integrals in Curved Space
10.6.1 Free and Interacting Part s of Action
10.6.2 Zero Temperature
10.7 Model Study of Coordinate Invariance
10.7.1 Diagrammatic Expansion
10.7.2 Diagrammatic Expansion in d Time Dimensions.
10.8 Calculating Loop Diagrams
10.8.1 Reformulation in Configuration Space
10.8.2 Integrals over Products of Two Distributions
10.8.3 Integrals over Products of Four Distributions
10.9 Distributions as Limits of Bessel Function
10.9.1 Correlation Function and Derivatives.
10.9.2 Integrals over Products of Two Distributions
10.9.3 Integrals over Products of Four Distributions
10.10 Simple Rules for Calculating Singular Integrals
10.11 Perturbative Calculation on Finite Time Intervals
10.11.1 Diagrammatic Elements
10.11.2 Cumulant Expansion of D-Dimensional Free-Particle Amplitude in Curvilinear Coordinates
10.1 1.3 Propagator in 1– Time Dimensions
10.1 1.4 Coordinate Independence for Dirichlet Boundary Conditions
10.11.5 Time Evolution Amplitude in Curved Space
10.1 1.6 Covariant Results for Arbit rary Coordinates.
10.12 Effective Classical Potential in Curved Space
10.12. 1 Covariant Fluctuation Expansion
10.12.2 Arbitrariness of qb
10.12.3 Zero-Mode Properties.
10.12.4 Covariant Perturbation Expansion
10.12.5 Covariant Result from Noncovariant Expansion
10.12.6 Particle on Unit Sphere
10.13 Covariant Effective Action for Quantum Particle with Coordinate- Dependent Mass
10.13.1 Formulating the Problem
10.13.2 Gradient Expansion
Appendix 10A Nonholonomic Gauge Transformations in Electromagnetism
10A.1 Gradient Representation of Magnetic Field of Current Loops
10A.2 Generating Magnetic Fields by Multivalued Gauge TrIlJ1&- formations
10A.3 Magnetic Monopoles.
10A.4 Minimal Magnetic Coupling of Particles from Multivalued Gauge Transformations
10A.5 Gauge Ficld Representation of Current Loops and Monopoles
Appendix 10B Comparison of Multivalucd Basis Tetrads with Vierbein Fields
Appendix 10C Cancellation of Powers of 5(0)
Notes and References
11 Schrodinger Equation in General Metric-Affine Spaces
11.1 Integral Equation for Time Evolution Amplitude.
11.1.1 From Recursion Relation to Schrodinger Equation.
11. 1.2 Alternative Evaluation
11.2 Equivalent Path Integral Representations
11.3 Potentials and Vcctor Potentials.
11.4 Unitarity Problem.
11.5 Alternative Attempts
11.6 DeWitt--Seeley Expansion of Time Evolution Amplitude.
Appendix 11A Cancellations in Effective Potential.
Appendix 11B DeWitt's Amplitude.
Notes and References
12 New Path Integral Formula for Singular Potentials
12.1 Path Collapse in Feynman's formula for the Coulomb System
12.2 Stable Path Integral with Singular Potentials
12.3 Time-Dependent Regularization
12.4 Relation to Schr&lingcr Theory. Wave Functions.
Notes and References
13 Path Integral of Coulomb System
13.1 Pseudotime Evolution Amplitude
13.2 Solution for the Two-Dimensional Coulomb System
13.3 Absence of Time Slicing Corrections for D = 2
13.4 Solution for the Three-Dimensional Coulomb System
13.5 Absence of Time Slicing Corrections for D = 3
13.6 Geometric Argument for Absence of Time Slicing Corrections
13.7 Comparison with Schrodinger Theory
13.8 Angular Decomposition of Amplitude, and Radial Wave Functions
13.9 Remarks on Geometry of Four-Dimensional ul"-Space
13.10 Runge-Lenz-Pauli Group of Degeneracy.
13.11 Solution in Momentum Space
13.11.1 Another Form of Action
Appendix 13A Dynamical Group of Coulomb States.
Notes and References.
14 Solution of Further Path Integrals by Duru-Kleinert Method
14.1 One-Dimensional Systems
14.2 Derivation of the Effective Potential.
14.3 Comparison with SchrOdinger Quantum Mechanics.
14.4 Applications
14.4. 1 Radial Harmonic Oscillator and Morse System
14.4.2 Radial Coulomb System and Morse System
14.4.3 Equivalence of Radial Coulomb System and Radial Oscilla.- tor
14.4.4 Angular Barrier ncar Sphere, and Rosen-Morse Potential
14.4.5 Angular Barrier near Four-Dimensional Sphere, and General Rosen-Morse Potential
14.4.6 Hulthen Potential and General Rosen-Morse Potential
14.4.7 Extended Hulthen Potential and General Rosen-Morse Potential
14.5 D-Dimensional Systems.
14.6 Path Integral of the Dionium Atom
14.6. 1 Formal Solution
14.6.2 Absence of Time Slicing Corrections
14.7 Time-Dependent Duru-Kleinert 'Transformation
Appendix 14A Affine Connection of Dionium Atom
Appendix 14B Algebraic Aspects of Dionium States.
Notes and References
15 Path Integrals in Polymer Physics
15.1 Polymers and Ideal Random Chains.
15.2 Moments of End-to-End Distribution
15.3 Exact End-ta-End Distribution in Three Dimensions.
15.4 Short- Distance Expansion for Long Polymer
15.5 Saddle Point Approximation to Three-Dimensional End-to-End Distribution
15.6 Path Integral for Continuous Gaussian Distribution
15.7 Stiff Polymers
15.7.1 Sliced Path Integral
15.7.2 Relation to Classical Heisenberg Model.
15.7.3 End-to-End Distribution
15.7.4 Momcnts of End-to-End Distribution
15.8 Continuum Formulation
15.8.1 Path Integral
15.8.2 Correlation Functions and Moments
15.9 SchrOdinger Equation and Recursive Solution for Moments
15.9.1 Setting up the SchrOdinger Equation.
15.9.2 Recursive Solution of SchrOdingcr Equation.
15.9.3 From Moments to End-to-End Distribution for D = 3
15.9.4 Large-Stiffness Approximation to End-to-End Distribution
15.9.5 Higher Loop Corrections
15.10 Excluded-Volume Effects
15.11 Flory's Argument
15.12 Polymer Field Theory.
15.13 Fermi Fields for Self-Avoiding Lines
Appendix 15A Basic Integrals
Appendix 15B Loop Integrals
Appendix 15C Integrals Involving Modified Green Function
Notes and References
16 Polymers and Particle Orbits in Multiply Connected Spaces
16.1 Simple Model for Entangled Polymers.
16.2 Entangled Fluctuating Particle Orbit: Aharonov-Bohm Effect
16.3 Aharonov-Bohm Effect and Fractional Statistics
16.4 Self-Entanglement of Polymer
16.5 The Gauss Invariant of Two Curves
16.6 Bound States of Polymers and Ribbons
16.7 Chern-Simons Theory of Entanglements
16.8 Entangled Pair of Polymers
16.8.1 Polymer Field Theory for Probabilities
16.8.2 Calculation of Partition Function
16.8.3 Calculation of Numerat or in Second Moment
16.8.4 First Diagram in Fig. 16.23
16.8.5 Second and Third Diagrams in Fig. 16.23
16.8.6 Fourt h Diagram in Fig. 16.23
16.8. 7 Second Topological Moment
16.9 Chcrn-Simons Theory of Statistical Interaction
16.10 Second-Quantized Anyon Fields
16.11 Fractional Quantum Hall Effect
16.12 Anyonic Superconductivity
16.13 Non-Abelian Chern-Simons Theory
Appendix 16A Calculation of Feynman Diagrams in Polymer Entanglement
Appendix 168 Kauffman and 8LM/Ho polynomials
Appendix 16C Skein Relation between Wilson Loop Integrals
Appendix 160 London Equations
Appendix 16E Hall Effoct in Eioctean Gas
Notes and References
17 Tunneling
17. 1 Double-Well Potential
17.2 Classical Solutions — Kinks and Antikinks
17.3 Quadratic Fluctuations
17.3.1 Zero-Eigenvalue Mode
17.3.2 Continuum Part of Fluctuation Factor
17.4 General Formula for Eigenvalue Ratios
17.5 Fluctuation Determinant fro m Classical Solution.
17.6 Wave Functions of Double-Well
17.7 Gas of Kinks and Antikinks and Level Splitting Formula
17.8 Fluctuation Correction to Level Splitting
17.9 Thnneling and Decay
17.10 Large-Order Behavior of Perturbation Expansions
17.10. 1 Growth Properties of Expansion Coefficients.
17.10.2 Semiclassical Large-Order Behavior.
17.10.3 Fluctuation Correction to the Imaginary Part and Large- Order Behavior.
17.10.4 Variational Approach to Tunneling. Perturbation Coefficients to All Orders
17.10.5 Convergence of Variational Perturbation Expansion.
17.11 Decay of Super current in Thin Closed Wire.
17.12 Decay of Metastable Thermodynamic Phases.
17.13 Decay of Metastable Vacuum State in Quantum Field Theory
17.14 Crossover from Quantum Tunneling to Thermally Driven Decay
Appendix 17A Feynman Integrals for Fluctuation Correction.
Notes and References.
18 Nonequilibrium Quantum Statistics
18.1 Linear Response and Time-Dependent Green Functions for T = 0
18.2 Spectral Representations of Green Functions for T = 0
18.3 Other Important Green Functions
18.4 Hermitian Adjoint Operators.
18.5 Harmonic Oscillator Green Functions for T = 0
18.5. 1 Creation Annihilation Operators
18.5.2 Real Field Operators.
18.6 Nonequilibrium Green Functions.
18.7 Perturbation Theory for Nonequilibrium Green Functions
18.8 Path Integral Coupled to Thermal Reservoir
18.9 Fokker-Planck Equation
18.9.1 Canonical Path Integral for Probability Distribution
18.9.2 Solving the Operator Ordering Problem
18.9.3 Strong Damping
18.10 Langevin Equations
18.11 Path Integral Solution of Klein-Kramers Equation
18.12 Stochastic Quantization
18.13 Stochastic Calculus
18. 13.1 Kubo's stochastic Liouville equation
18.13.2 From Kubo's to Fokker-Planck Equations
18.13.3 Ito's Lemma
18.14 Solving the Langevin Equation.
18.15 Heisenberg Picture for Probability Evolution
18.16 Supcrsymmetry
18.17 Stochastic Quantum Liouville Equation
18.18 Master Equation for Time Evolution
18. 19 Relation to Quantum Langevin Equation
18.20 Electromagnetic Dissipation and Decoherence
18.20.1 Forward- Backward Path Integral
18.20.2 Master Equation for Time Evolution in Photon Bath
18.20.3 Line Width
18.20.4 Lamb shift
18.20. 5 Langevin Equations
18.21 Fokker-Planck Equation in Spaces with Curvature and Torsion
18.22 Stochastic Interpretation of Quantum-Mechanical Amplitudes
18.23 Stochastic Equation for SchrOdingcr Wave Function
18.24 Real Stochastic and Deterministic Equation for SchrOdinger Wave Function
18.24.1 Stochastic Differential Equation
18.24.2 Equation for Noise Average
18.24.3 Harmonic Oscillator
18.24.4 General Potential.
18.24.5 Deterministic Equation
Appendix 18A Inequalities for Diagonal Green Functions
Appendix 18B General Generating Functional
Appendix 18C Wick Decomposit ion of Operator Products
Notes and References
19 Relativistic Particle Orbits
19.1 Special Features of Relativistic Pat h Integra1s
19. 1.1 Simplest Gauge Fixing.
19.1.2 Partition Function of Ensemble of Closed Particle Loops
19. 1.3 Fixed-Energy Amplitude.
19.2 Thnneling in Relativistic Physics.
19.2. 1 Decay Rate of Vacuum in Electric Field
19.2.2 Birth of Universe
19.2.3 Friedmann Model
19.2.4 TUnneling of Expanding Universe
19.3 Relativistic Coulomb System
19.4 Relativistic Particle in Electromagnetic Field.
19.4. 1 Action and Partition Function
19.4.2 Perturbation Expansion
19.4.3 Lowest-Order Vacuum Polarization
19.5 Path Integral for Spin-l/2 Particle.
19. 5.1 Dirac Theory
19. 5.2 Path Integral
19.5.3 Amplitude with Electromagnetic Interaction.
19.5.4 Effective Action in Electromagnetic Field
19.5.5 Perturbation Expansion
19.5.6 Vacuum Polarization
19.6 Supersymmetry
19.6. 1 Global Invariance.
19.6.2 Local Invariance
Appendix 19A Proof of Same Quantum Physics of Modified Action
Notes and References
20 Path Integrals and Financial Markets
20.1 Fluctuation Properties of Financial Assets
20.1.1 Harmonic Approximation to Fluctuations
20.1.2 Levy Distributions.
20.1.3 Truncated Levy Distributions
20.1.4 Asymmetric Truncated Levy Distributions.
20.1.5 Gamma Distribution
20.1.6 Boltzmann Distribution
20.1.7 Student or Tsallis Distribution
20.1.8 Tsallis Distribution in Momentum Space
20.1.9 Relativistic Particle Boltzmann Distribution.
20.1.10 Meixner Distributions
20.1.11 Generalized Hyperbolic Distributions
20.1.12 Debye-Waller Factor for Non-Gaussian Fluctuations
20.1.13 Path Integral for Non-Gaussian Distribution.
20.1.14 Time Evolutio n of Distribution
20.1.15 Central Limit Theorem
20.1.16 Additivity Property of Noises and Hamiltonians.
20.1.17 Uvy-Khintchine Formula
20.1.18 Semigroup Property of Asset Distributions.
20.1.19 Time Evolutio n of Moments of Distribution
20.1.20 Boltzmann Distribution
20.1.21 Fourier-Transformed Tsallis Distribution.
20.1.22 Superposition of Gaussian Distributions
20.1.23 Fokker-Planck-Type Equation
20.1.24 Kramcrs-Moyal Equation
20.2 ito-like Formula for Non-Gaussian Distributions
20.2.1 Continuous Time.
20.2.2 Discrete Times
20.3 Martingales.
20.3.1 Gaussian Martingales
20.3.2 Non-Gaussian Martingale Distributions
20.4 Origin of Semi-Heavy Tails.
20.4.1 Pair of Stochastic Differential Equations
20.4.2 Fokker-Planck Equation
20.4.3 Solution of Fokker-Planck Equation.
20.4.4 Pure x-Distribution
20.4.5 Long-Time Behavior.
20.4.6 Tail Behavior for all Times
20.4.7 Path Integral Calculation
20.4.8 Natural Martingale Distribution
20.5 Time Series
20.6 Spectral Decomposition of Power Behaviors.
20.7 Option Pricing
20.7.1 Black-Scholes Option Pricing Model
20.7.2 Evolution Equations of Portfolios with Options
20.7.3 Option Pricing for Gaussian Fluctuations
20.7.4 Option Pricing for Boltzmann Distribution.
20.7.5 Option Pricing for General Non-Gaussian Fluctuations
20.7.6 Option Pricing for Fluctuating Variance.
20.7.7 Perturbation Expansion and Smile
Appendix 20A Large-x Behavior of Truncated Levy Distribution
Appendix 20B Gaussian Weight..
Appendix 20C Comparison with Dow-Jones Data
Notes and References.
Index
Recommend Papers

Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets [5 ed.]
 9814273562, 9789814273565

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

I' ... . . .

PATH INTEGRALS in Quantum Mechanics, Statistics, Polynter Physics, and Financ ial Markets 5th Edition

-I

I

7305tp.kim.Q2.09.Is.indd 1

3J3O,OO

10;-4{I:14 AM

I 1-

This page intentionally left blank

I' ... . ..

• • ••

• •

PATH INTEGRALS in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets 51h Edilion

Hagen KLEINERT Professor of Physics freie Universirot Berlin, Germany !CRANe! Pescara, /mfy, and Nice, France

,~ World Scientific NeW JERSEY · LONDO N· SI NGAPORE · BEIJI NG · SHANGHAI · HO NG KONG · TAIPEI · CHE NN AI

-I

I

7305tp.kim.Q2.09.Is.indd 2

3J3O,OO

10;-4{I:14 AM

I 1-

Published by World Scientific Publishing Co. Pte. Ltd. 5 Toh Tuck Link, Singapore 596224 USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601 UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

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

PATH INTEGRALS IN QUANTUM MECHANICS, STATISTICS, POLYMER PHYSICS, AND FINANCIAL MARKETS (5th Edition) Copyright © 2009 by World Scientific Publishing Co. Pte. Ltd. All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from the publisher.

ISBN-13 978-981-4273-55-8 ISBN-13 978-981-4273-56-5 (pbk)

Printed in Singapore by Mainland Press Pte Ltd.

Kim - Path Integrals (5th Edn).pmd

1

3/31/2009, 4:21 PM

To Annemarie and Hagen 11

This page intentionally left blank

Nature alone knows what she wants . GOETHE

Preface The paperback version of the fourth edition of this book was sold out in Fall 2008. This gave me a chance to revise it at many places. In particular, I improved considerably Chapter 20 on fi nancial markets and removed some technical sections of Chapter 5. Among the many people who spotted printing erron; and suggested changes of text passages are Dr. A. Pelster, Dr. A. Redondo, and especially Dr. Annemarie Kleinert .

H. Kleinert Berlin, January 2009

vii

Preface to Fourth Edition The third edition of t his book appeared in 2004 and was reprinted in the same year without improvements. The present fourt h edition contains several extensions. Chapter 4 includes now semiclassical expansions of higher order. Chapter 8 offers an additional path integral formulation of spinning particles whose action contains a vector field and a Wess-Zumino term. From this, the Landau-Lifshitz equation for spin precession is derived which governs the behavior of quantum spin liquids. The path integral demonstrates that fermions can be described by Bose fields- the basis of Skyrmion theories. A further new section introduces the Berry phase, a useful tool to explain many interesting physical phenomena. Chapter to gives more details on magnetic monopoles and multivalued fields . Another feature is new in t his edition: sections of a more technical nature are printed in smaller font size. They can well be omitted in a first reading of the book. Among the many people who spott ed printing errors and helped me improve various text passages arc Dr. A. Chervyakov, Dr. A. Pelster, Dr. F. Nogueira, Dr. M. Weyrauch , Dr. H. Ba ur, Dr. T . Iguchi, V. Bezerra, D. Jahn , S. Overesch, and especially Dr. Annemarie Kleinert.

H. Kleinert Berlin, June 2006

viii

Preface to Third Edition This t hird edition of the hook improves and extends considerably the second edition of 1995: • Chapter 2 now contains a path integral representation of t he scattering amplitude and new methods of calculating functional determinants for timedependent second-order d ifferential operators. Most importantly, it introduces the quantum field-theoretic definition of path integrals, based on pert ur bation expansions around t he trivial harmonic theory.

• Chapter 3 presents morc exactly solvable path integrals than in the previous editions. It also extends the Bcnder-Wu recursion relations for calculating perturbation expansions to morc general types of potentials. • Chapter 4 discusses now in detail the quasiclassical approximation to the scattering a mplitude and Thomas-Fermi approximation to atoms. • Chapter 5 proves the convergence of variational perturba tion theory. It also discusses atoms in strong magnetic fields and the polaron problem. • Chapter 6 shows how to obtain the spectrum of systems with infinitely high walls from perturbation expansions. • Chapter 7 offers a many-path t reatment of Boso-Einst ein condensation and degenerate Fermi gases. • Chapter 10 develops the quantum theory of a particle in curved space, treated before only in the time-sliced formalism, to perturbatively defined path integrals. Their reparametrization invariance imposes severe constraints upon integrals over products of distributions. We derive unique rules for evaluating these integrals, thus extending the linear space of distributions to a semigroup. • Chapter 15 offers a closed expression for the end-to-end distribution of stiff polymers valid for all persistence lengths . • Chapter 18 derives the operator Langevin equation a nd t he Fokker-Planck equation from the forward- backward path integral. The derivation in the literature was incomplet e, and the gap was elosed only roccntly by an elegant calculation of the J acobian functional determinant of a second-order d ifferential operator with dissipation. ix

x • Chapter 20 is completely new. It introduces t he reader into t he applications of path integrals to the fascinating new field of econophysics. For a few years, the third edition has been freely available on the internet , and several readers have sent useful comments, fo r instance E. Babaev, H. Baur, B. Budnyj , Chen Li-ming, A.A. Dragulescu, K. Claum, L Grigorenko, T .S. Hatamian, P. Hollist er, P. Jizba, B. Kastening, M. Kramer, W.-F. Lu, S. Mukhin, A. Pelster, C. Dpllr, M.B. Pinto, C. Schubert, S. Schmidt, R. Scalettar, C. Tangui, and M. van Vugt. Re ported errors are corrected in the internet edition. When writing the new part of Chapter 2 on the path integral representation of t he scattering amplitude I profited from discussions with R. Rosenfelder. In t he new parts of Chapter 5 on poiarons, many useful comments came from J .T . Devreese, F.M. Peeters, and F. Brosens. In the new Chapter 20, I profited from discussions with F. Nogueira, A.A . Dragulescu, E. Eberlein, J . Kallsen, M. Schweizer, P. Bank, M. Tenney, and E.C. Chang. As in all my books, many printing errors were detected by my secretary S. Endrias and many improvements are due to my wife Annemarie without whose permanent encouragement this book would never have been fi nished.

H. Kleinert Berlin, August 2003

Preface to Second Edition Since this book first a ppeared three years ago, a number of important developments have taken place calling for various extensions to the text. Chapter 4 now contains a discussion of the feat ures of the semiclassical quantization which are relevant for multidimensional chaotic systems. Chapter 3 derives perturbation expansions in terms of Feynman graphs, whose usc is customary in quantum field theory. Correspondence is established with Rayleigh-SchrOdinger perturbation theory. Graphical expansions are used in Chapter 5 to extend t he Fcynman-Klcinert variat ional approach into a systematic variational perturbation theory. Analytically inaccessible path integrals can now be evaluated with arbitrary accuracy. In contrast to ordinary perturbation expansions which always diverge, the new expansions are convergent for all coupling strengths, including the strong-coupling limit. Chapter 10 contains now a new action principle which is necessary to derive the correct classical equations of motion in spaces with curvature and a certain class of torsion (gradicnt torsion). Chapter 19 is new. It deals with relativistic path integrals, which were previously discussed only briefly in two sections at the end of Chapter 15. As an application, the path integral of the relativistic hydrogen a tom is solved.. Chapter 16 is extended by a theory of particles with fract ional statistics (anyons), from which I develop a theory of polymer entanglement. For t his r introduce nonabelian Chern-Simons fields and show their relationship with various knot polynomia ls (Jones, HOMFLY). The successful explanation of the fractio nal quantum Hall effect by anyon theory is discussed - also the failure to explain high-temperature superconductivity via a Chern-Simons interaction. Chapter 17 offers a novel variational approach to tunneling amplitudes. It extend;; the semiclassical range of validity from high to low barriers. As an application, I increase the range of validity of the currently used large-order perturbation theory far into the regime of low orders. This suggests a possibility of greatly improving existing resummation procedures for divergent perturbation series of quantum fie ld theories. The Index now also contains the names of authors cited in the text. This may help the reader searching for topics associated with t hese names. Due to their great number, it was impossible to cite all the authors who have made important contributions. I apologize to all those who vainly search for t heir names. xi

xii In writing t he new sections in Chapters 4 and 16, discussions with Dr. D. Wintgen and , in particular, Dr. A. Schakel have been extremely useful. I also thank Professors G. Gerlich, P. Hanggi, H. Grabert, M. Roncadelli, as well as Dr. A. Pelster, and Mr. R. Karrlein for many relevant comments. Printing errors were corrected by my secretary Ms. S. Endrias and by my editor Ms. Lim Feng Nee of World Scientific. Many improvements arc duc to my wife Anncmarie.

H. Kleinert Berlin, December 1994

Preface to First Ed ition These are extended lecture notes of a course on path integrals which I delivered at the Freie Universitat Berlin during winter 1989/1990. My interest in this subject dates back to 1972 when t he late R. P. Feynman drew my attention to the unsolved path integral of the hydrogen atom. I was then 5pending my sabbatical year at Caltech, where Feynman told me during a discussion how embarrassed he was, not being able t o solve t he path integral of this moot fundament al quantum system. In fact, t his had made him quit teaching t his subject in his course on quantum mechanics as he had initially done.! Feynman challenged me: "Kleinert, you figured out all that grouptheoretic stuff of the hydrogen atom, why don't you solve t he path integraW He was referring to my 1967 Ph.D. thesis2 where I had demonstrated that all dynamical questions on the hydrogen atom could be answered using only operations within a dynamical group 0(4 ,2). Indeed, in that work, t he four-dimensional oscillator played a crucial role and the missing steps to the solution of the path integral were latcr found to be very few. After returning to Berlin, I forgot about the problem since I was busy applying path integrals in another context, developing a field-theoretic passage from quark theories to a collective field t heory of hadrons. 3 Later, I carried these techniques over into condensed matter (superconductors, superfluid 3He) and nuclear physics. Path integrals have made it possible to build a unified field theory of collective phenomena in quite different physical systems. 4 The hydrogen problem came up again in 1978 as I was teaching a course on quantum mechanics. To explain the concept of quantum fluctuations, I gave an introduction to path integrals. At t he same time, a postdoc from Turkey, I. H. Duru, joined my group as a Humboldt fellow. Since he was fam iliar with quantum mechanics, I suggested that we should try solving t he path integral of the hydrogen atom. He quickly acquired the basic techniques, and soon we found the most important ingredient to the solution: The transformation of time in the path int egral to a new path-dependent pseudotime, combined with a transformation of the coordinates to i Quoting from t he preface of the textbook by n.p. Feynman and A.n. Hibb6, Quantum Mechanics and Path Integrals, 1IIcGraw-HiU, New York, 1965: "Over t he succeeding years, . Dr. Fcynman's approach to teaching thc subject of quantum mechanics evolved Il()mcwhat a.way from t he initial path integral approoch .~ 2H. Kleinert, Fortschr. Phys. 6, 1, (1968), and Group Dynamics of the Hydrogen Atom, Lectures presented at the 19671loulder Summer School, published in Lectures in Theoretical Physics, Vol. X B, Pi>. 427-482, ed. by A.O. Barut and W.E. Brittin, Gordon and Breach, New York, 1968. 3Scc my 1976 Erice lectures, Hadronization of Quark Theories, publ ished in Understanding the Fundamental Constituenl.i! of Matter, Plenu m press, New York, 1978, p. 289, ed. by A. Zichichi. 4H. Kleinert, Phys. Leu. B 69, 9 (1977); Fortschr. Phys. 26, 565 (1978); 30, 187, 351 (1982).

xiii

xiv "square root coordinates" (to be explained in Chapters 13 and 14).5 These transformat ions led to t he correct result, however, only due to good fort une. In fact, our procedure was immediately criticized for its sloppy treatment of the time slicing.6 A proper treatment could, in principle, have rendered unwanted extra terms which our treatment would have missed . Other authors went t hrough the detailed timeslicing procedure,1 but the correct result emerged only by transforming the measure of path integration inconsistently. When I calculated t he extra terms according to the standard rules I found them to be zero only in two spacc dimensions. s The same t reatment in three dimensions gave nonzero "corrections" which spoiled the beautiful result, leaving me puzzled. Only recently I happened to loca te t he place where the t hree-dimensional treatment went wrong. I had just finished a book on t he use of gauge fields in condensed matter physics. 9 The second volume deals with eru;embles of defects which are defi ned and classified by means of operational cutting and pasting procedures on an ideal crystal. Mathematically, t hese procedures correspond to nonholonomic mappings. Geometrically, they lead from a flat space to a space with curvature and torsion. While proofreading that book, I realized t hat t he transformation by which t he path integral of t he hydrogen atom is solved also produces a certain type of torsion (gradient torsion). Moreover, this happens only in three dimensions. In two dimensions, where the t ime-sliced path integral had been solved without problems, torsion is absent. Thus I realized that t he transformat ion of the time-sliced measure had a hitherto unknown sensit ivity to torsion. It was t herefore essential to find a. correct path integral for a. particle in a space with curvature and gradient torsion . This was a nontrivial task si nce the literature was ambiguous already for a purely curved space, offering several prescriptions to choose from. The corresponding equivalent SchrOdinger equations differ by multiples of the curvature scalar. 10 T he ambiguities are path integral analogs of t he so-called opemtor-omering problem in quantum mechanics. When trying to apply t he existing prescript ions to spaces with torsion, I always ran into a disaster, some even yielding noncovariant answers. So, something had to be wrong with all of t hem. Guided by t he idea that in spaces with constant curvature the path integral should produce t he same result as an operator quantum mechanics based on a quantization of angular momenta, I was eventually able to find a consistent quantum equivalence principle 5I. H. Duru and H. Kleinert, P hys. Lett. B 84 , 30 (1 979), Fortschr. Phys. 30 , 401 (1982). 6C.A. Ringwood and J.T. Devreese, J . Math. Phys. 21, 1390 (1980) . 1R. Ho and A. inomata, Phys. Rev. Lett. 48, 231 (1982); A. inomata, Phys. Lett. A 87, 387 (1981). 8H. Kleinert, Phys. Lett. B 189, 187 (1987); contains also a crit icism of Ref. 7. 9H. Kleinert, Gauge FieltJ.j in Condensed Matter, World Scientific, Singapore, 1989, Vol. i, pp. 1- 744, Super/fow and Vortex Lines , and Vol. II, pp. 745- 1456, Stre8ses and Defects. 10B.S. DeWitt, Rev. Mod. Phys. 1!9, 377 (1957); K.S. Cheng, J. Math. P hys. 13, 1723 (1 972), H. Kamo and T. Kawai, Prog. Thoor. Phys. 50 , 680, (1973); T . Kawai , Found. P hys. 5, 143 (1975), H. Dekker, Physica A 103, 586 (1980), C.M. Cavazzi, Nuovo Cimento 101 A, 241 (1 981); M.s. Marinov, Physia; Reports 60, I (1 980) .

xv for path integrals in spaces with curvature and gradient torsion,1l thus offeri ng also a unique solution to the operator-orderi ng problem. T his was the key to the leftover problem in the Coulomb path integral in three dimensions - the proof of the absence of t he extra time slicing contributions prcscnted in C hapter 13. Chapter 14 solves a variety of one-dimensional systems by the new techniques. Special emphasis is given in Chapter 8 t o instability (path collapse) problems in the Euclidean version of Feynman's time-sliced path integraL These arise for actions containing bottomless potentials. A general stabilization procedure is developed in Chapter 12. It must be applied whenever centrifugal barriers, angular barriers, or Coulomb potentials are prescntYl Another project suggested to me by Feynman, the improvement of a variational approach to path integrals explained in his book on statistical mecha nics l 3 , found a faster solution. We started work d uring my sabbatical stay at the University of California at Santa Barbara in 1982. After a few meetings and discussions, the problem was solved and the preprint drafted. Unfortunately, Feynman 's illness prevented him from reading the fin al proof of t he paper. He was able to do this only t hree years later when T came to the University of California at San Diego for another sabbatical leave. Only then could the paper be submitted. 14 Due to recent int erest in lattice theories, I have found it useful t o exhibit the solution of several path integrals for a finite number of time slices, without going immediately to the continuum limit . T his should hclp identify typical lattice effects seen in t he Monte Carlo simulation data of various systems. The path integral description of polymers is introduced in Chapter 15 where stiffness as well as the famous excluded-volume problem are discussed. P arallels are drawn to path integrals of relativistic particle orbits. This chapter is a preparation for ongoing research in t he theory of fluctuating tmrfaces with extrinsic curvature stiffness, and their application to world sheets of strings in particle physics. I:> I have a lso int roduced the field-t heoretic description of a polymer to account for its increasing relevance to the understanding of various phase transitions driven by fluctuating line-like excitations (vortex lines in superfluicls and superconductors, defect lines in crystals and liquid crystals) .16 Special attention has been devoted in C ha pter 16 to simple topological questions of polymers and particle orbits, the latter arising by the presence of magnetic flux tubes (Aharonov-Bohm effect) . Their relationship to Bose and Fermi statistics of particles is pointed out and the recently popular top ic of fractional statistics is introduced. A survey of entanglement phenomena of single orbits and pairs of t hem (ribbons) is given and their application to biophysics is indicated . I1 H. Klei nert, Mod. P hys. Lett. A 4, 2329 (1 989); Phys. Lett. B 296, 315 (1990). 12H. Kleinert, Phys. Lett. B 224,313 (1989). 13R.p. Feynman, Statistical Mechanics, Benja.min, Read ing, 1912, Section 3.5. t° R.P. Feynman and H. Kleinert, Phy:l. Rev. A 34,5080, (1986). IS A.M. Polyakov, Nuel. P hys. B 268, 406 (1986), H. Kleinert , Phy:l. Lett. B 174,335 (1986). 16Sec Ref. 9.

xvi

Finally, Chapter 18 contains a brief introduction to the path integral approach of nonequilibrium quantum-statistical mechanics, deriving from it the standard Langevin and Fokker-Planck equat ions. I want to thank several students in my class, my graduate students, and my postdocs for many useful discussions. In particular, T. Eris, F. Langhammer, B. Meller, I. Mustapic, T . Sauer, L. Semig, J. Zaun, and Drs. G. German, C. Holm, D. Johnston, and P. Kornilovitch have all cont ributed with constructive criticism. Dr. U. Eckern from Karlsruhe University clarified some points in the path integral derivation of the Fokker-Planck equation in Chapter 18. Useful comments are due to Dr. P.A . Horvathy, Dr. J . Whit enton, and to my colleague Prof. W. Theis. T heir careful reading uncovered many shortcomings in the first draft of t he manuscript. Special t hanks go to Dr. W. Janke with whom I had a fert ile collaboration over the years and many discussions on various aspect s of path integration. T hanks go also to my secretary S. Endrias for her help in preparing the manuscript in WTBX, thus making it readable at an early stage, and to U. Grimm for drawing the figures. Finally, a nd most importantly, I am grateful to my wife Dr. Annemarie Kleinert for her inexhaustible patience and constant encouragement.

H. Kleinert Berlin, January 1990

Contents

Preface

vii

Preface to Fourth Edition

viii

Preface to Third Edition

ix

Preface to Second Edition

xi

Preface to First Edition

xiii

1 Fundamentals 1. 1 Classical Mechanics 1.2 Relativistic Mechanics in Curved Spacetime l.3 Quantum Mechanics 1.3.1 Bragg Reflections and Interference 1.3.2 Matter Waves . SchrOdinger Equation 1.3.3 1.3.4 Particle Current Conservation . 1.4 Dirac's Bra- Ket Formalism. 1.4.1 Basis Transformations 1.4.2 Bracket Notation 1.4.3 Continuum Limit . . . 1.4.4 Generalized Functions 1.4.5 SchrOdinger Equation in Dirac Notation 1.4.6 Momentum States 1.4.7 Incompleteness and Poisson's Summation Formula 1. 5 Obscrvables 1.5.1 Uncertainty Relation. 1.5.2 Density Matrix and Wigner Function. 1.5.3 Generalization to Many Particles 1.6 Time Evolution Operator . l.7 Properties of the Time Evolution Operator 1.8 Heisenberg Picture of Quant um Mechanics l.9 Interaction Picture and Perturbation Expansion 1.10 Time Evolution Amplitude 1.11 Fixed-Energy Amplitude xvii

1 1 10 11

12 13 15 17 18 18

20 22 23 25 26 28 31 32 33 34 34 37 39

42 43 45

xviii 1.12 1.13 1.14 1.15 1.16

Free-Particle Amplitudes. . . . . . . . . . . . . . . . Quantum Mechanics of General Lagrangian Systems. Particle on the Surface of a Sphere Spinning Top . . . . . . Scattering 1. 16.1 Scattering Matrix 1.16.2 Cross Section . . . 1.16.3 Born Approximation 1. 16.4 Partial Wave Expansion and Eikonal Approximation 1. 16.5 Scattering Amplitude from Time Evolution Amplitude 1. 16.6 Lippmann-Schwinger Equation 1.17 Classical and Quantum Statistics 1. 17.1 Canonical Ensemble. . 1. 17.2 Grand-Canonical Ensemble 1.18 Density of States and Ttacelog . Appendix l A Simple Time Evolution Operator. Appendix I B Convergence of the Fresnel Integral Appendix l C The Asymmetric Top Notes and References. . . . . . . . . . . . . . . .

47 51 57 59 67 67 68 70 70 72 72 76 77 77 82 84 84 85 87

2 P ath Integrals - E le mentary Properties and Simple Solutions 2.1 Path Integral Representation of Time Evolution Amplitudcs 2.1.1 Sliced Time Evolution Amplitude. 2.1.2 Zero-Hamiltonian Path Integral. . 2.1.3 SchrOdinger Equation for Time Evolution Amplitude 2.1.4 Convergence of of the Time-Sliced Evolution Amplitude 2.1.5 Time Evolution Amplitude in Momentum Space. 2.1.6 Quantum-Mechanical Partition Function. . . 2.1.7 Feynman's Configurat ion Space Pat h Integral 2.2 Exact Solution for the Frec Particle 2.2. 1 Direct Solution 2.2.2 Fluctuations around the Classical Path . 2.2.3 Fluctuation Factor . . . . . . . . . . . . 2.2.4 Finite Slicing Properties of Free-Particle Amplitude. 2.3 Exact Solution for Harmonic Oscillator . 2.3.1 Fluctuations around the Classical Path. . . . 2.3.2 Fluctuation Factor . 2.3.3 The i7]-Prescription and Maslov-Morse Index 2.3.4 Continuum Limit. . . . . . . . . . . . . 2.3.5 Useful Fluctuation Formulas. 2.3.6 Oscillator Amplitude on Finite Time Lattice . 2.4 Gelfand-Yaglom Formula. . . 2.4.1 Recursive Calculation of Fluctuation Detcrminant. 2.4.2 Examples

89 89 89 91 92 93 94 96 97 101 101 102 104 111 112 112 114 115 116 117 119 120 121 121

xix

2.5

2.6

2.7 2.8 2.9 2.10

2.11 2.12 2.13 2.14 2.15

2.16 2.17 2.18

2.19 2.20 2.21 2.22

2.23

2.4.3 Calculation on Unsliced Time Axis 123 2.4.4 D'Alembert 's Construction 124 2.4 .5 Another Simple Formula. 125 2.4.6 Generalization t o D Dimensions 127 Harmonic Oscillator with Time-Dependent Frequency 127 2.5.1 Coordinate Space. 128 2.5.2 Momentum Space .. . . . . . . 130 Free-Particle and Oscillator Wave Functions 132 General Time-Dependent Harmonic Action 134 Path Integrals and Quantum Statistics 135 Density Matrix 138 Quantum Statistics of the Harmonic Oscillator 143 Time-Dependent Harmonic Potential 148 Functional Measure in Fourier Space .. 151 154 Classical Limit. . . . . . . . . . . . . . . Calculation Techniques on Sliced Time Axis via the Poisson Formula 155 Ficld·Theorctic Definition of Harmonic Path Integrals by Analytic Regularization . . . . .. . . . . . . . . . . 158 2. 15.1 Zero--Temperature Evaluation of the Frequency Sum. 159 2. 15.2 Finite-Temperature Evaluation of the Frequency Sum. 162 2. 15.3 Quantum.Mechanical Harmonic Oscillator . 164 2. 15.4 Tracelog of the First·Order Differential Operator 165 2. 15.5 Gradient Expansion of the One-Dimensional Tracclog . 167 2. 15.6 Duality Transformation and Low· Temperature Expansion 168 Finite-N Behavior of Thermodynamic Quantities 175 Time Evolution Amplitude of Freely Falling Particle. 177 Charged Particle in Magnetic Field 179 2. 18.1 Action. 179 2. 18.2 Gauge Properties. 182 2. 18.3 Time-Sliced Path Integration 182 2. 18.4 Classical Action 184 2. 18.5 Translational Invariancc 185 Charged Particle in Magnetic Field plus Harmonic Potentia] 186 Gauge Invariance and Alternative Path Integral Representation 188 Velocity Path Integral. . . . . . . . . . . .. . . . . . 189 Path Integral Representation of the Scattering Matrix . 190 2.22.1 General Development 190 2.22.2 Improved Formulation . 193 2.22.3 Eikonal Approximation to the Scattering Amplitude 194 Heisenberg Operator Approach to Time Evolution Amplitude. 194 2.23.1 Free Particle . . . . . . . . . . . . 195 2.23.2 Harmonic Oscillator .. 197 2.23.3 Charged Particle in Magnetic Field 197

xx Appendix 2A Baker-Campbell-Hausdorff Formula and Magnus Expansion . 201 Appendix 28 Direct Calculation of the Time-Sliced Oscillator Amplitudc204 Appendix 2C Derivation of Mehler Formula. 205 Notes and References 206

3 External Sources, Correlations, and Perturbation Theory 209 3.1 External Sources .. . . . . . . . . . 209 3.2 Green Function of Harmonic Oscillator 213 3.2.1 Wronski Construction 213 3.2.2 Spectral Representation . . . . . 217 3.3 Green Functions of First-Order Differential Equation 219 3.3.1 Time-Independent Frequency 219 3.3.2 Time-Dependent Frequency . . . . . . . . . . . 226 3.4 Summing Spectral Representation of Green Function 229 3.5 Wronski Construction for Periodic and Antiperiodic Green F\mctions 231 3.6 Time Evolution Amplitude in Presence of Source Term 232 3.7 T ime Evolution Amplitude at Fixed Path Average 236 3.8 External Source in Quantum-Statistical Path Integral 237 3.8.1 Continuation of Real-Time Result 238 3.8.2 Calculation at Imaginary Time 242 3.9 Lattice Green Function .. 249 3.10 Correlation Functions, Generating Functional, a nd Wick Expansion 249 3.10.1 Real-Time Correlation Functions. . . . . . . . . . . . 252 3.11 Correlation Functions of Charged Particle in Magnetic Field 254 3.12 Correlation Functions in Canonical Path Integral. 255 3.12.1 Harmonic Correlation Functions 256 3.12.2 Relations between Various Amplitudes 258 3.12.3 Harmonic Generating Functionals. 259 3.13 Particle in Heat Bath . . . . . 262 Heat Bath of Photons. 266 3.14 3.15 Harmonic Oscillator in Ohmic Heat Bath 268 3.16 Harmonic Oscillator in Photon Heat Bath 271 3.17 Perturbation Expansion of Anharmonic Systems 272 3.18 Rayleigh-SchrOdinger and Brillouin-Wigner Perturbation Expansion 276 Level-Shifts and Perturbed Wave Functions from SchrOdinger 3.19 Equation . 280 3.20 CrucuiatioIl of Perturbation Series via Fcynman Diagrams. 282 287 3.21 Perturbative Definition of Interacting Path Integrals. 3.22 Generating Functional of Connected Correlation Functions 288 3.22.1 Connectedness Structure of Correlation Functions. 289 3.22.2 Correlation Functions versus Connected Correlation Functions . 292

xxi 3.22.3 FUnctional Generation of Vacuum Diagrams. . . . . . .. 294 3.22.4 Correlation Functions from Vacuum Diagra.ms . 298 3.22.5 Generating Functional for Vertex Functions. Effective Action 300 3.22.6 Ginzburg-Landau Approximation to Generating Functional 305 3.22.7 Composite Fields. 306 3.23 Path Integral Calculation of Effective Action by Loop Expansion 307 3.23.1 General Formalism 307 3.23.2 Mean-Field Approximation 308 3.23.3 Corrections from Quadratic Fluctuations. 312 3.23.4 Effective Action to Second Order in Ii 315 3.23.5 Finite-Temperature Two-Loop Effective Action 319 3.23.6 Background Field Method for Effective Action 321 3.24 Nambu-Goldstone Theorem 324 3.25 Effective Classical Potential . . . 326 3.25.1 Effective Classical Boltzmann Factor 327 3.25.2 Effective Classical Hamiltonian . . . 330 3.25.3 High- and Low-Temperature Behavior 331 3.25.4 Alternative Candidate for Effective Classical Potential 332 3.25.5 Harmonic Correlation FUnction without Zero Mode 333 3.25.6 Perturbation Expansion 334 3.25.7 Effective Potential and Magnetization Curves 336 3.25.8 First-Order Perturbative Result. . . . . . 338 3.26 Perturbativc Approach to Scattering Amplitude 340 3.26.1 Generating Functional 340 3.26.2 Application to Scattering Amplitude . . . 341 3.26.3 First Correction to Eikonal Approximation 341 3.26.4 Rayleigh-SehrOdinger Expansion of Scattering Amplitude. 342 3.27 Functional Determinants from Green FUnctions 344 Appendix 3A Matrix Elements for General Potential. . 350 4 Appendix 3B Energy Shifts for gx /4-Interaction. 351 Appendix 3C Recursion Relations for Perturbation Coefficients. 353 3C.l One-Dimensional Interaction X4 • 353 3C.2 General One-Dimensional Interaction. 356 3C.3 Cumulative Treatment of Interactions X4 and x 3 . 356 Ground-State Energy with External Current . . . 3C.4 358 3C.5 Recursion Relation for Effective Potential 360 Interaction r 4 in D-Dimensional Radial Oscillator. 3C.6 363 3C.7 Interaction r2q in D Dimensions. 36' 3C.8 Polynomial Interaction in D Dimensions 364 Appendix 3D Feynman Integrals for T =f 0 36' Notes and References . 367

xxii 4

Semiclassical Time Evolution Amplitude 369 4. 1 Wentzel-Kramers-Brillouin (WKB) Approximation. 369 4.2 Saddle Point Approximation 376 4.2. 1 Ordinary Integrals . . . . . . . . 376 4.2.2 Path Integrals 379 4.3 Van Vlcek-Pauli-Morette Determinant. 385 404 Fundamental Composition Law for Semiclassical Time Evolution Amplitude 389 4.5 Semiclassical Fixed-Energy Amplitude . 391 4.6 Semiclassical Amplitude in Momentum Space. 393 4.7 Semiclassical Quantum-Mechanical Partition Function . 395 4.8 Multi-Dimensional Systems . . . . . 400 4.9 Quantum Corrections to Classical Density of States 405 4.9. 1 One-Dimensional Case . 406 4.9.2 Arbitrary Dimensions . . . . . . . . . . . . . 408 4.9.3 Bilocal Density of States. 409 4.904 Gradient Expansion of Tracelog of Hamiltonian Operator. 411 4.9.5 Local Density of States on Circle . . . . .. . . . . . . 415 4.9.6 Quantum Corrections to Bohr-Sommerfeld Approximation 416 4. 10 T homas-Fermi Model of Neutral Atoms. 419 4.10.1 Semiclassical Limit. 419 4.10.2 Self-Consistent Field Equation 421 4.10.3 Encrgy Functional of Thomas-Fermi Atom. 423 4.1004 Calculation of Energies 424 4.10.5 Virial Theorem 427 4.10.6 Exchange Energy. . 428 429 4.10.7 Quantum Correction Near Origin 4.10.8 Systematic Quant um Corrections to T homas-Fermi Energies 432 4. 11 Classical Action of Coulomb System. 436 4. 12 Semiclassical Scattering. . 444 4.12.1 General Formulation 444 4.12.2 Semiclassical Cross Section of Mott Scatt ering . 448 Appendix 4A Semiclassical Quantization for Pure Power Potentials . 449 Appendix 4B Derivation of Semiclassical T ime Evolution Amplitude 451 Notes and References. . . . . .. ................... 455

5 Variation al Perturbation T heory 458 5.1 Variational Approach to Effective Classical Partition Function 458 5.2 Local Harmonic Trial Partition Function 459 5.3 Optimal Upper Bound . . 464 504 Accuracy of Variational Approximation . 465 5.5 Weakly Bound Ground State Energy in Finite-Range Potential Well 468 5.6 Possible Direct Generalizations . 469 5.7 Effective Classical Potential for Anharmonic Oscillator 470

xxiii Particle Densities . . Ext ension to D Dimensions Application to Coulomb and Yukawa Potentials Hydrogen Atom in Strong Magnetic Field . 5. 11.1 Weak-Field Behavior .. 5.11.2 Effective Classical Hamiltonian .. 5.12 Variational Approach to Excitation Energies 5.13 SY5tematic Improvement of Feynman-Kleinert Approximation ... 5.14 Application5 of Variational Perturbation Expansion 5.14. 1 Anharmonic Oscillator at T = O. 5.14.2 Anharmonic Oscillator for T > 0 . . . . . . . 5.15 Convergence of Variational Perturbation Expansion 5.16 Variational Perturbation T heory for Strong-Coupling Expansion 5.17 General Strong-Coupling Expansions . 5.18 Variational Interpolation betwccn Weak and Strong-Coupling Expansions . . . . . . . . . . . . . . . . . . . . . . 5.19 Systematic Improvement of Excited Energies 5.20 Variational Treat ment of Double-Well Potential 5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions .. . . . . . . . 5.21.1 Evaluation of Path Integrals . 5.21. 2 Higher-Order Smearing Formula in D Dimensions. 5.21.3 Isotropic Second-Order Approximation to Coulomb Problem 5.21.4 Anisotropic Second-Order Approximation to Coulomb Problem . . . 5.21.5 Zero-Temperature Limit 5.22 Polarons 5.22.1 Partition Function 5.22.2 Harmonic Trial System 5.22.3 Effective Mass 5.22.4 Second-Order Correction. 5.22.5 Polaron in Magnetic Field, Bipoiarons, etc. 5.22.6 Variational Interpolation for Polaron Energy and Mass 5.23 Density Matrices . 5.23.1 Harmonic Oscillator . . . . . . . . . . . . . . . . . . . 5.23.2 Variational Perturbation Theory for Density Matrices. 5.23.3 Smearing Formula for Density Matrices .. 5.23.4 First-Order Variational Approximation . 5.23.5 Smearing Formula in Higher Spatial Dimensions. Appendix 5A Feynman Integrals for T =f 0 without Zero Frequency Appendix 5B Proof of Scaling Relation for the Extrema of W N Appendix 5C Second-Order Shift of Polaron Energy Notes and References . 5.8 5.9 5.10 5.11

475 479 48 1 484 488 488 492 496 498 499 501 505 512 515

518 520 521 523 524 525 527 528 529 533 535 537 542 543 544 545 548 548 550 552 554 558 560 562 564 565

xxiv 6 Path Integrals with Topological Constraints 6.1 Point Particle on Circle. 6.2 Infinite Wall 6.3 Point Particle in Box . . . 6.4 Strong-Coupling Theory for Particle in Box. 6.4. 1 Partition Function . 6.4.2 Perturbation Expansion 6.4.3 Variational Strong-Coupling Approximations 6.4.4 Special Properties of Expansion . 6.4.5 ExponentiaJly Fast Convergence . Notes and References .

571

7 Many Particle Orbits - Statistics and Second Quantization 7.1 Ensembles of Bose and Fermi Particle Orbits 7.2 Bose-Einstein Condensation . 7.2. 1 Free Bose Gas 7.2.2 Bose Gas in Finite Box 7.2.3 Effect of Interactions . 7.2.4 Bose-Einstein Condensation in Harmonic 'frap 7.2.5 Thermodynamic Functions . 7.2.6 Critical Temperature . 7.2.7 More General Anisotropic Trap 7.2.8 Rotating Bose-Einstein Gas 7.2.9 Finite-Size Corrections . . . . 7.2.10 Entropy and Specific Heat .. 7.2. 11 Interactions in Harmonic Trap Gas of Free Fermions 7.3 Statistics Interaction . . . 7.4 7.5 Fractional Statistics . Second-Quantized Bose Fields 7.6 7.7 Fluctuating Bose Fields. 7.8 Coherent St ates Second-Quantized Fermi Fields 7.9 Fluctuating Fermi Fields . 7.10 7.10.1 Grassmann Variables . 7.10.2 Fermionic Functional Determinant 7.10.3 Coherent States for Fermions 7.11 Hilbert SPac

(P., F1) (Q"Q')

0, ~

o.

(1.47)

8

1 Fundamentals

This follows from the fact that the 2N x 2N matrix formed from the Lagrange brackets

-(Q; ,Q,) ) (P; ,Qj) can be written as Poisson brackets

(£_1)_1

p

E? ) -1,

=(

(1.48)

while an analogous matrix formed from the

{P;,Qj } {Q;, Qj }

- {P;, Pj

}

)

- {Q;, Pj}

(1.49)

is equal to J(E- 1JE)T. Hence C = p - l, so that (1.43) and (1.47) arc equivalent to each other. Note that the Lagrange brackets (1.43) [and thus the Poisson brackets (1.47)1 ens ure p/J; - P/:tj to be a total differential of some function of Pj and Qj in the 2N-dimensional phase space: (1.50)

The Poisson brackets (1.47) for P;, Q. have the same form as those in Eqs. (1.25) for the original phase space variables Pi, q;. The other two equations (1.42) relate the new Hamiltonian to t he old one. They can always be used to construct H'(Pj , Qj, t) from H(p;, q;, t}. The Lagrange brackets ( 1.43) or Poisson brackets (1.47) a.rc therefore both nc (t)).



(LJ36)

This is, of course, just an a pproximation to the integral

Jd'xh· (x)" (x l'l>(t)) .

(1.137)

The completeness of the basis hn(x) may therefore be expressed via t he abstract relation (1.138)



The approximate sign turns into an equality sign in t he limit of zero mesh size, 1:-0.

22

1.4.3

1 Fundamentals

Continuum Limit

In ordinary calculus, fine r and fine r sums are eventually replaced by integrals. T he same thing is done here. We define new continuous scalar products (1.139)

where X n are t he lattice points closest to x. With (1.1 34), the right-hand side is equal to lI1(xn,t) . In the limit t. ---- 0, x and X n coincide a.nd we have

(xlw(,» = w(x,').

(1.140)

The completeness relation can be used to write

(al,,(t»

~

2)alx. )(x. I,,(t» •

~ L"(a l x)(x l w('» I "~""

(1.141)



which in t he limit

t. ____

0 becomes

Jd'x (alx)(x IW(t».

(a lw(,» ~

(1.142)

This may be viewed as t he result of inserting the formal completeness relation of the limiting local bra. and ket basis vectors (x l and Ix) ,

Jd'x Ix)(xl ~

1,

(1.143)

evaluated between the vect ors (al and 11l«t)). With the limiting local basis, the wave functions can be treated as components of t he st ate vectors 11lr(t)) wit h respect to t he local basis Ix) in the same way as a.ny other set of components in an arbitrary basis la). In fact , the expansion

Jd'x (alx)(x l"'('»

(al"'(t» ~

(1.144)

may be viewed as a re-expansion of a component of 11l«t)) in one basis, la), into those of another basis, Ix) , just as in (1.129). In order to express all t hese transformation properties in a most compact nota.tion, it has become customary to deal with an arbitrary physical state vector in a basis-independent way and denote it by a ket vect or IIlr(t) . T his vector may be specified in any convenient basis by multiplying it with the corresponding completeness relation Da)(al ~ 1, (1.145)

,

resulting in t he expansion

Iw(t»

~

Da)(alw(,». ,

(1.146)

23

1.4 Dirac's Bra-Kct Formalism

This can be multiplied with any bra vector, say (bl, from the left to obtain the expansion formula (1. 131):

(bl"' (')) ~ Dbla)(al"'('» ,



(1.1 47)

The continuum version of the completeness relation (1.1 38) reads

Ji'x Ix )(x l ~

1,

(1.1 48)

and leads to t he expansion

1"' ('))

J

~ i'x Ix )(x l"' (')) ,

(1.1 49)

in which the wave function 1lr(x, t) = (x lw(t)) plays t he role of an xth component of the state vector II11(t)} in the local basis Ix). This, in turn, is t he limit of the discrete basis vectors Ix n), 1 (1.150) Ix ) ~ 0< Ix.) , V"

with Xn being the lattice points closest to x. A vector can be described equally well in bra or in ket form. To apply t he above formalism consistently, we observe that t he scalar products

(a lb) ~ (b la)

Jd'x f"( x )"f'(x ), Jd'x l'(x), f"(x)

(Ll5!)

satisfy the identity

(bla) " (alb)",

(1.1 52)

Therefore, when expanding a ket vector as

L: la)(a l"' (')),

(1.1 53)

("' (' )1 ~ D"' (t) la)(al,

(1.1 54)

1"'(')) ~ or a bra vector as



a multiplication of the first equation with the bra (x l and of the second with the ket Ix ) produces equations which are complex-conjugate to each other.

1.4.4

G e n eralize d Functions

Dirac's bra-ket formalism is elegant and easy to handle. As far as t he vectors Ix) are concerned t here is, however, one inconsistency with some fundame ntal postulates of quantum mechanics: When introducing state vectors, the norm was required to be unity in order to permit a proper probability interpretation of single-particle states.

24

1 Fundamentals

The limiting states Ix ) introduced above do not satisfy this requirement. In fact , the scalar product between two different states (x l and Ix') is

where Xn and X n, are the lattice points closest to x a nd x'. For x of x' , the states are orthogonal. For x = X , on t he other hand, the limit t: --+ 0 is infinite, approached in such a way that (1.156)

Therefore, the limiting state Ix) is not a properly normalizablc vector in the Hilbert space. For the sake of elegance, it is useful to weaken the requirement of normaliza bility (1.96) by admitting the limiting states Ix) to the physical Hilbert space. In fact , one admits all states which can be obtained by a limiting sequence from properly normalized state vectors. The scalar product between states (xlx') is not a proper function. It is denoted by the symbol 6(3)(X - x) and called Dimc J-junction: (x ix') =: 6(3)(x - x').

(1.157)

The right-hand side vanishes everywhere, except in the infinitely small box of width f around x Rj x. Thus the J-function satisfies J(3)(x -

x) =

0

for

x =F

x.

(1.158)

At x = x', it is so large that its volume integral is unity: (1.159)

Obviously, t here exists no proper functio n that can satisfy both requirements, (1 .158) and (1.159). Only the finite-f approximation in (1.155) to the 6-function are proper functions. In this respect, t he scalar product (xix') behaves just like the states Ix ) themselves: Both are f --+ 0 -limit.s of properly defined mathematical objects. Notc that thc integral Eq. (1.159) implies the following propcrty of the J function: o(3)(a(x - x'))

~ I~t)(x -

x').

(1.160)

In one dimension, this leads to the more general relation 1

o(f(x))

where

~ ~ II'(x,)lo(x - x;},

x. are the simple zeros of f(x).

(1.161 )

25

1.4 Dirac's Bra-Kct Formalism

In mathematics, one calls the a-function a generalized function or a distribution. It defines a linear functional of arbitrary smooth test functions f(x) which yields its value at any desired place x :

bl!, xl "

J

Jd'x (plx)(xl w(t)) J

(plw(t))

(1.1 91)

d 3 x e- ;px/hW(x , t) .

=

T he amplitudes (p llJ1 (t)) are referred to as momentum space wave functions. By inserting the completeness relation

Jd'xlx)(xl ~

1

(1.192)

between t he moment um states on t he left-hand side of the orthogonality relation (1. 186), we obtain the Fourier representation of the a-function

Jd'x (pl x)(xlp') J

(pip') =

1.4.7

(1.193)

dJxe- ·( p- p'}x /h.

Incomple t e n ess a n d P oisson 's Summa tion Formula

For many physical a pplications it is important to find out what happens to t he completeness relation (1. 148) if one restrict the int egral so a subset of positions. Most relevant will be the one-dimensional integral,

Jdx Ix)(x l ~ I , restrict ed to a sum over equally spaced points

(1.194)

Xn = na:

N

l:

n=- N

Ix.)(x. l·

(1.195)

1.4 Dirac's Bra-Kct Formalism

29

Taking this sum between momentum eigenstates Ip}, we obtain N

I:

(P1".)(x.IP') ~

n :_ N

For N

----> 00

N

I:

(Plx.)(x.IP')

~

N

I:

,'~-")=I'

(1.196)

n :_ N

n =_ N

we can perform t he sum with t he help of Poisson's summation fonnula (1.1 97) n", -oo

m "' - OO

Identifying p, wit h (p - rI)aj211"fl, we find using Eq. (1.160):

.t(Plx.)(x.IP') ~mt'C ;:2a-m) ~mt p

2:"

o(.-P' -2,:m).

(1.1 98)

In order to prove the Poisson formula (1.197), we observe that the sum 8(p,) == Lm 8(p, - m) on the right-hand side is periodic in p, with a unit period and has the Fourier series 8(1-') = L:::": _oo8 ne2".il'n. T he Fourier coefficients are given by 2lfi Sn = f~{~2 dp. s(l-')e- l'n == 1. T hese are precisely the Fouricr coefficients on the left-hand side. For a finite N, the sum over n on the left-hand sidc of (1.197) yields

- 1+ ( =

1+

1 _ e2lfiJl (N+1 ) 1

e2".il'

+ C.c.

)

(1.1 99)

e21fi l' _ e21fi l'(N+l) sin 1I"Jt(2N + 1) + c.C. = 1 e 2.n1' sin 11"11

T his function is well known in wave optics (see Fig. 2.4). It tion pattern of light behind a grating with 2N + 1 slits. Jt = 0, ± 1, ± 2, ±3,.. . and N - 1 small maxima between ing peaks, at II = (1 + 4k)/2(2N + 1) for k = 1, ... , N v ~ (1+ 2k)/(2N + 1) fo' k ~ 1, ... , N - 1. Inserting p, = (p - rI)a/211"n into (1.1 99), we obtain

determines t he diffracIt has large peaks at each pair of neighbor1. There are zeros at

~ (PI )( I ') ~ 'in (P - p')a(2N + 1)/21i x. x. P . (P p ') a /2 It' . Sill

L"

(1.200)

n :-N

Let us see how the right-hand side of (1.1 99) t urns into the right-hand side of (1.197) in the limit N ----> 00. In this limit , the area under each large peak can be calculated by an integral over the central large peak plus a number n of small maxima next to it: n/2N d sin 1I"p.(2N + 1) = ;n/ 2N d sin 211"p.N cos 1I"p.+ cos211"/lN sin 1I"/l . ; - n/2N I-' sin 1I"p. - n/2N I-' sin 1I"p.

(1.201)

30

1 Fundamentals

";nlfl'(2N +1)

"n If"

, Figure 1.2 Relevant function E;:,.,- Ne2ripn in Poisson's summation for mula. In the limit N --> 00 , Jl. is squeezed to the integer values.

Keeping a fixed ratio n/N «: 1, we may replace in the integrand sin 1r1l by 1f1l and COS1rll by 1. T hen t he integral becomes, for N --+ 00 at fixed njN, "J2N Sin1TJ.I(2N + 1) N_ oo ;n/2N sin27rJ1-N . --> dl' ; - n / 'lN dp. sm 7rJ.l. - n/2N 7rJ.l. N_oo ~

sinx -1 ;"".. dx11: -If" x

+ -1- ;

.."

21l'N _,....

dxcosx

+

N_oo -->

;n/'lN - n/2N

1

'

dJl,cos21rI-tN

(1.202)

where we have used the integral formula 00 ;

- 00

dxsinx x

=71".

(1.203)

In the limit N ____ 00, we find indeed (1.197) and thus (1.205), 3.'; well as the expression (2.464) for t he free energy. There exists another useful way of expressing Poisson's formula. Consider an arbit rary smooth fu nction f(p,) which possesses a convergent sum 00

2:

f(m) .

(1.204)

m=-oo

T hen Poisson's formula. (1.197) implies t hat the sum can be rewritten as an integral and an a uxiliary sum:

(1.205) T he auxiliary sum over n squeezes J1. to t he integer numbers.

1.5 Observablcs

1.5

31

Observables

Cha nges of basis vectors a rc an important tool in analyzing t he physically observable content of a wave vector. Let A = A{p, x) be an arbitrary time-independent real function of the phase space variables p and x. Important examples for such an A are p and x themselves, the Hamiltonian H( p , x) , and the angular momentum L = x x p . Quantum-mechanically, there will be an observable operator associated with each such quantity. It is obtained by simply replacing the variables p and x in A by the corresponding operators f> and x:

A = A(p, x).

(1.206)

T his replacement rule is the extension of the correspondence principle for the Hamiltonian operator (1.92) to more general functions in phase space, converting them into observable operators. It must be assumed that the replacement leads to a unique Hermitian operator, i.e. , that there is no ordering problem of the type discussed in context with the Hamiltonian (1. 101).8 If there are ambiguities, the naive correspondence principle is insufficient to determine the observable operator. Then the correct ordering must be decided by comparison with experiment, unless it can be specified by means of simple geometric principles. This will be done for the Hamiltonian operator in Chapter 8. Once an observable operator A is Hermitian, it has the useful property that t he set of all eigenvectors la) obtained by solving the equation

A la) ~ ala)

(1.207)

can be used as a basis to span the Hilbert spacc. Among the eigenvectors, there is always a choice of orthonormal vectors la) fulfilling the complet enes.,> relation

z= la)(al ~ 1.

(1.208)

T he vectors la} can be used to extract physical information concerning the observable A from arbitrary state vector IIIr(t)). For t his we expand this vector in the basis la}: (1.209) 1'"

,

XN)(X l, .. . , XNI = 1,

(1.229)

and define

(1.230) {XJ, ... , x Nlx"

=

x" {x l> ", , x NI·

T he SchrOdingcr equation for N particles (1.107) follows from (1.228) by multiplying it from the left with the bra vectors (Xl, ... , xN I. In t he same way, all other formulas given above can be generalized to N · body state vectors.

1.6

Time Evolution Operator

If the Hamiltonian operator posse&;es no explicit time dependence, the basisindependent SchrOdinger equation (1.163) can be integrated to find t he wave function IIV (t») at any time tb from the state at any other time ta

(1.231 ) The operator

O(tb,ta )

= e-i{to-t$)Hlh

(1.232)

is called the time evolution opemtor. It satisfies t he differential equation irI8t .U(tI>, ta) = H U(tb' tal.

(1.233)

Its inverse is obtained by interchanging t he order of tb and ta :

O-I(tb,ta) =: ei{t.-t~) iflh

=

O(ta,tl».

(1.234)

As an exponential of i times a Hermitian operator, 0 is a unitary operator satisfying

Of =

(;-1 .

(1.235)

lndood ,

(1.236)

35

1.6 Time Evolution Operator

If H(p,x, t) depends explicitly on t ime, t he integration of the SchrOdinger equation (1.163) is somewhat more involved. The solution may be found iteratively: For tb > t a , the time interval is sliced into a large number N + 1 of small pieces of thickness f. with f. == (tb - ta)/(N + 1), slicing once at each time t,. = ta + nf. for n = 0, ... , N + 1. We t hen use the SchrOdinger equation (1. 163) to relate the wave function in each slice approximately to t he previous one:

IW('o + ,))

~

(1- ;. t



IIli (ta + 2£))

~

(l-*l:::'ltal also for the anticausal (or advanced) case where tb lies before tao To be consistent with the above composition law (1.254), we must have (1.256)

Indeed, when considering two states at successive times (1.257)

the order of succession is inverted by multiplying both sides by (;-I(t", tb): I ~(@ ~ O(to . t,) - 'I ~(to)) .

t, < to.

(1.258)

The operator on the right-hand side is defined to be the time evolution operator U(tb' ta) from the later time ta to the earlier time tbIf the Hamiltonian is independent of time, with the time evolution operator being (1.259)

t he unitarity of the operator U(tb' ta) is obvious: A

U

t

(t,. to) ~ U(t,. to) A

_ I

.

(1.260)

Let us verify this property for a general time-dependent Hamiltonian. There, a direct solution of the SchrOdinger equation (1.163) for the state vector shows that the operator U(Lb , La) for tb < La has a representation just like (1 .252), except for a reversed time order of its arguments. One writes this in the form [compare (1.252)] (1.261)

where T denotes the t ime-ant iordering operator, with an obvious definition analogous to (1.241), (1.242). This operator satisfies the relation (1.262)

1.8 Heisenberg Picture of Quantum Mechanics

39

with an obvious generalization to the product of n operators. We can therefore conclude right away that (1.263)

With O(ta, tb) =: O(tb, ta)- I, this proves the unitarity relation (1.260), in general.

c) Schrodinger equation for U(tb, tal Since the operator U(tb' ta) rules the relation between a rbitrary wave functions at different times, 1"' ( t a , on the other hand, the poles in the lower half-plane give, via Cauchy'S residue

47

1.12 Free-Particle Amplitudes

theorem, the spectral representation (1.3 19) of the propagator. An i1}-prescription will appear in another context in Section 2.3. If the eigenstates are nondegenerate, the residues at the poles of (1.321) render directly the products of eigenfunctions (barring degeneracies which must be discussed separately). For a system with a continuum of energy eigenvalues, there is a cut in the complex energy plane which may be thought of as a closely spaced seQuence of poles. In general, the wave functions are recovered from the discontinuity of the amplitudes (JIQ, I ~)E across the cut, using the formula disc

in

(E iftEn) == "E;--ii~C-o-'+-:i="

"E'---;::---:'= ~ 2nM(E - Eo). E"

tT/

(1.324)

Here we have employed the reiationiO, valid inside integrals over E: E

1

En±iT/

~

E

PE~ ~ i7ro(E ..

(1.325)

E,,},

where P indicates that the principal value of the integral has to be taken. The energy integral over the discontinuity of the fixed-energy amplitude (1.321) (JIQ,lxah: reproduces t he completeness relation (1.318) taken between the local states (",I ""d I",,), (00

2d~ di" (",I""),, ~ I: >l'o(",)>I';(x o) ~ (",I",,) ~ ,(DJ(x, -

~oo 7r1~

x o)'

(1.326)

"

The completeness relation reflects the fonowing property of the resolvent operator: (00 dE



J-OIO 2Jrh discR (E)

• = 1.

(1.327)

In general, the system possesses also a cont inuous spectrum, in which case the completeness relation contains a spectral integral and (1.318) has the form LI>I'o)(>I'ol o

+

Jdv 1>1'")(>1'"1 ~

1.

(1.328)

The continuum causes a branch cut along in the complex energy plane, and (1.326) includes an integral over the discontinuity along the cut. The cut will mostly be omitted, for brevity.

1.12

Free-Particle Amplitudes

For a free particle with a Hamiltonian operator if = p2 j 2M , the spectrum is continuous. The eigenfunctions are (1.189) with energies E(p ) = p 2j2M. Inserting IOThis is often referred to as Sochocki's JOffllula. It is the beginning of an expansion in + 1'/ ['lf6'(x) ± idzPlx] + 0(1'/2).

of 1'/ > 0: 1/(x ± il'/) = Pix 'fi 'lf6(x)

p0WCn!

48

1 Fundamentals

the completeness relation (1. 187) into Eq. (1.296) , we obtain for the time evolution amplit ude of a free particle t he Fourier representation (1.329)

The momentum integrals can easily be done. First we perform a quadratic completion in the exponent and rewrite it as

1

P(Xb - x..) - -

2M

,

1 (

p (tb - ta) = -

2M

1

Xb- x..)2 (tb -

p - ---

M tl> - La

M{Xb - x..f . 2 il> ta

tal - -

(1.330) Then we replace the integration variables by the shifted momenta p' = p (xt. - Xa) /( tb - t".)M, and the amplitude (1.329) becomes (1.331)

where F(tb -

tal

is the integral over the shifted momenta (1.332)

This can be performed using t he Fresnel integml jO'fmula

f

OO

- 00

(.a ') vfaI 1 { Vi, l / vt,

dp

.;r;:;cxP ~ 2P

=

a> 0, a < O.

(1.333)

Here the square root .jt denotes the phase factor ei1r / 4 : This follows from t he Gauss formula 1 Reo> 0, (1.334) _oo ,;21rcxp - "2 P =

f OOdP

(0,) va'

by continuing a analytically from positive values into the right complex half-plane. As long as Rea> 0, this is straightforward. On the boundaries, i.e. , on the positive and negative imaginary axes, one has to be careful. At a = ±ia + 1] with a ~ and infinitesimal 1] > 0, the integral i;; certainly convergent yielding ( 1. 333). But t he integral also converges for 1] = 0, as can easily be seen by substituting x 2 = z. See Appendix lB. Note t hat differentiation of Eq. (1.334) with respect to a yields the more general Gaussian integral formula

°

f

oo

- 00

dp .j21f

-- p

2n

(0 ') = 1- (2n - 1)1!

exp - - p

2..;a

an

Rea> 0,

(1.335)

whcre (2n - I)!! is defined as t he product (2n - 1) · (2n - 3) ···1. For odd powers n +1, the integral vanishes. In the Fresnel formula (1.333) , an extra integrand p2n produces a factor (ija)n .

rl

49

1.12 Free-Particle Amplitudes

Since t he Fresnel formula is a special analytically continued case of the Gauss formula, we shall in the sequel always speak of Gaussian integrations and use Fresnel's name only if t he imaginary nature of the quadratic exponent is to be emphasized. Applying this formula to (1. 332), we obtain

F(tb-ta)

1

=

(1.336)

D'

/2rri;,(', - 'o)IM so t hat the full time evolution amplitude of a free massive point particle is 1

(Xbtb lxata) =

D

/2rri;,(', - 'o)IM

exp

[iM(",-X.)'] h2 t t . '

(1.337)

0

In the limit tb -+ t a , the left-hand side becomes the scalar product - x..), implying the following limiting formula for the a-function

(",Ix.)

~

a(D) (X{.

a(D) ( X b -

x,,) =

1

lim

t. -t~_O ..j27rifi.(tb _ ta)IM

D

[i

exp -M - ( X b - x..f ] . Ii 2 tb ta

(1.338)

Inserting Eq. (1.331) into (1.313) , we have for the fixed-e nergy amplitude the integral representation

(",Ixo)'

~

f

d(t, - to)

J(2~~D exp { ~ [P('" - xo) + (t, - to) (E - :~)] } . (1.339)

Performing the time integration yields iii

dVp.

(Xblx..)s =

J(21Th)D exp [t p (Xb - x a) ] E _ p2/2M + iT!,

(1.340)

where we have inserted a damping facto r e - '1{tb - t..) into the integral to ensure convergence at large tb - tao For a more explicit result it is more convenient to calculate the Fourier transform (1.337): 00

(XbIXa)s = in1 d(tb - tal o

1

..j21Tih(tb _ ta)/M

D exp

{ifi [E(tb - tal + 2"M (X.t - x.)'] } b

l

t

a

.

(1.34 1)

For E < 0, we set

0,

(1.347)

to obtain (1.348)

This result can be continued analytically to D > 2, which will be needed later (for example in Subsection 4.9.4) . ForE>Oweset (1.349) ls and use the formula (1.350)

where HPl(z) is the Hankel function, to find (1.351)

T he relation 16 (1.352) 12ibid., Formula 8.486.16 13/1,1. Abramowitz and I. Stcgun, Handbook 0/ Mathematical FUnctions, Dover, New York, 1965, Formula 10.2.17. 14ibid., Formula 9.6.9. 15ibid., Formulas 3.471.11 and 8.421.7. 16ibid., Formula 8.407.1.

1.13 Quantum Mechwljcs of Gelleral Lagrangian Systems

51

connects the two formulas with each other when continuing the energy from negative to positive values, which replaces ,.. by C i " J2 k = - ik . For large distances, the asymptotic behavior 17 (1.353)

shows t hat the fixed-energy amplitude behaves for E < 0 like

(>Co IXa) E ~

. M D- ' (2lr)(D --;::"'; ;::,,,, ,=ci:'=",e- "RI' , -tr;"" 1)/ 2 (""R)(D 1)/2

(1.354)

and for E > 0 like M

Ii.

kD-2",..."~'=",,,,='iL=e;kRJ" (21ri)(D I)J2 (kR)(D 1)/2 .

(1.355)

For D = 1 and 3, these asymptotic expressions hold for all R.

1.13

Quantum Mechanics of General Lagrangian Syste ms

An extension of the quantum-mechanical formalis m to systems described by a set of completely general Lagrange coordinates ql, ... , qN is not straightforward. Only in the special case of qi (i = 1, ... , N) being merely a curvilinear reparametrization of a D-dimensional Euclidean space are t he above correspondence rules sufficient to quantize the system. Then N = D and a variable change from X i to Qj in the SchrOdinger equation leads to the correct quantum mechanics. It will be useful to label the curvilinear coordinates by Greck superscripts and write ql' instead of Qj . This will help when we write all ensuing equations in a form that is manifestly covariant under coordinate transformations. In the original definition of generalized coordinates in Eq. (1. 1), t his was unnecessary since transformation properties were ignored. For the Cartesian coordinates we shall use Latin indices a lternatively as sub- or superscripts. The coordinate transformation Xi = Xi(ql') implies the relation between t he derivatives 01' == ojfJql' and Oi == fJjfJx i : (1.356)

with the transformation matrix (1.357)

called basis D-ad (in 3 dimensions t riad , in 4 dimensions tetrad, etc.). Let e;l'(q) = fJql' jfJxi be the inverse matrix (assuming it exists) called t he reciprocal D-ad, satisfying with e;1' the orthogonality and completeness relations (1.358)

52

1 Fundamentals

Then, (1.356) is inverted to (1.359)

fJj = e;"'(q)fJ"

and yields the curvilinear transform of the Carteian quantum-mecha.nical momentum operators Pi = -ihfJ; = -ilie/,(q)8,.. (1.360) T he free-particle Hamiltonian operator

iI = T =_I_ P,2 =_ !!....V2

(1.36 1)

• h' Ho = -2M!:::..'

(1.362)

o

2M

2M

goes over into

where t1 is the Laplacian expressed in curvilinear coordinates:

o? =

!:::. =

ei l'fJl' c{ov

eil'e{fJpfJ"

+ (e i P.(}pet)8".

(1.363)

At this point one int roduces the metric tensor (1.364)

its inverse

gI'V(q) = eil'(q)e{(q) ,

(1.365)

defined by 91'''9.,>. = 6")' , a nd the so-called affine connection (1.366)

Then the Laplacian takes t he form (1.367)

with

r I-' >.v being defined as the contmction r !' >v -~ g"r.." ".

(1.368)

T he reason why (1.364) is called a metric tensor is obvious: An infinitesimal square distance between two points in the original Cartesian coordinates ds'! =: dx?

(1.369)

becomes in curvilinear coordinates (1.370)

1.13 Quantum MechwljCS of GeIleral Lagrangian Systems

53

The infinitesimal volume element dDx is given by (1.371)

where (1.372)

is the determinant of the metric tensor. Using this determinant, we form the quantity (1.373)

and sec t hat it is equal to the once-contracted connection (1.374)

With the inverse metric (1.365) we have furthe rmore (1.375)

ov

We now take advantage of the fact t hat the derivatives fJ", applied to the coordinate transformation Xi{q) commute causing r"v~ to be symmetric in /-tv , i.e. , r"v A = rv/ and hence r"v" = rv. Together with (1.373) we find the rotation (1.376)

which allows the Laplace operator t:. to be rewritten in the more compact form (1.377)

This expression is called the Laplace-Beltrumi operutOT".18 Thus we have shown that for a Hamiltonian in a Euclidean space 1

H(p, x) = 2M P'l + V(x) ,

(1.378)

the SchrOdinger equation in curvilinear coordinates becomes

.

["'

1

H,p(q , t) '" - 2Mt. + V(q) ,p(q, t) ~ ih8,,p(q, t) ,

(1.379)

where V (q) is short for V(x(q)). T he scalar product of two wave functions which determines the transition amplitudes of thc system, transforms into (1.380) dDq.Jij ,p;(q,t),p,(q,t).

f dDxtf;;(x, t) tf;l(X, t),

J

--;'··'~!~ o'~'~d~"~~~ · .~w ~;~ll~"" C-:g~ ;'~'n:C;: I.~"~'~;n:-;Eqs. (11.12)- (11.18).

54

1 Fundamentals

It is important to realize t hat t his SchrOdinger equation would not be obtained by a straightforward application of t he canonical formalism to the coordinatetransformed version of the Cartesian Lagrangian

L(x,x) =

~ x.'l - V(x).

(1.381)

With the velocities transforming as

x' = e\,(q)q''',

(1.382)

the Lagrangian becomes

L(q, q)

~

M

,9,"(q)q'q" - V(q).

(1.383)

Up to a factor M , t he metric is equal to the Hessian metric of t he system, which depends here only on q" [recall (1.1 2)1: (1.384)

The canonical momenta a re

aL

PI' == 8it' =

M

."

9t",q·

(1.385)

T he associated quantum-mechanical momentum operators PI' have to he Hermitian in the scalar product (1.380) and must satisfy the canonical commutation rules (1.268),

IP"

q"] [q", q"]

IP",,"]

~

- iM.. ", (1.386)

0, ~

O.

An obvious solution is (1.387)

The commutation rules are t rue for _ilig- Z 8I' gZ with any power z, but only z = 1/4 produces a Hermit ian momentum operator:

JJlq..;gw;(q , t)[_ilig- 1/48",gl/4W1(q, t)1 ! Jlq JJlq..;g [-ilig-I/481,gl /4WZ(q, t) ]* w](q, t), =

g l/4 w; (q, t)[-i1i8",gl/4WI(q , t)1

=

(1.388)

as is easily verified by partial integration. In terms of the quantity (l.373), this can also be rewritten as (1.389)

1.13 Quantum MechwljCS of GeIleral Lagrangian Systems

55

Consider now the classical Hamiltonian associated with the Lagrangian (1. 383) , which by (1.385) is simply

H

=

Pl"ql" - L

=

2~gl",,(q)pI'p" + V(q).

(1.390)

When trying t o turn this expression into a Hamiltonian operator, we encounter the operator-ordering problem discussed in connection with Eq. (1.101 ). The correspondence principle requires replacing the momenta PI" by the momentum operators PI" but it does not specify the position of these operators with respect to the coordinates ql" contained in the inverse metric gl""(q). An important constraint is provided by the required Hermiticity of the Hamiltonian operator, but this is not sufficient for a unique specification. We may, for instance, define the canonicaJ Hamiltonian operator as (1.391)

in which the momentum operators have been arranged symmetrically around the inverse metric to achieve Hermiticity. This operator, however, is not equal to the corrcct SchrOdinger operator in (1.379). The kinetic term contains what we may call the canonical Laplacian .6.a

n

=

(01"+ ! rl")gl""(q)(o"

+ ! r,,).

(1.392)

It differs from the Laplace-Beltrami operator (1.377) in (1.379) by .6. - .6.ean =

-~ 81"(gl"" r,,)

- igl""r "r,...

(1.393)

The correct Hamiltonia n operator could be obtained by suitably distributing pairs of dummy factors of gl/4 and g- I/4 symmetrically between the canonical operators [5]:

fl = 2~g-I/4pl"gl/4g1""(q)gl/4p"g-1/4 + V(q).

(1.394)

This operator has the same classical limit (1.390) as (1.391). Unfortunately, the correspondence principle docs not specify how the classical factors have to be ordered before being replaced by operators. The simplest system exhibiting the breakdown of the canonical quantization rules is a free particle in a plane described by radial coordinates ql = r, q2 = '(): (1.395)

Since the infinitesimal square distance is ds 2 = dr 2 + r 2 dr, the metric reads

~) It has a determinant

g=

r2

.

(1.396)

"" (1.397)

56 and an inverse

1 Fundamentals

g~V= (~ ~-2

rv.

(1.398)

The Laplace-Beltrami operator becomes 1

,

6. = - (}rrfJr

1

+ "2(}cp ,

,

.

(1.399)

The canonical La placian, on the other hand, reads

(1.400)

The discrepancy (1.393) is t herefore (1.401)

Note that this discrepancy arises even though t here is no apparent ordering problem in the naively quantized canonical expression {JI'9p,,(q) if' in (1.400). Only the need to introduce dummy 9 1/ 4_ and g- I/ 4_factors creates such problems, and a specification of the order is required to obtain the correct result. If the Lagrangian coordinates qi do not merely reparametrize a Euclidean space

but specify the points of a general geometry, we cannot proceed as above and derive the Laplace- Beltrami operator by a coordinate t ransformation of a Cartesian Laplacian. With the canonical quantization rules being unreliable in curvilinear coordinates there a re, at first sight , severe difficulties in quantizing such a system. This is why t he literature contains many proposals for handling this problem (6]. Fortunately, a large class of non-Cartesian systems allows for a unique quantummechanical description on completely different grounds. These syst ems have t he common property that their Hamiltonian can be expressed in terms of the generators of a group of motion in the general coordinate frame. For symmetry reasons, the correspondence principle must then be imposed not on the Poisson brackets of the canonical variables p and q, but on those of the group generators and the coordinates. The bracket s containing two group gcnerators specify the structure of the group, those containing a generator and a coordinate specify thc defining representation of the group in configuration space. The replacement of t hese brackets by commutation rules constitutes the proper generalization of the canonical quantizat ion from Cartesian to non-Cartesian coordinates. It is called group quantization. The replacement rule will be referred to as the group correspondence principle. T he canonical commutation rules in Euclidean space may be viewed as a special case of the commutation rules between group generators, i.e., of the Lie algebm of t he group. In a Cartesian coordinate frame, the group of motion is the Euclidean group containing translations a nd rotations. The generators of translations and rotations arc the momenta and thc angular momenta, res~ti vely. According to thc group

57

1.14 Particle on the Surface of a Sphere

correspondence principle, the Poisson brackets between t he generators and the coordinates are to be replaced by commutation rules. Thus, in a Euclidean space, t he commutation rules between group generators and coordinates lead to the canonical quantization rules, and this appears to be the deeper reason why the canonical rules are correct. In systems whose energy depends on generators of t he group of motion other than those of translations, for instance on t he angular momenta, the commutators between the generators have to be used for quantization rather than the canonical commutators between positions and momenta. The prime examples for such systems are a particle on t he surface of a sphere or a spinning top whose quantization will now be discussed.

1. 14

Particle on the Surface of a Sphere

For a particle moving on the surface of a sphere of radius r with coordinates rsin Ocos4', x 2

Xl =

=

rsin Osin 4', x 3

=

rcosO ,

(1.402)

the Lagrangian reads (1.403)

The canonical momenta are (1.404)

and t he classical Hamiltonian is given by

1 (p'e +--p' 1) . sin20

H ~--

2Mr2

(1.405)

'P

According to the canonical quantization rules, the momenta should become opera..

to"' •

_

Po -

a

·to 1-,· 1/ 2lJ - tw-:-i 2 - II Sill (1 , Sill 0

• P'P

=

·ton

- tl«J'P'

(1406) •

But as explained in the previous section , these momentum operators are not expected to give the correct Hamiltonian operator when inserted into the Hamiltonian (1.405) . Moreover, there exists no proper coordinate transformation from the surface of the sphere to Cartesian coordinates lll such that a particle on a sphere cannot be treated via the safe Cartesian quantization rules (1.268);

w;,

X' ] ]x\X' ]

W;,p, ]

~

-iM/ , 0,

(1.407)

o.

19There exist, however, certain infinitesimal nonholonomic coordinate transformations which arc multivaluoo and can be used to t ransform infinitesimal distances in a curvlXl space into those in a flat one. They are introduced and applied in Sections 10.2 and Appendix lOA, leading once more to the same quantum mechanics 88 the one described here.

58

1 Fundamentals

The only help comes from the group properties of the motion on the surface of the sphere. The angular momentum

L =xx p

(1.408)

can be quantized uniquely in Cartesian coordinates and becomes an operator (1.409)

whose components satisfy the commutation rules of the Lie algebra of the rotation group

(i,j, k cyclic).

(1.410)

Note that there is no factor-ordering problem since the x"s and the Pi'S appear with different iudices in each Lk • An important property of the angular momentum operator is its homogeneity in x. It has the consequence that when going fro m Cartesian to spherical coordinates Xl =

rsinOcos'P, x 2 = rsin ()sin rp, x 3 = r cos 0,

(1.4 11 )

the radial coordinate cancels making the angular momentum a differential operator involving only t he angles fJ,



'P, t he

63

1.15 Spimling Top

For j = 1/2, t hese form the spinor representation of t he rotations around the y-axis d1/'1 (/3) = ( COS{3/2 -sin {3/2 ) m'm sin{3/2 cos{3/2 .

(1.444)

The indices have t he order + 1/2, - 1/2. The full spinor representation function DI/'1(a,{3, "y) in (1.442) is most easily obtained by inserting into the general expression (1.429) the representation matrices of spin 1/2 for the generators £. with the commutation rules (1.410) , t he famous Pauli spin matrices:

u' -_

(0 o' )' ,(O-i) ,( ' 0) 1

(1

=

i

0

'

(1

0 - 1

=

.

(1.445)

Thus we can writ e (1.446)

The first and the third factor yield the pure phase factors in (1.442). The function 2 d!!.,'!,.({3) is obtained by a simple power series expansion of e- ·/kJ / 2 , using the fact that «(12)'ln = 1 and «(12frt+l = (12: (1.447)

which is equal to (1.444). For j = 1, t he representation functions (1.443) form the vector representation

d:"'m(j3) =

!(1 +cos /3) -7zsin {3 ! (1 -cos/3) ) ~sin {3 cos{3 -~sin {3 . ( !(1-cos{3) -32sin {3!(1 +cos{3)

(1.448)

where the indices have the order +1/2,-1/2. The vector representation goes over into the ordinary rotation matrices R;j({3) by mapping the states 11m) onto the spherical unit vectors 10(0) = i , E(±I) = =f(x ± iy )/2 using the matrix elements (illm) = f'(m) . Hence R({3)E(m) = E:"'=_ l E(m')d:"'m(j3)· The rcpresentation functions DI(a,{3,"Y) can also be obtained by inserting into the general exponential (1.429) the representation matrices of spin 1 for t he generators L; with the commutation rules (1.410). In Cartesian coordinates, these arc simply (£i)jk = -if;j k, where fijk is the completely antisymmetric tensor with fl'13 = 1. In t he spherical basis, these become (L;)mm' = (m li)(L.:)ij(j lm') = fi(m)(L');jEj(m') . The exponential (e- itJL2 )mm' is equal to (1.448). The functions p,(oJJ\z) are the Jacobi polynomials [8], which can be expressed in terms of hypergeometric functions as

(ofl,_(- 1)'r(I+P+l)

P,

~-l!-

.

.

r (p+ 1) F(-I ,I+ l+ a+p,l+p,(1 H)/2) ,

(1.449)

64

1 Fundamentals

where

ab c

F(a ,bjc;z)=l+ -z +

a(a+l)b(b+l)z2 ( ) ,+ cc + l 2.

...

(1.450)

The rotation functions df:,m,({3) satisfy the differential equation

T he scalar products of two wave functions have to be calculated with a. measure of integration that is invariant under rotations;

(tf12I1J!I) == 121< [ l~" dadfJsinf3d"'( 'I/J'2 (0:, {J, "Y)'1/7J(a,{3,"'().

(1.452)

The above eigenstates (1.442) satisfy t he orthogonality relation

(1.453)

Let us also contrast in this example the correct q uantization via the commutation rules between group generators with the canonical approach which would start out with the classical Lagrangian. In terms of Euler angles, the Lagrangian reads (1.454)

where w{, w'I'w( are the angular velocities measured along t he principal axes of t he top. To find t hese we note that the components in the rest system Wl>~,W3 are obta.ined from the rela.tion (1.455)

WI

W2 W3

-/3sincr + tsin{jcoscr, = /3coscr+tsin (jsincr, = tcos(j+a. =

(1.456)

After the rotation (1.436) into the body-fixed system, these become

w, w, w(

~

~

/3 sin-y - a sin (j 008'Y, /3 cos 'Y + a sin {j sin 'Y, acos(j + 1'.

(1.457)

Explicit ly, the Lagrangian is L =

4[/{(jj2 + a

2

sin 2(j) + It;{a cos (j + t)2J .

(1.458)

65

1.15 Spimling Top

Considering 0:',/3, "f as Lagrange coordinates ql' with IJ. = 1, 2,3, this can be written in the fo rm (1.383) with the Hessian metric [recall (1.1 2) and (1.384)1: I{ sin2 /3 + I, cos2 /3 0 ~v =

(

whose determinant is

0

~

I(cos /3

0

(1.459)

9 = Iile sin2 /3.

(1.460) Hence the measure f rflq..j9 in the scalar product (1.380) agrees with the rotationinvariant measure (1.452) up to a trivial constant factor. Incidentally, t his is also true for the asymmetric top with h =f II) =f 1( , where 9 = Iii, sin 2 /3, although the mctric gl'v is then much morc complicated (sec Appendix I C) . T he canonical momenta associated with t he Lagrangian (1.454) are, according to (1.383) , =

Po

8LI8o = i{osin2 /3+ I(cos /3(ocos/3+1'),

i!L/i!/3 ~ [, /3,

p,

P"/ = 8L I81' = Jc(ocos /3 + 1') .

(1.461)

After inverting the metric to

(1.462) we find the classical Hamiltonian

[2. ,+ (000' [ . P 2.)' + /

H - ~ - 2 IV,

[ {sm2a+ fJ'

{

V,



P .1 2a Pa , _ [200, . 2a PaP"/ {sm fJ

{Sill fJ

(1.463)

This Hamiltonian has no apparent ordering problem. One is therefore tempted to replace the momenta simply by the corresponding Hermitian operators which are, according to (1.387),

Po

=

-ifi.8a ,

pp

1 - in(sinj3) - 1/28p(sin/3) 1/2 = -ih(8p + 2cot /3),

p"/ =

-ina"/.

(IA64)

Inserting these into (1.463) gives the canonical Hamiltonian operator

Han = H + Hdisc"

(1.465)

with

(1.466)

66 and

1 Fundamentals

_ 1 1 2 1 3 H,· =-(8" cotf.l)+ - cot f.I= _ _ __ • '''''' 2 " JJ 4 fJ 4sin 2 {3 4 A

(1.467)

T he first term if agrees with the correct quantum-mechanical operator derived above. Indeed, inserting the differential operators for the body-fixed angular momenta (1.437) into the Hamiltonian (1.423), we find f.J. The term Hdi1>cr is t he discrepancy between the canonical and the correct Hamiltonian operator. It exists even though t here is no apparent orderi ng problem, j ust as in t he radial coordinate expression (1.400). T he correct Hamiltonian could be obtained by replacing the classical pi term in H by the operator g- I/4 ppgl f 2fjpg- l/4, by analogy with the treatment of the radial coordinates in if of Eq. (1.394). As another similarity with t he two-dimensional system in radial coordinates and the particle on t he surface of the sphere, we observe that while the canonical quantization fails, the Hamiltonian operator of the symmetric spinning top is correctly given by t he Laplace-Beltrami operator (1.377) after inserting the metric (1.459) and the inverse (1.462). It is straightforward although tedious to verify that this is also true for the completely asymmetric top [which has quite a complicated metric given in Appendix l C, see Eqs. (lC.2), and (lC.4)]. This is an important nontrivial result, since for a spinning top, the Lagrangian cannot be obtained by reparametrizing a particle in a Euclidean space with curvilinear coordinates. The result suggests that a replacement (1.468)

produces the correct Hamiltonian operator in any non- Euclidean space.:!! W hat is t he characteristic non-Euclidean property of the a, (3, I space? As we shall see in detail in Chapter 10, t he relevant quantity is the curvature scalar R. T he exact definition will be found in Eq. (10.42) . For the asymmetric spinning top we find (see Appendix lC) (1.469)

Thus, just like a particle on the surface of a sphere, the spinning top corresponds to a particle moving in a space with constant curvature. In this space, the correct correspondence principle can also be deduced from symmetry arguments. The gcometry is most easily understood by observing t hat the a ,{3, "f space may be considered as the surface of a sphere in four dimensions, as we shall see in more detail in Chapter 8. An important non-Euclidean space of physical interest is encountered in t he context of general relativity. Originally, gravitating matter was assumed to move in a spacetime with an arbitrary local curvature. In newer developments of the t heory one also allows for the presence of a nonvanishing torsion. In such a general situation, 21 If the space has curvature and no torsion, t his is t he correct ansWtJr. If torsion is present, t he corroct answer will be given in Chapters 10 and 8.

1.16 Scattering

67

where the group quantization rule is ina pplicable, t he correspondence principle has always been a matter of controversy [see the references after (1.401)1to be resolved in this text . In Chapters 10 a nd 8 we shall present a new quantum equivalence principle which is based on an application of simple geometrical principles to path integrals and which will specify a natural and unique passage from classical to quantum mechanics in any coordina te frame. 22 The configuration space may carry curvature and a certain class of torsions (gradient torsion). Several arguments suggest t hat our principle is correct. For t he above systems with a Hamiltonian which can be expressed entirely in t erms of generators of a group of motion in t he underlying space, the new quantum equivalence principle will give t he same results as t he group quantization rule.

1.16

Scattering

Most observations of quantum phenomena arc obtained from scatt ering processes of fundamental particles.

1.16.1

Scattering Matrix

Consider a particle impinging with a moment um p" a nd energy E = E" = p!/2M upon a nonzero potent ial concentrated around t he origin. After a long t ime, it will be found far from the potential with some momentum Pb. The energy will be uncha nged: E = Eb = pV 2M. The probability amplitude for such a process is given by the t ime evolution amplitude in the moment um representation (1.470)

where the limit tb ----> 00 and t" ----> -00 has to be taken. Long before and after t he collision, t his amplitude oscillates with a frequency w = Elf! characteristic for frcc particles of energy E. In order to have a time-independent limit , we remove these oscillations, from (1.470), and define the scattering matrix (S-matrix) by t he limit (1.471 )

Most of the impinging particles will not scatter at all, so that this amplitude must contain a leading term, which is separated as follows: (1.472)

where (1.473)

shows the normalization of the states [recall (1.186)1. T his leading term is commonly subtracted from (1.471 ) to find the true scatt ering amplitude. Morcover, Z2 H.

Kleinert, Mod. Phys. Lett. A 4, 2329 (1989) (http: //vww .physik .fu- ber lin.de/ Phys. Lett. B 236, 3U'i (1990) (i~d. http/202) .

~kleinert/ 199) ;

68

1 Fundamentals

since potential scattering conserves energy, the remaining amplit ude cont ains a 6function ensuring energy conservation, and it is useful to divide this out, defining the so-called T-matrix by t he decomposition (1.474)

From t he definition (1.471) and the hermiticity of fl it follows that the scattering matrix is a unitary ma trix. This expresses the physical fact that t he total probability of an incident particle to fe-emerge at some t ime is unity (ill quantum field t heory the situation is morc complicated d ue to emission and absorption processes). In the basis states IpID) introduced in Eq. (1.180) which satisfy the completeness relation (1. 182) and are normalized to unity in a finite volume V, the unitarity is expressed as (1.475) m'

m'

Remembering the relation (1.185) between the discrete states IpIn) and their cont inuous limits Ip), we see t hat (1.476)

where £3 is t he spatial volume, and Pb and P:;' a re the discrete momenta closest to Pt> a nd p". In the continuous basis Ip), the unitarity relation reads (1.477)

1.16.2

Cross Section

The absolute square of (pt>ISl p,,) gives the probability PPb-P~ for t he scattering from the initial momentum state p" to t he final momentum state Pt>. Omitting the unscattercd particles, we have

p,.-"

1

~ L'

"

,

2.M(0)2.M(E, - Eo)l(p, ITI Po) 1 .

(1.478)

The factor 0(0) at zero energy is made finite by imagining the scattering process to take place with a n incident t ime--indcpendent plane wave over a. finite total time T . Then 21TM(0) = J dteil>'t/Ti le=o = T, and the probability is proportional to the time To (1.479)

By summing this over all discrete final momenta, or equivalently, by integrating this over the phase space of t he final momenta [reca ll (1.184 )], we find t he total probability per unit time for t he scattering to take place (1.480)

69

1,16 Scattering

The momentum integral can be split into an integral over the fi nal energy and the final solid angle, For non-relativistic particles, this goes as follows (1.481)

where dO = d¢>bd cos Bb is the element of solid angle into which the particle is scattered, The energy integral removes the a-function in (1.480), and makes Pb equal to Pa ' The differential scattering cross section du I dO is defined as the probability t hat a single impinging particle ends up in a solid angle dO per unit time and unit current density, From (1.480) we identify

du dF l i M p zl dO = dO] = £3(21r!i)321rrIITp~pJ ],

(1.482)

where we have set (1.483)

for brevity, In a volume £3 , t he current density of a single impinging particle is given by the velocity v = plM as

.

1 P

(1.484)

J = £3 M '

so that the differential cross section becomes (1.485)

If the scattered particle moves relativistically, we have to replace t he constant mass M in (1.481) by E = Ji? + M2 inside the moment um integral, where P = Ipl, so that =

J {JO (2:!i)3 J 10"'" dEEp, (2:!i)3

dO dO

dpp2

(1.486)

In the relativistic case, the initial current density is not proportional to plM but to the relativistic velocity v = pi E so that

. J =

1 P

£3£'

(1.487)

Hence the cross section becomes (1.488)

70

1.16.3

1 Fundamentals

Born Approximation

To lowest order in t he interaction strength, the operator

S'"

1-

S in

iV /n.

(1.471) is (1.489)

For a time-independent scattering potential , this implies (1.490)

where (1.491 )

is a function of the momentum transfer q =: Pb - p" only. Then (1.488) reduces to the so called Darn approximation (Dorn 1926) (1.492)

T he amplit ude whose square is equal to the differential cross section is usually denoted by fp~ P6' i.e. , one writes

dpe - il>' = e- i"'/4,J;,

(1B.6)

which goes into Fresnel's integral formula (1.333) by substituting p -- pM.

Appendix

Ie

The Asymmetric Top

The Lagrangian of the asymmetric top with three different moments of inertia reads

(IC.I)

86

1 Fundamentals

It has the Hessian metric [recall (1.12) and (1.384)] gil

!hI

U31

Ie sin2 fJ + I( cos2 f3 - (Ie - I.,l sin 2 thin2 'Y, - (Ie - 1'l)sin .Bsin,),cos')', I, oos{3, + (Ie - 1'1) 5in7 ,)"

977

1'1

932

0,

(IC.2) rather than (1.459). The determinant is

(IC.3) and the inverse metric has the components

~{I'l +(h - I'l)Sin7'Y}I< ,

9 11 9

21

g

1 - sin!3sin1C061'(1e - I,,) / {,

=

9

931

!{cosl3l-siU7 1'(Ie -

l2

!{Sin2 Plh

9 9

-

I1}l - I"llI"

sin 2 "'1(Ie - I")]}!,,

~{sin.6cosl3sin-YOOS 1'(I'I - I{)}I(,

y32

9 9

1

_ {mnl PIe!.,

33

9

+ coo 2 PI,,!, + cos7 {3sin2 i (Ie - I'I)I,}.

(ICA)

From t his we find the components of t he llicmann connection, the Christoffel symbol defined in Eq. (1.70):

Tl + [{fell/hI."

fnl

[cos.6cosl'siwy(l~ - I'll, -

f211

{cosfJIsin 2 -y(Il - I~ - (Ie - It)lId + I'I(Ie

f3J

1

[':n l ['32 1'33

{COS1'sin 1'[I~ -

II + (Ie -

+ [ '1 -

Idllj2sin .6 I e I '1 ' I,,)I(]} j2Ie I'I'

0,

1

[sin 2 "({ll - I~ - (Ie - J'I)Id - [qUe - 1'1 + IdIl2sin.BI~I",

1

0,

I' n 2 = (cosfJsin ,8[sin1 "'f(ll - I~ - Idle - I,,)) - Ie(le - I dl} / leI" , 1'21 2 1':11 2

{cosfJcos "'fsill"'f[Il - I~ - Idle - I'I)]}/21eI'I' {sinfJ[sin2 "'f (ll - I~ - I d l e - 1'1» - Idle - I'I - IdJ}/2IeI'I'

f':n 2 f'al

0,

2

0,

1'33

[cos"'fsin"'f(ll - I~ - I ((Ie - I 'I))]/2IeI'I'

f'n 3 = {cos "'fsin "'f[sin2 fJ( I eI 'I( I e - 1'1) - IdIl - I~) + IlUe - I~» + Ul - I~)I( - fl(Ie - I'I)]}/I eI'Ih, 1'21

3

{sil? fJ[sin 2 "'f(2I{I"(I,, - Ie) + IdIl - I~ ) - Il(Ie - I ,,)) +IeI,,(Ie - 1'1)

+ I'IId I "

- Id] - sin "f«(Il - I~)I( - Il(Ie - 1'1» 2

Notes 8Jld

87

RcfcrellCCS

1'31 3

-['lie (I! + 1'1 - l d}/2SiUp l!I'1 I" [C06PC06'YSin'y(I1- l~ - ld/{ - l'1» I/21{/'1'

006'Ysin'Y(I'1-/e)/lc, {cosp[sin 2 'Y (I~ - 11 + (l{ - 1'1)/d

+ l'1(1{

- 1'1 + Id]} /2sin p I'1 l{, (IC.5)

O.

The other components follow from the sy mmetry in the first two indices 1'1' and r 1''' \ respectively.

Notes and References For more details see some standard textbooks: I. Newton, Mathematische Prinzipien der Naturkhre, Wiss. Buchgcscllschaft, Darmstadt, 1963, J.L. Lagrange, Analytische Mechanik, Springer, Berlin, 1887; G. Hamel, Theoretische Mechanik , Springer, Berlin, 1949; A. Sommerfeld, Mechanik, Harri Deutsch, Frankfurt, 1977; W. Weizel, Lehrbuch der Theoretischen Physik, Springer, Berlin, 1963; H. Goldstein, Classical Mechanics , Addison-Wu'x(t) "

~Ix(t) , -

x(' -

,)1 ·

(2 .90)

Thcy are two differcnt discrete vcrsions of the time derivative Ot, to which both reduce in the continuum limit € --+ 0: (2.91)

if they act upon differentiable functions. Since the discretized timc axis with N + 1 steps constitutes a one-dimensional lattice, the difference operators 'V , 'V are also called lllttia derivatives.

2.2 Exact Solution for the.Free Particle

105

For the coordinates x .. = x(t .. ) at the discrete t imes t.. we write 1

, 1 -(x ,

-(X .. +l n -

N

x .. ),

~

n

~

0,

N+l~n~ 1.

x .. -t} ,

(2.92)

The time-sliced action (2.89) can then be expressed in terms of 'Vxn or 'Vx .. as (writing x .. instead of oXn)

(2.93)

°

In this notation, the limit f ...... is most obvious: The sum f 2::.. goes into the integral ft~dt, whereas both ('VX .. )2 and ('Vx .. )2 tend to i;2, so that

~

. . . 1~· dt~

(2.94)

i;2.

Thus, the time-sliced action becomes t he Lagrangian action. Lattice derivatives have properties Quite similar to ordinary derivatives. One only has to be careful in distinguishing 'V and 'V . For example, they a1low for the useful operation summation by parts which is analogous to integration by parts. Recall the rule for the integration by parts

f

d'g(,)j(,)

~ 9(')f(')I:: -

f

d'g(t)f(')·

(2.95)

On the lattice, this relation yields for functions f(t) ...... x .. and g(t) ...... P.. : N

N+l f

,,

L P..-'\i'x ..

= p.. x ..

" loN -H f L ('VPn)x ...

(2.96)

This follows directly by rewriting (2.35) . For functions vanishing at the endpoints, i.e., for XN+ l = Xo = 0, we can omit the surface terms and shift the range of the sum on t he right-hand side to obtain the simple formula [see also Eq. (2.47)] N+l

N

N+l

L P. '''". ~ - L("P.)x. ~ - L ("P.)x •.

(2.97)

The same thing holds if both pet) and x(t) are periodic in the interval tt> - t a , 50 that 1'1 = PN+l, Xo = XN+l. In this case, it is possible to shift the sum on the right-hand side by one unit arriving at the more symmetric-looking formula N+l

.-,

N+l

L P. "x. ~ - L ("P.)x • . n= 1

(2.98)

2 Path Integrals -

106

Elementary Properties and Simple Solutions

In the time-sliced action (2.89) t he qua ntum fluctuations xn (=ox,,) vanish at t he ends, so that (2.97) can be used to rewrite N+l

L

N

('\7xn)2 = -

n: l

L

(2.99)

xn'V'Vx".

n: l

In the 'Vx" -form of t he action (2.93), the same expression is obtained by applying formula (2.97) from the right- to the left-hand side and using the vanishing of Xo and IN+!: N+l

N

N

L ('lx. )' ~ - L x.'ii'lx. ~ - L x.'ii'lx•.

(2 .100)

n= ]

n= O

The right-hand sides in (2.99) and (2.100) can be written in mat rix form as N

- L xn'V\lx"

N

- -

n=1

L

x"(VV),,n,Xnl,

",n'=]

N

- L In \7\i'x"

N ~

,,=1

-

L X"{VV')"n'X,,I, n,,,' = 1

(2.1Ol)

with the same N x N -matrix

_ \7\7

_

==

- 2 1 0 1 - 2

1

,

0 0

0 0

0 0 (2. 102)

'\7\7 =: 2"

0 0

1 -2 1 1 -2 0

0 0 0 0

This is obviously the lattice version of the double time derivative fit, to which it reduces in the continuum limit l -+ O. It will therefore be called the lattice Laplacian. A furt her common property of lattice and ordinary derivatives is that thcy can both be diagonalized by going to Fourier components. When decomposing

(2. 103)

and applying the lattice derivative V, we find

I:

Vx(t n }

~

dw

~ (e- iw(t,,+ - TQ this gives directly (2.345) . Physically speaking, thc path has at high temperatures "no (imaginary) time" to fluctuate , and only one term in the product of integrals needs to be considered .

139

2.9 DellSity Matrix

If H (p, x) has the standard form

1" + V(x) ,

Hlp, x) ~ 2M

(2.348)

the momentum integral is Gaussian in p and can be done using the formula 00 / - 00

dp

e- ar /211

21rn

___ 1_

(2.349)

- J21rlia'

This leads to the pure x-integral for the classical partition function

(2.350) In the second expression we have introduced the length

(2.351) It is t he thermal (or Euclidean) analog of the characteristic length l(tb - t a ) intro-duced before in (2.126) . It is called the de Broglie wavelength associated with the tempemture T = l/k n f3 or, in short, t he thermal de Broglie wavelength. Omit ting t he x-integration in (2.350) renders t he hU'ge-T limit p(x), the classical particle distribution ~ Z- l cl

1

Ie (liP) e

- 9(.,)

(2.352)

For a free particle, the integral over x in (2.350) diverges. If we imagine the length of t he x-axis to be very large but finite , say equal to L, the partition function is equal to

(2.353) In D dimensions, this becomes

(2.354) where Vo is t he volume of t he D-dimensional system. For a harmonic oscillator with potential MW 2 X 2 /2, the integral over x in (2.350) is finite and yields, in the D-dimensional generalization ID (2.355) Z" ~ IDti,fi)' where

C ==

{lj; f3Mw 2

(2.356)

140

2 Path Integrals -

Elementary Properties and Simple Solutions

denotes t he classical length scale defined by the frequency of the harmonic oscillator. It is related to t he quantum-mechanical one -\.., of Eq. (2.301 ) by (2.357) Thus we obtain the mnemonic rule for going over from the partition function of a harmonic oscillator to that of a free particle: we must simply replace __ 0 l.,--L,

oc

(2.358)

"~!fiM

w

- 0

21f

L •

(2.359)

The real-time version of this is, of course, 1

(2.360)

-~

w

w_ o

Let us write down a path integral representation for p(x). Omitting in (2.337) the final trace integration over XI> == Xa and normalizing the expression by a factor Z - l , we obtain

(2.361) The thermal equilibrium expectation of an arbitrary Hermitian operator 6 is given by (2 .362)

In the local basis Ix), this becomes

" (2.363)

An arbitrary function of the position operator

x has the expectation

T he particle density p(x,,) determines the therma l averages of local obscrvablcs. If f depends also on t he momentum operator p, then the off-diagonal matrix element s (xb le- Plf lx,,) arc also needed. They arc contained in the density matrix introduced for pure quantum systems in &t. (1.221), and reads now in a thermal ensemble of temperature T: (2.365)

141

2.9 DellSity Matrix

whose diagonal values coincide with t he above particle density p(x,,). It is useful to keep the a nalogy between qua ntum mechanics and qua ntum statistics as close as possible and to introduce the time t ranslation operator along the imaginary time axis (2.366)

defining its local matrix elements as imaginary or Euclidean t ime evolution a mplitudes (2.367)

As in the real-time case, we shall only considcr t he causal time-ordering r~ > T". Otherwise the partition function and the density matrix do not exist in systems with energies up to infinity. Given the imaginary-time amplitudes, the partition function is found by integrating over the diagonal elements

z ~ [: dx(x hPlx 0),

(2.368)

and the density matrix (2.369)

For the sake of generality we may sometimes also consider the imaginary-time evolution operators for time-dependent Hamiltonians a nd the associated amplitudes. They are obtained by time-slicing the local matrix elements of the operator (2.370)

Here T.. is an ordering operator along the imaginary-time axis. It must be emphasized t hat the usefulness of the operator (2.370) in describing thermodynamic phenomena is restricted to the Hamiltonian operator H(t) depending very weakly on the physical time t . T he system has to remain close to equilibrium at all times. T his is t he range of validity of the so-called linear response theory (see Chapter 18 for more dctails). The imaginary-time evolution amplitude (2.367) has a path integral representation which is obtained by dropping t he final integration in (2.334) and relaxing the condition X~ = X,,: (2.371)

The time-sliced Euclidean action is

A;' ~

N+'

z= (- ip"(x" - x"_.J + 1°O

dx

Dx~ II

n=l

-CQ

•.

(2.377)

J2rrfll'.jM

It contains no extra 1/J27r1if./M factor , as in (2.374), due to t he trace integration over t he exterior x. The condition x{Ii(3) = x(O) is most easily enforced by expanding x{r) into a Fourier series X

()r = ;c.. ~

m= - CQ

1

I7ff'"'7"1 e

yN

+I

- >w""

x m,

(2 .378)

143

2.10 Quantum Statistics of the HarmoIlic Oscillator

with the Matsubara frequencies 2.m wm=27rmkBT/Ii= 1if3'

m=O,± I,±2, ....

(2.379)

When considered as functions on the entire T-axis, the paths are periodic in hf3 at any T , i.e., (2.380) Thus the path integral for the quantum-statistical partition function comprises all periodic paths with a period r~f3. In the time-sliced path integral (2.374), the coordinates X(T) arc needed only at the discrete times T" = nf. Correspondingly, the sum over m in (2.378) can be restricted to run from m = - N/2 to N/2 for even N and from - (N - 1)/2 to (N + 1)/2 for odd N (see Fig. 2.3). In order to have a real X(T,,), we must require that

Xm = x :m

(modulo N

+ I ).

(2.381)

Note t hat the Matsubara frequencies in t he expansion of the paths x( T) are now twicc as big as the frequencies 11m in t he quantum fluctuations (2. 110) (after analytic continuation of tb - ta to -ili/kBT). Still, t hey have about the same total number, since they run over positive llnd negative integers. An exccption is the zero freque ncy Wm = 0, which is included here, in contrast to the frequencies 11m in (2.110) which . This is necessary to describe paths wit h run only over positive m = 1, 2, 3, . arbitrary nonzero endpoints XI> = Xa = X (included in the trace) .

2.10

Quantum Statist ics of the Harmonic Oscillator

The harmonic oscillator is a good example for solving the quantum-statistical path integral. The T-axis is sliced at T" = m, with ( :::::: hj3/(N + I) (n = 0, ... , N + 1), ILIld the partition function is given by the N -- 00 -limit of the product of integrals (2.382) where ~ is the time-sliCtJd Euclidcan oscillator action (2.383) Integrating out t he x" 's, we find immediately ZN =

....

I VdetN+l (_(2 \1\1 + ( 2w 2 ) .

(2.384)

Let us evaluate the fluctuation determinant via the product of eigenvalues which diagonali1.e the matrix _€z \11i7' + E2 W Z in the s liced act ion (2.383). They are (2.385)

144

2 Path Integrals -

Sin fW

Elementary Properties and Simple Solutions

sin (W~

2

2

N"'ewll

N z odd

Nn l - cos€:lJ)

-,

- NO

N+L 2

-

-2

I-cos

(lJ)

-

N- ' 2

- 2

F ig ure 2.3 Illu~t ration of the eigenvalues (2.385) of t he fluct uation matrix in the action (2.383) for even and odd N. with the Matsubara frequencies

Wm .

For w = 0, the eigenvalues arc pictured in Fig. 2.3. The

action (2.383) becomes diagonal after going to the Fourier components X m . To do this we arrange the real and imaginary parts Rex", and Irnx", in a row vector (Re x\, 1m

XI;

Re x2, 1m %2; ... ; Re :1:", 1m x .. ; . .. j,

a nd see t hat it is related to the time-sliced positions x" = X(T.. ) by a traru;formation matrix with the rows

= (Tm)n x "

JN:

1

(~, ~

N:

12lf . 1,sin

N:

12,.. . 1,

c~~21f'2 sin~21r·2 . N+l

'N+l

m

'

m

)

... ,cos N+127f.n,siO N+ 12'1f.n, ... "x" .

(2.386)

For each row index m = 0, ... ,N, t he column index n runs from zero to N / 2 for even N, and to (N + 1)/2 for odd N. In the odd case, the last column sin ffi 211"· n with n = (N + 1)/2 vanishes

identically and must be dropped, so t hat the number of columus in T mn is in both CIlSCll N + 1 , as it should be. For odd N, t he second-last column of Tmn is an alternating sequence ± I. Thus, for a proper normalization, it ha:; to be multipli(.x1 by an extra normalization factor I/.[i, ju~t as the elements in the first column. An argument similar to (2.1 15), (2.116) shows that the resulting matrix is o rthogonal. Th us, we can diagonalize the sliced action in (2.383) as folloW1j for

N = even,

(2.387) for Thanks to t he orthogonality of T mn, t he measure

fl n f~ dx{Tn)

transforms simply into for

1- 1-

dxo

-

(N -'1/'1-

dxu'+I)n Q I

_ 00

1-

N = odd.

dRex m _oc dlmxm

N = even,

(2.388) for

N = odd.

145

2.10 Quantum Statistics of the HarmoIlic Oscillator By performing the Gaussian integrals '\\ll ohtain the partition function

(2.389) T hanks to the periodicity of t he eigenvalues under the replacement n __ n + N + I, the result hIlS become a unique product expression for both even and odd N . It is important to realize t ha.t contrary to the fluct uation factor (2.160) in the real-time ampli t ude, the partition function (2.389) contains t he square root of only positive eigcnmodes lIS a unique result of Gaussian integrations. There are no phase subtleties as in the F"rcsnel integral

(1.333). To calculate the product, wc observe that upon decomposing

sin~ W;f = (I + ~ W;f) (I - COS W;f) ,

(2.390)

the sequence of first factors l + cosW",f = l + cos~

(2.391)

N+l

2

runs for m = I , . . . N through the same values as the sequence of second factors

, - = -Wmf -= 2

Jrffi N+I- m 1 - cos - - := 1 + cos Jr "-",'ceo""

N+ l

N+I

(2.392)

'

except in an opposite order. ThIlS, separating Ollt the m = 0 -term, we rewrite (2.389) in the form

["

z~ = ~ II 2(1_~WmE) (W

,., _ 1

2

]-' U"II (1 + .. I

~~w~

4 Slll'2

)]-'" .

!o!m.!

2

(2.393)

Thc first factor on the right- hand side is the quantum·mechanical fluctua.tion determ inant of the free-part icle determinant detN( - f 2V'V ) = N + 1 [sec (2. 123)], so that we obtai n for both even and oddN (2.394) To evaluate the remaining product, we must distinguish again between even and odd For even N, where every eigenvalue occurs twice (see Fig. 2.3), we obtain

ca8e!i

of N .

(2.395) For odd N, the term with m = (N + 1)/2 occurs only once lI.Ild must be treated separately so t hat

II

•. T [( l + fW " ) '" (" - "1' ( 1+ fW ")] - ' Z~ = ~ 1iw 4 ",.1 4 sin~ Nfl

.

(2.396)

We now introduce t he parameter We, the Euclidean analog of (2.161), via the equations (2.391)

146

2 Path Integrals -

Elementary Properties and Simple Solutions

In the odd case, the product formu la l6

., 1

(N_.", [

2 slu[(N + l)xJ sin2x (N+ l )

sm x 1 - si n 2 ..."

II

7i"1m

... .. 1

(2.398)

[similar to (2. 163)] yields, with x = (;V/2, ZN = kaT [

'"

fiw

Sinh[(N+l)W.~/2Il - · .

1

Sinh(W. f/2)

N

+1

(2.399)

In the even case, the formula l ?

1 sin[(N + l)x] sinx (N + 1)

(2.400)

produces once more t he same ruml t !IS in Eq. (2.399 ). Im;crting Eq. (2.397) leads to the partition function on the sliced imaginary time axis:

1

N

(2.401)

Z'" = 2 sin h (liWeP/ 2) " T he partition function can be ex pandod into t he following series

Z:: =

e - row. nkBT

+ e - 3ft,;.,· n k

IJ

T

+ e - M i;;. ! 2k s T + .

By comparison with t he general spectral expansion (2.326), the system:

W{)

display the energy eigenvalues of

(n+DliWe.

En =

(2.402)

(2.403)

T hey show the ty pica l linearly rising O6Cillator sequence wi th _ 2 . h ""f w. = ,€8.I"SLll "'2

(2.404)

playing the role of t he frequency on t he sliced timc axis, and hWe /2 being the zero-point energy. In the continuum limit € ..... 0, t he t ime-sliccd partition function goes over into the usual OlSCil1ator partition fUIlction 1 (2.405) Z", = 2sinh(.81iw/2) .

Z:

In D dimensions this becomes, of course, [2 sin h(.BtlW/2)[- J), due to the additivity ofthe action in each component of x . Note that the continuum limit of the product in (2.394) can also be taken factor hy factor. Then Z", becomes (2.406)

According to formula (2.171 ), t he product a nd we find with :r: =

rr: ",1(1+ ",(, ) converges rapidly against sin hx/x

fu.J.B/2 Z", = kaT

ru..;/ 2k a T

hw sinh{Iiw/2knT ) """:-;;:::-:=-:c:-:;-;.,.,-;;:~ 161.S. Gradshtcyn and I. ~I. Ryzhik , op.

17ibid., formula 1.391.3.

2si nh{.BIiw/2)·

c it., Formula 1.391.1.

(2.407)

14 7

2.11 Time-Dependent Harmonic PotelltiaI

As discussed after &j. (2. 181 ), the continuum limit can be taken in each factor since the product in (2.394) contains only ratios of frequencies. Just lIS in the quantum-mechanical case, this procedure of obtaining the continuum limit can be summarized in the sequence of equations arriving at a ratio of differential operators 12 ZN [detN+I( _ f2VV + f2W2) r /

"

(_ f2VV)r 1/2 [det N+ L(_ e V'fJ + ~2W2)l - l/1 2

[det'

det'",+1( t2VV)

N+l

k.T

[det(-a~ det'(

h

2 +w )] - 1/2 = kBT

an

rr [w~ + oo

hw m=l

2 w ] -'

(2.408)

w~

In the W = 0 -determinants, the zero Matsubara frequency is excluded to obtain a finite expression. This is indicated by a prime. The differential operator - ~ a.ct~ on real functions which are periodic under the replacement T -> T+ hJ3. Remember that ca.ch eigenvalue w~ of - ~ occurs twice, except for the zero frequency Wo = 0, which appears only once. Let us finally mention that the resu lts of this section could also have hccn obtained directly from the quantu m-mochanical amplitude (2.l73) [or wit h the discrete t imes from (2.197)1 by an analytic continuation of t he time difference t~ - ta to imaginary values - i(Tb - Ta):

(X~TbI XaTa)

=

~VSinhW(:

x

exP{ - 2~sinhw~: T)(x~+X!)COSh W(Tb - Ta) - 2X~Xa] }.

ra) (2.409)

By setting x = X~ = Xa and integrating over x, we obtain [compare (2.331 )]

2sinh[w(Tb

ra)/2]'

(2.410)

Upon equating Tb - Ta = hJ3, we retrieve the partition function (2.405). A similar treatment of the discrete-time version (2. 197) would have led to (2.401). The main reason for presenting an independent direct evaluation in the space of real periodic functions was to display t he frequency structure of periodic paths and to see the difference with respect to the quantum-mechanical paths with fixed ends. \Ve also wanted to show how to handle the ensui ng product expressions. For applications in polymcr physics (see Chapter 15) one also needs the partition function of all path fluctuations with open ends

/

00 dxb / 00 dxa (xbTb IXara) =

_ oc

_ oc

V2,,-h

W(Tb

I

\/2,,-h{Tb

ra)/M

sin h[w(Tb

Ta) 21Th Ta)] Mw

1

Mw.,Isinh [w(Tb

(2.411)

Ta)['

The prefactor is .jfi times the length scale >..., of &j. (2.301).

2.11

T ime-Dep e nde nt Harmonic P ote ntia l

It is often necessary to calculate thermal fluctuation determinants for the case of s time-dependent

frequency

neT) wh ich is periodic under T __ r + hJ3. (XbTb IXaTa )

=

~

JV X

J'D

As in Section 2.3.6,

v.'C

consider t he anlplitude

P - r'dT[-ip~+p'nM + M{l'(T)r'nl/~

21Tn e

b a a xo'

(2.663)

After a quadratic completion in xo , the total exponent in (2.663) reads

iMwt. - 2" [- (Xb,

, tan IWt. (tl> + xa)

, ta)/2I + (XI> - xa).

5lfi WL

(1 tb

.MWL ( X b + xa Yb - Ya )' -'-h- tan [wI.{t b - ta)/2] Xo - - 2 - - 02t;-:a-=nf[w~,-';(t~,"-ot'}"/2"1 o +i Mwt, [(Xb + xa)2 tan[wI.{tb _ t a)/2] + (Yb - Ya)2 21i 2 2tan[wL(tb - t a)/2] .MWL (2.664) +t21i(Xb + Xa)(Yb - Ya).

1

The integration MWL f~oo dxo/21r1i removes the second term and results in a factor

MWt. ,I 1r1i 21r1i ViMwL tan[wL(tb

(2.665)

ta)/2]·

By rearranging the remaining terms, we arrive at t he a mplitude 1 M 3 Wdtb - t a)/2 a (X!>tb lx..t ) = V21rili(tb _ tal sin[wd tb _ ta)/2] exp h(AcI + .A..rl ,

[i

1

(2.666) with a n action

M { (z, - zo}' Acl = -2 tb

ta

+ WL -2 cot[WL(tb +

ta)/2] [ (Xb - xa) '

W...(XaYb - XbYa)}

(2.667)

I'

(2 .668)

and the surface term

A.f = -MWL 2- (XbYb - Xa Ya)

+ (Yb - Ya) 'J

e

= 2cBxy a·

184

2.18.4

2 Path Integrals -

Elementary Properties and Simple Solutions

C lassical Action

Since t he action is harmonic, the amplitude is again a product of a phase ei..A", and a fluct uation factor. A comparison with (2.639) and (2 .656) shows that t he surface term would be absent if the amplitude (Xbtb lx..t"h were calculat ed with the vect or potential in the axially symmetric gauge (2.638). Thus Acl must be equal to t he classical act ion in this gauge. Indeed, t he orthogonal part can be rewritten as

AJ = i~~ dt { ~ ~(XX + YiJ) + ~ [x(- x + Wl,y) + y(- y -

Wl,X)] }.

(2.669)

T he equations of motion are (2.670)

redueing the action of a classical orbit to M(. "la="2 M ([. . J+ [ . - Y,,Y,, . [) · "'",' M. Sw.uki, Physica A 191 , 501 (1992).

The path integral representation of the scattering amplltudc is developed in W.B. Campbell, P. Finkler, C.E. Jones, and M.N. Mishcloff, Phys. Rev. D 12, 12, 2363 (1975). See also: H.D.I. Abarband and C. Itzykson, Phys. Rev , k tt. 23, 53 (1 969); R. Ro3cnfelder, sec Footnote 37. The alternative path integral representation in Section 2.18 is dlle to M. Roncadelli, Europhys. Lett. 16,609 (1991 ); J. Phys. A 25, L997 (1992); A. Defendi and M. RoncadelIi, Europhys. Lett. 21 , 127 (\993).

AI>iV1/'Tfi I>Wtv;.

You stir what should not be stirred. H EROOOTUS

3 External Sources, Correlations, and Perturbation Theory Important informatio n on every quantum-mechanical system is carried by thc correlation functions of the path x(t). They are defined as the expectation values of products of path positions at different times, x(td . .. x(t n ), to be calculated as functional averages. Quantities of this type are observable in simple scattering experiments. The most efficient extraction of correlation functions from a path integral proceeds by adding to the Lagrangian an external time-dependent mechanical force term disturbing the system linearly, and by studying the response to t he disturbance. A similar linear term is used extensively in quantum field theory, for instance in quantum electrodynamics where it is no longer a mechanical force, but a source of fields , i.e., a charge or a current dcnsity. For t his reason wc shall call this term generically source or current term . In this chapter, the procedure is developed for the harmonic action, where a linear source term does not destroy the solvability of the path integral. The resulting amplitude is a simple functional of t he current. Its functional derivatives will supply all correlation fu nct ions of t he system, and for t his reason it is called the genemting junctional of t he theory. It serves to derive the celebrated Wick rule for calculat ing the correlation functions of an arbitrary number of x(t). This forms the basis for pcrt urbat ion expansions of anharmonic t heories.

3 .1

External Sources

Consider a harmonic oscillator with an action

(3.1) Let it be disturbed by an external source or current j(t) coupled linearly to t he particle coordinate x(t) . The source action is A;

~

', 1'.

dt x(t)j(t).

209

(3.2)

210

3 External Sources, Correlations, and Perturbation Theory

The total action

A = A.+Aj

(3.3)

is still harmonic in x and X, which makes it is easy to solve the path integral in t he presence of a source t erm. In particular, the source t erm does not destroy the factorization property (2.151) of the time evolution amplitude into a classical amplitude eiA ;, _ t) t~

+ Xbsi n w(t - ta) ]i(t).

211

3.1 ExtCrIlW SOllI"Cffi

Consider now t he fluctuating part of t he action, A ft = ..4..,,8 + A i,H. Since Xcl(t) extremizes the action without the source, A ft contains a term linear in ox(t) . After a partial integration [making use of the vanishing of ox(t) at the ends] it can be written as

(3. 12) where D",2(t,t') is t he differential operator

Dw,(t, t!) ~ (- if, - w'),(t - t!) ~ ,(t - t')( - a?,

-

w') ,

t, t'

E

(to, to) .

(3. 13)

It may be considered as a functional matrix in the space of the t-dependent functions va nishing at ta, tb. The equality of the two expressions is seen as follows. By partial integrations one has

f,_I" dtf(t)if,g(t ) ~ f,_I" dtif,f(t)g(t) ,

(3. 14)

for any f(t) and g(t) vanishing at t he boundaries (or for periodic functions in t he interval) . The left-hand side can directly be rewritten as It~ dtdt! f(t)o(t - t')8;9{t!), the right-hand side as If: dtdt' 8i f(t)o(t - t')9(t') , and after further partial integratiom;, os J dtdt' f(t)if,'(t - ng(t) . The inverse D:!(t, t') of t he functional matrix. (3.13) is formally defined by the relation

i'-"dt' D",,{t",

t')D~!(t' , t) = o(t" - t) ,

(3.15)

which shows that it is the standard classical Green function of the harmonic oscillator of frequency w;

This definition is not unique since it leaves room for an additional arbitrary solution H (t , t') of t he homogeneous equation II~ dt' Dw2(t", t') H (t' , t) = O. This freedom will be removed below by imposing appropriate boundary condit ions. In the fluctuation action (3. 12), we now perform a quadmtic completion by a shift of ox(t) to

,x(t)

1 (',

= 'x(t) + M f,

,-

(3. 17)

dt'Gw,(t , t')j(t').

Then the action becomes quadrat ic in both

ox and j:

[M

1

I" dt f,_I" de "2,x(t )Dw,(t,t'),x(t') - 2M 1 j (t)Gw,(t, t')j(t') . A . ~ f,_

(3.18)

The Green function obeys the same boundary cond it ion as the fluct uations ox(t): G",.(t , t')=O

for

f

t=

tb ,

1 t arbitrary,

t' arbit rary, t' = ta .

(3.19)

212

3 External Sources, Correlations, and Perturbation Theory

Thus, the shifted fluctuations ox(t) of (3. 17) also vanish at the ends and run through the same functional space as the original ox(t). T he measure of path integration f Vox(l) is obviously unchanged by the simple shift (3.1 7) . Hence t he path integral f V ox over eiAft / h with the action (3.18) gives, via the first term in A ft, the harmonic fluctuat ion factor F",(t b - tal calculated in (2. 169):

w

1

(3.20)

The source part in (3. 18) contributes only a trivial exponential factor PUt = cxp

{*A,f1 } ,

(3.21)

j

whose exponent is quadratic in j(t) : A j,H = - 2M 1

1" 1" Ie

dt t~ dt'j(t)Gw~(t , t')j(t').

(3.22)

The total time evolution amplitude in the presence of a source term can therefore be written as t he product (3.23)

where (Xbtb lxat)", is t he source-free t ime evolution amplitude

(3.24)

and Pj,cl is an amplitude containing the classical action (3. 11 ): Pj,cI =

e(i/f "

(3.43)

the second line in (3.41) can be rewritten as

(3.44) By choosing the initial conditions

6.(t, t)

=

0,

Li(t, t') I,':=, = -1 ,

(3.45)

we satisfy the inhomogeneous differential equation (3.27) provided 6.(t , t') obeys the homogeneous differential equation

[-if - n'(t)I£>(t, t')

~ 0,

fo, t > t .

(3.46)

This equation is solved by a linear combination

£>(t, t')

~

o(f)«t) + ptt')q(t)

of any two independent solutions 1J(t) and

[-if - n'(t)I((t)

~ 0,

~(t)

(3.47)

of the homogeneous equation

[-if - n'(t)Mt) ~ O.

(3.48)

T heir Wronski determinant W = ~(t)1j( t ) - ~(t)1J( t) is nonzero and, of course, timeindependent, so that we can dctermine t he coefficients in t he linear combination (3.47) from (3.45) and find

£>(t,

n ~ ~ [«tMt') - «t')"(t )l·

(3.49)

The right-hand side contains the so-called Jacobi commutator of the two funct ions ~(t) and 1J(t). Here we list a few useful algebraic properties of 6.(t, t'):

(3.50) (3.51)

216

3 External Sources, Correlations, and Perturbation Theory

Il(', 'o)8"Il(t" to) - Il('" ,) ~ Il(t" 'o)8,1l(', (0).

(3.52)

The retarded Green function (3.40) is so far not the unique solution of t he differential equation (3.27), since one may always add a general solution of t he homogeneous differential equation (3.48):

Gn,(',

n ~ a(, - nll(t, n + a(nW) H(")"(' ),

(3.53)

with arbitrary coefficients a(t') and b(t'). This ambiguity is removed by the Dirichlet boundary conditions 0,

tb

=I- t, (3.54)

Imposing these upon (3.53) leads to a simple algebraic pair of equations

a(')Wo) a(')W,)

+ b(,)"(,o) + b(')"(t,)

~

0, Il(t ,t,).

(3.55) (3.56)

Denoting the 2 x 2 -coefficient matrix by

(3.57) we observe that under the condition (3 .58) the system (3.56) has a unique solution for the coefficients a(t} and b(t) in the Green function (3.53). Inserting this into (3.54) and using the identity (3.50), we obtain from t his Wronski's general formula corresponding to (3.36)

At t his point it is useful to realize that t he funct ions in the numerator coincide with the two specific linearly independent solutions Da(t) and Db(t) of the homogenous differential equations (3.48) which were introduced in Eqs. (2.226) and (2.227). Comparing the initial conditions of Da(t ) and Db(t) with t hat of the function .6..(t, t') in Eq. (3.45) , we readily identify (3.60) and formula (3.59) can be rewritten as (3.61)

3. 2 Green FUnction of Harmonic Oscillator

217

It should be pointed out t hat this equation renders a unique and well-defined Green function if the differential equat ion [-0; - ffl(t) [y(t) = 0 has no solutions with Dirichlet boundary conditions y(t a) = y(tb) = 0, generally called zero-modes . A zero mode would cause problems since it would certainly be one of t he independent solutions of (3.49), say 1/(t). Due to the property 1/(ta) = 1/(tb) = 0, however, the determinant of A would vanish, thus destroying t he condition (3.58) which was necessary to find (3.59). Indeed, t he funct ion A(t,t') in (3.49) would remain undetermined since t he boundary condit ion 17(ta ) = 0 together with (3.55) impHes that also ~(ta) = 0, making W = ~(t)Jj(t) - {(t)17(t) vanish at t he initial time t a, and t hus for all times.

3 .2 .2

Spectral Representat ion

A second way of specifying the Green function explicitly is via its spectral representation. Cons tant Freque n cy For constant frequency O(t) == equat ion and vanish at the ends t = orthonormal functions:

W,

t he fl uctuations ox(t) which satisfy the differential

(- if - w') ;x(t) ~ 0, (3.62) ta and t = tb , are expanded into a complete set of

(3.63)

with t he frequencies [compare (2. 110)] (3.64)

These functions satisfy the orthonormality relations (3.65)

Since the operator -8i - w 2 is diagonal on x,,(t) , this is also true for the Green function CoAt, t') = (- 8i - w2) - 16(t - t'). Let C n be it s eigenvalues defined by (3.66)

T hen we expand Cwo(t, t') as follows:

..

00

Gw,(t, t') ~

L

"

G"x"(t)x"(t') .

(3.67)

218

3 External Sources, Correlations, and Perturbation Theory

By definition, the eigenvalues of Cwl(t , t') are the inverse eigenvalues of the differential operator (- 8; - w'l), which are v~ - w 2 • Thus (3.68)

and we arrive at the spectral representation of Cw,(t , t'l:

') ~ _ 2_ G.,2 (t, t tb -

~ sinvn(t - t,,)8inv,,(t' -

t"

w

2

,,= 1

2

t,,)

(3.69)

.

W

tin

We may use t he trigonometric relation sin V,,(tb - t) = - sin v,,[(t - t,,) - (lb - t,,)] = -(-1)" sin tI,,(t - t",) to reWTite (3.69) as

G",2 (t,l

') ~ _2_ ~(_ ),,+1 sin Vn(tb - t) sinvn(t'W 1 2 2 tb - ta "=l If" - W

tal

.

(3.70)

T hese expressions make sense only if tb - ta is not equal to an integer multiple of 1fjw, where one of the denominators in the sums vanishes. This is the same range of tb - ta as in t he Wronski expression (3.36).

Time-D e p e nde nt Freque ncy The spectral representation can also be written down for t he more general Green fu nction with a time-dependent frequency defined by t he differential equation (3.27). If y,,(t) are the eigenfunctions solving the d ifferential equation with eigenvalue ..\"

K(t)y.(t)

~

'.y.(t),

(3.71 )

and if these eigenfunctions satisfy the orthogonality and completeness relations to

l'.

dty"(t)y,,,(t) =

I: y.(t)y.(t')

(i"""

(3.72)

6(t - t') ,

(3.73)

and if, moreover, there exists no zero-mode for which ..\" = 0, then spectral representation

Go,(t, t') ~

I: y.(t~y.(t') . •



Gn~(t ,

t') has the (3.74)

This is easily verified by multiplication with K(t) using (3.71) and (3.73). It is instructive to prove t he equality between thc Wronskian construction and the spectral representations (3.36) alld (3.70). It will be useful to do this in several steps. In t he present context, some of these may appcar redundant. They will, however, yield intermediate results which will be needed in Chapters 7 and 18 when discussing path integrals occurring in quantum field theories.

219

3.3 Oreell Functions of First-Order Differential Equation

3.3

Green FUnc tions of First-Order Differentia l Equation

An important quantity of statistical mechanics arc the Green functions G:'J(t, t') which solve the first-order differential equation

[Wt

-

O(t) ]Go(t, t')

=

t - t' E [0, tb - t,,) ,

iJ(t - t') ,

(3.75)

or it s Euclidean version Gl;,,(r, r") which solves the differential equation, obtained ' from (3.75) for t = -iT: [0, - !l(T)[ Gn .• (T, T') ~ 8(T - T'),

r E [0, "fi).

T-

(3.76)

These can be calcula ted for an arbit rary funct ion O(t).

3.3.1

Time-Independent Frequency

Consider first t he simplest case of a Green function which solves the first-order differential equation (iiJ, - w)G~(t , n ~ ;o(t - f ),

~(t,

t') with fixed frequency W

t - f E [0, t. - to).

(3.77)

The equation determines ~(t, t') only up to a solution H (t , t') of t he homogeneous differential equation (Wt - w) H (t, t') = O. The ambiguity is removed by imposing t he periodic boundary condition G~(t , t')

" G~(t - t') ~ G~(t - f

+ t. -

to),

(3.78)

indicated by t he superscript p. With this boundary condition, the Green function G~(t,t') is translationally invariant in time. It depends only on the difference between t and t' and is periodic in it. T he spectral representation of ~(t, t') can immediately be written down, assuming t hat tb - ta does not coincide with an even multiple of 1r/w:

G~(t - t') = _ 1_

f

tb - ta m= - oo

e - ,:..,,,,(t- t, ) _ _i _ Wm - W

The frequencies Wm are twice as large as the previous Wm

2.m

== - -, t b - ta

rn =

11m'S

.

(3.79)

in (3.64):

0, ± 1, ± 2, ± 3, ....

(3.80)

As for the periodic orbits in Section 2.9, there arc "about as many" Wm as 11m, since there is an Wm for each positive and negative integer m , whereas the 11m arc all positive (see the last paragraph in that section). The frequencies (3 .80) are the real-time analogs of t he Matsubara frequencies (2.379) of quantum statistics wit h the usual correspondence tb - ta = - ih/k8T of Eq. (2.328).

220

3 External Sources, Correlations, and Perturbation Theory

To calculate the spectral sum, we use t he Poisson summation formula in the form (1.197) : (3.81)

Accordingly, we rewrite t he sum over wm as an integral over Wi, followed by an auxiliary sum over n which squeezes the variable Wi onto the proper discrete values Wm = 21rm/(t1> - tal : (3.82)

At this point it is useful to introduce another Green function Gw(t - t') associated with the first40rder differential equation (3. 77) on a n infinite time interval:

Gw(t)

f"'"

=

i_co

dw' e- iw't_i_. 271" w' - w

(3.83)

In terms of this function, t he periodic Green function (3.82) can be written as a sum which exhibits in a most obvious way t he periodicity under t -+ t + (tb - tal: 00

G~(t) ~

I:

Gw(t - (t. - to)n).

(3.84)

n : - oo

The advantage of using Gw(t - t') is that the integral over Wi in (3.83) can easily be done. We merely have to prescribe how to treat the singularity at w' = w. This also removes the freedom of adding a homogeneous solution H(t , t'). To make the integral unique, we replace w by w - iT} where T} is a very small positive number, i.e., by the iT}-prescription introduced after Eq. (2.166). This moves the pole in the integrand of (3.83) into the lower half of the complex wi-plane, making the integral over W i in G.,(t) fundamentally different for t < 0 and for t > o. For t < 0, the contour of integration can be closed in the complex wi-plane by a semicircle in t he upper half-plane at no extra cost, since e- iw't is exponentially small t here (see Fig. 3.1) . With t he integrand being analytic in the upper half-plane we can contract t he contour to zero and find that the integral vanishes. For t > 0, on the ot her hand, the contour is closed in the lower half-plane cont aining a pole at Wi = W - iT} . When contracting the contour to zero, the integral picks up the residue at this pole and yields a factor -27ri. At the point t = 0, fin ally, we can close the contour either way. The integral over the semicircles is now nOllzero, :r=1/ 2, which has to be subtracted from t he residues 0 and 1, respectively, yielding 1/ 2. Hence we find

-211" ,-wt w' (]W'

[.

- 00

=

. { I e- lWt x !

o

for for for

i W

+ iT} t > 0, t = 0, t < O.

(3.85)

3.3 Oreell Functions of First-Order Differential Equation

221

1m",'

F igure 3.1 Pole in Fourier transform of Green functions O~" (t), and infinite semicircles in the upper (lower) half-plane which extend the integrals to a closed contour for t < 0 (t > 0).

T he vanishing of thc Green function for t < 0 is the causality property of G..,(t) discussed in (1.306) and (1.307). It is a general property of functions whose Fouricr transforms are analytic in the upper half-plane. T he three cases in (3.85) can be collected into a single formula using the Heaviside function 6(t) of Eq. (1.309) :

Gw(t) ~ , - ""e(t).

(3.86)

The periodic Green function (3.84) can then be written as 00

L

~(t) =

e- iw ll - (lb - I~)nI9(t -

(tb - t,,)n).

(3.87)

n: - ""

Being periodic in tb - t", its explicit evaluation can be restricted to the basic interval

t E [O , tb - t"J

(3.88)

Inside the interval (0, tb - tOll , t he sum can be performed as follows:

0::,(') ~ e- iwlt - (tb - t~)/2l

- i 2sin [w( tb

ta )12l'

t E (0, tb - t,,).

(3.89)

At t he point t = 0, the initial term wit h 9(0) contributes only 1/2 SO that

G~(O) ~ G~(O+) - ~.

(3.90)

Outside the basic interval (3.88), the Green function is determined by its periodicity. For instance,

(3.91)

222

3 External Sources, Correlations, and Perturbation Theory

Note that as t crosses the upper end of the interval [0, tb - ta), the sum in (3.87) picks up an additional term (the term with n = 1). This causes a jump in C!:,(t) which enforces the periodicity. At t he upper point t = tb - t a , there is again a reduction by 1/2 so that C::,(tb - tal lies in t he middle of the jump, just as the value 1/2 lies in the middle of the jump of the Heaviside function 6(t). The periodic Green function is of great importance in the quantum statistics of Bose particles (see Chapter 7). After a continuation of the time to imaginary values, t ----> -iT, tt. - t" ----> -ifijkBT, it takes the form

Ct,e(T) = 1

e 1""'lkHT C - """,

T E (O,ht3),

(3.92)

where the subscript c records the Euclidean character of the time. The prefactor is related to the average boson occupation number of a particle state of energy nw, given by the Bose-Einstein distribution [unction nb = w

1

e",",,/knT

(3.93)



In terms of it, G~,.(T} ~ (l+n~Vw',

T E (O,hm .

(3.94)

The T-behavior of the subtracted periodic Green function Gt:',e( T) == G!,e(T) - l jht3w is shown in Fig. 3.2.

Figure 3.2 Subtracted periodic Grecn function Ce: e =:;; Ce,e(T)- l /h{3w and antiperiodic GreeD functioll G~,cCT) for frequencies w = (0, 5, 1O)/h{3 (with illcreasing dash length). The points show the values at the j umps of the three functions (with increasing point size) corresponding to the relation (3.90).

As a next step, we consider a Green funct ion order differential operator -vi - w2 ,

G~~(t)

associated with the second(3.95)

which satisfies the periodic boundary condition: (3.96)

223

3.3 Creell Functions of First-Order Differential Equation

1m",'

F igure 3.3 Two poles in Fourier t ransform of Green function

G~,a (t) .

J ust like Ct{t, t'l , this periodic Green function depends only t - t! . It obviously has the spectral representation

0 11

the time difference

(3.97) which makes sense as long as tb - ta is not equal to an even multiple of rr/w. At infinite tb - ta , the sum becomes a n integral over Wm with singularities at ±w which must be avoided by an i1]--prescr iption, which adds a negative imaginary part to the frequency w [compare the discussion after Eq. (2.166)]. This fixes also the continuation from small tb - ta beyond the multiple values of rr/w. By decomposing 1

1 (

w2 + if] = 2iw

,

-w'-c--w'--;+-,'C'"

-w~'~+-w'---'CC'1]) ,

(3.98)

t he calculation of the Green function (3.97) can be reduced to t he previous case. The posit ions of the two poles of (3.98) in the complex wi-plane are illustrated in Fig. 3.3. In this way we find, using (3.89) , 1 2wi [GC(t) - t > f > ta: (3 .122)

This equation is solved by a t'-independent C(t'): 1 C -- n IIP = - -",.-::::0:;:;;;--. r'b rk" 0(1") e )'4 - 1

(3 .123)

227

3.3 Oreell Functions of First-Order Differential Equation

Hence we obtain the periodic Green funct ion (3 .1 24)

For antiperiodic boundary conditions we obtain the same equation with n~ replaced by - n~ where • _ 1 ( )

no =

i f'bdHl(t )

e],6

3. 125

+1

Note that a sign change in t he time derivative of the firs t-order differential equation (3.75) to (3. 126) [-iiJ, - 0(')1 Gn ('" ,) ~ "(t - ") has the effect of interchanging in the time variable t and t' of the Green function Eq. (3.120). If the frequency O(t) is a matrix, all exponentials have to be replaced by timeordered exponentials [recall (1.252)[ (3. 127)

As remarked in Subsection 2.15.4, this integra.l cannot, in general, be calculated explicitly. A simple formula is obtained only if the matrix r2(t) varies only little around a fixed matrix 0 0 , For imaginary times 7 = it we generalize t he results (3.92) and (3. 110) for the periodic and antiperiodic imaginary-time Green functions of the first-order differential equation (3.76) to time-dependent periodic frequencies n(7) . Here the Green function (3. 120) becomes (3. 128)

and t he periodic Green function (3.124): (3. 129)

where

1

n b = ~~--''-----

-

(3. 130)

efo~fJ 01.,."0(.,." ) _ 1

is the generalization of the Bose distribution function in Eq. (3.93). For antiperiodic boundary conditions we obt ain the same equation, except that the generalized Bose distribution fu nction is replaced by the negative of t he generalized Fermi distribution function in Eq. (3. lIl ):

:,-,,---_ - --.,.,-,-,,1 r~" +1

nf =

e10

(3. 131)

d.,."!l(.,." )

For the opposite sign of the time derivative in (3. 128), the arguments interchanged.

7

and r' are

228

3 External Sources, Correlations, and Perturbation Theory

From the Green functions (3. 124) or (3.128) we may find directly the t race of the logarithm of the operators [- iOt + fl(t) ] or [8.- + 0(7)]. At imaginary time, we multiply 0(7") wit h a strength parameter g, and use the formula Tr log [a.. + 90(7)] =

fo9 dg' Gg'o(r, r) .

(3 .132)

Inserting on the right-hand side of Eq. (3.129), we find for 9 = 1: Trlog [8.. + fl(r) ]

=

!Og{2Sinh

[tf'P

dT'O(r")]}

~ fo"fJ dr" 0(7"") + log [1 -

e- 1:.8 d,..,'O(T" )], (3. 133)

which reduces at low temperature to

~ "2 j I" ,

dT" !l(T").

(3. 134)

The result is the same for the opposite sign of the time derivative and the t race of the logarithm is sensitive only to 6(r - r) at T = 7', where it is equal to 1/2. As a n exercise for dealing with distributions it is instructive to rederive t his result in the following perturbative way. For a positive 0(7) , we introduce an infinitesimal positive quantity 1J and decompose

Tdog [±a, + !l(T)] ~ '[i-iog [±a, + "[ + '[i-iog = Tr log [±8T

[1+ (±&, + ")- ' !l(T)[

(3.135)

+ 1J] + Tr log [1+ (±8 + 11)-10(7)] . T

The first term Tr log [±8 + 111 = Tr log [±8 + 111 = f~OQ dw log w vanishes since f~OQ dw logw = 0 in dimensional regularization by Veltman 's rule [see (2.506)]. Using the Green functions T

T

(3.136) the second t erm can be expanded in a Taylor series

For the lower sign of ±8T , the Heaviside functions have reversed arguments 72 72, ... , 71 - 7 n . The integrals over a cyclic product of Heaviside funct ions in (3 .1 37) are zero since the a rguments 71> ... ,Tn are time-ordered which makes t he argument of the last factor 9(7n - 71) [or 9(71 - 7n ) ] negative and t hus 9(7n - 71) = 0 [or 9 (7] - 7 n ) ]. Only t he first term survives yielding 71, 73 -

(3.138)

229

3.4 Summing Spectral Representation of Oreell Function

such that we re-obtain the result (3.134). This expansion (3. 133) can easily be generalized to an arbitrary matrix O(r) or a time-dependent operator, H(r). Since H(r) a nd H(r') do not necessarily commute, the generalization is Tr log[MT + H(r)] =

~ Tr

2h

[

r Jo

hfJ

dr H(r) ]-

f: .!.Tr [fe-nf: n

iJ

dr' li(r' )!"] , (3. 139)

n: 1

where f is the time ordering operator (1.241). Each term in the sum contains a power of the t ime evolution operator (1.255).

3.4

Summing Spectral Representation of Green Function

After these preparations we are ready to perform the spectral sum (3.70) for the Green function of the differential equation of second order with Dirichlet boundary conditions. Setting t2 =: tb - t, tl =: t' - t" , we rewrite (3.70) as

(3. 140)

We now separate even and odd frequencies lin and write these as bosonic and ferm ionic Matsubara frequencies Wm = 112m and wfn = I12m+! , respectively, recalling the definitions (3.80) and (3.104) . In this way we obtain

Cw,(t, t') = (3. 141)

Inserting on the right-hand side the periodic and antiperiodic Green functions (3.99) and (3. 108), we obtain the decomposition

aw,(t, t')

=

~ [a!:,(tz + td -

a:'(tz + t l )

-

a!:,(t2 - td

+ a:'(t2 - tdl. (3. 142)

Using (3.99) and (3. 11 3) we find that

'( ) _ G'( _ ) _ sin w[tz - (tb - t,,)/21sinwt l Gw t2 + tl w tz t] wsin[w(tb t,,)/2] ,

)-G'( _ G'( w t2 + t ] w tz

) _ _ cosw[t, - (tb - t.)/2[,in wt, wcOS[W(tb t,,)/2] ,

t\ -

(3. 143) (3. 144)

2 30

3 External Sources, Correlations, and Perturbation Theory

such t hat (3. 142) becomes GW~(t , t') =

.

I

(

WSlllW tb -

tIl

)

Sinwt2sinwt\ ,

(3. 145)

in agreement with the earlier result (3.36) . An important limiting case is

t"

---> -00 ,

(3. 146)

tb ----> 00.

T hen t he boundary conditions become irrelevant and the Green function reduces to G i(I,) 'f .i(t.)1 ~ - a,LI.(t" I) .

(3.162)

232

3 External Sources, Correlations, and Perturbation Theory

Defining now the constant 2 x 2 -matrices

iV ,O(t t) ~ ( {(to) 'f {(t.) .. 0 «to) 'f «t.)

(3. 163)

the condition analogous to (3.58), det AP,J,(t" , tb)

=

W dP"(t", tb) =f 0,

(3. 164)

with (3. 165)

enables us to obtain the unique solution to Eqs. (3. 162). After some algebra using the identities (3.51) and (3.52), the expression (3.159) for Green functions with periodic and antiperiodic boundary conditions can be cast into the form

') _ G 2 (

G" O(

n~ t, t -

0

')

t, t =F

[" (t, t.) ± ,,(to, t))[" (O.) ± ,,(to, t')[ .6,p,&(t t )"(t t) , ii ,

b

'"

(3. 166)

b

where C 02(t, t') is the Green function (3.59) with Dirichlet boundary conditions. As in (3.59) we may replace the fUllctions on the right-hand side by the solutions DB(t) and Db(t) defined in Eqs. (2.226) and (2.227) wit h t he help of (3.60). The right-hand side of (3.166) is well-defined unless the operator K(t) = - o'f (f.!(t) has a zero-mode, say 1](t) , with periodic or antiperiodic boundary conditions 7}(tb) = ±1](t,,), 1)(tb) = ±1)(t,,), which would make the determinant of the 2 x 2 -matrix ii,P,& vanish.

3.6

Time Evolution Amplitude in Prese nce of Source Term

Given thc Green function Cw,(t ,t'), we can write down an explicit expression for the timc evolution amplitude. Thc quadratic source contribut ion to the fluctuation factor (3 .21) is given explicitly by A j ,'

~

-

2

M i l" dtl" dfG,.(t,t')j(t)j(t') f~

t~

1 1 M wsin w(tb

to

)

(3. 167)

r·dtl1 dt' sin w(tb _ t)sinw(t' -t,,)j(t)j(t').

itA

t"

Altogether, the path integral in the presence of an external source j(t) reads

with a total classical action 1 ' ~ -2

Mw ' , ( )[ (Xb+Xa)COSW(tb-ta)-2xbXa tb to

Sill W

+

I

t

I ) • dt [xa sin W(tb _ t) sin w(tb t" j~

1

+ xbsinw(t - ta)b(t),

(3. 169)

233

3.6 Time Evolution Amplitude in Presence of Source Term

and the fluctuation factor composed of (2. 169) and a contribution from t he current term eiAj,R/":

Fw,jb,a-_ (t t) _ F. .(t" t.)eiAj,R/" _ xexp

{

TiM

1 J27ritt/ M

=

(W sin w tb

ta

)

}

'.'

. t. ( ) [ dt [ dt'sinw(tb - t)sinw(t' - ta)j(t)j(t') . (3. 170) wsm w tb - ta it~ it~

T his expression is easily generalized to arbitrary time-dependent frequencies . Using the two independent solutions Da(t) and Db(t) of the homogenous differential equations (3.48) , which were introduced in Eqs. (2.226) and (2.227) , we fi nd for the action (3.169) the general expression, composed of the harmonic action (2.266) and the current term fl~ dtxcl (t)j(t) with t he classical solution (2.246):

The fluctuation factor is composed of the expression (2.261) for the current-free action , and t he generalization of (3. 167) with the Green function (3.61): _

(

)iAj,R/" _

Fw,j(tb,ta) - Fw tb ,ta e X

-

r" dt iI't" dt' j(t) [Sit -

it~

1

1 {_ i ~ exp fiMD ( )

J21tih/M V Do.(tb)

t')D,(t) Do(t')

+ S(t! -

2~

a

tb

t)Do(t)D,(t')l j(t')}. (3.172)

For applications to statistical meclIanics which becomes possible after an analytic continuation to imaginary times, it is useful to write (3.169) and (3.170) in another form. We introduce the Fourier t ransforms of the current

A(w)

(3.173)

B(w)

(3 .1 74)

and see that t he classical source term in the exponent of (3.168) can be written as A-cl = -i . Mw }, smw(tb

tal

{ [xb(e.:w(t.-t~)A - B)l + Xa(e.:w(t·-t~)B - A)l.

(3. 175)

T he source contribution to the quadratic fluctuations in Eq. (3. 167) , on the other hand, can be rearranged to yield

234

3 External Sources, Correlations, and Perturbation Theory

This is seen as follows: We write the Creen function between j(t),j(t') in (3.168) as - [8inw(lb - t)sin w(t' - t,,)9(t - t') =

~

[ ( eiw (lb- t4 )e- iw (t-t')

+ c.c')

+ sinw(tb -

t')sin w(t - t,,)e(t' - t)l

- (e"",(t b+t4)e-iw(Hf')

+ c.c.) ]8(t -

t')

+{t_t'). Using

eCt -

t')

~ {_

+ e(t' - t)

=

1, this becomes

(eu...(lbHA)e- iw(t'H)

+eiw(t.-tAl

(3. 177)

+ c.c')

(e- iw(t-t'18(t _ t') + e-

+e- iw(to- I"l [eiw(/ - I')(l

-

SCt' -

iw {l'- t1



6(t' -

t»)

+ eu..(t'- t){1_ e(t -

(3. 178)

t'» ]}.

A multi plication by j(t),j(t') and an integration over the times t, I! yield

H-

eiw(t.-t"14M 2w2(B2

+ (eiw(t. -t.. 1 _

e-iw(t.-t A

+ A2)

»)

(3. 179)

r· dt I dt'e-iw1t-t'lj(t)j(t') + 4M w 2AB], itA I. t

2 2



thus leading to (3.176). If the source jet) is time-independent, the integrals in the current terms of the exponential of (3.169) and (3.170) can be done, yielding the j-dependent exponent

(3. 180)

Substituting (I - coso:) by sin a tan(a/2), this yields the total source action becomes

This result could aL'lO have been obtained more directly by taking the potential plus a constant-current term in the action

- 1~· dt (~ W2X2 -

Xj) ,

(3. 182)

and by completing it quadra.tically to t hc form (3. 183)

235

3.6 Time Evolution Amplitude in Presence of Source Term

This is a harmonic potential shifted in x by - jj Mw 2 . The time evolution amplitude can thus immediately be written down as

(3. 184)

i) }+ ~it.-t".,) i )(Xa - Mw2 2Mw 2 J .

(

-2 Xl> - Mw 2 In the free-particle limit w

->

0, the result becomes particularly simple:

As a cross check, we verify that the total exponent is equal to ijli t imes the classical action (3. 186)

calculated for the classical orbit Xj,d(t) connecting x" and XI> in the presence of t he constant current j. This satisfies the Euler-Lagrange equation Xj,d

=

jjM,

(3. 187)

which is solved by

Xj,cl () t

=

Xa

i ( ib - ta ),]ttb-_ t"ta + 2M i ( t - ta )' . + [Xb - Xa - 2M

(3. 188)

Inserting this into the action yields (3. 189)

just as in the exponent of (3 .185) . Let us remark that the calculation of the oscillator amplitude (xatblxa t )L in (3 .168) could have proceeded alternatively by using t he orbital separation

x(t) ~ x",,(t) + 'x(t) ,

(3. 190)

where Xj,cl(t) satisfies the Euler-Lagrange equations wit h t he time-dependent source term (3. 191)

236

3 External Sources, Correlations, and Perturbation Theory

rather t han the orbital separation of Eq. (3.7),

x(t)

~

x,,(t)

+ 6x('),

where Xd(t) satisfied the Euler-Lagrange equation with no source. For this inhomogeneous differential equa.tion we would have found the following solution passing through x" at t = ta and Xb at t = tb:

sin w(tb - t) xi.,,(t) ~ x. . (t t) smw b "

sinw(t - ta)

I1t~

+ x,.SlllW (t b t) +M a

la

I

I

.

dt Go,(t,' )J(t').

(3. 192)

T he Green fundion Gw 2(t, t') appears now at the classical level. T he separation (3.190) in the total action would have had the advantage over (3.7) that the source causes no linear term in !Sx(t). Thus, there would be no need for a quadratic completion; the classical action would be found from a pure surface term plus one half of the source part of the action

Inserting Xj,cl from (3 .1 92) and Cw,(t, t') from (3.36) leads once more to the exponent in (3.168). T he fluctuating action quadratic in ox(t) would have given the same fluctuation factor as in t he j = 0 -case, i.e., the prefactor in (3.168) with no further F (due to the absence of a quadratic completion).

3.7

Time Evolution Amplitude at Fixed P ath Average

Another interesting quantity to be needed in Cha pter 15 is the Fourier transform of the amplitude (3. 184):

(Xbtblxata)~

=

(tb - ta)

L: 2;'"

e-;j(t6-t~)"'oJ"(Xbtblxata)~.

(3. 194)

T his is the amplitude for a particle to run from Xa to Xb along restricted paths whose temporal average x == (tb - ta)-l Jt~ dt x(t) is held fixed at Xu :

(Xbtblxata):o =

JV xo(xo - x)

cxp

{ ~1:· dt ~ (±2 -

2

(3. 195)

w x2) } .

This property of t he paths follows directly from the fact t hat the integral over the time-independent source j (3.194) produces a o-function 8((tb - ta )xo dtx(t)) . Restrict ed amplitudes of this type will turn out to have important applications later in Subsection 3.25.1 and in Chapters 5, 10, and 15.

J::

237

3.8 External Source in Quantum-Statistical Path Integral

The integral over j in (3. 194) is done after a quadratic completion in A j - j(t/) ta)xo with A j of (3.181) :

A j - J.( t/) - ta )Xo

=

1 [W( tb - ta ) - 2 tan w(t, 2- t.)] (J. - Jo . )' + A:to, 2Mw 3

(3. 196)

with

.

Jo =

Mw' [ w(t, - t.) ] ~ W(tb - ta)xo - tan (x/) + x a) , w(tb-ta)- 2 tan W Z ~ 2

and

With t he completed quadratic exponent (3.196), the Gaussian integral over j in (3.194 ) can immediately be done, yielding "'0

(Xbtblx"ta)..,

=

(Xbtb lxata)

(i

iMw 3 /27rh % 0) . ~ exp;;-A w(tb-t,,)-2tan '" z ~ n

(3. 197)

If we set Xb = x" and integrate over Xb = Xa, we find the quantum-mechanical version of the partition function at fixed Xo: • Z,} =

) ,,] ( ) J27rh(tb 1- tall Mi .w(t,-t.)j2 [W(tb ta )j2[ exp [i( - 2"' tb - ta Mw xo' 3.198 I~ Sill

As a check we integrate this over Xo and recover the correct Zw of Eq. (2.410). We may also integrate over both ends independently to obtain the partition function

-

~n,"' and T< in the first line (3.206) denote the larger and the smaller of t he Euclidean times T and r, respectively.

3.8 External Source in Quantum-Statistical Path Integral

239

The source terms (3.204) and (3.205) can be rewritten as follows: (3.209) and

We have introduced the Euclidcan vcrsions of the functions A(w) and B(w) in Eqs. (3. 173) and (3. 174) as (3.211)

(3.212) From (3.201) we now calculate thc quantum-statistical partition function . Set;. ting Xb = Xu = x, t he firs t term in the action (3.202) becomes . ,(wfl/3 /2)x. ' A.,= . Mw hal< 2smb sm f.' llW

(3.213)

If we ignore t he second and t hird action terms in (3.202) and integrate (3.201) over x , we obtain, of course, the free partition function 1

Zw

~ 2,inh(p nw/2r

(3.21 4)

In the presence of j, we perform a quadratic completion in x and obtain a sourcedependent part of the action (3.202) : Aj =

~

where t he additional term reads

Ai

-"fI,e

+

Aj

(3.215)

"""",e'

At,c is the remainder left by a quadratic completion.

j Mw R1u.>( A re, = - 2 sm . h W f.'ae'" Ae

+ Be )' .

It

(3.216)

Combining this with A~ ,e of (3.210) gives i

A:J,c

j

_

+ ~,e -

fj 1 fh P fh I -,1 "-.-'1 . . r' Mw 8 1u.>/2 - 4Mw Jo dT Jo dT e )(T);() - sinh(/31iw/2) e'" AeBc·

(3.217)

240

3 External Sources, Correlations, and Perturbation Theory

This can be rearranged to the total source term

~ = __

1_

r"fJ dT rhP dr'cosh w( lr - 7'1 Jo

4Mw Jo

hfJ/2) j(r)j(T') .

sinh{/HIW/2)

(3.218)

This is proved by rewriting the latter integrand as 1

2sinh(phw/2)

{ [ewC"-- "")e - fJfow/2

+ (w ---> - w)] a(T _ T')

+ [e{r'- 1")e- Pllw/2 + (w -+ - w)j 8(7' -

T) } j(T)j(T').

In the second and fourth terms we replace e {Jhw/ 2 by e - (J1u.J/2 + 2sinh(j31lwj2) and integrate over T, r, with the m,ult (3.217). The expression between the currents in (3.218) is recognized as the Euclidean version of t he periodic Green function G~~ (T) in (3.99): G~2.e(T)

-

iG~2( - iT) l tb _j~= _;h.8

1 cosh w(r - h/3/2) 2w ,inh(phw/2)

T

E

10,lipl .

(3.219)

In terms of (3.218) , the partition function of an oscillator in the presence of the source term is

(3.220) For completeness, let us also calculate the partition funct ion of all paths with open ends in the presence of the 50urce j(t), t hus generalizing the result (2.4 11 ). Integrating (3.201) over initial and final positions x" and Xb we obtain

(3.221) whcre

(3.222) with 1 . 3 {3 {cosh wh{3[sinh w(1i{3- r) sinh w(1i{3- r') + sinh wr sinh wr'l 2wsmh wh +sinhw(1i{3-r) sinh wr ' + sinh w(fl{3- r') sinhwr} . (3.223)

By some trigonometric idcntities, this can be simplified to (3.224)

241

3.8 External Source in Quantum-Statistical Path Integral

The first step is to rewrite the curly brackets in (3.223) as sinh wT [ cosh wli(3 sinh WT ' + sinh w(Ii(3 - r')]

+ sinh w(f!(3 -T') [ coshwf~(3 sinh W(Ii(3 -T) + sinh w(Ii(3- «f~(3 - T»].

(3.225)

The first bracket is equal to sinh (3/kv cosh WT, t he second to sinh (3flW cosh w(li(3 -T'), so that we arrive at sinhwli(3[ sinh wT coshwr'

+ sinh w(Ii(3- T) oo;hw(fi.(3 - T') ].

(3.226)

The bracket is now rewritten as

~ [sinh w{T + T') + sinh w(T -

T')

+ sinh w(21i(3 - T - T') + sinh w(r' - T) ], (3 .227)

which is equal to

~ [ sinh w(li(3 + T + r' -

1i(3)

+ sinh w(li(3 + 1i(3 - T - T') ],

(3.228)

and thus to

~ [2sinh wli(3cosh w(li(3 - T - T')],

(3 .229)

such t hat we arrive indeed at (3.224). The source action in the exponent in (3.221) is t herefore:

(A~. e+ . .A~ e) =

- Ml fh{J dT dT' j(T)c:~.n(T, T')j(r'), 10)0 .

r

(3.230)

with (3. 205)

Cf''::: (T, T')

~

coshw(li(3 - IT - T'I) + cosh w{fl(3 - T - T') 2w sinh wli(3 coshw(li(3 - T» cosh wT< w sinh wfi.(3

(3.231)

This Green function coincides precisely with t he Euclidean version of Green function G~,(t, t') in Eq. (3.151 ) using t hc relation (3 .208) . T his coincidence should have been expected after having seen in Section 2.12 that t he partition function of all paths with open ends can be calculated, up to a trivial fact or lo(Ii(3 ) of Eq. (2 .351 ), as a sum over al1 paths satisfying Neumann boundary conditions (2.449), which is calculated using the measure (2.452) for t he Fourier components. In the limit of small-w, the Green function (3.231 ) reduces to _"(TT ') ::::: - 1 + P - -' 1T - T'I - -21 (T + T') w .e ' w'",o (3w2 3 2

L>' . ~,

which is the imaginary-time version of (3.157).

1 (T' +T") + -2(3

,

(3 .232)

2 42

3 .8 .2

3 External Sources, Correlations, and Perturbation Theory

Calculation a t Imaginar y Time

Let us now see how the partition function with a source term is calculated directly in the imaginary-time formulation, where the periodic boundary condition is used from t he outset. Thus we consider

Z",[j j =

JV x(r)e- A.lill",

(3.233)

with the Euclidean action (3.234)

Since x(r) satisfies the periodic boundary condition, we can perform a partial integration of the kinetic term without picking up a boundary term xii:: . T he action becomes (3.235)

Let De(r, r') be the functional matrix D"" ,~Jr, r')

== (- eJ: + w2 )o(r - r) , r - r' E [0, hpj .

(3.236)

Its functional inverse is t he Euclidean Green function,

with the periodic boundary condition. Next we perform a quadratic completion by shifting the path: X""" X

,

1

=x - M

cw2P ,cJ·.

(3.238)

This brings the Euclidean action to the form

Aelil =

(M dr M x'( 2

Jo

-eJ: + w2 )x' _ _2MJo 1_ ( "fj dr (M dr j(r)G (r _ r)j(r'). Jo w ,e P,

(3.239)

The fluctuations over the periodic paths x'(r) can now be integrated out and yield for j(r) == 0 Zw = Oct D:,i,~2 . (3.240) As in Subsection 2.1 5.2, we find the functional determinant by rewriting t he product of eigenvalues as Oct Dw2,c =

mD (w; + w2) = exp [m~.., log(w; + W2)] ,

(3.24 1)

3.8 External Source in Quantum-Statistical Path Integral

243

and evaluating the sum in the exponent according to t he rules of analytic regularization. This leads directly to the partition function of the harmonic oscillator as in Eq. (2.40n Z = 1 (3.242) w 2 einh(pfow/2) The generating functional for j(r)

f- 0 is therefore (3.243)

with the source term: (3.244)

T he Green function of imaginary time is calculated as follows. The eigenfunctions of the differential operator - 8; are e - iw",r with eigenvalues w~, and the periodic boundary condition forces Wm to be equal to the thermal Matsubara frequencies Wm = 27rm/h/3 with m = 0, ±1, ±2, . Hence we have the Fourier expansion (3.245)

In the zero-temperature limit, the Matsubara sum becomes an integral, yielding (3.246)

The frequency sum in (3.245) may be written as such an integral over W m , provided the integrand contains an additional Poisson sum (3.81): 00

00

L

o(m - m)

=

m=-oo

00

L

e,2"."m

=

"=-00

L

e'nwmhP.

(3.247)

"=-00

T his implies that the finite-temperature Green function (3.245) is obtained from (3.246) by a periodic repetition:

1 coshw(r - h/3/2) einh (pliw/2)

2w

, E [0, hpj.

(3.248)

A comparison with (3.97), (3.99) shows t hat G~2 ,e(r) coincides with G~2(t) at imaginary times, as it should. Note that for small w, the Green function has the expansion 1

p

Gw1,~(r)

=

li/3w2

r2

+ 21i/3

rh/3

-

'2 + 12 + ..

(3.249)

244

3 External Sources, Correlations, and Perturbation Theory

The first term diverges in the limit w --+ O. Comparison with the spectral representation (3.245) shows t hat it stems from the zero Matsubara frequency contribution to the sum. If this term is omitted, the subtracted Green function

G~~,e(T) has a well-defined w

-+

G~2,e(T) -

-

(3.250)

h;w2

0 limit

G"O,e (T ) =

1

'"'

2

1

h/3 _ ~

m_± 1,±2,..

-iwm'r _

w2 e

m

-

7

2ft{3

T

Ti{3

-"2 + 12'

(3.251)

the right-hand side being correct only for T E [0, h.8J. Outside t his interval it must be continued periodically. The subtracted Green function G~~,, (T) is plotted for different frequencies w in Fig. 3.4. '

Figure 3.4 Subtracted periodic Grccn function GP~ (7);::;: GP, (T) _ 1/1ij3w 2 and an'" ,e '" ,e tipcriodic Green fun ction G~2.,,(T) for freque ncies w = (0, 5, lO )/h{3 (with increasing dash

length). Compare Fig. 3.2. The limiting expression (3.251 ) can, incidentally, be derived using the methods developed in Subsection 2.15.6. We rewrite the sum as 1

hfi _L

(3.252)

m_±l,±2 ....

and expand (_1)m-' - -2 (nfi)' -'- L -1l [ -i-2rr (r-r.fij2) ]" L '--',b 2 00

lij3

2Jr

,,=0.2.4, .. n

lij3

m=l

m

(3.253)

n

The sum over m on the right-hand side is Riemann's eta function l

=L 00

"(z)

m=l

(

l)m-'

L-~_

m'

1M. Abramowitz and I. StCglm, op. cit., Formula. 23.2.19.

(3.254)

245

3.8 External Source in Quantum-Statistical Path Integral

which is related to the I'.eta function (2.519) by

"I') ~ (1 - 2' - (T"T)] (3280) T,7 T I\p,a ( ) , • '--' 7a,7b) u 'Ta(, 7b

6Sce H. Kleinert and A. Chervyakov, P hys. Lett. A 2~5, 345 (1998) (quant-ph/9803016); ~O , 6044 (1999) (physics/9712048).

J. Math. P hys. B

248

3 External Sources, Correlations, and Perturbation Theory

where C",~,e(T, 7') is the imaginary-time Green function with Dirichlet boundary conditions corresponding to (3.206): G

r') ~ 8(T-T')6.(T"T)6.(T', To) + 8(T-r')6.(T"r')6.(T,To)

( w',e 7,

"(

)

u Ta, Tb

,

(3.281 )

with (3.282)

and (3.283)

Let also write down the imaginary-time versions of the periodic or antiperiodic Green functions for time-dependent frequencies. Reca ll the expressions for constant frequency ~(t) and G!(t) of Eqs. (3.94) and (3.112) for 7 E (O,Iij3): G~,,,(T)

~

-1

1i{J

L ,-""",. m

- 1

iwm

-

= e- w(" - "fJ/2) W

1 2 einh(fiIiw12)

(l +n:)e- "'''' ,

(3.284)

and

C::',,,(T)

~

-1 h{J

~

L ,- it.l.,. . m

- 1

iw!,.

=

e- w(7 - r./Jf2j

1

2 cooh (filiw 12)

w

( l -n:)e- W "

(3.285)

the first sum ext ending over the even Matsubara frequencies, the second over t he odd ones. The Bose and Fermi distribution functions were defined in Eqs. (3.93) and (3. 111 ). For 7 < 0, periodicity or antiperiodicity determine

nY

G""(T) w,e

~

±G""(T + hfi). w,e

(3.286)

The generalil.ation of these expressions to t ime-dependent periodic and antiperiodic frequencies 11(7) satisfying t he differential equations

[- 8.,. has for

11(7)l G~: (7, 7') =

6P '''(7 - 7')

(3 .287)

/3 --- 00 the form G~,: (7, 7') = 9(7 _ r)e- fo' d""I1(,.'j .

Its periodic superposition yields for finite 00 e- Jo C'" ' ' d,.'O( ...'j P '" ( GO,e ') = '("" T,7 L ,. ",0

/3 a sum analogous

{I} (- I)"

,

(3.288)

to (3.277):

h/3>7>r>O,

which reduces to (3.284), (3.285) for a constant frequency 11(7) =: w.

(3.289)

249

3.9 Lattice Green FUn ction

3.9

Lattice Gree n FUnction

As in Chapter 2, it is easy to calculate thc above results also on a sliced time a.xis. This is useful when it comes to comparing ana.lytic results with Monte Carlo lattice simulations. We consider here only the Euclidean versions; the quantum-mechanica.l ones can be obtained by ana.lytic conti nuation to real times. The Green function Gw,(T, T') on an imaginary_time lattice with infinitely many lattice points of spacing ~ reads [instead of the Euclidean version of (3. 147)]: 1 1 _.;..1 ... _ ... '1 G ' (T,T') = 2 . fh - e - ';'. 1.. - .,.'1 = 2w h( - 12)' , sm €We coo €We where We is given,

118

in (2.404), hy

_ 2 1 'w "'e = -, arsin )-2 .

(3.290)

(3.291)

T his is derived from the spectral representation (3.292) by rewriting it

118

(3.293) with n ;;; (T' - T)/(, performing the ",'-integral which produces a Bessel function I("' _"") I« 28/(~) , and subsequently t he integra.l over s with the help of formula (2.473). The Green function (3.290) is defined only at discrete 7"" = nhp/(N + 1). If it is summed over a.ll periodic repetitions n ...... n + k(N + 1) with k = 0, ±1, ±2, ... , one obtains the lattice analog of the periodic Green function (3.248):

-

,~ """_00 I: 2(1 1

1

2t..> COSh(fW/2)

3.10

cosbw(T - hP/2) sinh(,",,",p/2)

TE [O,htJ] .

(3.294)

Correlation FUnctions, Generating FUnctional, and Wick Expa nsion

Equipped with the path integra.l of the harmonic oscillator in the presence of an external source it is easy to calculate the correlation functions of any number of position variables X(T). We consider here only a system in thermal equilibrium and study the behavior a t imaginary times. The real-time correlation functions can be discussed similarly. The precise relation between them will be worked out in Chapter 18. In general, i.e. , also for nonharmonic actions, the t herma l correlation functions of flo-varia bles X(T} a re defined as thc functional avcrages

G~~)(Tl, ... , Tn) _ _

(x{Tdxh)···X{Tn)) Z - lJVXX(Tdx(T2) ... X(Tn)exp( - ~A,).

(3.295)

250

3 External Sources, Correlations, and Perturbation Theory

They are also referred to as n-point functions. In operator quantum mechanics, the same quantities are obtained fro m the t hermal expectation values of t ime-ordered products of Heisenberg position operators iH(T):

G~~)(T" ... , Tn)

=

Z - ITr

{t" [XH(TdxH(T2)"

. XIl(Tn)e- il/knT]} ,

(3.296)

where Z is the partition function Z = e- F/k"T = Tr(e- lilk8T)

(3 .297)

and tT is t he time-ordering operator. Indeed, by slicing the imaginary-time evolution operator cJ'iTI" at discrete times in such a way t hat the t imes T; of the n position operators x( Ti) arc among them, we find t hat G~".{ (7\, .. . ,Tn) has precisely the path integral representation (3.295) . By definition, the path integral with the product of X(Ti} in the integrand is calculated as follows. First we sort the times 1'; according to their time order, denoting the reordered times by Tt(;) . We also set Tb == 1'1(..+1 ) and Ta == Tt(O) ' Assuming that the times Tt(;) are different from one another, we slice the time axis l' E [Ta , Tb ) into the intervals [7b , Tt(fI» ) , [Tt(fI),Tt(fI _ l»), !rt( .. - 2),Tt( ..- 3»)," " [TI(4),Tt(3» ) , [TI(2) , Tt(I) ], [Tt(l), Ta ]. For each of these intervals we calculate the time evolution amplitude (Xt(HI)Tt(i+1) IXt(i)Tt(i») as usual. Finally, we recombine the amplitudes by performing the intermediate x(Tt(;j)-integrations, with a n extra factor X(T;) at each Ti,

i.e.,

'(Xt(Hl)Tt(H l)IXt(i)Tt(i»)' X(Tt(i »)' (Xt(i)Tt(i) IXt(i_l)Tt(i_l»)'

X(Tt(i _ l))

(3. 298 )

•.... (Xt(2)Tt(2)IXt(I)Tt(I»)' X(Tt( I »)' (Xt(I)Tt(1) IXt(O)Ta).

We have set Xt (n+1) == Xb = Xa == xt(O), in accordance with the periodic boundary condition. If two or more of the times Ti are equal, the intermediate intcgrals arc accompanioo by t he corresponding power of X(Ti). Fortunately, t his rather complicated-looking expression can be rcplaced by a much simpler one involving functional derivatives of the t hermal part ition function Zfj ] in the presence of an external current j . From t he definition of Zfj ] in (3.233) it is easy to see that all correlation functions of the system are obtained by the functional formula

G~':'(,,, ...,'.) ~

[zur't. ;j~,.) ... t, ;jf,./U1Lo .

(3.299)

T his is why Zfj ] is called the gene1llting functional of the theory. In the present case of a harmonic action , Zfj ] has t he simple form (3.243), (3.244), and we can write

G~~)(Ti> ... ,T.. )

=

[t.-;j(,,) ;_... /l. -;j(,.) ;-

x

exp

{2;M 10

M dT

lo"tJ drj(T)G~1,e(T -

(3.300) r)j(T') }

L~

3.10 Correlation Functions, Generating FuIICtiOlllll, wid Wick Expansion

251

where G~2,e(T - T') is the Euclidean Green function (3.248) . Expanding the exponential into a Taylor series, the differentiations are easy to perform. Obviously, any odd number of derivatives vanishes. Differentia ting (3.243) twice yields the two-point function [recall (3.248)[

G~Z;(T,T') =

(x(T)x(r» =

~G~2,e(T -

T') .

(3.301)

Thus, up to the constant prefactor, t he two-point function coincides with t he Euclidean Green function (3.248). Inserting (3.301 ) into (3.300), all n-point functions are expressed in terms of t he two-point function G~/(T, r): Expanding the exponential into a power series, the expansion term of order nj2 earries the numeric prefactors Ij(nj2)! . Ij2 n/ Z and consists of a product of nj2 factors fohP dT'j(T)G~;(T, r)j(T')j1i? The n-point function is obtained by functionally differentiating this term n times. The result is a sum over products of nj2 factors G~Z;(T, T') with n ! permutations of t he n time arguments. Most of these products coincide, for symmetry reasons. First, G~2;(T,T') is symmetric in its arguments. Hence 2n / Z of t he permutations correspond to identical terms, their number canceling one of t he prefactors. Second, t he nj2 Green functions G~;(T, r) in t he produet are identical. Of the n ! permutations, subsets of (nj2)! permuta.tions produce identical terms, their number canceling the other prefactor. Only n!j[(nj2)!2n/Z] = (n - 1)· (n - 3)···1 = (n - I )! ! terms are different. T hey all carry a unit prefactor and their sum is given by the so-called Wick rule or Wick expanswn:

G~~)(Ti> . .. , Tn) =

L

G~/(Tp(l) ' Tp(2»··· G~2;(Tp(n_l}' Tp(n}) .

(3.302)

"';'

Each term is characterized by a different pair configurations of the time arguments in the Green funct ions. These pair configura tions are found most simply by t he following rule: Write down all time arguments in the n-point function T1TZT3T4 ... Tn . Indicate a pair by a common symbol, say 7"p(;)7"P(Hl)' and call it a paiT contraction to symbolize a Green function G~;(Tp(;), TP(H l » . T he desired (n - 1)!! pair configurations in t he Wick expansion (3.302) are t hen found iteratively by forming n - 1 singlc contractions (3.303)

and by t reating t he remaining n - 2 uncontracted variables in each of these terms likewise, using a different contraction symbol. The procedure is continued until all variables are contracted. In the literature, one sometimes another shorter formula. under the name of Wick's rule, stating that a single harmonically fluctua ting variable satisfies t he equality of expectations: (3.304)

252

3 External Sources, Correlations, and Perturbation Theory

This follows from the observation that the generating functional (3.233) may also be viewed as Z", times the expectation value of t he source exponential (3.305)

Thus we can express the result (3.243) also as (3.306)

Since

(h/M)G~~,e(T,r')

G~z;(T, r')

=

in t he exponent is equal to the correlation function

(x(r)x( r'» by Eq. (3.301 ), we may also write (ef dTj(r)z(r)/")

=

ef dr J dr'j{r)(z{r}z(r'J);{T'1/2h'.

(3.307)

Considering now a discrete time axis sliced at t = tn. and inserting the special source current j(Tn} = K6n,o, for instance, we find directly (3.304). The Wick theorem in this form has an important physical application. The intensity of the sharp diffraction peaks observed in Emgg scattering of X -rays on crystal planes is reduced by therma1 fluctuations of the atoms in the periodic lattice. T he roouction factor is usually written as e-:2W and called the Debye- Waller factor. In the Gaussian approximation it is given by (3.308)

where u (x) is the atomic displacement field. If the fluctuations take place around (x(r)} into

i

0, then (3.304) goes obviously over

(3.309)

3 .10.1

R eal-Time Corr elation Functions

T he translation of these results to real times is simple. Consider, for example, the harmonic fluctuations ox(t) with Dirichlet boundary conditions, which vanish at tb and to.. Their correlation funct ions can be found by using t he amplitude (3.23) as a generating funct ional, if we replace x(t) -+ ox(t) and Xb = Xa -+ O. Differentiating twice with respect to the external currents j(t) we obtain

c~i(t, n ~ (x(t)x(t')) ~ i !Cw,(t - r),

(3.310)

with the Green funct ion Gw 2(t - t') of Eq. (3.36), which vanishes if t = tb or t = to.' T he correlation function of i:(t) is

.().(')) (x txt

.hcosw(t.-t»cosw(t,D,ZI>,

(X(T)p,,(T'» = iMa~,G~2J,B,z:r,(T,T') - MWBG~J,B,%~(T,T'),

(3.3M)

2 ) d "",H,~1>y ( T ,'T) 2J d...',B,.", (T,T ')

(X(T)p~(r'» = iMl}T,G~J,B,,,,~(T, T')

(3.355)

C ..'"',B,p.,,_ (T,'T)

+ MWBG~lJ,B,n{T,r'),

(3.356) (Z(T)P.(T'» = iMl}T, G~l,B, .. (T,r'), 2 2 (Pz (T )p:z( T'» = - M l}T8T•G~lJ,B,u (T, :r') - 2iM w B8TG~1,B ,:z~ (T, r')

+ M2w~ G~l,B,u(T, T') + hM6(T - T'l, C...'"',B,p. ". (T,T ')

(3.357)

(P.,(T)p~(T')}= -M2 8TaT,G~1.B,:z~(T,T') + iM28TG~1.B ,n(T, T') 2 ) + M 2 w2B d"",B.z~ (n') , ,

(3.358)

G~l,B.p,p, (T,T') =: (P.(T)P.(r')}= - M2aTaT,G~1.R, .. (T,r') + hM6(T - T').

(3.359)

Only diagonal correlations between momenta contain the extra 6-function on the right-hand side according to t he ru le (3.352). Note that 8TIJT'G~1'BloI>(T' T') = -1J:G~2J,B,"b{T, T'). Each correlation function is, of course, invariant under time trans ations, depending on ly on the time difference T - T'.

The correlation functions (X(T)X(T')} and (X( T)Y(T'» are the same as before in Eqs. (3.324) a nd (3.325).

3.12.2

R elations between Various Amplitudes

A slight generahation of the genera ting functional (3.330) contains paths with fixed end points rather than all periodic paths. If the endpoints are held fixed in configura tion space, one defines (x~h.Blx .. O) [j,kl =

1""'-"

{l

Dp~ exp - "i"A,,[j,kj Dx-,

"'(O)=z.

1f«

«

}.

(3.360)

If t he e ndpoints arc held fixed in momentum space, one defincs

(3.361) The two are related by a Fourier transformation

(3.362) We now observe that in the canonical pa.th integral, the am plitudes (3.360) and (3.361) with fixed endpoints can be reduced to those with vanishing endpoints with modified sourecs. T he modification consists in shifting the current k(T) in t he action by the source term ix~6(1"b - T) -

259

3.12 Correlation Functions in C8.ll0Ilical Path Integral

iX"J(T - '1,,) and observe that this produces in (3.361) an overall phase factor in the limit 1), 1 hf3 and Ta J 0: lim lim (Pb hf3IPa O)li(T), k(r)

... 1~~r. IO

= exp

+ iXbJ(1), - '1) - 1x"J(r - Ta) 1

{* (PbXb - Pax,,) } (Pb hPiPa O)li(r), k(T)].

(3.363)

By inserting (3.363) into the inverse of the Fourier transformation (3.362) , (3.364) "''il

obtain

(Xb hPlx a O)li, k]= lim lim (OTif3100)li(T), k(T) ... lh/Jr. IO

+ ixbJ(1), -r ) - ix"J(T - ra )] .

(3.365)

In th is way, the fixed-endpoint path integral (3.360) can be rWuced to a path integral with vanish ing endpoints but addi t ional J-terms in the current k(T) coupled to the momentum P{T). T hen: is also a simple n:lation betwccn path integrals with fixed equal endpoints and periodic path intcgrals. T hc measuI"CS of integration an: related by "(11/1)=,,

1

"(0) _,,

V V ~=

f

21ft,

V Vp

- "- J(x(O) - x) . 21fh

(3.366)

Using the Fourier decomposition of the delta function, we rewrite (3.366) as (3.367) Inserting nOW (3.367) into (3.365) leads to t he announced dcsirW relation

(Xb li/JIX" O)[k,j]

=

lim lim lim

roo

... 111/J .... IO ...!IO} _oo

xZ li(T) - ip"J(T -

,dP: 1f"

T~), k(T)

+ iXbJ(1), -

'0) - ix"J( '0 - '0,,)] ,

(3.368)

where Zli, kJ is the thennodynamic part ition function (3.330) summi ng all periodic paths. When using (3.368) we must oc careful in evaluati ng the thf(lC limits. T he limit T~ J 0 has to oc evaluated prior to the other limits 1), 1 hf3 and '1" 1 o.

3.12.3

Harmonic Generating Functionals

Here we write down explicitly the harmonic generating funetionals with the above shifted source terms:

f(T) = k(T) + ixbJ(1), - '0) - iX"J(T - Ta),

J(T) = jeT) - ipJ(T - ~),

(3.369)

leadi ng to the factorized generating functional

Z", [k,Ji = z~O)[O,OJ Z~I) [k , jiZ![k,jJ. T he respective terms on the right-hand side of (3.370) read in detail

Z~O)[O,OJ

=

Z",exp

C~2 { _p2G~,,(~,~) -

2p

[x"G~p(~,T,,) + xbG~,,(~, 1),)]

(3.370)

260

3 External Sources, Correlations, and Perturbation Theory -z! G~(1"c, 'Te) - x~G~(1)" 1),) + 2xQXbG~h, Til) }) , exp

(3.371)

(:1 [ '/3 dr {j(T)[-ipG~%(T, T~) + iX~G~p(T,1"b) - iX"G~p(T,T,,)J

+k(T) [-ipG~k,r,;,) + ixbCl:,,cr, 1),) - ix"G~(T,T,,)J} ) ,

~p { ~ ["tJ dT) r~tJ dT2[Uhl,k(T2)) 2h Jo Jo

ZE lk,;]

x ( G~r(T" rJ) Gi;,,{Tl> 1"2) G~(Tl,Tl)

G~(Tl,Tl)

(3.372)

1( l]} j(7"2) k(1',)

(3.373)

,

where Z.., is given by (3.342) and G~p(TJ,T2) etc. are the periodic E uclidean Green functions defined in Eqs. (3.346)- (3.349) in an abbrev iated notation. Inserting (3.370) into

all} beTh 7"2)

(3~3;)

and performing the Gaussian momentum integration, over the exponentials in and Z~l) [ k,jl , the result is

(Xb 1i.Blxa O)[k, j] = (Xb 1i.Blxa 0)[0, 01 x cxp { xexp

{ 1""r dT)

--;.

21i

0

dT:!lU(TJ),k(T2»

)0

Z~O) [0, 0]

~ ['tJ dT [Xel(T)j(T) + PCI(T) k(T) ]}

(

101 101 ) ( .J(Tl) G(~)(Tt,Tl) G(:;t1'1'2)

Gp~ (T!,1'2)

G"" (Tl ,T2)

k (1'2)

l]}

,

(3.374) where t he Green funct ions G~?)(Tl' T:l) have now Dirichlet boundary conditions. In particular, the Green function G~?\T!,Tl) is equal to (3,36) continued to imaginary timc. Thc Green funct ions G~~)(Tl ' 1'2) and G~)(Tl' 1'2) arc Dirichlet versions of Eqs, (3,346)- (3,349) which arise from t he above Gaussian momentum integrals. After performing t he integrals, the first factor without eurrents is

Performing the limits using h lim lim G~p«, T,,) = -i - ,

~.IO~~IO

2

(3.376)

where t he order of t he respect ive li mits turns out to be important, we obtain t he amplitude (2.409):

Mw 21fhsinh h{3w

x cx p {

2h::h\3w[(X~+X~) COSh li.8w -2xax~1 } .

The first exponential in (3.374) contains a complicated representation of the classical path

(3.377)

261

3.12 Correlation Functions in C8.ll0Ilical Path Integral

(3.378)

and of the classical momentum

(3.379)

Indeed, lI\lierting the explicit periodic Green functions (3.346)-(3.349) and going to the limits we obtain X B sinh..,(h/3

- T) + sinhh{3w

xb~inh"'T

(3.380)

(3.381)

the first being the imaginary-time version of the elassical path (3.6), the second being related to it by the classical relation PdCr) = iMd:rd(T)/dT. The second exponential in (3.374) quadratic in the currents contains the Green functions with Dirichlet boundary condit ions (3.382) (3.383)

(3.384) (3.3SS) After applying some trigonomtric identities, thc:;c take the form h

G~~)(T!, Ta) = 2M.., sinh h/L [cosh..,(IiP-IT) - T21) - cosh..,(h/3 - T) - T2) ],

G~~)(T!, Ta)

2

si:

h{3w (O(T) - Tal sinh ..,(h/3- IT) -Tall

-B(T2-Tilsinh..,(h/3 - ITZ - T)i)+sinh..,(h/3 - T) - T2)},

G~)(TJ,72)

(3.386)

(3.387)

.h

2sinhh/L {O{T) - 72)smh..,{h/3 - IT) - T21) -D{1'J - TJ) sinh..,(h/3 - 172 - T)il - sinh ..,{h/3 - T) - 1'J)},

At"'"h{3w [cosh ..,{h/3 - ITI -Tal) + cosh..,(h/3-TI - Ta)] . G~)(TJ, Ta) = 2 sinh

(3.388) (3.389)

T he first correlation function is, of course, the imagi nary-time version of the Groen function (3.206). Observe t he symmetry properties under interchange of the time argumenl.$:

G~~){TL'Tz) = G~~){T2,Tll,

G~~)h,T2) = - G~~)(Tz, TI) ,

(3.390)

G~~){TL' Ta) = -G~){T2,Tll,

G~)(ThT2) = G~){Tz, TI),

(3.391)

262

3 External Sources, Correlations, and Perturbation Theory

and t he identity (3.392)

In addition, there are the following derivative relations between the Groon functions with Dirichlet boundary conditions: (3 ,393) (3 ,394) (3 ,395) Note that Eq. (3.382) is It. nonlinear alternative to the additive decomposition (3. 142) of It. Green function with Dirichlet boundary condit ions: into Green functions with periodic boundary

conditions.

3.13

Particle in Heat Bath

The results of Section 3.8 are the key to understanding t he behavior of a quantummechanical particle moving through a dissipative medium a t a. fixed temperature T . We imagine the coordinate x(t) a particle of mass M to be coupled linearly to a heat bath consisting of a great number of ha rmonic oscillators X i(T) (i = 1, 2,3, . .. ) with various masses M; and frequencies 0 ;. The imaginary-t ime path integral in t his heat bath is given by

(x,"Plx. O) ~ x

I>" "'" f.I f VX;(r) 1>(0)=>. V x(r)

exp{-~fo"i3 dr ~ [~i(X/+ O~Xl)]}

x exp {

(3. 396 )

-~ fo"fJ dr [~ :i;2 + V(x(r)) - ~ C;X;(r)x(r) ]} x n;lz;'

where we have allowed for an arbitrary potential V (x ). The partition functions of t he individual bat h oscillators

Z; == ~

f VX;(r)exp { - ~ fo"P dr [~;(X/ + O~X;)]} 1 2 sinh("j3il;(2)

(3.397)

have been divided out, since their t hermal behavior is trivial and will be of no interest in the sequcl. The path intcgrals ovcr Xi(r) can be performed as in Section 3.1 leading for each oscillator label i to a source expression like (3.243), in which c;x(r) plays t hc role of a current j(r). T he result can be written as

(Xb h.6lx"O)

=

r(~)=Zb V x(r)cxp { _ _"I r"fJ dr [M:i;2 + V (x(r » ]- _,1, .Abat.b [XI} ,

Jz(O )_zG

Jo

2

(3.398)

263

3.13 Partide in Heat Bath

where Abath [x] is a nonlocal action for the particle motion generated by the bath

d7 Io'P d7'X(7)O:(7 - r)x(r). 'Io'P 2

Abatb[X] =

- -

0

0

(3.399)

The function a(7 - 7') is t he weighted periodic correlation function (3.248) :

a(7 - 7')

=

~~ ~i G:;~.e (7 -

~

L: --=L w;h fl,( IT -

7')

T'I - oN 2) , mh(fl,op /2)

,2M,fl,

Its Fourier expansion has the Matsubara frequenc ies Wm

=

(3.400)

21rk n T/h

(3.401) with the coefficients

"

am = L. •

d

1

~.• W2m + w~• ·

(3.402)

Alternatively, we can write the bath action in the form corresponding to (3.276)

Abath[X] =

00 d7 1 d7'X(7)ao(7 11," 2

--

0

- 00

7')x(r),

(3.403)

with the weighted nonperiodic correlation function [recall (3.277)]

(3.404) T he bath properties are conveniently summarized by the spectml density of the bath

Po(w') " 2.

L: --=Lo(w' 2M;n; i

fl,) .

(3.405)

T he frequencies n; are by definition positive numbers. The spect.ral densit.y allows us to express 0:0(7 - r) as t he spectral integral

(3.406) and similarly

(

') 1,00 -Pow ,,",' (,)coshw'( IT -

a7-7 -

0

211"

,-'1- oN 2) sinh(w'MJ/2)·

(3.407)

264

3 External Sources, Correlations, and Perturbation Theory

For t he Fourier coefficients (3.402), t he spectral integral reads (3.408)

It is useful to subtract from these coefficients t he first term 0'0 , and to invert the sign of t he remainder making it positive definite. Thus we split Ct m

=2

dd ",,(w') ( W!) -- 1wI2 1,o00 21f w' w",+ 2

=0'0-9",·

(3.409)

T hen the Fourier expansion (3.401) separates as a(T - T') = O'oQP(r - r) - 9(r - T'),

(3.410)

where JP(r - r) is t he periodic o-function (3.279) :

b'(r - r) ~ -

1

ti/3

0000

L

e-"'-('-" ) ~

"' = - 00

L

b(r - r' -

""/3),

(3.4 11)

1>= - 00

the right-hand sum following from Poisson's summation formula (1.197). The subtracted correlation function

has t he coefficients

(3.4 13)

The corresponding decomposition of the bath action (3.399) is (3.4 14)

where

11," dr 1," dT'X(T)g(T - r)x(T'), 2 , ,

A~th [xl = -

(3.4 15)

and (3.4 16)

is a local action which can be added to t he origina l action in Eq. (3.398) , changing merely t he curvature of t he potential V (x). Because of t his effect, it is useful to introduce a frequency shift ~'l via t he equation M!:::.w

, =-Qo= _ 1,00 ",,'-",,(w') Ci -2 - = - 2: --, . o

21f

w'

i

Mini

(3.4 17)

265

3.13 Partide in Heat Bath

Then the local action (3.41 6) becomes A loe

M fM 2 = "2AW2 Jo dTX (T) .

(3.418)

This can be absorbed into the potential of the path integral (3.398) , yielding a renormalized potential (3.4 19)

With the decomposition (3.414), the path integral (3.398) acquires the for m

(x/,1i.8lx aO) =

_ VX(T)exp (-~1 1" dT [M _ ±2 + Vn;,n(X(T))11'('''~'' 2 0: (0)-0:4

/I

0

-h1 Ab..th[X]} . (3.420)

The subtracted correlation function (3.4 12) has the property

r"

Jo

(3.421)

dTg(T - T') = 0.

Thus, if we rewrite in (3.415) (3.422)

the first two terms do not contribute, and we remain with

'1" dT 1, " drg(T - r)[X(T) - X(T'W .

Ab..th [x] = - -

4 ,

(3.423)

If the oscillator frequencies fl; are densely distributed, the function Pb(w') is continuous. As will be shown later in Eqs. (18.208) and (18.317) , an oscillator bath introduces in general a friction force into classical equations of motion. If t his is to have t he usual form - M -yx(t) , the spectral density of the bath must have the approximation Pb(W')::::,: 2M,,{w' (3.424)

[see Eqs. (18.208), (18.31 7)]. T his approximation is characteristic for Ohmic dissipation. In general, a typical friction force increases with W only for small frequencies ; for larger w , it decreases again. An often applicable phenomenological approximation is the so-called Drude form

Pb(w')::::,: 2M,,{w'

2

w'0

Wo + w'

2'

(3.425)

where l / wf) == TO is Drude 's relaxation time . For times much shorter than the Drude time TO , there is no dissipation. In the limit of large Wv , the Drude form describes again Ohmic dissipation.

266

3 External Sources, Correlations, and Perturbation Theory

Inserting (3 .425) into (3.41 3) , we obtain the Fourier coefficients for Drude d issipation

, 10

9m = 2M"(wl)

00

o

dw -2

2w~

1 2

2

2

IrWo + w wm +w

2

= MIWm l'Yl

WD Wm

I

+wO

.

(3.426)

It is customary, to factorize 9m == M lwml'Ym .

(3.427)

so that Drude dissipation corresponds to (3.428)

and Ohmic dissipation to 1m =: 'Y. The Drude form of the spectral density gives rise to a frequency shift (3.4 17) (3.429) which goes to infinity in the Ohmic limit Wo -- 00.

3.14

Heat Bath of Photons

The heat bath in the last section was a convenient phenomenological tool to reproduce the Ohmic friction observed in many physical systems. In nature, t here can be various different sources of dissipation. The most elementary of these is the deexcitation of atoms by radiation, which at zero temperature gives rise t o t he natural line width of atoms. The photons may form a t hermally equilibrated gas, the most famous example being the cosmic bl ack~body radiation which is a gas of the photons of 3 K left over from the big bang 15 billion years ago (and which create a sizable fraction of the blips on our television screens) . The theoretical description is quite simple. We decompose the vector potential A (x , t) of electromagnetism into Fourier components of wave vector k

A(x, t) ~

L: ",,(x)X, (t),

,

(3.430)

The Fourier components Xk (t) can be considered as a sum of harmonic oscillators of frequency l1k = elk l, where e is the light velocity. A photon of wave vector k is a quantum of Xk(t ). A certain number N of photons with the same wave vector can be described as the Nth excited state of the oscillator Xk (t). The statistical sum of thesc harmonic oscillators led Planck to his famous formula for the energy of black-body radiation for photons in an otherwise empty cavity whose walls have a temperature T. These will form the bath, and we shall now study its effect on the quantum mechanics of a charged point particle. Its coupling to the vector potential is given by the interaction (2.632) . Comparison with the coupling to t he

267

3.14 Heat Bath of Photons

heat bath in Eq. (3.396) shows t hat we simply have t o replace - E,c; X ,(r)x(r) by - EkCJ. Xk (r):ic(r). T he bath action (3.399) takes then t he form Aba~h[xl

= --1 2

i" dr i"~ dr'x'(T)a"( . ..x (T),T; X(T') , r').i'(T') . , 0

(3.431)

0

where a,j(x,T;X',r') is a 3 x 3 matrix generalization of the correlation func tion (3.400),

,

Q;j(X,T; x' ,r) ~ ;c'I:C-k(x ),.(x')(X:k(T)XI(r)).

(3.432)

k

We now have to account for the fact that there arc two polarization states for each photon , which are transverse to the momentum d irection. We therefore introduce a transverse Kronecker symbol (3 .433)

and write the correlation function of a single oscillator X~ k (T) as

G;j_ k'k (T - T')

=

(X';- k' () T X'jk' (')) T

=

hTOj , G'",,2 ,., k (T - T') , 0kOkk'

(3.434)

with (3.435)

Thus we fi nd ii(

. '

') _ "

(t x,r, x ,T - dl

J(271")3 fi3k

1b":i

e,k (x - x')

k

20 k

cosh Ok(IT - T'I - hfJj2) sinh(l1k ;tfJj2) .

(3.436)

At zero temperature, and expressing 11k = elkl, this simplifies to

.. (t"(x, r;x,r')

e fi3k c1! (271")3 i5'~ 2

=

T . . eik (x - x') - cl k ll.,. - .,.'I

21kl

.

(3.437)

Forgetting for a moment the transverse Kronecker symbol and the prefactor e2 jc?, the integral yields ) 1 GcR ( X,T; X" ,T = 47T2dl(r

1

T')2 + (X

x ,)2/dl'

(3.438)

which is t he imaginary-time version of t he well-known retarded Green function used in electromagnetism. If the system is small compared to the average wavelengths in the bath we can neglect the retardation and omit the term (x - x')2/r:? In t he finite-temperature expression (3.437) this amounts to neglecting the x-dependencc.

268

3 External Sources, Correlations, and Perturbation Theory

The t ransverse Kronecker symbol can then be averaged over all directions of the wave vector and yields simply 26;j /3, and we obtain the a pproximate function . x'

ii( 0'

x , T,

2

') _ 2e {/i _ 1_ ,T 3& 2Jrc2

J21T w coshw(lr - 7'1 - ftf3/ 2) sinh(wh.8/2) . dw

(3.439)

This has the generic form (3.407) with the spectral function of the photon bath (3.440)

T his has precisely the Ohmic form (3.424), but there is now an important difference: the bath action (3.431) contains now the time derivatives of the paths x(r). This gives rise to an extra factor wI'.! in (3.424), so that we may define a spectral density for the photon bath: (3.44 1)

In contrast to t he usual friction constant J in the previous section, this has the dimension I Jfrequcncy.

3.15

Harmonic Oscillator in Ohmic Heat Bath

For a harmonic oscillator in an Ohmic heat bath, the partition function can be calculated as follows. Setting

M , , V""(x) ~ ,w x ,

(3.442)

the Fourier decomposition of the action (3.420) reads

A.,

=

Mh klJT

{W', ~ ["+ "'2xo + ;;:1 Wm

W

+ Wm 'Ym1IXm

I'}

.

(3.443)

The harmonic potential is the full renormalized potential (3.4 19). Performing the Gaussian integrals using the measure (2.445), we obtain the partition function for the damped harmonic oscillator of frequency W [compare (2.406») (3.444)

For the Drude dissipation (3.426) , t his can be written as ZdlUIlp

'" Let

WI, W'l, W3

fi

= koT w~(wm + WD) 1iw m= Iw;. + W;'WO + wm(W'l + 'YWD) + WOW'l·

(3.445)

be the roots of t he cubic equation w 3 _ W'lWD

+ w(w'l + 'YWD) -

W'lWD

= O.

(3.446)

269

3.15 Harmonic Oscillator in Ohmic Heat Bath

Then we can rewrite (3.445) as z""mp

= knT

liw

w

fi

Wm

Wm

Wm

Wm

m = IWm + W\Wm + W2Wm + W3

+ WD

(3.447)

Wm

Using the product representation of t he Gamma functionS

nZ r(Z) ~ lim -

II -m+-z "

(3.448)

.. _ 00 Z m = 1 m

and t he fact that (3.449)

t he partition function (3.447) becomes zdam p =

'"

~ W f(w';wd f ("Il'2/wdr (W3/wd 2Jr WI r(wD/wd '

where WI = 2Jrk n T/h is the first Matsubara frequency, such that In the Ohmic limit WD --lo 00 , the roots WI, W2 , W3 reduce to

WI = ,,(/2 - i6,

WI = 1/2 + i6,

W3

(3.450) Wi/WI

= w;(3/2Jr.

(3.451)

= WD - 1 ,

with

J " JW'-1' /4 ,

(3.452)

and (3.450) simplifies further to

z~amp =

1 -2 ":::" r (wdwdf(wdwd. rrw,

For vanishing fr iction, the roots WI and the formula9

W2

become simply WI = iw,

(3.453) W2

= - iw, and

rr sm Jrz

r( 1 - z)r(z) ~ -.-

(3.454)

can be used to calculate (3.455)

showing that (3.450) goes properly over into the partition function (3.214) of the undamped harmonic oscillator. The free energy of the system is

F(T) =

- knT [log(w/2Jrwd - IOgr(WD/wd + log f (wdwd + log f(W2/WI) + log f(W3/wd l .

----;:;;;-;;:::-=:-:-=~:::;:;;:-

sI.S. Gradshtcyn and I.M. Ryzhik , op. cit. , Formula 8.322. "ibid., Formula 8.334.3.

(3.456)

270

3 External Sources, Correlations, and Perturbation Theory

Using t he large-z behavior of log r(Z)lO

1) 1 1 1 log r (z}= ( z-2" logz-z + 2"log21r+ 12z - 360z3 - O (l/z't

(3.457)

we find the free energy at low temperature

1 F(T) - E,- ( WI

+ -1 + -1 W2

WI

W2)1I" 2 '111" 2 - (kRT) ~ Eo- - (kRT) , W\W2W3

6w 2h

6h

(3.458)

where

Eo

= - -

Ii

2.

[WJlog(wdwD)

+ 1V21og(w~dwD) + W3!Og(W3/WV)]

(3.459)

is t he ground state energy. For small fr iction, this reduces to

'Y 21T

flW

Wv

"r'~

W

16w

Eo~ - + - Iog - - -

2

( 1+ -4w

)

7rWo

_> + Ob)·

(3.460)

The P-behavior of F(T) in Eq. (3.458) is typical for Ohmic dissipation. At zero temperature, t he Matsubara frequencies Wm = 21fmknTjh move arbitrarily close together, so t hat Matsubara sums become integrals according to t he rule (3.461) Applying t his limiting procedure to t he logarit hm of the producl. formula (3.445), t he ground state energy can also be written as an integral

E _ h roo dJ..J I [W~ + W!.WD+wm(wZ+l'WD)+WDWZl 0 - 21!"Jo m og w;,(wm+WO) ,

(3.462)

which 5hows that thc encrgy Eo incrcases with thc friction coefficient 'l'It is instructive to calculate the density of states defined in (1.580). Inverting the Laplace transform (l. 579) , we have to evaluate

p(.;:)

1 1"+;00 d{3 eit{i z~"mp({3), 21!"t ,,- ;00

= -.

(3.463)

whcre TJ is an infinitesimally sma.lI posit ive number. In thc a.bsence of friction , t he integral over Z.,({3) = L~=o e - {ihw(n+ l /2j y ields 00

p(,) ~

l: b(, -

(n + 1/2)1iw) .

(3.464)

n=O

In the presence of friction, we expect t he sharp 6-function spikes to be broadened. The calculation is done as follows: The vertical line of integration in the complex {3-planc in (3.463) is moved all the way to the left, thereby picking up the poles

271

3.16 Harmonic Oscillator in Photon Heat B ath



c

~

R.,

Figure 3.5 Poles ill complex p-pJane of Fourier integral (3.463) coming from the Gamma functions of (3.450) of the Gamma funct ions which lie at negative integer values of w;(J/27r. From the representation of the Gamma function l l

f (z)

=

1.

00

dt e - 1e- t

1

1)"

+ L: -+,---,:,---, n=O n !(z + n) 00

(

(3.465)

we see the size of the residues. Thus we obtain the sum (3.466)

R,. _ w (_ I)n- 1 f (- nwdwdr(- nw3/ wd ,I - w~ (n I)! r ( 'T!WO/WI)

,

(3.467)

with analogous expressions for R..,2 and R..,3' T he sum can be done numerically and yields the curves shown in Fig. 3.6 for typical underdamped and overdamped situations. T here is an isolated a-function at the ground stat e energy Eo of (3.459) which is not widened by t he friction. Right above Eo , the curve continues from a finite value p( Eo + 0) = '""(1r /&i determined by the first expansion term in (3.458).

3.16

Harmonic Oscillator in Photon Heat Bath

It is straightforward to extend this result to a photon bath where the spectral density is given by (3.441) and (3.468) becomes

Z;~, ~ lOibid., Formula 8.327. 11 ibid., Formula 8.3l4.

kBT Iiw

{IT [w~ +w:wm+w::.0j }-' m= l

(3,468)

272

3 External Sources, Correlations, and Perturbation Theory

,

3'

,

pet ) 2.5

3.'

,

p(e) 2_5

,

,.,

",

I - ----

0.'

0.'

0,LJec-o,e-'3--., -"C-"C-',--;s tjw O,L, _,\--;,e-., -',c-', ---;''--'8 € /w

Figure 3.6 Density of states for weak and strong damping in natural units. On the left, t he parameters are ,,(/w = 0.2, WD / W = 10, on the right 'Y/w = 5, wo /w = 10. For more details see Hanke and Zwerger in Notes and References.

with"( = e 2 j621rM . The power of Wm accompanying the friction constant is increased by two units. Adding a Drude correction for the high-frequency behavior we replace I by wm/(Wm + wo) and obtain instead of (3.445)

zdamp = kBT

'"

Iiw

fi: ma l

W~(wm+ wD)(l+'YwD) ~(1

.

+ ')'WD) + W~WD + wm w 2 + WDW 2

(3.469)

The resulting partition function has again t he form (3.450), except that W123 are t he solutions of the cubic equation

(3.470) Since the electromagnetic coupling is small, we can solve this equation to lowest order in -y. If we also assume Wv to be large compared to w, we find t he roots

where we have introduced an effective friction constant of the photon bath ~ff 1'pb

=

e2 '). 6c27rM w ,

(3.472)

which has the dimension of a frequency, just as the usual friction constant l' in the previous heat bath equations (3.451) .

3.17

Perturbation Expansion of Anharmonic Systems

If a harmonic system is disturbed by an additional anharmonic potential V(x ), to be called interaction, the path integral can be solved exactly only in exceptional cases. These will be treated in Chapters 8, 13, and 14. For sufficiently smoot h and small potent ials V(x) , it is always possible to expa nd the full partition in powers of t he interaction strength. The result is the so-called perturbation series. Unfortunately, it only renders reliable numerical results for very small V (x) since, as we shall prove

3.17 Perturbation Expansion of Anharmonic Systems

273

in Chapter 17, the expansion coefficients grow for large orders k like k!, making t he series strongly divergent. The can only be used for extremely small perturbations. Such expansions arc called asymptotic (more in Subsection 17.10.1). For this reason we arc forced to develop a more powerful technique of studying anharmonic systems in Chapter 5. It combines the perturbation series with a variational approach and will yield very accurate energy levels up to arbitrarily large interaction strengths. It is t herefore worthwhile to find the formal expansion in spite of its divergence. Consider the quantum-mechanical amplitude

(Xbtblxata)

=/.%(!bl "'%b Vxexp { ~Ii. ltbt~ dt [M2 :t2 _ Mw'l2 x %(t~l "'%~

2-

V(X)] }

(3.473)

(3.475)

A similar expansion can be given for the Euclidean path integral of a partition function (3.476)

(3.477)

The individual terms arc obviously expectation values of powers of the Euclidean interaction A nte ,

,

== 10" drV(x(r)),

(3.478)

274

3 External Sources, Correlations, and Perturbation Theory

calculated within t he harmonic-oscillator partition function Z",. The expectation values are defined by (3.479)

With t hese, the perturbation series can be written in the form

Z

(1-

=

~(Ain~,c)", + 2!~2{A~,t.c)w - 3!~3{A~nt'Dh" + ... )Z",.

(3.480)

As we shall see immediately, it is preferable to resum the prefactor into an exponential of a series 1=

~(A;n~,")w + 2!~2 (A;~~,c)w - 3!~J (Al

nl ,,,),,,

exp { - *(AiDt,e}w

+ ...

+ 2!~2 (Afnt,,,}w,e - 3!~3 (A~nt.e}w,c + .. .} .

(3.481)

T he expectation values {}",,c are called cumulanls. They are related to the original expectation vaJues by the cumulant expansionY

(A~nt,e}w,c -

(A~nt,e}."'c =

{Afnl, .. )", - (Aint,,,): {[A nt,., - {A nt,e)wfl ) ... , (A~Dt,e)w - 3{A~nt,e)w(Aint,e)w {[A im,., - {Aint,e)...

(3.482)

+ 2{Aint,e)!

(3.483)

f')...,

T he eumulants contribute directly to the free energy F = - (l/,8)logZ. From (3.481) and (3.480) we conclude that the anha rmonic potential V(x) shifts t he free energy of t he harmonic oscillator Fw = (1/,8) log[2sinh (h,8w/2)] by fJ.F =

~ (~(Aillt,C)'" - 2!~2(A~nt,e ) ...,c + 3!~3(A~nt,e)w,c + ... ).

(3.484)

Whereas the original expectation values ( A [;.t,e) ... grow for large ,8 with the nth power of ,8, d ue to contributions of n disconnected diagrams of first order in 9 which are integrated independently over 1" from 0 to A,8, t he cumulants (A~t"e) ... are proportional to ,8, thus ensuring that the free energy F has a fin ite limit, t he ground state energy Eo. In comparison with the ground state energy of the unperturbed harmonic system, the energy Eo is shifted by

!leo = J!.?'. ;, P1 (1h(A;nt,e)w -

1

2

1

3

2!1i2 (Ai"t,e)w,c + 3!IiJ (Ai"t,e)w,c

) + ....

(3.485)

12Note that the subtracted expI'CS8iollll in the second lines of these equations are particularly simple only for the lowest two cumulants given here.

3.17 Perturbation Expansion of Anharmonic Systems

275

There exists a simple functional formula for the perturbation expansion of the partition function in terms of the generating functional Z... [j] of the unperturbed harmonic system. Adding a source term into the action of the path integral (3.476) , we define the generating functional of the interacting theory:

Z[j]

=

f Vx exp { -~ fo" Pdr [~ (X2 +

w

2X2) + V(x) - jX] }.

(3.486)

T he interaction can be brought outside the path integral in the for m

Z[j]

=

e- i

f:

1l

dr V(J/ Jj(r»

Z... [j ]

(3.487)

The interacting partition fu nction is obviously Z~Z[OI ·

(3.488)

Indeed, after inserting on the right-hand side the explicit path integral expression for Z[j ] from (3.233): (3.489)

and expanding the exponential in t he prefactor • Jo r Ail drV(6J6J(r» . e- K

+;;;;:; 1 2.n

i'" i"

=

0

dT,V(ojJj(T,))

0

- ;;;;:;. 1 3.llO

Ii" drV(o!oj(r» i" i"

1- Ii

(3.490)

dT, V(OjOj(T,))

0

dT3V(OjOj(T3))

dT,V(ojo(T,))

0

i"

dT, V(OjO(T.»)

0

+.

the functional derivatives of Z[j] with respect to the source j(r) generate inside the path integral precisely the expansion (3.480) , whose cumulants lead to formula (3.484) for the shift in t he free energy. Before continuing, let us mention that the partition function (3.4 76) can, of course, be viewed as a generating functional for the calculation of the expectation values of the action and its powers. We simply have to form the derivatives with respect to /i -I : (3.491)

For a. harmonic oscillat or where Z is given by (3.242) , this yields (3.492)

The same result is, incidenta.lly, obtained by calculating the expectation value of the action with analytic regularization:

2 (x2(r»).., + w (x2(r»)..,=

I

dw

l

wt2 271"w12+w2+

l,w w' Id"" 2;w12+w2=

The integral vanishes by Veltman's rule (2.506).

271" =0.(3.493)

276

3 External Sources, Correlations, and Perturbation Theory

3_18

Rayle igh-Schrodinger and Brillouin-Wigne r P e rturbation E x p a n sion

T he expectation values in formula (3.484) can be evaluated by means of the socalled Rayleigh-Schriidinger perturbation expansion, also referred to as old-fashioned perturbation expansion. This expansion is particularly useful if the potential V(x) is not a polynomial in x. Examples are V( x) = 6(x) and V(x) = l/x. In these two cases the perturbation expansions can be summed to all orders, as will be shown for the first example in Section 9.5. For the second example the reader is referred to t he literature. 13 We shall explicitly demonstrat e t he procedure for the ground state and the excited energies of an anharmonic oscillator. Later we shall also give expansions for scattering amplitudes. To calculate the free-energy shift ~F in Eq. (3.484) to first order in V(x), we need t he expectation (3.494)

The time evolution amplitude on the right describes the temporal development of the harmonic oscillator locat ed initially at the point x , fro m the imaginary time 0 up to 7]. At the time 7], the state is subject to the interaction depending on its position XI = x(7d with the amplitude V(xd. After that, the state is carried to the final state at the point x by the other time evolution amplitude. To second order we have to calculate the expectation in V(x): 1( ,

"2 A int ,.,

)

_

'"

=

(3.495)

T he integrat ion over 71 is taken only up to 72 since the contribution with 7] > 72 would merely render a factor 2. The explicit evaluation of t he integrals is facilitated by t he spectral expa nsion (2.298). T he t ime evolution amplitude at imaginary times is given in terms of the eigenstatcs l/Jn(X) of the harmonic oscilla tor with the energy En = Tiw(n + 1/ 2): 00

(Xb7b lx"7,,),,,

=

L

1j.on (xb) 1j.>~(x,,) e- £" ("" -T~)/h .

(3.496)

n=O

The same type of expansion exists also for the real-time evolution amplitude. T his leads to the Rayleigh-SchrOdinger perturbation expansion for the energy shifts of all excited states, as we now show. The amplitude can be projected onto the eigenstates of the harmonic oscillator. For this, the two sides are multiplied by the harmonic wave functions 1j.o~(Xb) and llM.J. Goo~rts and J .T. Devr~, J. Math. Phys. 13, 1070 (1 972).

3.18

Rayleigll~SchrOdinger

277

and Brillouin-Wigner Perturbation Expansioll

tP.,(x,,) of quantum number n and integrated over in the expansion

Xb

and x"' respectively, resulting

(3.497)

with the interaction

A;••

1'."dt V(x(t)).

=-

(3.498)

The expectation values arc defined by

where

Z

=

QM,w ,n

e- iw (n+1/Z}{t. - I

Q )

(3.500)

is the projection of the quantum-mechanical partition function of the harmonic oscillator

[see (2.40)] onto the nth excited state. The expectation values are calculated as in (3.494), (3.495) . To first order in V(x), one has (nIA nt ln)w =: - ZQ:"',w ,n

1:"dt Jdxbdx"dxltP~(Xb)(Xbtblxltdw l

x V{Xl)(Xlt 1Ix"ta)wtPn(Xa).

(3.501)

The time evolution amplitude on the right-hand side describes the temporal development of the initial state tPn{x,,) from t he time t" to the time t l , where t he interaction takes place with an amplitude - V(xd. After that, the time evolution amplitude on the left~hand side carries the state to tP~(Xb). To second order in V(x), the expectation value is given by the double integral

(3.502)

As in (3.495) , the integral over t, ends at t z. By analogy with (3.481), we resum the corrections in (3.497) to bring t hem into the exponent: i 1 z i 3 1 + 'h(nIAint ln)w - 2!1i2(n l ~ntln)w - 3!h3 (nIAntln)w + . =

exp {*(nIAint[n)W -

(3.503)

2!~2 (nIA~,t ln)w,e - 3!~3 (nIA~>tln)w,c + ... } .

278

3 External Sources, Correlations, and Perturbation Theory

The cumula nts in the exponent are

(nIAi.. ln). - (nIA'odn)~

(nI Ai.. ln)."

~ (nIlA o, - (nIA,"dn).I'ln).,

(nI AI", ln)."

(3.504)

_ (nIAi.. ln). - 3(nIAio, ln).(nIA,"dn). + 2(n I A"dn)~ (nl[A int - (nIA indn)",]3In)"" (3.505)

=

From (3.503) , we obtain the energy shift of the nth oscillator energy dE.. =

{i

. ih 1 2 j.J:~co tb _ ta h(nIA;nd n )", - 2!hz {nIA ;nt ln)""c -

3!~3(nIA~nt ln)""c + .. .},

(3.506)

which is a generalization of formula (3.485) which was valid only for the ground state energy. At n = 0, the new formula goes over into (3.485), after the usual analytic continuation of the time variable. The cumulants can be evaluated furthe r with the help of the rcal-time version of the spectral expansion (3.496): 00

(Xt,lb lx"ta)", =

L ¢n(Xb)t,b;(xa)e- E"(t6- I,,l /h. i

(3.507)

.. =0

Th first order in V (x), it leads to (3.508)

To second order in V(x) , it yields (3.509)

The right-hand side ean also be written as (3.510)

and bocomes, after the time integrations,

(3 .511)

3.18

Rayleigll~SchrOdinger

and Brillouin-Wigner Perturbation Expansioll

279

As it stands, the sum makes sense only for the Elc #- En -terms. In these, the second term in the curly bracket s can be neglect ed in the limit of large t ime differences tb - ta o T he term with Ek = En must be treated separately by doing the integral directly in (3.51O). This yields (3.512)

so t hat

I ) =- '" 'I( 2 n lA'intn"' L

",,0: ..

m

•. (- ) EV.mV Elhtb ta + VnnVnn m

(t. -2 'oJ' .

(3.513)

n

The same result could have been obtained without the special treatment of the Ek = En -term by introducing artificially an infin itesimal energy difference Ek - En = (; in (3.511), and by expanding the curly brackets in powers of tt> - tao When going over to the cumulants !(nIA?ntln)..,.c according to (3.504) , t he k = nterm is eliminated a nd we obtain (3.514)

For the energy shifts up to second order in V(x), we thus arrive at the simple formula (3.515)

The higher expansion coeffi cients become rapidly complicated. T he correction of third order in V(x) , for example, is (3.516)

For comparison, we recall the well-known formula of Brillouin- Wigner equation 14

ll. En = R...n(En + ll.En),

(3.517)

~here R..n(E) are the diagonal matrix elements (n IR(E) ln) of the level shift operator R(E) which solves the integral equation

R(E)

~ II + II 1- p.

E - H",

R(E)

(3.518)

The operator Pn =: In)(nl is the projection operator onto the state In}. The factors 1 - P ensure that t he sums over the intermediate states exclude the quantum R

14L. Brillouin and E.P. Wigner, J . Ph~. Radium 4 , 1 (1933); M.L. Goldberger and K.M. Watson, Collision Theory , John Wiley & Sons, New York, 1964, pp. 425-430.

280

3 External Sources, Correlations, and Perturbation Theory

number n of the state under consideration. The integral equation is solved by t he series expansion in powers of V:

R(E)

=

V + V 1 - ~" V + if E - H",

Up to the third order in

V, Eq.

I - ~" V 1 - ~" V + .. E - H", E - H",

(3 .519)

(3.5 17) leads to the Brillouin- Wigner perturbation

expansion

Vnk Vkl Vi.. E,)(E E,)

+ ... ,

() 3.520

which is an implicit equation for t:::.E" = E - En. The Brillouin-Wigner equation (3.517) may be converted into a n explicit equa tion for the level shift t::.E,,:

C> E" ~ R."(E") + R."(E")R;."(E")+ [R."(E")R;."(E")' + 1il!"(E")R;;"(E") [ + [R." ( E") R;." (E")' + I il!" (E") R;." (E") R;;" (E") + ,R!" (E" )R;7" ( E") [+ .. . .(3 .521) Inserting (3.520) on the right-hand side, we recover the standard RayleighSchriidinger perturbation expansion of Quantum mechanics, which coincides precisely with the above perturbation expansion of the path integral whose fi rst three terms were given in (3.515) a nd (3.516). Note that starting from t he t hird order, t hc explicit solution (3.52 1) for t he level shift introduces more and morc extra disconnccted terms with respect to t he simple systematics in t he Brillouin-Wigncr expansion (3.520). For arbitrary potentials, the calculation of the matrix elements VnJ; can become quite tedious. A simple technique to find them is presented in Appendix 3A. The calculation of t he energy shifts for the particular interaction V ex) = gx 4 /4 is described in Appendix 3B. Up to order if, t he result is

n;

!:lEn

(2n

+ 1) + ~3(2n2 + 2n + l)a4

(~r 2(34n 3 + 51n2 + 59n + 21)a8~ +

(3 .522)

(~r 4· 3(125n4 + 250n3 + 472n2 + 347n + 111 )a 12 n2~2'

The perturbation series for this as well as arbitrary polynomial potentials can be carried out to high orders via recursion relations for the expansion coefficients. This is done in Appendix 3C.

3 .19

Leve l-Shifts and P ertur b ed W ave Funct io n s from Schrodinger E quation

It is illStructive to rederive the perturbation eXp8l1Sion from ordinary operator SchrOdinger theory. T his derivation provides us a1so with the perturbed eigenstates to any desired order.

3.19 LeveJ-Sllifts and Perturbed Wave FunctioIlS from Schriidinger Equ8tioIl

fl

T he Hamilton ian operator

is split into a free and an interacting part

fl = flo+v. Let

In)

be the eigelllltates of

281

(3.523)

those of fl :

flo and I,,(n»

(3.524) We shall assume that the two sets of states In) and 1,,("» a re orthogonal sets, the first with unit norm, t he latter normalized by scalar products a~») ::::::

Due to t he completeness of t he states

In),

(nl,,(»» = I.

(3.525)

the states 11b(n» can be expanded as

L

11b(n» = In) +

a~)lm),

(3.526)

m#

where

are the components of the interacting states in the free basis. P rojecting the right-hand Schrooinger equation in (3.524) onto (ml a nd using (3.527), we obtain

(3.528) Inserting here (3.526), th is becomes

+ (mW ln) + L

E~m)(l~) and for m =

fl.,

a~"){mWlk) = E(n)a;;;),

(3.529)

""

due to the special normahation (3.525),

E~n)

+ (nIVln) + L a~"){n I V l k) =

E (nJ.

(3.530)

"" ;\Iulti plying th is equation with a;;;) and subtract ing it from (3.529), we eliminate the unknown exact energy E(n) , and obtain a set of coupled algebraic equations for a;;;):

a;;;) =

E(") 1 0 -

e(m)

[(m - a;;;JnIVln) + L ainJ(m -

0

k,ton

a~)nIVlk)l '

(3.531)

where we have introduced the notation {m - a;;;)nl for the combina tion of states (ml- a;;;)(nl , for brevity. This equation can now easily be solved pcrturbatively order by order in powers of t he interaction strength. To count these, we replace V by gV a nd expand a;;;) as well as the energies E(n) in powers of 9 as: ~

a;;;)(g) =

L a~~/( - .q)'

(m#n),

(3.532)

1=[

and ~

E(n) = E~n) - L (- g)'Ei n ).

,-,

(3.533)

282

3 External Sources, Correlations, and Perturbation Theory

Inserting these expansions into (3.530), and equating the coefficients of 9, we immediately find t he perturbation ex pansion of the energy of the nth level

(3.534)

(nIVln),

" ,.)

-

L,.. a k,I _ 1 (n lV lk)

I

".

> l.

(3.535)

T he expansion coefficients a~~ are now determined by inserting t he ansatz (3.532) into (3.531). T his yields '

(mIV ln)

o(n) _

I.I.lldforl> 1: (n) a ... ,I=

(m)

Eo

1

(nl

Eo

[

(n)

-

(3.536)

E'.) E(n)' o - 0

m,1 -

- Om,l_l(nIVln)+

L

(n) (1.1:,1 _ 1 (mlV lk)

10'1"

-

,-,L

I' ~ l

(n) 0 ",,1'

L (lk,l _ l_I,(n lV lk)1. (n)

,

k,. ..

(3.53 7)

Using (3.534) and (3.535), this can be si mplified to

(3.538) Together with (3.534), (3.535), and (3.536), this is a set of recursion relations for the coefficients a (n) ... ,1

and

E(n) . I

T he recursion rela tions allow us to recover the perturbation expansions (3.515) and (3.51 6) for the energy s hift. T he seoond-order result (3.515), for example, follows directly from (3.537) and (3.538), the latter giving

E(n) = ' " a(n)(nIV lk) = ' " (kIVln){nIV lk) . 2

L

k;in

k,t

L

~....

E(k) 0

E (n) 0

(3.539)

If the potential V = V(X) ilia polynomial in X, its matrix elements (n IVlk) are nonzero only for n in a finite neighborhood of k, and the recursion relations consist of finite sums which ean be solved exactly.

3.20

Calculation of Perturbation Series via Feynman Diagrams

The expectation values in formula (3.484) can be evaluated also in anot her way which can be applied to all potent ials which are simple polynomials pf x. T hen t he partition function can be expanded into a sum of integrals associated with certain Feynman diagrams . The procedure is rooted in t he Wick expansion of correlation functions in Sect ion 3.10. To be specific, we assume t he anharmonic potential to have the form (3.540)

The graphical expansion terms to be found will be typical for all so-called cp4. theories of quantum field theory.

283

3.20 Calculation of PerturbatiolJ Series via Feynman Diagrams

To calculate the free energy shift (3.484) to first order in 9, we have to evaluate t he harmonic expectation of A int ,,,. This is written as (A;o~,c}.., =

'4 ior"fJ dT(X 4 (Tn..,·

9

(3.54 1)

T he integrand contains the correlation function

(X(TJ)X(TZ)xh)x(T4))..,

=

G~4J(Th TZ, 1'3 , 1'4)

at identical time arguments. According to the Wick rule (3.302), this can be expanded into the sum of t hree pair terms

G~l(TJ, Tz)G~j(T3' 1'4)

+ G~j(TJ, T3)G~~(T2' 1'4) + G~2j(T!, T4)G~l(T2' 1'3),

where G~l(T, 1") are the periodic Euclidean Green functions of the harmonic oscillator [see (3.301) and (3.248)]. T he expectation (3.541) is therefore equal to the integral

,',( 1', l' )' . '4 ioI"~ dTG..,~

(Ant,c ).., -- 3 9

(3.542)

The right-hand side is pictured by the Feynman diagram

CD 3

Because of its shape this is called a two-loop diagram. In general, a Feynman diagram consists of lines meeting at Pzjints called vertices . A line connecting two points represent-s the Green function G..,21(T}, TZ). A vertex indicates a factor 9/4n and a variable l' to be integrated over the interval (0, h{3) . The present simple diagram has ollly olle point, and the T-arguments of the Green functions coincide. The number underneath counts how often the int egral occurs. It is called the multiplicity of t he diagram. To second order in V(x), the harmonic expcct ation to be evaluated is (3.543)

The integral now contains the correlation fu nction G~lh , . .. ,1'8) with eight time arguments. According to the Wick rule, it decomposes into a sum of 7!! = 105 products of four Green functions G~21(T, 1") . Due to t he coincidence of the time arguments, there are only three different types of contributions to the integral (3.543):

,',( TZ,T2 )G..,. ,',(1'2,1'1 )'G.."TJ,TI ",( ) ( ') (9)' '4 ioI"~ dTZ io['" dTI [72G",2 A int,e'" =

+24G~21(Tz, Td 4 + 9G~1(T2' T2)2G~21h, Td 2].

(3.544)

The integrals are pict ured by the following Feynman diagrams composed of three loops:

284

3 External Sources, Correlations, and Perturbation Theory

coco 24

72

9

They contain two vertices indicating two integration variables 1"1> 1"2. The first two diagrams with the shape of t hree bubbles in a chain and of a wat ermelon, respectively, are connected diagrams, the third is disconnected. When going over to the cumuiant {Afnt,e)w,,;, the d isconnected d iagram is eliminated. To higher orders, the counting becomes increasingly tedious a nd it is worth developing computer-algebraic tochniqucs for this purpose. Figure 3.7 shows t he diagrams for the free-energy shift up to four loops. The cumulants eliminate precisely all disconnected diagrams. This diagram-rearranging property of the logarithm is very general and happens to every order in g, as can be shown with the help of functional differential equations.

~-~(o;p + g) ,. +M

Cf:rD + V' + § 2592

1728

""

+

© )+ m,

Figure 3.7 Perturbation expansion of free energy up to order

rI

(four loops).

The lowest-order term {3Fw containing the free energy of t he harmonic oscillator [,ecrul Eq,. (3. 242) and (2.524)[

Fw

=

1 log ( 2 sinh -Pt~) P 2-

(3.545)

is often represented by the one-loop diagram (3.546)

With it, the graphical expansion in Fig. 3.7 starts more systematically with one loop rather than two. The systematics is, however, not perfect since t he line in t he one-loop diagram does not show that integrand contains a logarithm. In addition, the line is not connected to any vertex. All r-variables in t he diagrams are integrated out. The diagrams have no open lines a nd arc called vacuum diagrams . The calculation of t he diagrams in Fig. 3.7 is simplified with the help of a factorizat ion property: If a diagram consists of two subdiagrams touching each other

285

3.20 Calculation of Perturbatioll Series via Feynman Diagrams

at a single vertex, its Feynman integral factori:t.es into those of the subdiagrams. Thanks t o this property, we only have to evaluate the following integrals (omitting the factors g/41i for each vertex)

0 0

0 g g

~

~

fo"/3dTG~}(T' T)

= h{3a 2,

l"'oloMdT1dT2 G",,(TI,T2) =: 1i{3-1 a:;., l"'olo"'ln"PdTldT2dT3 G""(Tl,T2)G",,(T2,T3)G,,,,(T3,Td (2)

2

0

~

0

4

W

(2)

0

(2)

(2)

- h{3 (~f a~, ~

l"'o ln"PdTldT2 G ,(TI,T2) (2) '"

0

~

4

=: h{3-1 a82, w

1"10"10"/ 3dTldT2dT.~ G",1(TI,T2)G""(T2,T3)G,,,, (T3,Td o 0

0

(2)

(2)

(2)

3

- h{3 (~f a~O,

@

~

1h{11"1oh{1 o 0 0 dT1dT2dT3 0 Z"I} > = Z - lli]OjtTJ) { [Wj(TJ)j(TOJ) = Z - Ili] {Wj {1'l.)j{TOJ)jh)

+ Wj{TOJ) Wj(TJ)] Zli]}

(3.580)

+ (Wj('l)Wj{TOJ)jh) + Wj (TOJ) Wj {1'l.)j{TJ)

+ Wj {n)Wj (1l) Wi (TJ) } Zb] + [Cii)(Td Ci2)(T2, T3) + 2 perm] + C?)(T])C~I)(T2)Cii)(T3)' + Wj {TJ) Wj ('l)j{1l»)

Gi3 ) (TJ, T2, T3)

=

and for the four-point function C(4) (Til " . , 74) =

ci4) (7J, ... , T4) + [C~3) (TI' T2 , TJ) Gi]) (T4) + [c i 2) (TJ, T2) c i 2) (T3, T4) + 2 perm] + [ci2) (TJ, T2) ci1)(T3)Cil )(T4)

+

+ 3 perm]

+ 5 perm]

Gil )(TJ) ... cil)(T4)'

(3.581)

In the pure x 4 -theory there are no odd correlation functions , because of the symmetry of the potential. For t he general correlation function G{n} , t he total number of terms is most easily retrieved by dropping all indices and differentiating with respect to j (the arguments TI, ... , Tn of the currents arc again suppressed) : =

= =

G(4) =

e-W(eW)j = Wi = Gil) e- W (e W ) = W .. + W ·2 = G(2) + C{l}2 e- (e jij = W· ·· + 3W··W· + W·3 = C(3) + 3G(2)C{l) + G{l)3 1JJJ cc e- W (e W ) J]]J .... = Wjjjj + 4WjjjWj + 3Wj / + 6W W/ + W/ jj

W

W

)

JJJCC

11J

C

= G~4)

+ 4Gi3)C~I) + 3Gi +

C

jj

2 )2

6Gi2 )Gil )2

+ G?}4.

(3.582)

All equations follow from the recursion relation G(n) = G~n- I)

+ G(n- i)G~I),

n 2:

2,

(3.583)

if one uses G~n - l ) = Gin) and t he initial relation G( l ) = G~I). By comparing the first four relations with the explicit expressions (3.579)- (3.581) we see that the numerical factors on the right-hand side of (3.582) refer to the permutations of the arguments TI , T2, T3, ... of otherwise equal expressions. Since there is no problem in

294

3 External Sources, Correlations, and Perturbation Theory

reconstructing the explicit permutations we shall henceforth write all composition laws in the short-hand notation (3.582) . The formula (3.582) and its generalization is often referred to as cluster decomposition, or also as the cumulant expansion, of t he correlation functions. We can now prove that the connected correlation functions collect precisely all connected diagrams in t he n-point functions. For this we observe that the decomposition rules can be inverted by repeatedly differentiating both sides of t he equation WIi] = log ZIJ] fWlctionally with respect to the current j; c(1)

=

a(2) _ C(l)C(i) e(3) _ 3C(2)C(1)

+ 2C(I)3

C(4j _ 4G(3)G(i)

+ 12C(2)C(I)2 _

3C(::!)::! _ 6C(I}4.

(3.584)

Each equation follows from t he previous one by one more derivative with respect to j, and by replacing the derivatives on the right-hand side according to the rule

(3.585)

Again the numerical factors imply different permutations of thc arguments and the subscript j denotes functional differentiations with respect to j. Note that Eqs. (3.584) for t he connected correlation functions arc valid for symmetric as well as asymmetric potentials V ex). For symmetric potentials, the equations simplify, since all terms involving G{l) = X = (x) vanish. It is obvious that any connected diagram contained in G(n) must also be contained in Gin), since all the terms added or subtracted in (3.584) are products of Gjn)s, and thus necessarily disconnected. Together with the proof in Section 3.22.1 that the correlation functions Gin) contain only t he connected parts of G("), we can now be sure that Gi") contains precisely the connected parts of G(n).

3.22.3

Functional Generation of Vacuum Diagrams

The functional differential equation (3.570) for Wfj ] contains all information on the connected correlation functions of t he system. However, it does not t ell us anything about the vacuum diagrams of the theory. These are contained in WIO] , which remains an undetermined constant of functional integration of these equations. In order to gain information on the vacuum diagrams, we consider a modificat ion of t he generating functional (3.558), in which we set the external source j equal to zero, but gencralize the source j(r) in (3.558) coupled linearly to x(r) to a bilocal form K(r,r') coupled linearly to x(r) x (r'): ZIK] =

J

Vx(r)e - A.,[z,Kl,

(3.586)

where Ae [x, K ] is the Euclidean action

A.[x,K]

=A.[x]+ A;"· [x]+ ~J dr J

dr'x(r)K(r,r')x(r).

(3.587)

295

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

When forming t he functional derivative with respect to K(7, 7') we obtain the correlation function in the presence of K(7, 7'): ;Z G"'( 7,7') = -2Z - ' [K ] OK(7,T')'

(3.588)

At the end we shaH set K(7,7') = 0, just as previously the source j. When differentiating Z[K] twice, we obtain the four-point function (3.589) As before, we introduce the functional W[K] == 10gZ[K]. Inserting this into (3.588) and (3.589), we find

(3.590) (3.591)

With the same short notat ion as before, we shall use again a subscript K to denote functional differentiation with respect to K, and write (3.592)

From Eq. (3.582) we know that in the absence of a source j and for a symmetric potential, G(4) has the connectedness structure (3.593)

This shows t hat in contrast to W jjjj , the derivative W KK does not directly yield a connected four-point function, but two disconnected parts: (3.594)

the two-point functions being automatically connected for a symmetric potential. More explicitly, (3.594) reads 40'W

OK(7J> 72)oK(73 , 74) = G~4)h, 72, 73, 74)

+ G~2)(7J, 73)G~2)(72' 74) + G~2)h , 74)CF)(72, 73) ' (3.595)

Let us derive functional differential equations for Z [K ) and W[K). By analogy with (3.560) we start out with t he trivial functional differential equation

I

VX x(T)_'- c A.I. 1 (3.600) Go ZK + 3ZKK = 2Z. Inserting Z[K] = e W1KI, this becomes (3.601) It is useful to reconsider the functional W(KJ as a functional W[Go]. JG%K = Gl, a nd t he derivatives of W[KJ become

Then

(3.602)

and (3.601) takes the form A( , , , ) 1 GoWGo + 3" GoWGoGo + 2GoWGo + GoWGo WGo = 2'

(3.603)

This equation is represented diagra mmatically in Fig. 3.9. The zeroth-order solution

Figure 3.9 Diagrammatic reprcscntation of funct ional differential equation (3.603). For t he purpose of finding the multiplicities of the diagrams, it is convenient to reprcscnt here by a vertcx the coupling strength - )"/4! rather than g/4 in Section 3.20.

297

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

to this equation is obtained by setting ,\ = 0:

W (O' (Go] ~

1 ,TIlog(Go).

(3.604)

Explicitly, the right-hand side is equal to the one-loop contribution to the free energy in Eq. (3.546) , apart from a factor -{J. T he corrections are found by iteration. For systematic treatment, we write W [Gol as a sum of a free and an interacting part,

(3.605) insert this into Eq. (3.603) , and find the differential equation fo r the interacting part:

(3.606) This equation is solved it eratively. Setting wint [Gol = 0 in all terms proportional to'\ , we obtain the first-order contribution to wint[Gol:

wint (G ] =3 - ).Gz 4!

o

(3.607)

o'

T his is precisely the contribut ion (3.542) of the two-loop Feynman diagram (apart from the different normalization of g). In order to sec how t he iteration of Eq. (3.606) may be solved systematically, let us ignore for the moment the functional nature of Eq. (3.606) , and treat Go as an ordinary real variable rather than a functional matrix. We expand W [Gol in a Taylor series:

wint [Gol =

f AW (~; )P (GO )2p,

(3.608)

p

p=lP,

.

and find for the expansion coefficients the recursion relation Wp+l = 4 { [2P (2 P - 1)

+ 3(2p)1Wp + ~ ( : )

Solving this with the initial number WI connected vacuum diagrams of pth order:

=

2q Wq x 2(P - q)Wp _ q } . (3.609)

3, we obtain the multiplicities of the

3, 96,9504, 1880064, 616108032, 301093355520, 205062331760640, 185587468924354560, 215430701800551874560, 312052349085504377978880.(3.610) To check these numbers, we go over to ZIG] = eWIGol, and find the expansion:

Z [Gol = exp

[~Tr log Go + ~ ~Wp (~t

D,,'I' (Go]

r

{GO)2P]

[1+ f: ~z, (-;)' (Go)"]. p=!

p.

4.

(3.611)

298

3 External Sources, Correlations, and Perturbation Theory

The expansion coeffi cients zp count the total number of vacuum diagrams of order p. T he exponentiation (3.611) yields zp = (4p - I) !!, which is the correct number of Wick contractions of p interactions X4. In fact , by comparing coefficients in t he two expansions in (3.611 ), we may derive another recursion relation for Wp:

which is fulfilled by the solutions of (3.609). In order to find the associated Fcynman diagrams, we must perform the differentiations in Eq. (3.606) functionally. The numbers Wp become then a sum of diagraJDl;, for which the recursion relation (3.609) reads

WI>+1= 4 [G~d~Z W p+ 3·Cg

ddn W

p+ ~(:) (d~Wq)G~ ' ~CdnWp-q)l ' (3.6 13)

where the differentiation dldn removes one line connecting two vertices in all possible ways. This equation is solved diagrammatically, as shown in Fig. 3.10.

Wp+1= 4[ G~d~~WP

+3

~ /nWp + ~(:)(/n Wq)G~.G~ (ddn Wp-q)]

Figure 3.10 Diagrammatic representation of recursion relation (3.609). A vertex repre-

sents the coupling strength - A. Starting t he iteration with WI = 3 CD , we have dWp/dn = 6 0 and iPw p/ d n 1 = 6 X. Proceeding to order five loops and going back to t he usual vertex notation -.x, we find the vacuum diagrams with their weight factors as shown in Fig. 3.11 . For more t han five loops, the reader is referred to t he paper quoted in Notes and References, and to t he internet address from which Mathematica programs can be downloaded which solve the recursion relations and plot all diagrams of W[O] and the resulting two- and four-point functions.

3.22.4

Correlation Functions from Vacuum Diagrams

The vacuum diagrams contain information on all correlation functions of the thcory. One may rightly say that t he vacuum is the world. The two- and four-point funct ions are given by t he functional derivatives (3.592) of the vacuum functional W[K]. Diagrammatically, a derivative with respect to K corresponds to cutting one line of a vacuum diagram in all possible way:;. T hus, all diagrams of the two.-point function

299

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

II

diagrams and mnlt iplicities

g'

300

g'

~( 1128 ©

i'

g'

~ (62208 0

+ 66296

165888~ +

~(24 §

+

3456@

+

+

qp +

124416

248832

o::xx:o

@

12

ceo

2592

I

)

CXXXJ

1728~

)

g+ ~+ cxxg + ~

+ 491664

+ 248832

+

165888

62208

248832

~

)

Vacuum diagrams up to five loops and their multiplicities. The total numbers to orders gn are 3, 96,9504, 1880064, respectively. In contrast to Fig. 3.10, and to the previous diagrammatic notation in Fig. 3.7, a vertex stands here for -Aj4! for brevity. For more than five loops see the tables on the internet

Figure 3.11

(http ://www .physik.fu- berlin/Nkleinert/b3/programs).

can be derived from such cuts, multiplied by a factor 2. As an example, consider the first-order vacuum diagram of W[K ] in Fig. 3.11. Cutting one line, which is possible in two ways, and recalling that in Fig. 3.11 a vertex stands for - ,\/41 rather than - A, as in the other diagrams, we find

C (Z)

(3.614)

The second equation in (3.592) tells us that all connected contributions to the four-point function C (4) may be obtained by cutting two lines in all combinations, and multiplying the result by a factor 4. As an example, take the second-order vacuum diagrams of W[O] with the proper translation of vertices by a factor 4! , which are (3.6 15)

Cutting two lines in all possible ways yields the following contributions to the connected diagrams of the two-point function: C(4} =

4 x (2.1 . ~ 16

+ 4 . 3 .~) ;()< 48

(3.616)

3 00

3 External Sources, Correlations, and Perturbation Theory

It is also possible to find all diagrams of the four-point function from the vacuum diagrams by forming a derivative of W[O] with respect to t he coupling constant - >., and multiplying the result by a factor 4!. This follows directly from the fact that this differentiation applied to Z[O] yields the correlation function f dr{x 4 }. As an example, take the first diagram of order rT in Table 3.11 (with t he same vertex convention as in Fig. 3.11 ):

W, [O[ ~

4~

©

(3.6 17)

Removing one vertex in the t hree possible ways and multiplying by a factor 4! yields C(4) =

3.22.5

1 4! x 48 3

U .

(3.618)

G e nerating Functional for Vertex FUnctions. Effective Action

Apart from the connectedness structure, the most important step in economizing the calculation of Feynman diagrams consists in the decomposition of higher connected correlation functions into one-particle irreducible vertex functions and one-particle irreducible two-particle correlation functions, from which the full amplitudes can easily be reconstructed. A diagram is called one-particle irreducible if it cannot be decomposed into two disconnected pieces by cutting a single line. There is, in fact , a simple algorithm which supplies us in general with such a decomposition. For t his purpose let us introduce a new generating functional fiX ], to be called thc effective action of t he theory. It is defined via a Lcgendre transform ation of WU ]:

- fiX [ = Wb[ - Wjj.

(3.619)

Here and in the following , we use a short-hand notation for the functional multiplication, Wjj = fd7Wj (7)j(7), which considers fields as vectors with a continuous index 7 . The new variable X is the functional derivative of WU ] with respect to j(T) [cecali (3.569)[, (3.620)

and thus gives t he ground state expectation of the field operator in t he presence of the current j. When rcwriting (3.619) as

- fiX [ = Wb[ - X j,

(3.62 1)

and fUllctionally differentiating this with respect to X, we obtain the equation

fx [X[

~ j.

(3.622)

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

301

This equation shows that t he physical path expectat ion X(7) = (X(7)), where t he ext ernal current is zero, extremizes the effective action:

f x iXi

~ O.

(3.623)

We shall study here only physical systems for which the path expectation value is a constant X(7) =: Xo. Thus we shall not consider systems which possess a timedependent XO(7), although such systems can also be described by x 4-theories by admitting more general types of gradient terms, for instance x({P - k~)2X. The ensuing 7-dependence of XO(7) may be oscillatory. 11 Thus we shall assume a constant

(3.624) which may be zero or non-zero, depending on the phase of the system. Let us now demonstrate that the effective action contains all the information on the proper vertex functions of the theory. These can be found directly from t he functional derivatives:

(3.625) We shall sec that the proper vertex functions are obtained from these functions by a Fouricr transform and a simple removal of an overall factor (211,)DJ (2:::=1Wi) to ensure moment um conservation. The functions r{n) (71, .. . , 7n ) will therefore be called vertex junctions , without t he adjective proper which indicates t he absence of the a-function. In particular, the Fourier transforms of t he vertex functions r (2) (71, 72) and r (1) (71, 72,73 , 74) are related to their proper versions by r (2)(Wl, W;!) = r (4}(Wl,W;!, WJ,W4) =

21rJ (WI 21rJ

+ W;!) r (2}(wd ,

(~ W;) r (4}(Wl,W;!, WJ,W4).

(3.626) (3 .627)

For the functional derivatives (3.625) we shall use t he same short-hand notation as for the functional derivatives (3.569) of WliJ, setting

(3 .628) The argumcnt s 71, .. . ,7n will usually be suppressed. In order to derive relations between the derivatives of the effective action and the connected correlation functions, we first observe t hat the connected one-point function Gil) at a nonzero source j is simply the path expectation X [recall (3.578)]:

(3.629) 17In higher dimensions t here can be crystal- or qua.sicrystal-likc modulations. Sec, for example, H. Kleinert and K. Maki, Fortscl!r. P hys. 29, I (1981) (bttp;f/""".pbyai k.fu-berlin.der kleinert/75). T hill paper was the first to investigate in detail icosahedral quasicrystal line structures discovered later in aluminum.

3 02

3 External Sources, Correlations, and Perturbation Theory

Second , we see that the connected two-point function at a nonzero source j is given by 0(2) = c

C(l) = ]

W . , = oX =

oj

]1

(OJ )-1= r-I oX xx-

(3.630)

T he inverse symbols OIl the right-hand side arc to be understood in the functional sense, i.e., rxi denotes the functional matrix:

8'r

r X(~)X(y) == [JX(T)JX (r')

]-'

(3.631)

which satisfies

(3.632)

Relation (3.630) states that the second derivative of the effective action determines directly the connected correlation function G~:l)(w) of the interacting theory in the presence of t he external source j. Since j is an auxiliary quantity, which eventually be set equal to zero t hus making X equal to X o, the actual physica1 propagator is given by

a("C 1 ~ r xx -'IX = X o . j=O

(3.633)

By Fourier-transforming th b relation and removing a a-function for the overall momentum conservation, the full propagator Gw 2(W) is related to the vertex function r (2)(w), defined in (3.626) by _ _ ("(

Gw 2(W ) = G

_ 1 k) - f(2)(W)'

(3.634)

T he t hird derivative of the generating functional Wli] is obtained by functionally differentiating Wjj in Eq. (3.630) once more with respect to j , and applying t he chain rule: W jjj =

-rx~rxxx ~~

=

-rx~rxxx = _G~2)3 rxxx .

(3.635)

This equation has a simple physical meaning. The third derivative of Wfj] on the left-hand side is the full t hree-point function at a nonzero source j, so that (3.636)

This equation states that the full three-point function arises from a t hird derivative of r[X] by attaching to each derivation a full propagator, apart from a minus sign. We shall express Eq. (3.636) diagrammatically as follows:

3 03

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

where II

,, -1

~

I

·: - C{")

:

'~

2

denotes the connected n-point fu nction, and

n~

1~2

... - - [ XI ... X"

the negative n-point vertex function. For the general analysis of the diagrammatic content of the effective action, we observe that according to Eq. (3.635), the functional derivative of the correlation function G with respect to t he current j satisfies (3.637)

This is pictured diagrammatically as follows: (3.638)

T his equation may be differentiated furt her with respect to j in a diagrammatic way. From the defin ition (3.557) we deduce the trivial re 2. By applying 6j6j repeatedly to the left-hand side of Eq. (3.637), we generate all higher connected correlation functions. On t he right-hand side of (3.637), the chain rule leads to a derivative of all correlation functions G = G12) with respect to j, thereby changing a line into a line with an extra t hree-point vertex as indicated in the diagrammatic equation (3.638). On the ot her hand, t he vertex function rxxx must be differentiated with respect to j. Using the chain rule, we obtain for any n-point vertex function: (3.640)

3 04

3 External Sources, Correlations, and Perturbation Theory

which may be represented diagrammatically as

n

n- l

t 1~2 "rh, .

n+~~ l

1~2

With these diagrammatic rules, we can differentiate (3.635) any number of times, and derive the diagrammatic structure of the connected correlation functions with an arbitrary number of external legs. The result up to n = 5 is shown in Fig. 3.12.

$YMDOLS

COMP OSITION

=i;

:0: -C~.,~

I 14 p'nn )

+ (~19P"m) + ~ Fig ure 3 .12 Diagrammatic differentiations for deriving t ree decomposition of connected correlation functions. The last term in each decomposition yields, after amputation and removal of an overall .,-function of momentum conservation, precisely all one-particle irreducible diagrams.

The diagrams generated in t his way have a t ree-like structure, and for t his reason they are called tree diagrams. The tree decomposition reduces all diagrams to their one-particle irreducible contents.

3.22 Generating Functiollal of COllllectoo Correlation Fullctions

305

The effective action r[X ] can be used to prove an important composition theorem: T he full propagator G can be expressed as a geometric series involving the so-called self-eneryy. Let us decompose the vertex function as f(2)

--

a-' 0 +

rint

xx,

(3.641)

such that the full propagator (3.633) can be rewritten as

G = (1 + Gor~\-

r

1

Go.

(3.642)

Expanding t he denominator, this can also be expressed in the form of an integral equation:

G = Go - Gof'~\:Go + Go f'~\-Gof'~\:Go - ...

(3 .643)

The quantity -f~~ is called the self-energy, commonly denoted by E: (3 .644)

i.e., the self-energy is given by the interacting part of t he second functional derivative of t he effective action, except for t he opposite sign. According to Eq. (3.643) , all diagranls in G can be obtained from a repetit ion of self-energy diagrams connected by a single line. In terms of E, the full propagator reads, according to Eq. (3.642): (3 .645)

This equation can, incidentally, be rewritten in the form of an integral equation for the correlation function G:

G = Go + GoEG. 3.22.6

(3.646)

Ginzburg-Landau Approximation to Generating Functional

Since the vertex functions are the functional derivatives of the effective action [sec (3.625)], we can expand the effective action into a functional Taylor series

r[X]

~ f: ~ n=on.

Jdr, ... dr. r "'(r" ... , r.)X(r,) ... X(r.).

(3.647)

The expansion in the number of loops of the generating functional r[X ] collects systematically the contributions of fluctuations. To zeroth order, all fluctuations are neglected, and the cffective action reduces to the initial action, which is t he mean-field approximation to the effective action. In fact, in the absence of loop diagrams, t he vertex fu nctions contain only the lowest-order terms in r {2) and r (4) :

r&2)(T"T2) r &4)(T"T2 , T3 ,T4) =

M(-a~ +W2)J(Tl-T2),

(3.648)

AJ(Tl - T2)J(TI - T3)J(Tl - T4) .

(3.649)

306

3 External Sources, Correlations, and Perturbation Theory

Inserted into (3.647), this yields the zero-loop approximation to r[X ]: (3.650)

This is precisely the original action functional (3.559). By generalizing X(T) to be a magnetization vector field, X(r) --+ M (x ), which depends on the three-dimensional space variables x rather t han the Euclidean t ime, t he functional (3.650) coincides with the phenomenological energy functional set up by Ginzburg a nd La ndau to describe t he behavior of magnetic materials near the Curie point, which they wrote

os"

(3.65 1)

The use of this functional is also referred to as mean-field theory or mean-field approximation to the full theory.

3 .22.7

Composite Fields

Sometimes it is of interest to study also correlation functions in which two fields

coincide at one point , for instance (3.652)

If multiplied by a facto r Mw 2, the composite operator Mw 2x 2(r)/2 is precisely the frequency term in the action energy functional (3.559) . For this reason one speaks of a frequency insertion, or, since in the Ginzburg- Landau action (3.651) the frequency w is denoted by t he mass symbol m , one speaks of a mass insertion into the correlation function c(n)(rl>' .. , rn). Actually, we shall never make use of the full correlation function (3.652), but only of thc integral over r in (3.652) . This can be obtained directly from the generating functional Zlil of all correlation fu nctions by differentiation with respect to t he square mass in addition to the sourcc tcrms

Z - l () J J zr ~ll jdTG(I,n)(T,T" .. . , T) n - M8w2 Jj(rd ... Jj(rn) lJ j = o ' By going ovcr to the generating functional connected parts:

Wli J,

- __ a___, _ jdTGc(I ,n)(T"T, .. . , T) n M8w2 Jj(rd ISL.O. Landau, J. E.T.P. 7, 627 (1937).

(3 .653)

we obtain in a similar way the

-'-W "ll

... Jj(rn}

lJ j=o'

(3.654)

3.23 Path Integr81 C81culatioll of Effective Action by Loop ExpllJlsion

307

The right-hand side can be rewritten as

J

dT cil ,n) (T, TI, . .. ,Tn) = -

M~2Cin)(TJ,'"

, Tn) .

(3.655)

The connected correlation funct ions Gil ,n)(T, TI, . . . ,Tn ) can be decomposed into tree diagrams consisting of lines and one-particle irreducible vertex functions r (1,n)(T, TI," " Tn). If integrated over T, these are defined from Legendre transform (3.6 19) by a further differentiation with respect to Mw 2: (3.656)

implying the relation

J

dT

3.23

r (I,n)(T, TI,

. . . ,Tn) = - M:w 2 r

Cn) (Tl, .. .

, Tn) .

(3.657)

Path Integral Calculation of Effective Action by Loop Expansion

Path integral;; givc the most direct access to the effective action of a thcory avoiding the cumbersome Legendre transforms. The derivation will proceed diagrammaticruly loop by loop, which will turn out to be organized by the powers of the P lanck constant h. This will now be kept explicit in all formulas. For later applications to quantum mechanics we shall work with real time.

3.23.1

General Formalism

Consider the generating functional of all Green functions ZUJ ~ "w Olf ' ,

(3.658)

where Wlil is the generating functional of all C()nnected Green functions. The vac· uum expectation of the field , the average X (') " (x(t)),

(3.659)

is given by the first functional dcrivative

XI')

~ ;W Ul/;j(t).

(3.660)

This can be inverted to yield j(t) as a functional of X(t): j(t)

~

j [X](t),

(3.661)

which leads to the Legendre transform of Wlil:

r[XJ " W UJ -

Jd'j(t)X('),

(3.662)

308

3 External Sources, Correlations, and Perturbation Theory

where the right-hand side is replaced by (3.661 ). T his is the effective action of the theory. The effective action for time independent X (t) == X defines the effective potential

V "" (X)

=-

-1- fiX ]. (3.663) tb - t" The first functional derivative of the effective action gives back the current 'fiX] . H(.) ~ - J(t).

(3.664)

The generating functional of all connected Green functions can be recovered from the effective action by the inverse Legendre transform

WU] ~ fi X ] +

Jdtj(t)X(t).

(3.665)

We now calculate these quantities from the path integral formula (3.558) for the generating functional Zlil:

Zlil =

J

V x(t)e(iJh){A[a:1 + jd!j(t)z(t) }.

(3.666)

With (3.658) , t his amounts to the pat h integral formula for r[X):

J

V x(t) e(i/h){ A(zl+ f dtj(tlz(tl }.

eHqxl+f dt;(t)X(t)} =

(3 .667)

The action quantum fl is a measure for the size of quantum fluctuat ions. Under many physical circumstances, quantum fluctuations arc small, which makes it desirable to develop a method of evaluating (3.667) as an expansion in powers of Ii.

3.23.2

Mean-Field Approximation

For Ii ----t 0, the path integral over the path x(t) in (3.666) is dominated by the classical solution Xcl(t) which cxt remizes t he exponent

M ix] I ox(t)

_ -j(t)

%=", 0, we perform the fluctuation expansion around the minimum of the potential (3.663) at X = 0, where the two Green functions coincide, both having t he same frequency w: (3.690)

For w2 < 0, however, where the minimum lies at the vector Xn of length (3.680), they are different: (3.691)

Since the curvature of the potential at the minimum in radial direction of X is positive at the minimum, the longitudinal part has now the positive frequency _ 2w 2 . The movement along the valley of the minimum, Oil the other hand, does not increase the energy. For this reason, the transverse part has zero frequency. This feature, observed here in lowest order of the fluct uation expansion, is a very general one, and can be fou nd in the effective action to a ny loop order. In quantum field t heory, there exists a theorem asserting this called Nambu-Coldstone theorem. It states t hat if a quantum field t heory without long-range interactions has a continuous symmetry which is broken by a nonzero expectation value of t he field corresponding to the present X [recall (3 .659)]' then the fl uctuations transverse to it have a zero

3 12

3 External Sources, Correlations, and Perturbation Theory

mass. They are called Nambu-GoldBtone modes or, because of their bosonic nature, Nambu-Goldstone bosoM. T he exclusion of long-range interactions is necessary, since these can mix with the zero-mass modes and make it massive. This happens, for example, in a superconductor where they make the magnetic field massive, giving it a finite penetration depth, the famous Meissner effect. One expresses this pictorially by saying t hat the long-range mode can eat up the Nambu-Goldstone modes and become massive. The same mechanism is used in elementary particle physics to explain t he mass of the W ± and ZO vector bowns as a consequence of having eaten up a would be Nambu-Goldstone boson of an auxiliary Higgs-field theory. In quantum-mechanical systems, however, a nonzero expedation value with t he associated zero frequency mode in the transverse direction is found only as an artifact of perturbation theory. If all fluctuation correction;; are summed, the minimum of the effective potential lies always at the origin. For example, it is well known, that the ground state wave functions of a particle in a double-well potential is symmetric, implying a zero expectation value of the particle position. This symmetry is caused by quantum-mechanical tunneling, a phenomenon which will be discussed in detail in Chapter 17. This phenomenon is of a nonperturbative nature which cannot be described by an effective potential calculated order by order in the fluct uation expansion. Such a potential does, in general, posses a nonzero minimum at some Xo somewhere near the zero-order minimum (3.680). Due to this shortcoming, it is possible to derive the Nambu-Goldstone theorem from the quantum-mechanical effective action in the loop expansion, even t hough the nonzero expectation value Xo assumed in the derivation of the zero-frequency mode does not really exist in quantum mechanics. The derivation will be given in Section 3.24 . The use of the initial action to approximate the effective action neglecting corrections caused by the fl uctuations is referred to as mean-field approximation.

3.23.3

Corrections from Quadratic Fluctuations

In order to find the first /i-correction to the mean-field approximation we expand the action in powers of the fluct uations of the paths around t he classical solution (3.692)

6x(t) " x(t) - Xcl(t),

and perform a perturbation expansion. The quadratic term in 8x(t) is taken to be the free-part icle action, the higher powers in 8x(t) are the interactions. Up to second order in the fl uctuation;; 8x(t), the action is expanded as follows:

Jdtj(t) [Xcl(t) +6x(t)[ A[xcl] + Jdtj(t)Xcl(t) + Jdt { j(t) + 6!1,) I . ~J 6x(t) + Udtde 6x(t) 6X(~:'(t') I..... 6x(t') + ((6x)') .

A [xcl +6x[ +

0

(3.693)

3.23 Path Integr81 C81culatioll of Effective Action by Loop ExpllJlsion

313

The curly bracket mult iplying t he linear terms in the variation ox(t) vanish due to the extremality property of the classical pat h X ci expressed by the equation of motion (3.668). Inserting this expansion into (3.667) , we obtain the approximate expression

ZIi]

~ e(i/lIl {Aj"" I+ j dtj(tl%,,(ll } jVoxexp { ~ jdtdti ox(t) Ii

I

oZA OX(t') } . 8x(,)8x(") " -%01 _

(3.694) We now observe that the fluctuations Ox(t ) will be of average size.,fh due to the /i-denominator in the Fresnel exponent. Thus the fluctuations (Jx)n are of average size .,fhn. The approximate path integral (3.694) is of the Fresnel type and my be integrated to yield e(i/hl {A j%ol l+ j dtj(t)"

t",

) log{2"'in [w(t, - toll /Mw }.

(3.783)

If the boundary conditions are periodic, so that the analytic continuation of the result can be used for quantum statistical calculations, the result is

V'ff(X) ~ V(X) -

(tt>

' ) log{2"in [w(t, - to)/2[}. t"

(3.784)

It is important to keep in mind that a line in the above diagrams contains an infinite series of fu ndamental Feynman diagrams of the original perturbation expansion, as can be seen by expanding the denominators in the propagator go.l> in Eqs. (3 .686)- (3.688) in powers of X 2 . This expansion produces a sum of d iagrams which can be obtained from the loop diagrams in t he expansion of the trace of the logarithm in (3.707) by cutting the loop. If t he potential is a polynomial in X , the effective potentia l at zero temperature can be solved most efficiently t o high loop orders with t he help of recursion relations. T his is shown in detail in Appendix 3C.5 .

3 .24

Nambu- Goldstone Theorem

T he appearance of a zero-frequency mode as a consequence of a nonzero expecta· t ion value X can easily be proved for any continuous symmetry and to all orders in perturbation theory by using the full effective action. To be more specific we consider as before the case of O(N)-symmetry, and perform infinit esimal symmetry transformations on the currents j in the generating functiona l WU]: (3.785)

where Lcd a re the N(N - 1)/2 generators of O(N)·rotations with t he matrix elements (3.786)

and £1> are the infinitesimal angles of the rotations. Under these, the generating functional is asSllmed to be invariant: (3.787)

Expressing the integrand in terms of Legend re-transformed quantities via Eqs. (3.620) a nd (3.622), we obtain (3.788)

3 25

3.24 Nambu-GoldstollC Tlworcm

This expresses t he infinitesimal invariance of the effective action fiX] under infinitesimal rotations

The invariance property (3.788) is called the Wam-Takakashi identity for the functional r[X ]. It can be used to find an infinite set of equally named identities for all vertex functions by forming all r[X ] functional derivatives of r[X ] and setting X equal to the expectation vaJue at the minimum of r[X]. The fi rst derivative of r[X ] gives directly from (3.788) (dropping t he infinitesimal parameter fed)

(3.789)

Denoting the expectation value at the minimum of the effective potential by X , this yields

I ,- ,

J'r[X]

dt Xo,(t) (L",)o" OX (t'),X (t) b

a

I

_ ~ O.

(3.790)

X(/) = X

Now the second derivative is simply the vertex function r(2) (t' , t) which is the functional inverse of the correlation function C(2)(t' , t). The integral over t selects the zero-frequency component of the Fourier transform (3.791)

If we define the Fourier components of r (2)(t' , t) accordingly, we can write (3.790) in Fourier spacc as (3.792)

Inserting the matrix elcments (3.786) of the generators of the rotations, this equation shows that for X =F 0, the fully interacting transverse propagator has to possess a singularity at w' = O. In quantum field theory, this implies the existence of N - 1 massless particles, t he Nambu-Goldstone boson. The conclusion may be drawn only if there are no massless particles in the theory from the outset, which may be "eaten up" by the Nambu-Goldstone boson, as explained earlier in the context of Eq. (3.688) . As mentioned before at the end of Subsection 3.23.1 , the Nambu-Goldstone theorem does not have any consequences for quantum mechanics since fluctuations are too violent to allow for the existence of a nonzero expectation value X . The effective action calculated to a ny finite order in perturbation theory, however, is incapable of reproducing this physical property and does have a nonzero extremum and ensuing transverse zero-frequency modes.

3 26

3.25

3 External Sources, Correlations, and Perturbation Theory

Effective Classica l Potential

T he loop expansion of the effective action f[X ] in (3.768), consisting of the trace of the logarithm (3.702) and the one-particle irreducible diagrams (3.741), (3.781) and the associated effective potential V(X) in Eq. (3.663), can be continued in a straightforward way to imaginary times setting tb - ta -+ - ihi3 to form the Euclidean effective potential r e[X]. For the harmonic oscillator, where the expansion stops after the trace of t he logarithm and the effective pot ential reduces to t he simple expression (3.782), we find the imaginary-time version (3.793)

Since the effective action contains the effect of all fl uctuations, the minimum of the effective potential V(X) should yield directly the full quantum statistical partition functio n of a system: (3.794) z ~ exp(- pV(X) 1 ). moo Inserting the harmonic oscillator expression (3.793) we find indeed the correct result (2.405). For anharmonic systems, we expect t he loop expansion to be able to approximate V(X) rather well to yield a good approximation for the partition function via Eq. (3.794). It is easy to realize that this cannot be true. We have shown in Section 2.9 that for high temperatures, the partition function is given by the integral (,ecall (2.351)) (3.795)

T his integral can in principle be t reated by the same background field method as the path integral, albeit in a much simpler way. We may write x = X + Jx and find a loop expansion for an effective potential. T his expansion evaluated at the extremum will yield a good approximation to the integral (3.795) only if the potential is very close to a harmonic one. For any more complicated shape, the integral at small i3 will cover the entire range of x and can therefore only be evaluated numerically. T hus we can never expect a good result for the partition function of anharmonic systems at high temperatures, if it is calculated from Eq. (3.794). It is easy to find the culprit for this problem. In a one-dimensional system, the correlation functions of the fluctuations around X are given by the correlation function [compare (3.301), (3.248), and (3.687)] (JX(T)Jx(r)) = cgl{x)(T, T') = !Cb2{X),,,(T - T') h 1 coshll(X)((T - ,.' ( - hP/2) M 21l(X) 'inh(Il(X)hpj2)

(T -

,.'1E [0, hPJ,

(3.796)

3.25 Effective Classical Pot.ential

3 27

with the X -dependent frequency given by (3.797)

At equal times

7 = 7' ,

this specifies the square width of the fluctuations OX(7):

( [OX(T)]

')" 1 O(X)"P ~ M20(X) coth 2 .

(3.798)

The point is now that for large temperatures T, t his width grows linearly in T ')

(

[OX(T) ]

kBT

T _ oo

~ MO'.

(3.799)

T he linear behavior follows t he historic Dulong-Petit law for the classical fl uctuation width of a harmonic oscillator [compare with the Dulong-Petit law (2.601) for t he thermodynamic quantities]. It is a dircct consequence of the equiparlition theorem for purely thermal fluctuations, according to which the potential energy has an average keTj2: (3.800)

If we consider the spectral representation (3.245) of the correlation function ,

~ 1 e- ;"'",("- 1"') G'O',c (7 _ 7 ') -_ ~ liP L 2 , fJ m ,,",_oow + !v

(3.801 )

m

we see that the linear growth is entirely due to t erm with zero Matsubara frequency. The important observation is now that if we remove this zero frequency term from the correlation function and form t he subtmcted correlation function [recall (3.250)] (3.802)

we see that the subtracted square width (3.803) decrease for large T . This is shown in Fig. 3.14. Due to this decrease, there exists a method to substantially improve perturbation expansions with the help of the so-called effective classical potential.

3.25.1

Effective Class ical Boltzmann Factor

T he above considerations lead us to the conclusion that a useful approximation for partition function can. be obtained only by expanding the path integral in powers of

328

3 External Sources, Correlations, and Perturbation Theory

2

unrestricted ,/ " fluctuation width :::V'"

1. 5

,/

\

'"

".- '/' ' and yields {,OO

1

2sinh(,81i.w/2)

dA

L;oo 21Ti exp

(2Mw2{3 A')

1

=

1

2sinh(.8hwj2) lc(hfJ) wn/3.

(3.819)

Inserting t his into (3.816) we obtain the local Boltzmann factor

T he final int egral oyer Xo in (3.805) reproduces the correct partition function (2.407) of the harmonic oscillator.

3.25.2

Effective Class ical Hamiltonian

It is easy to generalize t he expression (3.813) to phase space, where we define t he effective classical Hamiltonian Heffcl(po,xo) a nd the associated Boltzmann factor B(Pu, xo) by t he path int egral

B(po, xo) == exp [_/3H effd (po,XO)] ==

f vx f ~:n J(xo - x)27rM(po - p)

e-A.[p,z)/h,

(3.821) where x = fo"PdTX(T)/ft/3 and p = fo"PdTp(r)/ft/3 are the temporal averages of position and momentum, and A[P, xl is the Euclidean action in phase space

A,fp,x] ~

10r" dT ]-;p(T);'(T) + Hfp(T) ,x(T))].

(3.822)

The full quantum-mechanical partition function is obtained from t he classicallooking expression [recall (2.344)] (3.823)

3.25 Effective Classical Pot.ential

331

The definition is such that in the classical limit, Hclfcl(Pu, xo)) becomes the ordinary Hamiltonian H(Pu,xo) . For a harmonic oscillator, the effective classical Ha miltonian ca n be directly deduced from Eq. (3.820) by "undoing" the Pu-integration :

B ,- x ) == e-H~" e'l.Po,z")lkHT w\Pl»

0

=

I (hP) f3tiw/ 2 e-P~/2M+Mw~zl) e sinh(f3tiw/2) .

(3.824)

Indeed , insert ing this into (3.823), we recover t he harmonic partition function (2.407) . Consider a particle in t hree dimensions moving in a constant magnetic fie ld B along t he z-axis. For the sake of generality, we allow for an additional harmonic oscillator centcred at t he origin with frequencies wil in z-direction and W.L in t he xy-plane (as in Section 2.19). It is t hen easy to calculate the effective classical Boltzmann factor for the Ha miltonian [recall (2 .687)] (3.825)

where lA p, x ) is t he z-componcnt of the angular momentum defined in Eq. (2.645). We have shifted the center of momentum integration to Po, for later convenience (see Subsection 5.11.2) . Thc vector x.L = (x, y) denotes t he orthogonal part of x. As in t he generalized magnetic field action (2.687), we have chosen different frequenc ies in front of t he harmonic oscillator potential and of the term proportional to lz , for generality. The effective classical Boltzmann factor follows immediately from (2.701 ) by "undoing" the momentum integrations in PZ ,PII' a.nd using (3.824) for t he motion in the z-direction:

B(Po Xo )=e-PH."e'(po,X(I )=l3(hf3) hf3w+/ 2 hf3w_/ 2 hf3wlI/2 e - PIl( pO,X(I ) , e sinh hf3w+/ 2 sinh hf3w_/ 2 sinh hf3wn/2 ' (3.826) where w± == W8 ± W.L, as in (2.697) . As in Eq. (3.820), the restrictions of the path integrals over x a nd p to the fixed averages Xo = x and Po = P give rise to the extra numerators in comparison to (2.701 ).

3.25_3

High- and Low-Temperature B e h avior

We have remarked before in Eq. (3.809) that in the limit T -+ 00, the effective classical potential Veffcl(XO) converges by construction against the initial potential V (xo) . There exists, in fact , a well-defined power series in tiw/kBT which describes t his approach. Let us study this limit explicitly for the effective classical potential of the harmonic oscillator calculated in (3.820), after rewriting it as (3.827)

332

3 External Sources, Correlations, and Perturbation Theory

Due to the subtracted logarithm of w in t he brackets, the effective classical potential has a power series M

effel

Vw

(xo)

=

TW

2 '2

[

1 fIw

1

( fiw )' + . .. 1.

Xo + fUJJ 24 kBT - 2880 k8 T

(3.828)

This pleasant high-temperature behavior is in contrast to that of the effective tential which reads for the harmonic oscillator

p0-

(3.829)

as we can sec from {3.793} . The logarithm of w prevents this from having a power series expansion in hw/k8T, reflecting the increasing width of the unsubtractcd fluct uations. Consider now the opposite limit T ..... 0, where the fi nal integral over the Boltzmann factor 8(xo) can be calculated exactly by the saddle-point method. In this limit, the effective classical potential Vcffcl(XO} coincides with the Euclidean version of the effective potential: VefFcl{XO)T-=O Veff{xo) =:

r c[Xl/,6lx=zo'

(3.830)

whose real-time definition was given in &t. (3.663). Let us study this limit again explicitly for the harmonic oscillator, where it becomes off" ) T_O 1i4J M 2 2 flW (3.83 1) VW (XO - + - w Xo - k8 T log k T' 2 2 B i.e., the additional constant tends to 1i4J/2. This is just the quantum-mechanical zero-point energy which guarantees the correct low-temperature limit

(3.832)

The limiting partition function is equal to t he Boltzmann factor with the zero-point energy tlW/2.

3.25.4

A lternative Candidate for Effective C lassica l Potential

It is instructive to compare this potential with a related expression which can be defined in terms of the partition function density defined in &t. (2.330): (3.833)

3.25 Effective Classical Pot.ential

333

This quantity shares with V:ffd(XO) the property that it also yields the partition function by forming the integral [compare (2.329)1: (3.834)

It may therefore be considered as an alternative candida te for an effective classical potential. For the harmonic oscillator, we find from Eq. (2.331) the explicit form

2"'" "'"

[ (

1

. Mw "'" x 2 .(3.835) V:ffd(x)= _ kaT __ log __ + _ + kBT log l _ e- 2"'" I kilT ) + -tanh-2 kBT 2 Ii kBT

This shares with the effective potential vcft"(X ) in Eq. (3.829) t he unpleasant property of possessing no power series representation in the high-temperature limit. The low-temperature limit of V:ffcl(x) looks at first sight quite similar to (3.83 1): -efFd( x, ) _ T-o-ruu kBT 2r"" + k aT -MW:2 x -- Iog V

n.

2

2

kBT'

(3.836)

and the integration leads to t he same result (3.832) in only a slightly different way:

,

- ""'/ 2k n T

J-kBT 21iw dx -Mw"~ - [ . ---, - 00

I"

lo (Ii~)

= e- ""'/ 2k nT .

(3.837)

There is, however, an important difference of (3.836) with respect to (3.831). The width of a local Boltzmann factor formed from the partition function density (2.330): (3 .838)

is much wider than that of the effective classical Boltzmann factor 8(xo) = Whereas 8(xo) has a finite width for T -+ 0, t he Boltzmann factor 8 (x) has a width growing to infinity in t his limit. Thus t he integral over x in (3.837) converges much more slowly t han t hat over Xo in (3.832) . This is t he principal reason for introducing Vclfd(xo) as an effective classical potential rather than Veffd(XO) .

e - v ",," . , the linear

term in >. yields, a fter a quadratic

(3.843) Combined with the second exponential in (3.84 1) t his leads to a generating functional for t he subtracted correlation functions (3.839):

~,",,/2 - pM_' .!/, exp 1_1 _ r"d r'P ,·( )G" ( _ ').( 'i ). Zw""ILJ - sin ({3hwj2) e 12MIi 10 T Jo dT J T w2,e T T J T (3.844)

For j(T)

3.25.6

== 0, this reduces to the local Boltl'.mann factor (3.820). P e rturbation E x pans ion

We can now apply the perturbation expansion (3.480) to the path integra1 over 1](r) in Eq. (3.813) for the effective classical Boltzmann factor 8(xo) . We take the action (3.845)

and rewrite it as (3.846)

3.25 Effective Classical Pot.ential

3 35

with an unperturbed action

O'(x,) " V"(x,)/M, and an interaction

Ant,,, [xO;1]]

=

I"

Jo

.

(3.847)

(3.848)

drvmt(xo ;1](r»,

containing the subtracted potential .

V'"'(x", "(r))

~

1

,

V(x, + "(r)) - V(x,) - V'(x,)ry(r) - 2V"(x,)" (r).

(3.849)

This has a. Taylor expansion starting with the cubic term

vint(xo; 1]) =

~ VIlt(xo)1]J + ~V(4) (XO)1]4 + .

(3.850)

Since 1](r) has a zero temporal average, the linear term ft fJ dr V'(xo)1](r) is absent in (3.847) . The effect ive classical Boltzmann fact or 8(xo) in (3.813) has then the perturbation expansion [compare (3.480)] '. B(xo) = ( 1 - h1 (Ant,")n

' (Ant, ' ).." + 2!fi? n -

' (A ';nt),e " 311i3 n

+ ... ) Bn(xo)·

(3.851)

The harmonic expect ation values are defined with respect to t he harmonic path integral (3.852)

For an arbitrary functional F [x] one has to calculate (3.853)

Some calculations of local expectation values are conveniently done with the explicit Fourier components of the path integral. Recalling (3.806) and expanding the action (3.814) in its Fourier components using (3.804), they are given by the product of integrals (3.854)

T his implies t he correlation functions for the Fourier components (3.855)

336

3 External Sources, Correlations, and Perturbation Theory

From these we can calculate once more the correlation fu nctions of the fluctuat ions 1](T) as follows:

("(T)"(r»;: ~ Performing the sum gives once more the subtracted correlation function Eq. (3.839), whose generating functiona.l was calculated in (3.844). The calculation of the harmonic averages in (3.851) leads to a similar loop expansion as for the effective potential in Subsection 3.23.6 using the background field method. The path average Xo takes over the role of the background X and the nonzero Matsubara frequency part of the paths 1](7) corresponds to the fluctuatio ns. The only difference with respect to the earlier calculations is that the correlation functions of 1](7) contain no zero-frequency contribution. T hus they are obtained from the subtracted Green functions GPo;,"'0 ),c(T) defined in Eq. (3.802). All Feynman diagrams in the loop expansion are one-particle irreducible, just as in t he loop expansion of the effective potential. T he reducible diagrams are absent since there is no linear term in the interaction (3.850). This trivial absence is an advantage with respect to the somewhat involved proof required for the effective action in Subsection 3.23.6. The diagrams in the two expansions are therefore precisely the same and can be read off from Eqs. (3.741) and (3.781). The only difference lies in the replacement X -+ Xo in the analytic expressions for the lines and vertices. In addition , there is the final integral over Xo to obtain the partition function Z in Eq. (3.808). This is in contrast to the partition function expressed in terms of the effective potential Veff(X ), where only t he extremum has to be taken.

3.25.7

Effective Pote ntia l a nd Magne tization Curves

The effective classical potential Veffd(XO) in the Boltzmann factor (3.807) allows US to estimate the effective potential defined in Eq. (3.663) . It can be derived from the generating functional Zlij restricted to time-independent external source j(T) =: j, in which case Zlij reduces to a mere function of j: (3.857)

where i is the path average of X(T). The function Z(j) is obtained from the effective classical potential by a simple integral over Xo:

Z( .)

(00

dxo

J = 1-00 ..j27r{3e

- ,8[v'''"

(3.894)

Th evaluate this we shall always change, as in (2.750), the t ime variables t l ,2 to length variables .1'1,2 =' Pl,2t/M along the direction of p. For spherically symmetric potentials V Cr) with r ;;; Ixl = ~, we may express the deriva tives parallel a.nd orthogonal to the incoming particle momentum p as follows:

V yV =zV'/r,

(3.895)

V l.V = bV' /r.

(3.896) The part of the integrand before the bracket is obviously symmetric under z ..... - z and under t he exchange ZI ..... Z2. For this reason we can rewrite (3.897) Now we use the relations (3.895) in the opposite direction as

zV'/r = 8.V, and performing a partial integration in

Zl

bV'jr = Ov..ers i;4(zdx4(z:d' .. X4(Zn) of the operator x(z) = (atz + az - I ) between the excited oscillator states (nl a.nd In). Here a and at are the usual creation and annihilation opera tors of t he harmonic oscillator, and In) = (at)n IO) /v'ni . To evaluate these expectations, we make repeated use of the commutation rules [a, all = 1 and of the ground sta.te property alO) = O. For n = 0 this gives (x 4 (z»", = 3,

(X 4(ZI)X 4(Z2»'" = 72z ;-2z1 + 24z;-4Z~ + 9, (X 4(zJ)X 4(Z2)X 4(Z3» ... = 27· 8 Z;-2z~ + 63· 32z;-2z22z~ + 351 . 8Z;-2Z~ + 9 . 8Z ;-44 + 63 . 32z;-4z~z~ + 369 . 8Z ;-4 z1 +27· 8Z;2Z~ + 9· 8z24Z~ + 27.

(3B.l)

T he cumulants a.re

(X 4(ZI)X 4(Z2)}""c = 72z ;-2z~ + 24z 14z~, (X 4(ZJ)X 4(Z2)X 4(Z3» ....C= 288(7z;-2Z;2Z1 + 9Z ;-2Z~ + 7Z;-4 4z~ + IOz;-4 z~).

(3B.2)

T he powers of z show by how many steps t he intermediate states have been excited. They determine the energy denominators in the formulas (3.515) and (3.516). Apart from a factor (g/4)" a nd a

352

3 External Sources, Correlations, and Perturbation Theory

factor 1 /(2w)~" which carries the correct length scaJe of xC:) , the energy shifts 6.E = A 1Eo t::. 2 Eo + 6. 3 £0 are thus found to be given by

6. 1 £0 = 3, 1l2Eo = -

(72. ~ + 24·

6.3 £0=288(7.

n'

+

(3B.3)

~ . ~ + 9 .~ . ~ + 7' ~. ~+ 10· ~.~) =333·4.

Between excited stales, the calculation is somewhat more tedious and yields {x4(Z)}", = fu,2 + On + 3, 4 4 (x (zdx (.zz»""c = (16'1 4 + 96,,3 + 212n2 + 204n + 72)z12z~ + (n 4 + 1On3 +35n2 + 5Ctn+24)z [ 44 + (n 4 _ 6n 3 +11112 _ 6n)::z24 + (16,,4 _ 32,,3 + 2(}n2 _ 4 n)z~Z;2,

[(16n6 + 24On~

(x4(zdx4(Z2lx4(za)) ...o

(3B.4)

(3B.5)

+ 1444n4 + 4440113 + 7324n 2 + 612011 + 2016)

+

4 X ( zl 2 zi z z~ +%1 z~z~) 4 2 (384,,5 + 288On + 8544,,3 + 12528n + 9072n + 2592):j'"2:j

+ +

(48n 5 (16n 6

+ GOOn4 + 288On3 + 66OOn 2 + 7152n + 2880)z I 4z~

144n5 + 484n4 - 744 n 3 + 508n2 - l2On)ztZ;2z;2 6 + (_ 48n + 360n4 _ 960n3 + 1080n2 _ 432n):tz;-4 + (16n6 + 48n 5 + 4n 4 _ 72n 3 - 2On 2 + 24n)z?z24z~ + ( _ 384n5 + 96On' - 864n 3 + 336n~ - 48n)zlz; 2 + (16n6 _ 144n5 + 484n' _ 744n 3 + 508n~ - l2On)zf~z;4 + (16n6 + 48n 5 + 4n' - 72n 3 - 2On 2 + 24n)z j'" 2z1z; 2]. -

(3B.6)

From these we obtain the reduced energy shifts:

d 1 E o = 6n 2 + 6n + 3, d 2Eo = _ (16n 4 + 96n 3 + 212n 2 + 204n + 72)· _ (n 4 + 1071 3 + 35n 2 + 50n + 24)· ! _ (n4 _ 6nl

(3B.7)

t

+ ll n 2 _

6n). ~l 3 2 - (I6n' - 32n + 2071 - 4n)· :t = 2· (34n 3 + 51n2 + 59n + 21), d 3EO = [(16n!> + 24071 5 + 1444n' + 444071 3 + 7324n~ + 6120n + 2016)· + (384n 5 + 288071 4 + 8544n 3 + 12528n 2 + 9072n + 2592)· t· + (48n 5 + 600n 4 + 288On 3 + 6600n2 + 7152n + 2880) ·1·1 +(16n6 - 144n5 + 484n4 - 744n 3 + 508n2 - 12071) . ! . ~ + ( _ 48n5 + 36On' - 960n 3 + 108On2 - 432n)· ! ·1 + (16n6 + 48n 5 + 4n4 - 72n3 - 2071 2 + 24n)· t·1 +( -384n 5 + 960n4 _ 864n 3 + 336n 2 - 48n) . ~ . ~ + (16n6 _ 144n5 + 484n4 - 744n3 + 508n 2 - 12071) . ! + (16n6 + 48n 5 + 4n4 - 72n 3 - 2071 2 + 24n)· t· ;!] = 4 ·3· (125n4 + 25On 3 + 472n 2 + 347n + 111).

n ·1+ 1· !)

(3B.8)

t

·1 (3B.9)

Appelldix 3C Recursion RelatiollS for Perturbation Coefficiellts

Appendix 3C

353

Recursion Relations for Perturbation Coefficients of Anharmonic Oscillator

Bender and WnZ4 were the first to solve to high orders recursion relations for the pert urbation coefficients of t he ground state energy of an anharmonic oscillator with a. potential xl/2 + 9x4/4. T heir relations are si milar to Eqs. (3.534), (3.535), and (3.536), but not t he same. Extending t heir method , we derive here a recursion relation for the perturbation coefficients of all energy levels of the anharmonic oscillator in a.ny uumber of dimensions D, where the radial poteutial is 1(1 + D - 2)/2r2 + r2/2 + (g/2)(a4r4 + !l6r6 + ... + a2 qx 2q ), where t he first term is the centrifugal barrier of angular momentu m I in D dimensions. We shall do t his in several steps.

3e.l

One-Dime nsional Interaction

X4

In natural p hysical units with h = l ,w = 1, M = 1, t he SchrOdinger equation to be solved reads (3C.l) At 9 = 0 , this is solved by the harmonic oscillator wave functions (3C.2) wit h p roper normalizat ion constant N", where H,,(x) are t he Hermite polynomial of nth degree

L" h!x".

H.. (x) =

(3C.3)

~o

Generalizing this to the anharmonic case, we solve the SchrOdingcr equation (3C.l) with the power series ansatz

t,l>(n)(x)

e- ,,"/2

f

(_ 9)k $~n)(x),

(3C A )

k=Q 00

E(n)

=

L lEi"J.

(3C.5)

k =O

To make room for derivative sy mbols, the superscript of $ ~n)(x) is now d ropped. Inserting (3C A ) and (3C.5) into (3C.l) a nd equating t he coefficients of equal powers of g, we obtai n the equa t ions

x$~(x) -

n $ k(x) =

~$~(x) -

L• (_ I )k' E!;) $ k_dx) ,

X4 $ k_ I(X) +

(3C.6)

k' s l

where we have inserted the u nperturbed energy

E~nJ = n

+ 1/2,

(3C.7)

and defined $ k(X) :::::: 0 for k < O. The funct ions $ k(X) are anharmon ic versions of the Hermite polynomials. T hey turn ont to be polynomials of (4k + n)th degree: 41:r-,

(3C.1O)

p~O

we see that t he f(!(;ursion starts with (3C.11 )

For levels with an even principal quantum number n, the functions ' = {h'f!"/h~ o 0

for O::; p' ::; n', otherwise.

(3C.23)

The energy expansion coefficients are given in either case by

Bin) = _(_ l)kC~.

(3C.24)

T he solution of the recursion relations proceeds in three steps as follows. Suppose ....'C have p' calculated for some value of k all coefficients Cr~l for an upper index in t he range 1

:s

:s

2(k-1)+n'.

:s :s

In a first step, we find Cr' for 1 pi 2k + n' by solving Eq. (3C.2Q) or (3C.22), starting with p' = 2k + n ' and lowering p' down to pi = n' + I . Note that the knowledge of the coefficients C~ (which determine the yet unknown energies and are contained in t he last tcrm of the recursion rola.tions) is not roquired for p' > n', sinoo t hey arc accompanied by factors which vanisb dne to (3C.23). Next we usc the recursion relation with rf = n' to find equa.tions for the coefficients Cl containl.'([ in the lalSt tcrm. The result is, for even k,

ct

(3C.25) For odd k , the factor (2n' + 1) is replaced hy (2n' + 3). These equations contain once more the coefficients C;' . Finally, we take the recursion relations for II < n', a nd relate the coefficients CI:' - \, ... , C~ to Cr'. Combining t he results ....'C determine from Eq. (3C.24) all expansion coefficients n ) . The relations can easily be extended to interactions which are a.n arbitrary linear combination

Ei

~

V (x) = L a~

..

fnX2n.

(3C.26)

.. =2

A short Mathematica program solving the relations can be downloaded from the internct.z~ The expansion coefficients have the roma.rkahle property of growing, for large order k, like

Ek(.. ) - - + - -1~12(. -1"-3)r(k + n + l/2). 11"

1I"n.

(3C.27)

T his will be shown in Eq. (17.323). Such a factorial growth implies the perturbation expansion to have a zero radius of convergence. The reason for this will be explai ned in Section 17.10. At the expansion point 9 = 0, the energies possess an CS'lCntial singularity. In order to extract meaningful numbers from a Tay lor series expansion around such a singularity, it will be nOCOSlSary to find a oonvergent resummation method. T his will be provided by the variational pert urbation t hoory to be developed in Section 5.14. 2~Sce http://www.physik.fu-berlin/-kleinert/b3/progra.m8 .

356

3 External Sources, Correlations, and Perturbation Theory

3C.2

G e ner a l One-Dimensional Interaction

Consid(lr now an arbirary inwraction which is (lxpandable in a power series

., ~

1)(x) = :L: 9k1)k+2Xk+2.

(3C.28)

Note t hat th(l coupling constant corresponds now to th(l square root of th(l previous one, the IOw3X3 + y21)4:t 4 + . The POwtlJ1; of 9 count the number of loops of the associated Feynman diagrams. Then Eqs. (3C.6) and (3C.I5) become

(p - n)A: =

~(p + 2)(P + I)At 2 -



2:)- 1)k'1)k'+2A:: r2 k' .1



+L

(- l)k' Ei~) A:_ k.. (3C.30)

k' . d

Th(l expansion oocffici(lnts of the energies are, as before, given by (3C.17) and (3C.IS) for even and odd n, respectively, but the recursion relation (3C.30) has to be solved now in fulL

3C.3

Cumulative Treatment of Interactions

x4

and

x3

T here exists a slightly different recursi ve treatment which we shall illustrate for the simplest mixed interaction potential

(3C.31) Instead of the ansatz (3C.4) we shall now factorize the wave fnnction of the ground state as follows:

(3C.32) i.e., we allow for powers series expansion in the exponent: ~

4>(n l (x) = :L:yk 4>~n>C:t ).

(3C.33)

k=1 \Ve shall find that t his expansion contains fewer terms than in th(l &nder·Wu (lxpansion of t he correction factor in Eq. (3C.4). For completeness, we keep here physical dimensions with explicit constants h,w, M. Inserting (3C.32) into t he SchrOdinger equat ion

[-

2~f :::2 + (~ "/X2 + 91>3X3 + l1)~x')

- E(n)] t/J(n)(x) =

0,

(3C.34)

wtl obtain, aft(lr dropping (lvcrywhere the superscript (n) , th(l differential equation for 4>(n)(x):

(3C.35) where ( denotes th(l correction to the harmonic energy

(3C.36)

Appelldix 3C Recursion RelatiollS for Perturbation Coefficiellts We shall calculate

£

357

as a power series in 9:

(3C.37) From now on we shall consuder only the ground state with n = O. Inserting expans ion (3C.33) into (3C.35), and comparing coefficients, we obtain the infinite set of differential equations for ¢.I:(x): "1

"1 ,\: _1

2~ ¢Z(x) + hwxoP~(x) - 2~ L oP~_I(X)oP;(x) + 05 k ,IV:iX 3 + 05t.2VtX4 =

-

,=,

fA: .

(3C.38)

Assuming that oPA:(x) is a polynomial, "''() can show by induction that it:! degree cannot be greater than k + 2, i.e., 00

oPt (x) =

L

withc!:: );;;; O fOrm > k + 2 ,

c!::) x"',

(3C.39)

",=1

T he lowest terms 4 .1:) have been omitted since they will be determined at the end the normalization of the wave function t{I(x) . Inserting (3C.39) into (3C.38) for k = I , v.'() find ~1) = 0,

£1

(3C.40)

= 0.

For k = 2, we obtain (1) c\

(2

3V4

7vj

(2)

= 0, Co2

r4"l) =

= SM2",4 - 4M",2'

0,

(3C.41)

llv~h"1 3V4h"1 8M3W4 + 4APw1 .

-

(3C.42)

For the higher-order terms we must solve the recursion relations

(m + 2)(m + l)h (t) _ ,_,_ ~ ~I ( 2mM", c"'+1 + 2mM", L L n m

(t ) _

em -

+

2 _

) (I) (.1:_ 1)

n

en

C",H_ .. '

(3C.43)

1=1 .. .. 1

"1

_ _

fA:-

~

"1 .1:- 1

(k) _

MCi

~" (I) (.1: _ 1) 2M L

CICI

(3C.44)

.

/= \

Evaluating t his for k = 3 yields 5v~h

(3)

c,

= -

M 4",7

+

6V:lv 4 h A{3",S'

~3) = 0 , (3)

c4

(3) _

= 0, c5

-

v~

- lOM 2 hw 5

V:lV4

+ 5Mfu.J3 '

f3

= 0,

(3C.45)

and for k = 4: (4)

305v~fl

Co2 = 32M s",9 -

123vjv4f' 8M4w1

21 v~h

+ 8M3",5' (3C.46) (3C.47)

358

3 External Sources, Correlations, and Perturbation Theory

The general form of the coefficients is, now in natural units with h = 1, At = 1,:

(3C.48) (3C.49) T his leads to the recursion relations (t) _ { m

em,)' -

k - lm + 1

{} + ' }m+ l k + 2,

~ (.1:) + 2X (k) + ....!!:..... ~ ( (k _ l) (I) + (.1:_ 1) (I») + -.!... ~'(X) Mw q q M"" L q c1 c1 q rn,., k , I_I

(30 .76) (30.77)

Appelldix 3C Recursion RelatiollS for Perturbation Coefficiellts

1

363

10 _1

_!!:.....'" (I) (10 - 1) . 2M~c I c i

(3C.78)

,.,

The results are Listed in Table 3.3,

3C.6

Inte r actio n r 4 in D -Dime n siona l Radial O scilla t or

It is easy to generalize these relations further to find the perturbation expansions for the eigenvalues of the radial SchrOdinger equation of an anharmonic oscillator in D dimensions

_! D [_~~ 2dr 2 2 r The case 9 =

°

l .!!.. + /(1 + D - 2) dr 2r2

+ ~r2 + !!.r4] 2

4

R.,(r) = E (")R.,(r) .

(3C.79)

will be solved in Soction 9,2, wi t h the energy eigenvalues

E(") = 2n'+I+ D/ 2 = n+D/2,

n = 0, 1,2,3, . .

,1 = 0,1, 2,3,.

(3C.80)

For a fixed princi pal quantum number n = 2n,. + I, t he angular momentnm runs through 1 = 0,2, ... , n for even, and I = 1, 3, . . . , n for odd n . There are (n + 1)(n + 2)/2 degenerate levels. Removing a factor r' from R.,( r ), and defining R.,(r) = r'wn(r), the SchrOdinger equation becomes +D- 1dI294) (.) dr2 - 2 r dr + 2r + 4"r w .. (r) = E w .. (r). ( -21~121

(3C.Sl)

The second term modifies the differential equation (3C.G) to ,

1 2

r k(r) - 2n'k(r ) = - ~ (r)+

(21 + D-1 ) 2r

1 0 , (n ;'(r)+r4k_1 (r)+ '" (_ 1)10 E k, )k_dr) . (3C.82) ~ 10'_1

The extra terms change t he recursion relation (3C.15) into

(p- 2n')A: =

~ [(p+ 2)(P +

1) + (p+ 2)(21 + D _ 1) ]A:+2 + A::~ +

L• (_ 1)10' Ek:')A:_ ,.(3C.83) k

10 ' ", 1

Fbr even n = 2n'

+ 1 with I = 0, 2, 4,. ,. ,n, we

C1 =

normalize the wave functions by setting

(21 + D)oSok ,

(3C.84)

rather than (3C.12), and obtain

2(p' -

n')Cr' = [(2p' + 1)(P' + 1) +(p' + 1)(I+D/2 -

1/2)]Cr'+1 + Cf~-12 -

L• c:' Cr'_k"

(3C.85)

instead of (3C.20). For odd n = 2n' + 1 with I = 1,3,5"." n, the equations analogous to (3C.13) and (3C.22) are

e! = 3(21 + D)6ok

(3C.56)

ru,d

2(P' -

n')er' = [(2p' + 3)(P' + 1)+(p' +3/2)(1+ D/ 2- 1/2)]er'H + er~-;2 -

L• el, er~ k"

(3C.87)

In either case, the expansion coefficients of the energy arc given by E(n ) _ _ k

-

(_1) 10 21+ D + lei

2

21

+D

k'

(3C.88)

364

3C.7

3 External Sources, Correlations, and Perturbation Theory

Interaction

r2q

in D Dimensions

A further extension of the recursion relation applies to interact ions gx 2q /4. Then Eqs. (3C. 20) and (3C.22) arc changtxj in the s(''COnd terms on the right-hand ~ide which bocome Cr' - q. In a first step, these equations are now solved for Cr' for I ::; P ::; qk + Il, starting with pi = qk + Il' and lowering pi down to pi = Il' + L As before, the knowlodge of the coefficients C~ (which determine the yet unknown energies and are contained in the last term of the recursion relations) is not required for pi > Il'. T he second and third steps are completely analogous to the case q = 2. T he same generalization applies to t he D-dimensional case.

3C.8

Polynomial Interaction in D Dimensions

If t he SchrOdinger equation has the general form

~ _~D[ _!2dr2 2 r

l~ dr

+ 1(I +

D - 2)

+ ~ r2

2r2

2

+ ~(a4r4 + a6r6 + ... + a2qx2q)j R,,(r) = E(n)R,,(r),

(3C.89)

we simply have to ruplaoo in t he recursion relations (3C.85) and (3C.87) the second term on t he

right-hand side as follows

2 C,,'- 3 C.,,'- 12 .... a4 C,,'~ _ ) + Clf, ~ _ ) + . . . +ll2q

c,,'-q k- l '

(3C.90)

a nd perform otherwise the same steps as for the potential gr 2q /4 alone.

Appendix 3D

Feynman Integrals for T-/ 0

T he calculation of the Feynman integrals (3.547) can be done straightforwardly with t he help of the symbolic p rogram Mathematica. T he first integral in Eqs. (3.547) is triv ial. The second and forth integrals arc simple, since one overall integration over, say, 1"3 yields merely a factor h{J , d ne 10 translational invariance of the integrand along the 1"-sxis. T he triple integrals can t hen 00 split ~

L"fjlo',fjlo',fjd1")d'l)d73 1(11"1 = h{J

Ll!.flo"8d1")d1"2 /(11") -

= h{J

(10

M d'l)

1"2 1, 1'1) - 731, 173 - 1") 1)

(3D.I)

'1)1, 11"2 1, l1"d )

L~ d1"t/('I) -1"1, '1), 1"1) + 10"8d'l)

1:8

d1"t/(1"1 - '1),'1), 1"J)) ,

10 ensure that the arguments of t he Green fWiction have the same sign in each term. T he lines

represent t he thennal correla tion function

C(2)(r r) = h ooshw ll1" - 1"'1 - h{J/21 , 2M,.., s in h{,..,h{J/2)

(30.2)

W ith the dimension less variable x :::::: ,..,h(j, the result for t he quantities O~L definod in (3.547) in the Feynman d iagrams with L lines and V vertices is

21 coth "2' 1

1

-, -.- , - (x + sinh x), smh ~

(3D.3) (3D.4)

Appelldix 3D Feynman Integrals for T

=I 0

365

1 - ( - 3 cosh -X +2x 2 cosh -X + 3 cosh -3x + 6xsmh . -x) -1 - 64sinh3~ 2 2 2 2 '

(30.5)

_1___\ _

(30 .6)

256 sinh

(6x+8 sinhx+sinh2x),

~

II (

x

x

3x

- - - - - - 40 cosh- + 24 x 2 OO8b -+ 35OO8h 4096 sinh~ 1 2 2 2 + 5 cosh 5; +

72xsinh~ +

1 1 - - - - , - ( - 48 + 32x2 - 3 cosh x 16384 sinh ~

12xsinh ~) ,

(30 .7)

+ 8X2 cosh x

+48 cosh2x + 3 cosh3x + 108xsinhx), 1 x), 24 ~(5+24cosh smh }

(30.8)

x x 3 xcosh"2+9sinh"2 +si nhT

(30.10)

1 1

1

72sinh3~ 1

(

3x)

1

'

. .

(30.11)

2304 -.- . - (lOx + 1048mhx + 5smh2x). smh

(30 .9)

~

For completeness, we have also listed the integrals ot o~, and o~9, corresponding to the three diagram.s

e ,

e ,

(30.12)

respectively, which occur in perturbation expansions with a cubic interaction potential x 3 . T hese will appear in a modified version in Chapter 5. In the low-temperature limit where x = wh{3 __ 00, the x-dependent factors o~L in Eqs. (30.3)(30 .11) converge towards the constants 1/2,

1/4,

3/ 16,

1/32,

5/(8.25 ),

3/(8.26),

1/12,

1/18,5/(9.25 ),

(30. 13)

respectively. From these numbers we deduce t he relations (3.550) and, in addition, ,,~

-.

__ _2 a6 3

'

0 ,,1 - J'

.....

~.'" 9 .

(30.1 4)

In the high- temperature limi t :r: ..... 0, t he Feynman integrals h{3(l/w)V - la~L witb L li ne; and V vertices diverge like {3V( l /{3)L. The first V factors are due to the V-integrals over T, the second are the consequence of t he product of n/2 factors a 2 . T hus, a~L behaves for x ..... 0 like (30.15) Indeed, the x-dependent factors o}L in (30.3) - (30 .11) grow like

a' al a! a! a'",

~ ~ ~ ~ ~

l/x + x/12 + 1/x + x 3/72O, 1/x+x 5/30240 + . .. , 1/x3 + x/ 12O - x 3/378O + x 5 /80640 + ... , 1/x3 + x/240 _ x 3/15120 + x 7 /6652800 +.

366

3 External Sources, Correlations, and Perturbation Theory a~~

~

.! ·1

~ ~

a1?

~

1/x4 + 1/240 + x~ /15120 - x6/4989600 + 701 x 8/34871316480 + ... , 1/x2 + x 2 j240 - x 4 /6048 + ... , l/r? + :r? /720 - x 6/518400 + ... , 1/x3 + x/31lO - x 5/1209600 + 629x9 /261534873600 +.. (30.16)

For the temperature behavior of t hClSC Fcynman integrals soo Fig. 3.16. We have plottod the reduced Fcynman integrals iitL(x) in which the low-temperature behaviors (3.550) a nd (3D.14) have been divided out of at"L.

1k-- --====-==--__ _

1.2

a'

0.8 0.6

0.4 0.2 -1---"'0-0 .1- -0 ""."2- - '0".30---= 0-' .4- -0 "" .5 L(x

Figure 3.16 Plot ofrcduccd Fcynman integrals o.~L(x) as a function of Llx = LknTjhI..J. The integrals (30Al- (3D,I I) arc indicated by decreasing dash-lengths. T he integrals (30.4) and (3D.5) for a~ and by the operation

n" (

"!Mn

8)"

- 8w2

ag can be obtained from the integral (30.3) for a2 n" (

= n!Mn

18)"

- 2w 8w

(30.17)

with n = 1 and n = 2, resp e .... (xcl)/h

1

00

-00

dJx

_ _ _ e ia"{Xd )(Jx]2/2h

=

eia(xc,)/h

.jall(xcl)

V211"itl

.

(4.45)

The right-hand side is t he saddle point approximation to the integral (4.41). The saddle point approximation may be viewed as the coru;cquenee of the classica! limit of the exponential funct ion: e

ia(x)/h

-

h_O

~ J( x -

r::;;;:::-;:. Va"(xcI)

Xci

)

.

(4.46)

Corrections can be calculated perturbatively by expanding t he integral in powers of Ii, lcading to what is called t he saddle point expansion . For this we expand the remaining exponent in powers of ox:

(4.47) and perform t he resulting integrals of the type

dJ __ 1 00

X_._eia"(Xcl)(6x)2/2h(Ox)fI =

-00 J21nli

{

(n - I)!! (. )"1' _ [a"(xcl)](l+fI)/2 tli , n - even, 0, n = odd.

(4.48)

Each factor ox in (4.47) introduces a power .jli/all(xcl). This is the average relative size of t he quantum fluctuations. The increasing powers of h ensure the decreasing

4 Semiclassical Time Evolution Amplitude

3 78

importance of the higher terms for small h. For instance, the fourth-orde r term a(41(xcl)(ox)4j4 ! is accompanied by fi , and the lowest correction amounts to a factor (4) 311 h 1 - ta (Xcl) 4! [a"(xcl)]2

a(31(xcJ)(ox)3j3!

The cubic term

1+

(4.49)

yields a factor

*

(3)

25

1

1

h

(4.50)

(Xcl)] 72 [a"(xcl) j3

Thus we obtain the saddle point expansion to the integral is dx

1

00

- -e

- (>0

>,(.)/"

v21rin

=

ei IPata)""

=

~[-fJ:l;6f)",~A{Xb' x,,; tb - t .. )l I/2ei[A(%~,z~;fb-ta)-PbZ.+p.,O:AlI",(4.163)

where H is the matrix

aZbaZAA(Xblx,,~tb ~AA(XblXa,tb

-

tal ) . tal

The exponent must be evaluated at the extremum with respect to lies at

XI>

(4.164)

and

Xa ,

which

(4. 165)

The exponent contains then the Legendre transform of the action A(Xb, X and p,,: (4. 166)

The inverse Legendre transformation to (4 .165) is (4. 167)

The important observation which greatly simplifies the result is that for a 2 x 2 matrix Hal> with (a, b = 1, 2) , the matrix element -H I 2/detH is equal to H 12 . By writing the matrix H and its inverse as

(4. 168)

we sec that, just as in the Eqs. (2.276) and (2.277):

H

"-

1 _

fJx a _ a'lA(PI»Pa ;t b - tal 8Pb8pa·

- -8Pb -

(4. 169)

As a result, the semiclassical time evolution amplitude in momentum space (4. 163) takes the simple form (4. 170)

In D dimensions, thb kcomes (4.171 )

4.7 $emic1l1SSical Quantum-Mechallical Partition Functioll

395

oc

(4.172)

these results being completely analogous to the x-space expression (4 .125) and (4.127), respectively. As before, the subscripts a and b can be interchanged in the determinant. If we apply these formulas to t he harmonic oscillator wit h a time-dependent frequency, we obtain precisely the amplit ude (2.283) . Thus in t his case, the semiclassical time evolution amplitude (Pbtb lp"t")",, happens to coincide with t he exact one. For a free particle wit h t he action A(Xb,X,,; tb - t,,) = M(Xb - x,,? j 2(tb - t,,) , t he formula (4. 163) cannot be applied since determinant of H vanishes, SO t hat the saddle point approximation is ina pplicable. The formal infinity one obtains when trying to a pply Eq. (4 .163) is a reflection of the 6-function in t he exact expression (2.138), which has no semiclassical approximation. T he Legendre transform of t he action can, however, be calculated correctly and yields via the derivatives Pa = Pb =: P = A(Xb, Xa; tb - t,,) = M(Xb - xa)j2(tb - tal the expression

A(Pb , p,,;tb - t,,)

=

r

-2(tb - t,,) ,

(4.173)

which agrees with the exponent of (2. 138).

4.7

Semiclassical Quantum-Mechanical Partition Function

From the result (4 .103) we can easily derive the qua ntum-mecha nical partition function (1.537) in semiclassical approximation:

ZQM(tb - tal =

JdXa(Xatb lxata)"" JdXaF(xa, Xa; tb - ta)e;A(z~,z~;tb-t,,)//i;. =

(4.174)

Within the semiclassical approximation t he path integral, as the final trace integral may be performed using t he saddle point approximation. At the saddle point one h", ['" in (4. 133)[

i.e" only classical orbits contribute whose momenta are equal at the coinciding endpoints. This restricts the orbits to periodic solutions of the equations of motion. The semiclassical limit selects, among aJl paths wit h x" = Xb, t he paths solving t he equation of motion, ensuring t he continuity of the internal momenta along these paths. The integration in (4.174) enforces the equality of the initial aud final momenta on these paths and permits a continuation of the equations of motion beyond

4 Semiclassical Time Evolution Amplitude

3 96

the final time tb in a periodic fashion , leading to periodic orbits. Along each of these orbits, t he energy E(x" , xo. , tb - tal and t he action A(xa, X a, tb - tal do not depend on t he choice of x" o The phase factor c,AI" in the integral (4.174) is t herefore a constant. T he integral must be performed over a full period between the turning points of each orbit in the fonvard and backward direction. It contains a nontrivial xa-depcndence only in t he fluctuation factor. Thus, (4.174) can be written as

(4.176) For the integration over the fl uctuation factor we use the expression (4. 123) and the equation

(4.177) fo llowing from (4.119) and (4.93), and have

F(Xb,xa;tb Inserting

Xa

=

Xb

tal =

1

[

82 Aj'/'

1

J2nin X(tb)X(t,,) 8t~

(4. 178)

leads to 1

1

[8'Aj'/'

F(xa ,x,,; tb - tal =..j . -:- i!t' 2Jrln Xa /)

(4. 179)

T he action of a periodic path does not depend on Xa,SO t hat the xo-integration in (4. 174) requires only integrating 1/3;" forward and back, which produces the total period: tb- t a=2

{x+dxa~= 2 {x+dx

1._

Xo

1._

M

J2M IE V( x)J

(4.180)

Hence we obtain from (4. 174) :

Z~ QM

(t _

b

t ) = to - ta I8'A Il/2eiA(t6- t..l/h- i". . a

..j2Jrin

&tl

(4. 181)

There is a phase factor e- i ". associated with a Maslov-Morse index /.I = 2, first introduced in the fluct uation factor (2.269). In the present context, this phase factor arises from the fact that when doing the integral (4.176), the periodic orbit passes t hrough the turning points x _ and x+ where the integrand of (4.180) becomes singular, even though the integral remains finite. Near the turning points, the semiclassical approximation breaks down, as discussed in Section 4.1 in the context of the WKB approximation to the SchrOdinger equation. This breakdown required special attention in t he derivation of t he connection formulas relating the wave functions on

4.7 $emic1l1SSical Quantum-Mechallical Partition Functioll

397

one side of the turning points to those on the other side. There, t he breakdown was circumvented by escaping into t he complex x-plane. When going around the singularity in the clockwise sense, the prefactor l /p(x) = 1/ J2M(E _ V(x» ]/2 acquired a phase factor e- i1f / 2 • For a periodic orbit, both turning points had to be encircled producing twice this phase factor, which is precisely the phase e- i 1f given in (4.181). The result (4.181) takes an especially simple form after a Fourier transform action:

(4. 182)

In the semiclassical approximation, the main contribution to the integral at a given energy E comes from the time where tb - ta is equal to the period of the particle orbit with this energy. It is determined as in (4.139) by the extremum of

A(t. - to)

+ (t. - tole.

(4.183)

Thus it satisfies (4.184)

As in (4. 140), the extremum determines the period tb - ta of the orbit with an energy E. It will be denoted by t(E). The second derivative of the exponent is (i/h)8 2A(tb - ta)/&t~. For t his reason, the quadratic correction in the saddle point approximation to the integral over tb cancels the corresponding prefactor in (4. 182) and leads t o t he simple expression (4. 185)

The exponent contains again the eikonal S(E) = A(t) + t(E)E, t he Legendre transform of the action A(t) defined by

S(E)

~ A(t) _ tV~~t),

(4.186)

where the variable t has to be replaced by E(t) = -8A(t)/8t. Via t he inverse Legendre transformation, the derivative 8S(E)/8E = t leads back to

A(t) ~ S(E) _ vS(E) E vE .

(4. 187)

Explicitly, S(E) is given by the integral (4.68):

S(E) ~ 2

1_ 0

10



dxp(x) ~ 2

1_ 0

10



dx "j2M [E - V(x)[ .

(4. 188)

4 Semiclassical Time Evolution Amplitude

398

Finally, we have to take into account t hat the periodic orbit is repeatedly traversed for an arbitrary number of times. Each period yields a phase factor eiS (E)l h- ;1r. The sum is Z~ (E) QM

is(E)/h

00

~ "t(E),;·,S(E)I' · V(x): p(E;x) =

{I- 1~~ VI/(x ) d~2

-

2:t~ [V'(xW d~3 + ... } Pd(E;x).

(4.254)

Inserting (4.253) and performing the differentiations with respect to V we obtain

E)

P( ;x

1

IM{ [E-V(x) 1 Ii' V"( )3 1 j1/2- 12M x 4 [E_V(X) ]5/2

=rrnV"2

r.' [ '

[2 15 - 24M V (x ) alE

1

V(x)J'/'

} ( + .... 4.255)

Note that t he proceeds in powers of higher gradients of the potential; it is a gradient expansion. The integral over (4 .255) yields a gradient expansion for p(E) [2J. The second term can be integrated by parts which, under the assumption that V(x) vanishes at the boundaries, simply changes the sign of the third term, so that we find

1

1M!

{

p() E ~

dt eil[E- V(Xl l/" [ dt' - ;e" / ~ dE _oo21rh(27riht/M)OnJo 7 e .

(4.282)

Integra.ting over the energy yields

[ .[ _ J :tJoroo Tr logH d _ _

d

D

dt

1

T (2trihtjM)Dl le

- iW(xl! "

.

(4.283)

Deforming the contour of integration by the substi tution t = - iT , "'() arrive at the inu'gTal roprt).. scntation of the Gamma. function (2.496) which rcproduCU'! immediately the result (4.281). T he quantwn corrections are obtained by multiplying this with the prdactor in curly brackets in the expa.nsion (4.258): Trlog if

(4.284) The !!COOnd term can 00 integrated by parUi, which replaces \7 2 V(x) ..... -['i7V(xWdjdV , so that we obtain the gradient expansion

TrlogH = -

M)DI' r{ - D/2) JdDx {h' l + - - [V V(xW - d', +. .} [V(x)]Ol l. (~ 2'1fh 24M dV

(4.285)

The curly brackets can obviously be replaced by

1'12 r(3 - D/ 2) [V V(x)J2 { 1 - 24M f( D/2) [V(x)J3

}

+. .. .

(4.286)

In one dimension and wi th M = 1/2, this amounts to the fonnula

(4.287) It is a useful exercise to rcderivc this with the help of the Gelfand-Yaglom method in Section 2.4. There exists another method for deriving the gradient expansion (4.284). We split V(x) into a coru;tant term V and a small x-depcndent term JV(xl, and rewrite Trlog [-

2~ V 2 + V(X)]

= Trlog [- 2":1 V 2 + V + OV(x)] = Trlog ( -

where 6. v denotcs the functional matrix

h' ) -' 6. v (x,x') = ( - 2M V 2 + V =

2~ V 2 + V) +Trlog (1 + 6. v oV) ,

J

dDp

e;p(X - x'j,A

(211"Iil1) p l/2M + V=6. v (x - x').

(4.288)

(4.289)

T his coincides with the fixed-energy am plitude (i/Ii)(x lx'l£ at E = - V [rocall Eq. (1.344) ]. T he first term in (4.288) is equal to (4.281) if we replace V (x) in that expression by t he constant V , so that ,,",'C may write Tr log if = [Tr log Hlell

V(xj_V

+ Tr log (1 +.6. v 5V) .

(4.290)

413

4.9 Quantum Corrections to ClllSSicai Density of States We now expand the remainder Tr log (1

+ .6. v 6V) = Trfl. v W - ~Tr(fl.V OV)2 +.

(4.291)

and evaluate the expansion terms. The first term is simply (4.292) where dDp 1 (21rh)D p2/2M

J

fl.V(O) =

"

(4.293)

+ V = llv [Trlog H lct!v(x)_ v'

The result of the integra.tion was given in Eq. (1.348). The second term in the rema.inder (4.291) reads explici t ly

- ~Tr We now make

(AVOV)2 =

U8C

-~

J J

dDx' Av(x,x')6V(x')Av{x', x)oV(x).

dDx

(4.294)

of t he operator relation

[/[A), B[ ~ f'(AIiA,B[ - ~f"(AIiA, [A, B[[ + .

(4.295)

to expand

WAy

=

AvW +At [t,ovj +A~ [t ,[t,WI] + .

(4.296)

where t is the operator of the kinetic energy p 2/2M. It commutes with any function I(x) as follows:

[T,I[

[1',[1',111

~

h'

,

- 'M il" 1[ + 2(" 11, " 1,

4~;2 {[(V 2)2 IJ + 4[VV2II· V + 4[ViVdl'ii'iV

(4.297) j ) ,

(4.298)

Inserting this into (4.294), we obtain a first contribution (4.299) The spatial integrals arc performed by going to momentum space, where we derive the general formula

(4.300)

This simpllfies (4.299) to (4.301) We may now combine the non-gradient terms of oV(x) consisting of the first term in (4.290), of (4.292), and of (4.301), and replace in the latter Av(O) according to (4.293), to obtain the first three expansion terms of [Trlog HId with the full x"dependent Vex) in Eq. (4.281).

414

4 Semiclassical Time Evolution Amplitude

The Ilext contribution to (4.294) coming from (4.297) is 4":f

J J J dDx

dDxl

dDx2L\.V(X, XJ)AV(Xl, X2 )Av(X2, X)

X ([V 2W(xl] W(x)

+ 2 [V6V(x)f + 2[V 5V(xl]6V(x)V } ,

(4.302)

where the last V acts on the first x in 6. V (X,X l), due to t he trace. It does not contribu te to the integral since it is odd in x - Xl' We now perform the integrals over Xl and Xz using formula (4.300) and find

(4.303) The first term can be integrated by parts, after which it removC8 half of the second term. A t hird contribution to (4.294) whi.:h contains only the lov.'CSt gradients of 6V(x) comes from

the third term in (4.298);

where the last 'iI;'Vj acts again on we mcn the first x in A v( x , xIl, as a consequence of the t race. in momentum space, we encounter the integral

(4.305) so that the third contribution to (4.294) reads, after an integration by parts,

h'J

- At

I .1)

d D x [V oV(x) ]2~(1"2 D 2"OV+i3Vuy

(4.306)

a v(O).

Combining all gradient terms in [V V(xW and replacing av(O) according to (4.293), we recover the previous result (4.285) with the curly brackets (4.286). For the one-dimensional tracelog, this leads to the formula

Trlog[- /i

a: + V (x)]

2

=

h1

J

c/:J: JV(x)

{ h Z [V'(xW } 1 + 32 V3(X) + ... .

(4.307)

It is a useful exen-:ise to rederive th is with t he help of the Gelfand-Yaglom method in Section 2.4. This expansion ca.n actually be deduced, and carried to much higher order, with the help of the gradient expansion of the t race of the logarithm of the operator - h 2 lP.. + W 2(T) derived in Subsection 2. 15.4. If we replace T by x/Ii and V(T) by V (x), we ohtain from (2.548)- (2.550): 1

h

J

{

V'

, (5V, 2

dx JV(x) 1 -Ii 4V3/2 -Ii

_

V" ) , ( 15V,3 9V'V" 32V3 - 8V2 - Ii 64V9j2 - 32V7/2

,(l105V,4 _ 221V,2V " Ii

2048V~

256V~

19V,,2

+ 128Vt +

V (3) )

+ 16V~j2

7V'V(3) _ V (4) ) } 32Vt 32V3 .

(4.308)

The h-term in the curly brackets vanishes if V (x) is the same at the boundaries, and the h2 term goes over into the til_term in (4.307).

415

4.9 Quantum Corrections to ClllSSicai Density of States

4.9.5

Local D e ns ity of States on Circle

Fbr future usc, let us also calculate t his d

pcl(b, E;x) = --; ~2L - (E) 11"1< 211"hlM .. = 0 p ;x

K 1/2{np(E;x)blh).

(4.31 3)

Inserting K 1/ 2 (z) = J'lr/2ze - ' [recall (2.559)], this becomes M -I PI(b E·x) = c

T he sum

"

1I"hp(E; x)

L::".. , a" is equal to al(1 -

( 1+2 L~

e- np(E,o:)b/h )

n_ 1

.

(4.31 4)

a) , so that we obtain

==th,-.j,;2~M~tEEi; V¥,(x:\,)Jr-b/"2,,h ( b E.x) = M coth[p(E;x)bI2h] = M CO Pel " 'lrh p{E;x) 'lrh J2M[E V ex)]

(4.315)

For b ..... 00, this reduces to t he previous density (4.253). If we include t he h igher powers of Tin (4.252), we ohtain the generalization of expression (4.254): pCb, E;x) = { 1 -

1~~ V" (x) d~2 - 2~l~ [V'(x) ]2 d~3 + ... } Pcl(b, E;x).

(4.31 6)

416

4 Semiclassical Time Evolution Amplitude

T he traoolog is obtained by integrat ing this over dElog E from Vex) to infinity. The integral d iverges, and we must employ analytic regularization. We proceed as in (4. 282), by usi ng t he real-time version of (4.311) a nd rewriting log E as a n integral - fo""(dt' /t')e - · Et '''', so that t he leading term in (4.316) is given by

, --1'

[TrlogH]Ol -

I)

1~ dx

1O..! !. . L

dE

_ 00

_ 00

2 Ii 11"

e - ;n'Mb'/2~'+j'(E - V(")I/~ (2

n~_ ""

'~

11"1 It

/Af)I/2

1"" 0

dt' - i E.' I"

t

J

e

. (4.317)

T he integral over E leads now to

(4.318) Deforming agai n the contour of integration by the substitution t = - iT, creating the denomi nator by a n integration over V(x) /h, we see that

[Tr logHJcl =

=

t dx ( to dVpc\(b,E;x)1 t ("" dV 1 coth 4.,jVJE;

-11"

)0

11" ) 0

lv(:_V(r)_V

dx iv(%) E6411"b

h/V/Eb

'

(4.319)

where E6 :::::: h/2Mb 2 is the energy associated with the length b. The integration over V produces a factor h/ T in the integrand. T hus we obtain

1>logH~ 1,'

,,+ +2:~[V'(X)['

:"

+p }~IO+iOh [~tg) ]}

(4.320)

For M = tI2 / 2, this give us the finite-b correction to formula (4.307). Replacing:r by the Euclidean time T, b by h/3, and V(x) by t he time-dependent square frequency W2(T), we obtain the gradient exparu;ion (4.321)

4.9 .6

Quantum Corrections to Bohr- Sommerfeld Approximation

T hc cxpansion (4.308) ca.n be used to obtain a higher-order expansion of the dcnsity of states peE), thereby extending Eq. (4.256). For this we rco-:a1l Eq. (1.592) according to which we can calculate the exact density of states from the formula 1 peE) = --8E 1m Trlog {-a!

,

+ [V ex) -

E I} .

(4.322)

Integrat ing thi~ over the energy yields, according to Eq. (1.593), t he number of Btatcs t imes a nd t hw; t he simple exact quantization condition for a nondegenerate ono-dimensional system: - h nTrlog {-a! By comparison with Eq. (4.209) eikonal:

"'1)

+ [v{x) -

EI} = 7t(n + 1/ 2).

7r,

(4.323)

may define a fullly quantum corrected version of the classical

Sqc(E) = - 2h lm Trlog {- a!

+ [v(x) - EJ}.

(4.324)

417

4.9 Quantum Corrections to ClllSSicai Density of States

The semiclassical expansion of t his can be obtained from our earlier result (4.308) by replacing V(x) ..... Vex) - E, so that JV(x) ..... -i JE Vex) , yielding Sqc(E) = 2 .[

-h

f

fix JE

V 2 V(X) { 1 + 1i [32(:,2 )J + 8(E "V)2] V

I I05V,4 22IV,2 V" 19V,,2 7V'V(J) V (4) ]) 2048(£ V )6 + 256(£ V)5 + 128(£ V )4 + 32(£ V )4 + 32(£ V )3 +... . (4.325)

The first term in the expansion corresponds to the Bohr-Sommerfeld approximation (4.27), the remain ing ones yield t he quantum corrections. The integrand agrees, of course, with the WKB expansion of the eikonal (4.14) after inserting the expansion terms So, S" . .. , obtained by integrating the relations (4.20) over the funct ions %(X) , qL(X) , . . . of Eq. (4.18). Using Eq. (4.322) we obtain from Sqc(E) the density of states

pee) = + h3

2 I I J I 3V"] -,.-Ii Sqe(E) = -, .-, dx -" -E-_V I\1- h [25V" 32 (E _ V )3 + ~8~(E:;.'--~V~)"

12155V,4 1989 V,2 V" 133V,,2 49V'V(3) 5V (~)] ) . (4.326) [ 2048(E V)6 + 256(E V) ~ + 128(E V )4 + 32(E V)4 + 32(E V)J

Let us calculate the quantum corrections to the semielassical energies for a purely quartic potential Vex) = gx4 /4, where the integral over the first term in (4.325) between the turning points ±Xf; = ±(4£/g)1 /4 gave the Bohr-Sommerfeld approximation (4.33). The integrals of the higher terms in (4.325) are divergent, but can be calculated in analyt ically regularized form using oncc more the integral formula (4.34). This extends t he Bohr·Sommerfeld oquation vee) = n+ 1/2 to the exact equation N(E):= Sqc/27rh = n + 1/2 [recall (1.593)1. If we express NCE) in terms of veE) defined in Eq. (4.33) rather than E, we obtain the ex pansion

N(v) _ v _ _1_ -

12lfv +

117r2

4

I0368ra(~)v3

46977r _ 3900657r + 1866240fSWv6 501645312f I6 (!)v 7

533528937r3 7739670528f 16 )v 9 + . . . = II + 1/2.

(!

(4.327)

The function is plotted in Fig. 4.327 for incre8lling orders in y. Given a solut ion equation, we obtain the energy E(") from Eq. (4.35) wit h

11(")

of t his

For large II, where 11(") ..... II, we recover the Bohr-Sommerfeld result (4.36). We can invert the series (4.327) and obtain v

' )[ 0.026525823 (.. ) _ ( - 11+ 2 1 + (n+~j2

0.002762954

(n+

D4

0.001299177

(n+ 1)6

+ o.oo(314009). 1 + 0.(007594)4,.97 + ... 11+ 1 11+ 1

J.

(4.329)

The results are compared with the exact ones ill Table 4. 1, which approach rapidly the BohrSommerfeld limit 0.688 253 702 ... The approach is illustrated in the right-hand part of Fig. 4.1 where log[II(")/(n + 1)4/ 3 - 1] = log[E(")/Ek1 - 11 is plotted oncc for the exact values and oncc for the semiclllSlSical expansion in Fig. 4. L The second cxeitL', it will be convenient to turn this into a local functionaL This is done as follows. We introduce an auxiliary local field · Let us introduce an auxiliary triplet of spherical coordinates =: (rb, Ob,rPb)' Then we factorize the determinant in (4.M3) as

PbPb, then Sb

We fu rt her calculate

Os.

det 8x., =

fh. 8x. 80. 8x. 8¢. 8x.

8,. 8y. 89. 8y. 8¢. 8y.

8,. 8,. 80. 8,. 8¢. 8,.

(4.544)

Long after the collision, for tb --+ 00, a small change of the starting point along t he trajectory dzo. will not affect the scattering angle. Thus we may approximate t he matrix clements in the third column by tho..o:::: &¢/8zo. ~ O. After the same amount of time the particle will only wi nd up at a slightly more distant 1'b, where drb ~ dzb. Thus we may replace the matrix element in the right upper corner by 1, so that the determinant (4.544) becomes in the limit

(4.545)

where dO. = sin ()bdOd¢b is the element of the solid angle of the emerging trajectories, and dn the area element in the x - y -plane, for which the trajectories arrive in an element of the final solid angle dO. Thus we obtain

&x.]-' ~ r~dO' 1 do [det &X"

(4.546)

The ratio dn / dO is precisely the classical differential cross section of the scattering process. Combining (4.543) and (4.545), we see that the contribution of a n individual trajectory to t he semiclassical amplitude is of the expected form [5]

Jdncl '"" f P.Pa=dOx~

.,is(p •• Pa;Ej/h -iVlf/2 .

...

(4.547)

cI ..... tr&j.

Note that this equation is also valid for some potentials which are not restricted to a finite regime around the origin , such as the Coulomb potentials. In the operator theory of quantum-mechanical scattering processes, such potentials always cause considerable problems since the outgoing wave funct ions remain d istorted even at large distances from the scattering center.

447

4.12 Semiclassical Scattering

Usually, there are only a few trajectories contribut ing t o a process with a given scatt ering angle. If the actions of t hese t rajectories differ by less than Ii, t he semiclassical approxima tion fails since the fluctuation integrals overlap. Examples arc the light scattering causing t he ordinary rainbow in nature, and glory effects seen at night around the moonlight. We now turn to a derivation of the amplitude (4.547) from t he more reliable formula (1.528) for t he interacting wave function

by isolating the factor of eip~rh Irb for large rb, as discussed a t the end of Section 1.16. On the right-hand side we now insert the x-space for m (4. 127) of the semiclassical amplitude, and use (4.87) to write

x ei rs(Xb,x..;E~)+iPGx. - i ""lzl/hl

. x~ =p~t~/M

(4.548)

Now we observe that

o)'/' [det3 (ilpo - th4 )]'/' -_[ det3 (ftxo th4 )]'/' '

-t (M

(4.549)

In Eq, (4 ,546) we have found that this determinant is equal to JdCYldfl.lrb, bringing Eq, (4 ,548) to t he form

(Xb IU[(O,t,,) lp,,)

=

lim .!. Vddcleirs(xp,,,,EG)+iP,,"" - iVlrIZlIhl

t~ __ oo

rb

dfl.

' (4.550)

x~= p.t~/M

For large Xb in the direction of the final momentum Pb, we can rewrite the exponent as [recalling (4, 240)1 (4.551)

so that (4.550) consist s of an outgoing spherical wave function eip"r.lh Iro multiplied by the scattering a mplitude (4.552)

t he same as in (4 .547).

448

4.12.2

4 Semiclassical Time Evolution Amplitude

Semiclassical Cross Section of Mott Scattering

If the scattering particle is distinguishable from the target particles, the extra phase in the semiclassical formula (4.552) docs not change the classical result (4.522). A quantum-mechanical effect becomes visible only if we consider electron-electron scattering, also referred to as Mott scattering . The potential is repulsive, and t he above Coulomb potential holds for the relative motion of the two identical particles in their center-of-mass frame. Moreover, the identity of particles requires us to add the amplitudes for the trajectories going to Pb and to - Pb [see Fig. 4.6]' so that t he differential cross section is (4.553)

The minus sign accounts for the Fermi statistics of the two electrons. For two identical bosons, we have to use a plus sign instead. Now the eikonal function S(p, p , E) enters into the result. According to Eq. (4.525), this is given by

Figure 4.6 Classical trajectories in Coulomb potential plotted in the center-of-mass frame. For identical particie:;, trajectories which merge with a scattering angle () and 7r - () arc indistinguishable. Their amplitudes must be subtracted from each other, yielding the differential cross section (4.553).

Mnoo ~+ 1 S( Pb, Pa; E) = ---p:- log ';1 + 6. _ l ' where Peo

=

(4.554)

v2M E is the impinging momentum at infinite distance, and 6. = (p~ - reo)(Pa'l - p~) r eolpb Palz .

(4.555)

The eikonai function is needed only for momenta Pb, Pa in the asymptotic regime where Pb ,Pa::::: Peo , so that d is small a nd

Mn",

S( Pb , Pa; E) -;;::: - - log 6., Poo

(4.556)

Appelldix 4A Semiclassical Quantization for Pure Power Potentials

449

which may be rewritten as Mfu:a

.

S(p , PaE) ~ 2(10 - - - log(sm 2 8/ 2) , Poo

with

Mt!CD: I [(P~ 0"0 = - 2- - og Poo

- P~)(Pa2 - p~)l ..-.4

'

(4.557)

(4.558)

1'00

and the scattering angle determined by cos8 = [Pb' P,,/PbPaj. The logarithmically diverging constant 0"0 for Pa = PbPoo does, fortunately, not depend on the scattering angle, and is t herefore the semiclassical approximation for t he phase shift at angular momentum l = O. It therefore drops, fortunately, out of the difference of the amplitudes in Eq. (4 .553) . Inserting (4.557) with (4.558) into (4 .552) and (4.553), we obtain the differential cross section for Mott scattering (sec Fig. 4.7 for a plot) dr' (1/;(r)'1j(r'» = , f"> dr , f"> dr'G;j(r, r') = 12hMAr(A r1 Jr. Jr. Jr.

J~.

1:

dr ('1i(r)'1j(r» = l : drGik, r) =

~

[t::.t

- r:] 6;j+

r;)o;j ,

~; t::.x; A Xj.

(48.14) (48.15)

T hus we ootain the se mi classical imaginary-time amplitude (4B.16)

T his agrees precisely with t he real- time amplitude (4. 265). For the partition function at inverse temperature f3 = (7b - r,,)/h, this implies the semiclassical approximation

(4Il.11) A partial integration simplifies this to

(48.18) Actually, it ill easy to calculate all terms in (4Il.16) proport ional to Vex) and its derivatives. Instead of the expansion (48.6), we evalnate

(48.19)

454

4 Semiclassical Time Evolution Amplitude

Dy rewriting V(X + 11 ) as a Fourier integral Vex

+ ,..) =

dDk (211")D V(k) cxp [i k (x + TtlJ,

J

(48.20)

we obtain

(48.2 1)

The expectation value can be ca.lculatcd using Wick's t heorem (3.307) as

f"

~

r'" dTe - k;.!('1;( ~)'1!(~»/2 =

dT(eikrj (~») =

~

f '" dTe - k;k;(A>(~,r)6;J + D'("',"')6"',6"'Jln .

~

(48.22)

where A2(r, T) , B2(T,r) a.re from (48. 12):

A2(r,r) =

M~1Jt::"T - T)T'

(r -

2

t/

B{T,T)=~ .

(48.23)

Inserting the inverse of the Fourier decomposition (48.20), (48.24)

where T] is now a time-independent variable of integration, we find

(48.25) After a quadratic completion of t he exponent, t he momentum integral can be performed and y ields

Using the transverse and longitudinal projection matrices p,L 'J

=

f),x i Llxj

(48.27)

(.6.x?'

satisfying p T2 = pT, pL2 = p T. , we can decompose Gij(r, T') as Gil' ;;; A~(T,r') Pi~

+ [A~(T, T'l + 8~(T, T')(.o.x)~l

Pi~·

(48.28)

It is t hen easy to find the determinant (48.29)

and t he inverse matrix

C- '( )

)

il' T = A2(T,T')

[,s

.o.Xi eu , ] i, -

(.6.xj2

1 .o.x,.o.Xj + A2(T, TI) + D2(T,T')(.6.xj2 (.o.xj2·

(48.30)

Inserting (4B.26) back into (48.21), and taking the correction into the exponent, we arrive at

(X1"(. Ix"Ta )

=

' /2 1"(. I(L.>X

m (X .... l ' / 2 Ta ) e - (I!hl!:·MV. '. ,

L.>X

(48.31)

Notes 8Jld

455

RcfcrellCCS

where v"m('i, '1) is the harmonically smeared potential

V"",(x, '1) == [det G('1, '1)r l /2

J

dD1) V(x

+ 'f) e-(l/2)'1,G;jL(~)'11.

(4B.32)

By expanding V (x + 'f) to second order in 'f) , the exponent in (4B.31 ) becomes (4B.33) According to Eq. (4B.15), we have (4B. 34)

so that we roobtain the first two correction terms in the curly brackets of (48.16) T he calculation of t he higher-order corrections bcoomcs quite tooious. One rewrites the cxpansion (4B.2) as V(X + 'f) (T» ) = e'l'(~)O· V (x) and (4B.19) as (4B.35) Now we apply Wick's rule (3.307) for harmonically fluct uating variables, to re-express (e'l(~)8,) = e (m{"-)'1J(~))/2 = eGjj(r,..-)8(OJ/2, (e'l.(~)O'e'l,(-r')B.) = e[('1'{"I")'1.J(r»B.0J + {'1'{"-')'1.J{..-'»O,8J+2(rn("I")'1J{..-'l)B.8J]1 2

= e [G'J ( ~,~)8,8J + G,.J (~', ~')B.O.J +2G'J ( ~. ~')8,O.J ]/2

(4B.36) Expanding the exponentials and performing the T-integrals in (4B.35) yields all desired higherorder corrections to (4B.16). For Ax = 0, the expansion has been driven to high orders in Ref. [6] (including a minimal interaction with a vector potential).

Notes and References For the eikonal expansion, see the original works by G. Wentzel, Z. Physik 38, 518 (1926); H.A. Kramers, Z. Physik 39, 828 (1926); L. Brillouin, C. R. Acad. Sci. Paris 183, 24 (1926); V.P. Mas!ov and M.V. Fcdoriuk, SemidMsirol A pproximation in Quantum Mechanics, Reidcl , Dordn,'Cht, 1982; J.B. Delos, Semida&ical Calculation of Quantum Mechanical Wave FUnctions, Adv. Chern. P hys. 65, 161 ( 1986); M.V. Berry and K.E. Mount, SemidMsirol Wave Mechanics, Rep. Prog. Phys. 35,315 (1 972); and the references quotoo in the footnotes. For the semiclassical exp8.l18ion of path integrals see R. Dashen, B. Hasslacher and A. Neveu, P hys. Rev. D 10, 4114, 4130 (1974) , R. Rajarama.n, Phys. Rep. 21C, 227 (1975); S. Coleman, Phys. Rev. D 15, 2929 (l977); and in The Whys of Su/mudear Physics , Erice Lect ures 1977, Plenum Press, 1979, 00. by A. Zichichi.

4 Semiclassical Time Evolution Amplitude

456

Recent semiclassical treatments of atomic systems are given in R.S. Manning and G.S. Ezra, Phys. Rev. 50,954 (1994). Chaos 2, 19 (1992). Scmiclassical scattering is tr 0,

(5.58)

we find an approximate ground state energy EiO) =

-~l. 27r

(5.59)

The exact value is (5.60)

469

5.6 Possible Direct Generalizations

The failure of the variational approximation is due to the fact that outside the range of the potential, the wave function is a simple exponential e- k1xl with k = ...; - 2E(0), and not a Gaussian. In fact, if we consider the expectation value of the Hamiltonian operator 1 d2 H = - - - - gt5(x) (5.61) 2dx 2 for a normalized trial wave function (5.62) we obtain a variational energy (5.63) whose minimum gives the exact ground state energy (5.60). Thus, problems of the present type call for the development of a variational perturbation theory for which Eq. (5.54) and (5.55) read (5.64) where (5.65) For an arbitrary attractive potential whose range is much shorter than a, this leads to the correct energy for a weakly bound ground state El(0) ~

5.6

-"21 [1 -

00

-00

dxoI V ( xo )] 2

(5.66)

Possible Direct Generalizations

Let us remark that there is a possible immediate generalization of the above variational procedure. One may treat higher components Xm with m > 0 accurately, say up to m = in - 1, where in is some integer> 1, using the ansatz

(5.67)

470

5 Variational Perturbation Theory

with the trial function Lm:

and a smearing square width of the potential a~ m

(5.69)

For the partition function alone the additional work turns out to be not very rewarding since it renders only small improvements. It turns out that in the low-temperature limit T -> 0, the free energy is still equal to the optimal expectation of the Hamiltonian operator in the Gaussian wave packet (5.49). Note that the ansatz (5.7) [as well as (5.67)] cannot be improved by allowing the trial frequency n(xo) to be a matrix nmm,(XO) in the space of Fourier components Xm [i.e., by using Lm,m' n mm, (XO)x;;'xm' instead of n(xo) Lm IxmI2 ]. This would also lead to an exactly integrable trial partition function. However, after going through the minimization procedure one would fall back to the diagonal solution nmm,(XO) = .=0

Inserting this into the right-hand side of (5.98), we find

VaL,aT (rO)

2

= [ v (rO

+ 0>. ) . /

1

2

1

VI - 2aLA /1 - 2afA V

D 1 e2r5>.2ai/(1-2ai>.l]

(5.101) >.=0

which has the expansion

+ v'(r5)[a"i + (D -l)a~l ~v"(r5)[3ai + 2(D - l)a"ia~ + (D2 -

VaL,aT(rO) = v(r5) +

l)af + 4r5a"il

+ ....

(5.102)

The prime abbreviates the derivative with respect to r5. In general it is useful to insert into (5.98) the Fourier representation for the potential

- J(27r)De dDk

V(x) -

ikx

(5.103)

V(k),

which makes the x-integration Gaussian, so that (5.99) becomes

Vai,aj,(ro)

=

dDk J(27r)D V(k)

(

a'i

2

a'f

2

.

)

exp -2"kL - 2"kT - zrOkL .

Exploiting the rotational symmetry of the potential by writing V(k) decompose the measure of integration as

(5.104)

== v(k 2 ),

we

(5.105)

481

5.10 Application to Coulomb and Yukawa Potentials

where

27rD/2 SD = f(D/2)

(5.106)

is the surface of a sphere in D dimensions, and further with kL = k cos ¢>, kT = k sin¢>: dDk SD-l D- l (5.107) (27r )D = (27r )D -1 d cos ¢> Jo dk k .

11

J

roo

This brings (5.104) to the form

(SD-)~ 27r

Va2 a2 (ro) = L' T

X

11

- 1

du1°O dkk D- 1V(k 2) - 1

exp {-~ [( aiu 2 + a~(1- u 2)) k 2 - irouJ} .

(5.108)

The final effective classical potential is found by minimizing W 1 (ro) at each ro in aL, aT, ~h , nT . To gain a rough idea about the solution, it is usually of advantage to study first the isotropic approximation obtained by assuming a~(ro) = ai(ro), and to proceed later to the anisotropic approximation.

5.10

Application to Coulomb and Yukawa Potentials

The effective classical potential can be useful also for singular potentials as long as the smearing procedure makes sense. An example is the Yukawa potential

V(r)

=

_(e 2/r)e- mr ,

which reduces to the Coulomb potential for m

(5.109)

== O. Using the Fourier representation (5.110)

we easily calculate the isotropic ally smeared potential

Va2(ro)

=

J.J27ra2

- e2

d3 x

3

[1

exp --(Xo - x) 2] V(r) 2a2

_e 2 27rem2a2 /21~ da'2 1 3 exp ( -r~/2a'2 - m 2a'2 /2) a .J27ra,2 2 2 _e2 2em a /2 _1 (0 / ..;20,2 a dte-(t2+m2r5/4t2).

Vir

ro Jo

(5.111)

In the Coulomb limit m ---.. 0, the smeared potential becomes equal to the Coulomb potential multiplied by an error function , (5.112)

482

5 Variational Perturbation Theory

where the error function is defined by erf(z)

=V7r2 iorz dxe-

X

2

(5.113)



The smeared potential is no longer singular at the origin, (5.114) The singularity has been removed by quantum fluctuations. In this way the effective classical potential explains the stability of matter in quantum physics, i.e. , the fact that atomic electrons do not fall into the origin. The effective classical potential of the Coulomb system is then by the isotropic version of (5.97) !l

3

WI (Xo) = 73 ln

sinh[n,60(xo)/2]

n,60(Xo)/2

e2

[ixol

1

3

2

2

- ~ erf J2a 2(xo) - "2 MO (xo)a (xo). (5.115)

Minimizing WI(ro) with respect to a2(ro) gives an equation analogous to Eq. (5.37) determining the frequency 02(ro) to be 2

_

o (ro) -

2~_1___1_ -r5/2a2

e 3 J21r (a2)3/2e

.

(5.116)

We have gone to atomic units in which e = M= n= kB= 1, so that energies and temperatures are measured in units of Eo = Me 4 /Mn 2 ~ 4.36 x lO-llerg ~ 27.21eV and To = Eo/kB ~ 31575K , respectively. Solving (5.116) together with (5.24) , we find a2 (ro) and the approximate effective classical potential (5.32). The result is shown in Fig. 5.9 as a dashed curve. The above approximation may now be improved by treating the fluctuations anisotropic ally, as described in the previous section, with different 02(XO) for radial and tangential fluctuations OXL and OXT, and the effective potential following Eqs. (5.97) and (5.98). For the anisotropically smeared Coulomb potential we calculate from (5.108):

Va L4(ro) =

-IfII

du J ai u2 +

e:f(l _u2)

exp

{-" 2-;-[a-; i'- u-: -2_+- ,r5'" 'a}".2.(""1-_-U-::-2~)]}' (5.117)

Introducing the variable ,\ = Jifu/ Jai,\2 + af(l - ,\2), which runs through the same interval [-1 ,1] as u, we rewrite this as (5.118)

483

5.10 Application to Coulomb and Yukawa Potentials

T=lxlO' K

~"-::;::;;.o"'~

4

Figure 5.9 Approximate effective classical potential W1(ro) of Coulomb system at various temperatures (in multiples of 104 K). It is calculated once in the isotropic (dashed curves) and once in the anisotropic approximation. The improvement is visible in the insert which shows WdV. The inverse temperature values of the different curves are j3 = 31.58, 15.78, 7.89, 3.945, 1.9725, 0.9863, 0 atomic units, respectively. Extremization of W1 (ro) in Eq. (5.97) yields the equations for the trial frequencies

~V2 84 aT,a

nT2 (ro)

2 L

= e2

raIl

y~

1 d)' ).2 exp [-(r5l 2a1J).2]

[aiC1 - ).2) + a}).2J2'

-1

82 + -31[n2L + 2nT2] = -31 [ uaL !l

(5 .119)

82 ] Va 2 a 2 uaT T' L

2!l

22 1 1 2 2) e -3 ~ r:2:4 exp( -rO/2a L . v27r yaia}

(5.120)

These equations have to be solved together with 2

(

) _

1

aL,T ro - /3ni,T(rO)

[/3n L,T(r o) 2

co

th /3n L,T(rO) _ 2

1] .

(5.121)

Upon inserting the solutions into (5.97) , we find the approximate effective classical potential plotted in Fig. 5.9 as a solid curve. Let us calculate the approximate particle distribution functions using a three-dimensional anisotropic version of Eq. (5.94) . With the potential W 1 (ro), we arrive at the integral [6]

(5.122)

484

5 Variational Perturbation Theory

2

4

8

6

10

12

14

r

Figure 5.10 Particle distribution g(r) == v27r(i p(r) in Coulomb potential at different temperatures T (the same as in Fig. 5.9), calculated once in the isotropic and once in the anisotropic approximation. The dotted curves show the classical distribution. For low and intermediate temperatures the exact distributions of R.G. Storer, J. Math. Phys. 9, 964 (1968) are well represented by the two lowest energy levels for which p(r)=7r- 1c 2T e,6/2+(1/87r)(1 - r + r2/2)e- Te,6/8+(1/387r)(243 - 324r + 216r2-48r3+4r4)e- 2T/3e,6/18.

The resulting curves to different temperatures are plotted in Fig. 5.10 and compared with the exact distribution as given by Storer. The distribution obtained from the earlier isotropic approximation (5.116) to the trial frequency [l2(ro) is also shown.

5.11

Hydrogen Atom in Strong Magnetic Field

The recent discovery of magnetars [7] has renewed interest in the behavior of charged particle systems in the presence of extremely strong external magnetic fields. In this new type of neutron stars, electrons and protons from decaying neutrons produce magnetic fields B reaching up to 1015 G, much larger than those in neutron stars and white dwarfs, where B is of order 1010 - 1012 G and 106 - 108 G, respectively. Analytic treatments of the strong-field properties of an atomic system are difficult, even in the zero-temperature limit. The reason is a logarithmic asymptotic behavior of the ground state energy to be derived in Eq. (5.132). In the weak-field limit , on the other hand, perturbative approaches yield well-known series expansions

485

5.11 Hydrogen Atom in Strong Magnetic Field

in powers of B2 up to B 60 [8]. These are useful, however, only for B « B o, where Bo is the atomic magnetic field strength Bo = e3 M2 In 3 ::::; 2.35 x 105 T = 2.35 X 109 G. So far, the most reliable values for strong uniform fields were obtained by numerical calculations [9]. The variational approach can be used to derive a single analytic expression for the effective classical potential applicable to all field strengths and temperatures [10]. The Hamiltonian of the electron in a hydrogen atom in a uniform external magnetic field pointing along the positive z-axis is the obvious extension of the expression in Eq. (2.644) by a Coulomb potential: H (p , x )

12M 2 2 = 2M P + 2WBX -

wBlz

e2

()

p , x -~ '

(5.123)

where WB denotes the B-dependent magnetic frequency wL/2 eB 12Mc of Eq. (2.649) , i.e., half the Landau or cyclotron frequency. The magnetic vector potential has been chosen in the symmetric gauge (2.638). Recall that lz is the z-com ponent of the or bi tal angular momentum l z (p, x) = (x x p) z [see (2.645)] . At first , we restrict ourselves here to zero temperature. From the imaginary-time version of the classical action (2.638) we see that the particle distribution function in the orthogonal direction of the magnetic field is, for Xb = 0, Yb = 0, proportional to (5.124) This is the same distribution as for a transverse harmonic oscillator with frequency Being at zero temperature, the first-order variational energy requires knowing the smeared potential at the origin. Allowing for a different smearing width an and al along an orthogonal to the magnetic field, we may use Eq. (5.118) to write

WB.

(5.125) Performing the integral yields 1

2

all

2) 2 arccosh-

27r (a ll - a.l

a.l

(5.126)

Since the ground state energies of the parallel and orthogonal oscillators are and 2 x 0.l/2, we obtain immediately the first-order variational energy

°11/2

(5.127)

°

with wil = and an,.l = 1/20Il,.l· In this expression we have ignored the second term in the Hamiltonian (5.123), since the angular momentum lz of the ground state must have a zero expectation value.

486

5 Variational Perturbation Theory

For very strong magnetic fields, the transverse variational frequency become equal to WB, such that in this limit

nT

will

(5.128) Extremizing this in

nil

yields n 4e4 HII :=::;-

7r

I

2 27rW B

og - , e4

(5.129)

and thus an approximate ground state energy E 1(O)

:=::; WB -

e4

-

7r

I

og

2 27rWB --4-'

e

(5.130)

The approach to very strong fields can be found by extremizing the energy (5.127) also in n1-. Going over to atomic units with e = 1 and m = 1, where energies are measured in units of fO = Me 4/!i2 == 2Ryd :=::; 27.21eV, temperatures in fo/kB :=::; 3.16 X 105 K, distances in Bohr radii aB = !i2/ M e2 :=::; 0.53 x 10- 8 cm, and magnetic field strengths in Bo = e3 M2 /!i 3 :=::; 2.35 X 105 T = 2.35 X 109 G, the extremization yields

~ y7r 2 ( 2a a2 ) yn1-= InB-2InlnB+lnB+ln2B+b

+ O(ln- 3 B) ,

(5.131 ) with abbreviations a = 2 - In2 :=::; 1.307 and b = In(7r/2) - 2 :=::; -1.548. The associated optimized ground state energy is, up to terms of order In- 2B, E(O)(B) = B _ .! {ln2 B - 4ln B lnln B + 4ln 2ln B - 4blnln B + 2(b + 2) In B + b2 2 7r -ln1B [8In2InB-8blnlnB+ 2b2] } +O(ln- 2B). (5.132)

The prefactor 1/7r of the leading In 2B-term using a variational ansatz of the type (5.64), (5.65) for the transverse degree of freedom is in contrast to the value 1/2 calculated in the textbook by Landau and Lifshitz [11] . The calculation of higher orders in variational perturbation theory would drive our value towards 1/2. The convergence of the expansion (5.132) is quite slow. At a magnetic field strength B = 105 B o, which corresponds to 2.35 x 1010 T = 2.35 X 1014 G, the contribution from the first six terms is 22.87 [2 Ryd]. The next three terms suppressed by a factor In- 1 B contribute -2.29 [2 Ryd], while an estimate for the In- 2B-terms yields nearly -0.3 [2 Ryd]. Thus we find c;(1)(10 5 ) = 20.58 ± 0.3 [2 Ryd]. Table 5.2 lists the values of the first six terms of Eq. (5.132). This shows in particular the significance of the second term in (5.132), which is of the same order of the leading first term, but with an opposite sign.

487

5.11 Hydrogen Atom in Strong Magnetic Field

Table 5.2 Example for competing leading six terms in large-B expansion (5.132) at B = 105 Bo ~ 2.35 X 1014 G. (l/rr)ln B

- (4/rr)lnBlnlnB

(4/rr) In InB

- (4b/rr)lnlnB

42.1912

-35.8181

7.6019

4.8173

[2(b+2)/rr]lnB 3.3098

b2 /7r 0.7632

30 ,-----------~------------~

Landau:-log 2 B2/2 ------

---

f(B)/2Ryd

10

20

10

,,(1) (B)/2Ryd

°O~----------~l~~=---------­

500

BIBo

1000 BIBo

1500

2000

Figure 5.11 First-order variational result for binding energy (5.133) as a function of the strength of the magnetic field. The dots indicate the values derived in the reference given in Ref. [12] . The long-dashed curve on the left-hand side shows the simple estimate 0.51n2 B of the textbook by Landau and Lifshitz [11]. The right-hand side shows the successive approximations from the strong-field expansion (5.134) for N = 0,1,2,3,4, with decreasing dash length. Fat curve is our variational approximation.

The field dependence of the binding energy

E(B)

=B

_

E(O)

(5.133)

2

is plotted in Fig. 5.11, where it is compared with the results of other authors who used completely different methods, with satisfactory agreement [12] . On the strongcoupling side we have plotted successive orders of a strong-field expansion [24]. The curves result from an iterative solution of the sequence of implicit equations for the quantity weB) = VE(4B)/2 for N = 1,2,3,4: (5.134) where

a4=

Pi y7rD VB 2 ' (5.135)

488

5 Variational Perturbation Theory

JY2+I

and D denotes the integral D == l' - 21000 dy (y / -1) log y ~ -0.03648, where l' ~ 0.5773 is the Euler-Mascheroni constant (2.467). Our results are of similar accuracy as those of other first-order calculations based on an operator optimization method [25] . The advantage of our variational approach is that it yields good results for all magnetic field strengths and temperatures, and that it can be improved systematically by methods to be developed in Section 5.13, with rapid convergence. The figure shows also the energy of Landau and Lifshitz which grossly overestimates the binding energies even at very large magnetic fields, such as 2000Bo ex: 10 12 G. Obviously, the nonleading terms in Eq. (5.132) give important contributions to the asymptotic behavior even at such large magnetic fields. As an peculiar property of the asymptotic behavior, the absolute value of the difference between the Landau-Lifshitz result and our approximation (5.132) diverges with increasing magnetic field strengths B. Only the relative difference decreases.

5.11.1

Weak-Field Behavior

Let us also calculate the weak-field behavior of the variational energy (5.127) . Setting 1] == 0 11 / 0 1- , we rewrite W1 (O) as W 1 (0 ) -- -01- ( 1 + -1])

2

2

B2 + V 1]011 - 1n 1 - J1=17 . + ---8017r J1=17 1 + J1=17

(5.136)

This is minimized in 1] and 01- by expanding 1](B) and O(B) in powers of B2 with unknown coefficients, and inserting these expansions into extremality equations. The expansion coefficients are then determined order by order. The optimal expansions are inserted into (5.136), yielding the optimized binding energy c(1)(B) as a power series 00

W 1 (O)

=

L

cnB2n.

(5.137)

n=O

The coefficients Cn are listed in Table 5.3 and compared with the exact ones. Of course, the higher-order coefficients of this first-order variational approximation become rapidly inaccurate, but the results can be improved, if desired, by going to higher orders in variational perturbation theory of Section 5.13.

5.11.2

Effective Classical Hamiltonian

The quantum statistical properties of the system at an arbitrary temperature are contained in the effective classical potential Heif cl (Po , xo) defined by the threedimensional version of Eq. (3.821):

B(po, xo) ==e-,l3Heffc1 {po ,x o) ==

f f (~~3 V 3x

J(3)(XO - X) (27rnlJ(po - p) e-A[p,xlln, (5.138)

489

5.11 Hydrogen Atom in Strong Magnetic Field

Table 5.3 Perturbation coefficients up to order B6 in weak-field expansions of variational parameters, and binding energy in comparison to exact ones (from J.E. Avron et al. and B.G. Adams et al. quoted in Notes and References). n

0 1.0

'In

fin en €~x

1 40571"2 - - - "" -0.5576 7168

32 - "" 1.1318 971" 4 - - "" - 0.4244 371"

9971" "" 1.3885 224 971" 128 "" 0.2209

- 0.5

2

_ 129397571"3 "" - 2.03982 19668992 801971"3 - - - "" - 0.1355 1835008

-~ "" 192

0.25

3 388699933207571"6 "" -4.2260 884272562962432

1682896571"4 "" 1.3023 1258815488

0.2760

52443166718771"5 "" 5.8077 27633517592576 25644980771"5 "" 0.2435 322256764928 5581 "" 1.2112 4608

where Ae[p , x] is the Euclidean action Ae[p,x]

(",(3 = Jo dT [-ip(T)X(T) + H(p(T),X(T))],

(5.139)

and x = fl,(3 dTX(T)/h(J and p = fOIi(3 dTp(T)lh(J are the temporal averages of position and momentum. Note that the deviations ofp(T) from the average Po share with x( T) - Xo the property that the averages of the squares go to zero with increasing temperatures like liT, and remains finite for T -+ O. while the expectation of p2 grows linearly with T (Dulong-Petit law). For T -+ 0, the averages of the squares of p( T) remain finite. This property is the basis for a reliable accuracy of the variational treatment. Thus we separate the action (5.139) (omitting the subscript e) as Ae[p, x]

= (JH(po, xo) + Aft XO [p, x] + A int [p, x],

(5.140)

where Aho,xO [p, x] is the most general harmonic trial action containing the magnetic field. It has the form (3.825), except that we use capital frequencies to emphasize that they are now variational parameters:

r (3

1

li

AhO,XO[P,x]

= Jo dT { - i[p(T) - Po] ' X(T) + 2M [p(T) - PO]2

+DBlAp(T)-Po,X(T)-Xo)

+ ~ Di [X.l(T)-xo.d 2+ ~ Drr[Z(T) - ZO]2}. (5.141)

The vector x.l = (x, y) is the projection of x orthogonal to B . The trial frequencies n = (DB, D.l, DII) are arbitrary functions of Po, XQ, and B. Inserting the decomposition (5 .140) into (5.138) , we expand the exponential of the interaction, exp {-AntlP, x] /h}, and obtain a series of expectation values of powers of the interaction (Annt[P, x] )hO ,XO, defined in general by the path integral ( O[p, x]

) .,~O,XO

=

zrt1

XO

f'1"\3

1)3p O[

v x(27rh)3

x exp { -~Aho,xO [p, x] } ,

] ,(3) (

p ,X u

-) (2 t 0 like (_l)N-l g N x (1 / 0)3N-I. Since the trial energy goes for large 0 to positive infinity, only the odd approximations are guaranteed to possess a minimum. The second-order approximation W 2 can nevertheless be used to find an improved energy value. As shown by Fig. 5.14, the frequency of least dependence O2 is well defined. It is the frequency where the O-dependence of W 2 has its minimal absolute value. Thus we optimize 0 with the condition

a2 wf

a0 2 = O.

(5 .193)

This leads to the energy values E~O)(g) listed in Table 5.6. They are more accurate than the values EiO) (g) by an order of magnitude.

5.14 Applications of Variational Perturbation Expansion

501

0.8 0.75

\

N=l

\

0.7

\

N=2 /

\

0.65

\

W 0.6 0.55 0.5 0.45 0.4

1

1.5

2

2.5

3

Figure 5.14 Typical O-dependence of approximations W1 ,2,3 at T = O. The coupling constant has the value 9 = 004. The second-order approximation wf has no extremum. Here the minimal O-dependence lies at the turning point, and the condition 8 2 Wf / 80 2 renders the best approximation to the energy (short dashes).

5.14.2

Anharmonic Oscillator for T

>0

Consider now the anharmonic oscillator at a finite temperature, where the expansion (5.183) consists of the sum of one-particle irreducible vacuum diagrams

Wr=-f3

1{

I?!

?!

--O+CD 2 3

?! --1 ( CJ:J:) +

2!

72

(5.194) 24

-

?!

2592

1728

+~( ciJ:o +V

?!

+

@ 3456

1728

648

648

The vertices represent the couplings 93(XO)/3!n, 94(xO)/4!n, whereas the lines stand for the correlation function G'~o (7, 7'). The numbers under each diagram are their multiplicities acting as factors . Only the five integrals associated with the diagrams

o

502

5 Variational Perturbation Theory

n

need to be evaluated explicitly; all others arise by the expansion of or by factorization. The explicit form of three of these integrals can be found in the first, forth, and sixth of Eqs. (3.547). The results of the integrations are listed in Appendix 5A. Only the second and fourth diagrams are new since they involve vertices with three legs. They can be found in Eqs. (3D.7) and (3D.8). In quantum field theory one usually calculates Feynman integrals in momentum space. At finite temperatures, this requires the evaluation of multiple sums over Matsubara frequencies. The present quantum-mechanical example corresponds to a D = 1 -dimensional quantum field theory. Here it is more convenient to evaluate the integrals in T-space. The diagrams

for example, are found by performing the integrals

Ti(3a 2 Ti(3

(~) 2 a~2

The factor Ti(3 on the left-hand side is due to an overall T-integral and reflects the temporal translation symmetry of the system; the factors l/w arise from the remaining T-integrations whose range is limited by the correlation time l/w. In general, the (3- and xo-dependent parameters a~L have the dimension of a length to the nth power and the associated diagrams consist of m vertices and n/2 lines (defining == 2 ). We now use the rule (5.188) to replace w by and expand everything in powers of 92 up to the third order. The expansion can be performed diagrammatically in each Feynman diagram. Letting a dot on a line indicate the coupling 92/2Ti, the one-loop diagram is expanded as follows:

ai a

n

fl

0=0-20+0

1

3



The other diagrams are expanded likewise: fl

CD

1 1 CD +-6 CD +-6 CD,

1

2 E3

1 - 6

9

1 -6

CD 1

fl

2 E3

-

-

fl

1

-

2

9

1

-

2

E3 ,

g,

(5.195)

503

5.14 Applications of Variational Perturbation Expansion fl

lo::D

2

In this way, we obtain from (5.194) the complete graphical expansion for W:P(xo) including all vertices associated with the coupling g2(XO) :

jJW:P(xo)

=

-~ 0 +

-~[~ 0

+~[

0

+3

G0

+~

+~

CD

+3G

CD )

+~

o::D + 2~

9

+~

E7 ]

CD +~ CD)

(~4 (XX) + ~4 c:x:::D + ~6 \::::f ~ + ~2 E3 )

+ (136

CXXD

+~ cg +~ @ +~@

+~O +~g)l

(5.196)

In the latter diagrams, the vertices represent directly the couplings gn/n. The denominators n! of the previous vertices gn/n!n have been combined with the multiplicities of the diagrams yielding the indicated prefactors. The corresponding analytic expression for W:P(xo) is

(5.197)

The quantities a~L are ordered in the same way as the associated diagrams in (5.196). As before, we have omitted the variable Xo in all but the first three terms, for brevity. The optimal trial frequency O(xo) is found numerically by searching, at each value of Xo , for the real roots of the first derivate of W:P(xo) with respect to O(xo). Just as for zero temperature, the solution happens to be unique.

504

5 Variational Perturbation Theory

By calculating the integral (5 .198)

we obtain the approximate free energy (5 .199)

The results are listed in Table 5.7 for various coupling constants 9 and temperatures. They are compared with the exact free energy (5.200) whose energies En were obtained by numerically solving the Schrodinger equation.

9 0.002 0.4

2.0

4.0 20

200 2000

80000

f3 2.0 1.0 5.0 1.0 5.0 10.0 1.0 5.0 1.0 5.0 10.0 5.0 0.1 1.0 10.0 0.1 3.0

Fl 0.427937 0.226084 0.559155 0.492685 0.699431 0.700934 0.657396 0.809835 1.18102 1.24158 1.24353 2.54587 2.6997 5.40827 5.4525 18.1517 18.50l

F3 0.427937 0.226075 0.558678 0.492578 0.696180 0.696285 0.6571051 0.803911 1.17864 1.22516 1.22515 2.50117 2.69834 5.32319 5.3225 18.0470 18.146

Fex 0.427741 0.226074 0.558675 0.492579 0.696118 0.696176 0.6571049 0.803758 1.17863 1.22459 1.22459 2.49971 2.69834 5.31989 5.3199 18.0451 18.137

Table 5.1 Free energy of anharmonic oscillator with potential V(x) various coupling strengths 9 and f3 = l/kT.

= x 2 /2 + gx 4 /4

for

We see that to third order, the new approximation yields energies which are better than those of FI by a factor of 30 to 50. The remaining difference with respect to the exact energies lies in the fourth digit . In the high-temperature limit, all approximations WN(xo) tend to the classical result V(xo), as they should. Thus, for small (3, the approximations W3 (xo) and WI (xo) are practically indistinguishable.

505

5.15 Convergence of Variational Perturbation Expansion

The accuracy is worst at zero temperature. Using the T --+ 0 -limits of the Feynman integrals a~L given in (3.550) and (3D.14), the approximation wf takes the simple form

W:f(xo) = V(xo)

+ liO;xo) + ~a2 + ~a4 [g~ a4 + g2g4 a6 + g~ ~a6 + gl ~a8]

_ _ 1_ 211,0

2

2

6 3

24 2

(5.201)

1 [ 33 6 24 8 2 8 2 13 10 235 10 337 12] +611,202 g22 a +g2g33a +g2 g4 3a +g3g43 a +g2g416a +g464 a , where a2 = li/2MO. As in (5.197), we have omitted the arguments Xo in 0 and gi , for brevity. At zero temperature, the remaining integral over Xo in the partition function (5.198) receives its only contribution from the point Xo = 0, where W3(0) is minimal. There it reduces to the energy W3 of Eq. (5.192) .

5.15

Convergence of Variational Perturbation Expansion

For a single interaction x P , the approximation W N at zero temperature can easily be carried to high orders [13, 14]. The perturbation coefficients are available exactly from recursion relations, which were derived for the anharmonic oscillator with p = 4 in Appendix 3C . The starting point is the ordinary perturbation expansion for the energy levels of the anharmonic oscillator

E(n) = w ~ E(n) ~ k=O

k

(L) k 4w

(5.202)

3

It was remarked in (3C.27) and will be proved in Section 17.10 [see Eq. (17.323)] that the coefficients ELn) grow for large k like

(n) Ek

1 f;12n

---+ - 7r

- - I (-3) 7r

n.

k r(k + n + 1/2) .

(5.203)

Using Stirling's formula3 (5.204) this amounts to

E(n) k

---+ _

2vf3 (_4)n 7r

n!

[-3(k + e

n)] n+k

Ek

(5.205)

Thus, n ) grows faster than any power in k. Such a strong growth implies that the expansion has a zero radius of convergence. It is a manifestation of the fact that the energy possesses an essential singularity in the complex g-plane at the expansion point g = O. The series is a so-called asymptotic series. The precise form of the 3M. Abramowitz and 1. Stegun, op. cit., Formulas 6.1.37 and 6.1.38.

506

5 Variational Perturbation Theory

I

I

\ I

N=l

:

\

\ I

0.7

N=3

\ I

\

W

\

0.65

: :

N=5

\ I \ II

\1I

0.6

N=ll

\

3.5

4

0.59 r-~-~-----ry---~'--~-r-~---'

N=2

0.58 0.57

W 0.56 . 0.55

1

2.5

2

1.5

3

n

3.5

4

Figure 5.15 Typical !1-dependence of Nth approximations WN at T = 0 for increasing orders N . The coupling constant has the value g/4 = 0.1. The dashed horizontal line indicates the exact energy.

singularity will be calculated in Section 17.10 with the help of the semiclassical approximation. If we want to extract meaningful numbers from a divergent perturbation series such as (5.202) , it is necessary to find a convergent resummation procedure. Such a procedure is supplied by the variational perturbation expansion, as we now demonstrate for the ground state energy of the anharmonic oscillator.

Truncating the infinite sum (5.202) after the Nth term, the replacement (5.188) followed by a re-expansion in powers of g up to order N leads to the approximation WN at zero temperature:

n_

~

(0) (

wN-n~Cl 1=0

g

4fP

)1 ,

(5.206)

507

5.15 Convergence of Variational Perturbation Expansion

w

N=21

0.55914625

1.4

N=l1 1.6

2.2

2

n

Figure 5.16 New plateaus in W N developing for higher orders N 2: 15 in addition to the minimum which now gives worse results. For N = 11 the new plateau is not yet extremal, but it is the proper region of least !1-dependence yielding the best approximation to the exact energy indicated by the dashed horizontal line. The minimum has fallen far below this value and is no longer useful. The figure looks similar for all couplings (in the plot, 9 = 0.4). The reason is the scaling property (5.215) proved in Appendix 5B.

with the re-expansion coefficients Era) =

t

EjO) ( (1

~~j!/2

) (-40-)/-j.

(5.207)

J

J=O

Here 0- denotes the dimensionless function of D

0- == _~ Dr2 = D(D2 - 1) . 2

9

(5.208)

In Fig. 5.15 we have plotted the D-dependence of W N for increasing N at the coupling constant g/4 = 0.1. For odd and even N, an increasingly flat plateau develops at the optimal energy. At larger orders N :2: 15, the initially flat plateau is deformed into a minimum with a larger curvature and is no longer a good approximation. However, a new plateau has developed yielding the best energy. This is seen on the high-resolution plot in Fig. 5.16. At N = 11, the new plateau is not yet extremal but close to the correct energy. The worsening extrema in Fig. 5.16 correspond here to points leaving the optimal dashed into the upward direction. The newly forming plateaus lie always on the dashed curve. The set of all extremal DN-values for odd N up to N = 91 is shown in Fig. 5.17. The optimal frequencies with smallest curvature are marked by a fat dot. In Subsection 17.10.5. we shall derive that

(5.209)

508

5 Variational Perturbation Theory

10

8

6 ON

4

/.~

.;:-e/ ---

2

00

20

40

60

80

100

N

Figure 5.17 Trial frequencies ON extremizing the variational approximation WN at T = 0 for odd N ~ 91. The coupling is g/4 = 1. The dashed curve corresponds to the approximation (5.211) [related to ON via (5.208)]. The frequencies on this curve produce the fastest convergence. The worsening extrema in Fig. 5.16 correspond here to points leaving the optimal dashed into the upward direction. The newly forming plateaus lie always on the dashed curve.

grows for large N like IJN :;::,

eN,

e = 0.186047 .. . ,

(5.210)

so that ON grows like ON :;::, (eN g)1/3. For smaller N, the best ON-values in Fig. 5.17 can be fitted with the help of the corrected formula (5.210): IJN :;::,

6.85) eN ( 1 + N2/3 .

(5.211)

The associated ON-curve is shown as a dashed line. It is the lower envelope of the extremal frequencies. The set of extremal and turning point frequencies ON is shown in Fig. 5.18 for even and odd N up to N = 30. The optimal extrema with smallest curvature are again marked by a fat dot. The theoretical curve for an optimal convergence calculated from (5.211) and (5.209) is again plotted as a dashed line. In Table 5.8, we illustrate the precision reached for large orders N at various coupling constants 9 by a comparison with accurate energies derived from numerical solutions of the Schrodinger equation. The approach to the exact energy values is illustrated in Fig. 5.19 which shows that a good convergence is achieved by using the lowest of all extremal frequencies, which lie roughly on the dashed theoretical curve in Fig. 5.17 and specify the position of the plateaus. The frequencies ON on the higher branches leaving the dashed curve in that figure, on the other hand, do not yield converging energy values. The

509

5.15 Convergence of Variational Perturbation Expansion 6

,-------------~--------------~------------_,

4

2

o

10

20

N

30

Figure 5.18 Extremal and turning point frequencies ON in variational approximation W N at T = 0 for even and odd N :S 30. The coupling is 9/4 = 1. The dashed curve corresponds to the approximation (5.211) [related to ON via (5.208)].

-\ 10. 5

9/4 = 1

~:::::::::::::::::::::::::::::::::: ......-~

~ . .

~ BBBBBBBBBBBBBD

----.......'.-.-.-.... .. -

10.20

o

20

40

N

60

80

100

Figure 5.19 Difference between approximate ground state energies E = WN and exact energies Eex for odd N corresponding to the ON-values shown in Fig. 5.17. The coupling is 9/4 = 1. The lower curve follows roughly the error estimate to be derived in Eq. (17.409). The extrema in Fig. (5.17) which move away from the dashed curve lie here on horizontal curves whose accuracy does not increase.

sharper minima in Fig. 5.16 correspond precisely to those branches which no longer determine the region of weakest O-dependence. The anharmonic oscillator has the remarkable property that a plot of the ONvalues in the N, (IN-plane is universal in the coupling strengths g; the plots do not depend on g. To see the reason for this, we reinsert explicitly the frequency w (which

510

5 Variational Perturbation Theory N 1 2 3 4 5 10 15 20 25

exact

N 1 2 3 4 5 10 15 20 25

exact

g/4 =0.1 0.5603073711 0.5591521393 0.5591542188 0.5591457408 0.5591461596 0.5591463266 0.5591463272 0.5591463272 0.5591463272 0.5591463272 g/4 =50 2.5475803996 2.5031213253 2.5006996279 2.4995980125 2.4996213227 2.4997071960 2.4997089403 2.4997089079 2.4997087731 2.4997087726

g/4 = 0.3 0.6416298621 0.6380887347 0.6380357598 0.6379878713 0.6379899084 0.6379917677 0.6379917838 0.6379917836 0.6379917832 0.6379917832 g/4 =200 4.0084608812 3.9365586048 3.9325538203 3.9307488127 3.9307857892 3.9309286743 3.9309316283 3.9309315732 3.9309313396 3.9309313391

g/4 =0.5 0.7016616429 0.6963769499 0.6962536326 0.6961684978 0.6961717475 0.6961757782 0.6961758231 0.6961743059 0.6961758208 0.6961758208 g/4 =1000 6.8279533136 6.7040032606 6.6970328638 6.6939036178 6.6939667971 6.6942161680 6.6942213631 6.6942212659 6.6942208522 6.6942208505

g/4 = 1.0 0.8125000000 0.8041900946 0.8039140528 0.8037563457 0.8037615232 0.8037705329 0.8037706596 0.8037706575 0.8037706513 0.8037706514 9/4 =8000 13.635282593 13.386598486 13.372561189 13.366269038 13.366395347 13.366898079 13.366908583 13.366908387 13.366907551 13.366907544

g/4 = 2.0 0.9644035598 0.9522936298 0.9517997694 0.9515444198 0.9515517450 0.9515682249 0.9515684933 0.9515684887 0.9515584121 0.9515684727 9/4 =20000 18.501658712 18.163979967 18.144908389 18.136361642 18.136533060 18.137216200 18.137230481 18.137230214 18.137230022 18.137229073

Table 5.8 Comparison of the variational approximations W N at T = 0 for increasing N with the exact ground state energy at various coupling constants g.

was earlier set equal to unity). Then the re-expanded energy WN in Eq. (5.206) has the general scaling form (5.212) where WN is a dimensionless function of the reduced coupling constant and frequency , g

g

== ~V'

, W

W

== n'

(5.213)

respectively. When differentiating (5.212), d WN dO

d - 2,2 d] = [ 1 - 3'g dg W c1W 2

WN

('g, W'2) ,

(5.214)

we discover that the right-hand side can be written as a product of gN and a dimensionless polynomial of order N depending only on a = 0(0 2 - w 2 )/g: d Wrl

dO

N

= g'N PN (a) .

(5.215)

A proof of this will be given in Appendix 5B for any interaction x p . The universal optimal aN-values are obtained from the zeros of pN(a). It is possible to achieve the same universality for the optimal frequencies of the even approximations WN by determining them from the extrema of pN(a) rather than from the turning points of W N as a function of o. The universal functions pN(a) are found most easily by replacing the variable a in the coefficients of the re-expansion (5.206) by its W = 0 -limit alw=o = 0 3 / g = l/g . This yields the simpler expression

dO)

W~ = OWN(g, 0) = 0 f. E}O) (~) 1, l=O

(5.216)

511

5.15 Convergence of Variational Perturbation Expansion

with c}O) =

t E?) ( (1 ~~j!/2 ) (-4/flY-j .

(5.217)

J

)=0

The derivative of WN with respect to

n yields (5.218)

In Section 17.10, we show the re-expansion coefficients cia) in (5.207) to be for large k proportional to EkO): e(O) ~ e-2(7NE(0) Ck ~ k ,

(5.219)

[see Eq. (17.396)]. Thus, at any fixed n, the re-expanded series has the same asymptotic growth as the original series with the same vanishing radius of convergence. The behavior (5.219) can be seen in Fig. 5.20(a) where we have plotted the logarithm of the absolute value of the kth term

Sk = cia)

(~r

(5.220)

of the re-expanded perturbation series (5.202) for various optimal values

nN

and

9 = 40. All curves show a growth ex: kk . The terms in the original series start growing immediately (precocious growth). Those in the re-expanded series, on the other hand, decrease initially and go through a minimum before they start growing (re-

tarded growth) . The dashed curves indicate the analytically calculated asymptotic behavior (5.219).

30

b)

Figure 5.20 Logarithmic plot of kth terms in re-expanded perturbation series at a coupling constant 9 / 4 = 1: (a) Frequencies ON extremizing the approximation WN. The dashed curves indicate the theoretical asymptotic behavior (5.219). (b) Frequencies ON corresponding to the dashed curve in Fig. 5.17. The minima lie for each N precisely at k = N, producing the fastest convergence. The curves labeled 0 = W indicate the kth term in the original perturbation series.

5 Variational Perturbation Theory

512

The increasingly retarded growth is the reason why energies obtained from the variational expansion converge towards the exact result. Consider the terms Sk of the resummed series with frequencies ON taken from the theoretical curve of optimal convergence in Fig. 5.17 (or 5.18). In Fig. 5.20(b) we see that the terms Sk are minimal at k = N, i.e., at the last term contained in the approximation W N . In general, a divergent series yields an optimal result if it is truncated after the smallest term Sk. The size of the last term gives the order of magnitude of the error in the truncated evaluation. The re-expansion makes it possible to find , for every N, a frequency ON which makes the truncation optimal in this sense.

5.16

Variational Perturbation Theory for Strong-Coupling Expansion

From the w -+ 0 -limit of (5.206) , we obtain directly the strong-coupling behavior of WN. Since 0 = (g/g)I /3, we can write (5.221) and evaluate this at the optimal value therefore

g=

l /uN . The large-g behavior of WN is

(5.222) with the coefficient (5.223)

exp( 6A I · 9A2Nl/3) 10.5

,

e.xp( 6.40 - 9.05 N' /3 )

o'~'::< eXp(8.2 - 9. 7 N!/3)

10-10

0

0

10- 10

.... ' "

°000: " ,

"

10- 15

10- 15

10-20

10- 20

10-30 ll---~--~---'o---~_--'o----'

8 1/ 3

Figure 5.21 bo and bl·

27 1 / 3 64 1 / 3 N ' /3

125 1 / 3

216 1 / 3

10-30 ll---~--~_--'o-_-~_--'o----' 8 /3 27 / 3 64 / 3 125 / 3 216 / 3

'

'

N ' /3

'

'

'

Logarithmic plot of N-behavior of strong-coupling expansion coefficients

513

5.16 Variational Perturbation Theory for Strong-Coupling Expansion

o

-1

2

3

5

6

Figure 5.22 Oscillations of approximate strong-coupling expansion coefficient bo as a function of N when approaching exponentially fast the exact limit. The exponential behavior has been factored out. The upper and lower points show the odd-N and even-N approximations, respectively.

The higher corrections to the leading behavior (5.222) are found just as easily. By expanding WN(g ,CP) in powers of 2 ,

w

(5.224) and inserting ,2 W

g2/3

(5.225)

= -;-(g--'/;:'-w""'3)""""2/"'3 '

we obtain the expansion

( g ) - 4/3 ] g)1/3 [ ( g ) - 2/3 WN = ( 4 bo + bi 4w 3 + b2 4w3 + . ..

,

(5.226)

with the coefficients (5.227)

The derivatives on the right-hand side have the expansions

d ( dW 2

)n

, ,2 _ WN(g , W) -

N (,)1 (0)

g

~Enl 4 '

(5.228)

514

5 Variational Perturbation Theory

Table 5.9 Coefficients bn of strong-coupling expansion of ground state energy of anharmonic oscillator obtained from a perturbation expansion of order 251. An extremely precise value for bo was given by F. Vinette and J. Cizek, J. Math. Phys. 32, 3392 (1991): bo = 0.66798625915577710827096201619860199430404936 .... n 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

-

bn 0.66798625915577710827096 0.1436687833808649100203 0.008627565680802279128 0.000818208905756349543 0.000082429217130077221 0.000008069494235040966 0.000000727977005945775 0.000000056145997222354 0.000000002949562732712 0.000000000064215331954 0.000000000048214263787 0.000000000008940319867 0.000000000001205637215 0.000000000000130347650 0.000000000000010 760 089 0.0000000000000004458901 0.0000000000000000589898 0.00000000000000001919600 0.00000000000000000328813 0.00000000000000000042962 0.000000000000000000044 438 0.0000000000000000000032305 0.0000000000000000000000314

with the re-expansion coefficients

(5.229)

For increasing N, the coefficients bo, bl, ... converge rapidly against the values shown in Table 5.9. From the logarithmic plot in Fig. 5.21 we extract a convergence

I 8.2- 9.7 NI/3 , Ib0 - bex 0 rve

(5.230)

This behavior will be derived in Subsection 17.10.5, where we shall find that for any 9 > 0, the error decreases at large N roughly like e -[9.7+(cg) - 2/ 3]N I /3 [see Eq. (17.409)]. The approach to this limiting behavior is oscillatory, as seen in Fig. 5.22 where we have removed the exponential falloff and plotted e- 6 .5+9.42N I/ 3 (bo - b1n against N [16].

515

5.17 General Strong-Coupling Expansions

For the proof of the convergence of the variational perturbation expansion in Subsection 17.10.5. it will be important to know that the strong-coupling expansion for the ground state energy o (g)1 /3 [ ( g E( ) = 4 bo + b1 4w3

)-2/3 + b (4w3g )-4/3 + . ..], 2

(5.231)

converges for large enough g > gs·

(5.232)

The same is true for the excited energies.

5.17

General Strong-Coupling Expansions

The coefficients of the strong-coupling expansion can be derived for any divergent perturbation series (5.233) n=O

for which we know that it behaves at large couplings g like M

E(g)

= gp/q

L

bm(g-2/ q)m.

(5.234)

m=O

The series (5.233) can trivially be rewritten as (5.235) with w = 1. We now apply the square-root trick (5.188) and replace w by the identical expression (5.236) containing a dummy scaling parameter O. The series (5.235) is then re-expanded in powers of g up to the order N, thereby treating w 2 - 0 2 as a quantity of order g. The result is most conveniently expressed in terms of dimensionless parameters 9 == gloq and a == (1 - C;2)/fJ, where C; == wlO. Then the replacement (5.236) amounts to w ---+ 0(1- ag)1/2, (5.237) so that the re-expanded series reads explicitly N

WN(g, a)

=

Op

L

cn(a) (gt,

(5.238)

n=O

with the coefficients: (5.239)

516

5 Variational Perturbation Theory

For any fixed g, we form the first and second derivatives of W}J(g) with respect to 0, calculate the O-values of the extrema and the turning points, and select the smallest of these as the optimal scaling parameter ON. The function W N (g) == W N(g, ON) constitutes the Nth variational approximation EN(g) to the function E(g) . We now take this approximation to the strong-coupling limit g --+ 00 . For this we observe that (5.238) has the general scaling form (5 .240)

For dimensional reasons, the optimal ON increases with g for large g like ON ~ gl/qCN, so that g = c"}/ and (J = l/g = c'fv remain finite in the strong-coupling limit, whereas C;2 goes to zero like l/[CN(g/W q)l/qj2. Hence (5 .241)

Here CN plays the role of the variational parameter to be determined by the lowest extremum or turning point of d;.wN(c"ii, 0). The full strong-coupling expansion is obtained by expanding WN(g , C;2) in powers of C;2 = (g/w qg) - 2/ q at a fixed g. The result is

/ [_ WN(g) = gP q bo(g)

_ ( g ) -2/q _ + b1(g) wq + b2(g)

( g ) -4/q wq + ... ] ,

(5 .242)

with

bn (') g

=

~'d'2 (~)n WN(n)(,g, w'2)1 n.

w

w2 =0

g'(2n- p)/q ,

(5 .243)

with respect to C;2. Explicitly:

Since g = c"ii, the coefficients bn(g) may be written as functions of the parameter c: (5 .245)

The values of c which optimize WN(g) for fixed g yield the desired values of CN. The optimization may be performed stepwise using directly the expansion coefficients bn ( c). First we optimize the leading coefficient bo(c) as a function of c and identifying the smallest of them as CN. Next we have to take into account that for large but finite C¥, the trial frequency 0 has corrections to the behavior gl/qC. The coefficient c will depend on g like

5.17 General Strong-Coupling Expansions

C(g) = C+ 1'1

~q A

(

)-2/

517 q

(

+ 1'2 ~q A

)-4/

q

+ ...,

(5.246)

requiring a re-expansion of c-dependent coefficients bn in (5.242). The expansion coefficients c and 1'n for n = 1,2,... are determined by extremizing b2n (c). The final result can again be written in the form (5.242) with b( c)n replaced by the final bn : g) -2/q ( g ) -4/q ] (5.247) + b2 wq + .... WN(g) = gp/q [bo + b1 ( wq The final bn are determined by the equations shown in Table 5.10. The two leading coefficients receive no correction and are omitted. The extremal values of g will have a strong-coupling expansion corresponding to (5.246):

g = c"i/ [1 +

g) - 2/q

=:;:j 2%, see Eq. (5.167)] . In Fig. 5.23 we have plotted the relative deviation of the variational approximation from the exact one in percent. The strong-coupling behavior is known from (5.226). It starts out like l/3, followed by powers of g- 1/3, g- l, g- 5/3. Comparison with (5.234) shows that this

519

5.18 Variational Interpolation between Weak and Strong-Coupling Expansions 1.02

1.0

-2

-3

3

-1

4

logg/4

0.995 Figure 5.23 Ratio of approximate and exact ground state energy of anharmonic oscillator from lowest-order variational interpolation. Dashed curve shows first-order FeynmanKleinert approximation WI (g). The accuracy is everywhere better than 99.5 %. For comparison, we also display the much worse (although quite good) variational perturbation result using the exact aix = 3/4.

corresponds to p = 1 and q = 3. The leading coefficient is given in Table 5.9 with extreme accuracy: bo = 0.667986259155777108270962016919860 . ... In a variational interpolation, this value is used to determine an approximate al (forgetting that we know the exact value a!X = 3/4). The energy (5.238) reads for N = 1 (with a = g/4 instead of g): W1(a,O) =

("20+ 201) ao + 02 a . al

(5.250)

Equation (5.245) yields, for n = 0: c al (5.251) bo = -ao +-. 2 c2 Minimizing bo with respect to c we find c = Cl == 2(aI/2ao)1/3 with bo = 3aocI/4 = 3(a5aI/2)1/3/2. Inserting this into (5.251) fixes al = 2(2/3bo)3/ a5 = 0.773970 ... , quite close to the exact value 3/4. With our approximate al we calculate W1(a , 0) at its minimum, where r.

_

~~-

{

1wcosh [~acosh(g/g(O))l 1w cos [~arccos(g/ g(O)) 1

fur

9 9

> g(O), < g(O),

(5.252)

with g(O) == 2w3ao/3v'3al' The result is shown in Fig. 5.23. Since the difference with respect to the exact solution would be too small to be visible on a direct plot ofthe energy, we display the ratio with respect to the exact energy WI (g) / E2x. The accuracy is everywhere better than 99.5 %.

520

5.19

5 Variational Perturbation Theory

Systematic Improvement of Excited Energies

The variational method for the energies of excited states developed in Section 5.12 can also be improved systematically. Recall the n-dependent level shift formulas (3.515) and (3.516), according to which

(5.253)

By applying the substitution rule (5.188) to the total energies

E(n) =

hJ),(n + 1/2)

and expanding each term in powers of 9 up to level shift

+ ilE(n) ,

l,

we find the contributions to the

~

[3(2n2

_

(~) 2 [2(34n 3 + 51n2 + 59n + 21)a8

+ 2n + 1)a4 + (2n + 1)a2r2],

+4· 3(2n2

1

+ 2n + 1)a6r2 + (2n + 1)a4 r4lnn'

(~) 3 [4. 3(125n4 + 250n3 + 472n2 + 347n + 111)a12 + 51n2 + 59n + 21)a10r2 1 +16· 3(2n2 + 2n + 1)a8r 4 + 2· (2n + 1)a6r6ln2n2 '

+4· 5(34n 3

which for n

(5.254)

= 0 reduce to the corresponding terms in (5.192). The extremization in

n leads to energies which lie only very little above the exact values for all n.

This is illustrated in Table 5.11 for n = 8 (compare with the energies in Table 5.5). A sum over the Boltzmann factors e-j3E~n) produces an approximate partition function Z3 which deviates from the exact one by less than 50.1%. It will be interesting to use the improved variational approach for the calculation of density matrices, particle distributions, and magnetization curves.

521

5.20 Variational 'neatment of Double-Well Potential

Table 5.11 Higher approximations to excited energy with n = 8 of anharmonic oscillator at various coupling constants g. The third-order approximation E~8l(g) is compared with the exact values E~~l(g), with the approximation Ei8 l (g) of the last section, and with the lower approximation of even order E~8l (g) (all in units of nw).

g/4 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 10 50 100 500 1000

5.20

E~~J(g) 13.3790 15.8222 17.6224 19.0889 20.3452 21.4542 22.4530 23.3658 24.2091 24.9950 51.9865 88.3143 111.128 189.756 239.Q12

Ei8 J(g)

E~8J(g)

13.3235257 15.7327929 17.5099190 18.9591071 20.2009502 21.2974258 22.2851972 23.1879959 24.0221820 24.7995745 51.5221384 87.5058600 110.105819 188.001018 236.799221

13.3766211 15.8135994 17.6099785 19.0742800 20.3287326 21.4361207 22.4335694 23.3451009 24.1872711 24.9720376 51.9301030 88.2154879 111.002842 189.540577 238.740320

E~8J(g) 13.3847643 15.8275802 17.6281810 19.0958388 20.3531080 21.4629384 22.4625543 23.3760415 24.2199988 25.0064145 51.9986710 88.3500454 111.173183 189.833415 239.109584

Variational Treatment of Double-Well Potential

Let us also calculate the approximate effective classical potential of third order W 3 (xo) for the double-well potential w2 9 M 2 w4 V(x) = M _x 2 + _x4 + - w2 = -1. (5.255) 2 4 4g' In the expression (5.197), the sign change of w 2 affects only the coupling g2(XO), which becomes (5.256)

[recall (5.176)]. Note the constant energy M 2 w4 /4g in V(x) which shifts the minima of the potential to zero [compare (5.78)] . To see the improved accuracy of W3 with respect to the first approximation W I (xo) discussed in Section 5.7 [corresponding to the first line of (5.197)], we study the limit of zero temperature where the accuracy is expected to be the worst. In this limit, W 3 (xo) reduces to (5.201) and is easily minimized in Xo and O. At larger coupling constants 9 > ge ~ 0.3, the energy has a minimum at Xo = O. For 9 ::::; ge, there is an additional symmetric pair of minima at Xo = ±xm oF 0 (recall Figs. 5.5 and 5.6) . The resulting W3(0) is plotted in Fig. 5.24 together with WI(O). The figure also contains the first excited energy which is obtained by setting w 2 = - 1 in r2 = 2M(w 2 - 02)/g of Eqs. (5.253)- (5.255). For small couplings g , the energies WI (O), W3(0) , .. . diverge and the minima at x = ±x m of Eq. (5.82) become relevant. Moreover, there is quantum tunneling across the central barrier from one minimum to the other which takes place for 9 ::::; ge ~ 0.3 and is unaccounted for by W3(0) and W3 (xm). Tunneling leads to a level splitting to be calculated in Chapter 17. In this chapter, we test the accuracy of WI (X m ) and W 3(x m ) by comparing them with the averages of the two lowest energies. Figure 5.24 shows that the accuracy of the approximation W 3 (x m ) is quite good.

522

5 Variational Perturbation Theory 0.5

l. 8

1.5

:1 1.4

II II II

II II II '\--

E

\

.....1.... • .~

0.6

O.9t-----..,----.,--------~ \ I \

\

',-\.. \

0.8

\

\

E

0.4

0.1

0.2

9

0.3

0.4

Figure 5.24 Lowest two energies in double-well potential as function of coupling strength 9. The approximations are WI(O) (dashed line) and W3(O) (solid line). The dots indicate numeric results of the Schriidinger equation. The lower part of the figure shows WI(x m ) and W3(O) in comparison with the average of the Schriidinger energies (small dots) . Note that WI misses the slope by 25%. Thnneling causes a level splitting to be calculated in Chapter 17 (dotted curves). Note that the approximation WI(x m ) does not possess the correct slope in 9, which is missed by 25%. In fact, a Taylor expansion of WI (Xm) reads (5.257) whereas the true expansion starts out with (5.258) The optimal frequency associated with (5.257) has the expansion !/1(Xm

In

)

= v2 -

3

27V2

4: 9 - 64 9

2

27 3 - 32 9

+ ...

Let us also compare the xo-behavior of W 3 (xo) with that of the true effective classical potential calculated numerically by Monte Carlo simulations. The curves are plotted in Fig. 5.5, and the

5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions

523

agreement is seen to be excellent. There are significant deviations only for low temperatures with f3 ;:: 20. At zero temperature, there exists a simple way of recovering the effective classical potential from the classical potential calculated up to two loops in Eq. (3.767). As we learned in Eq. (3.830), Xo coincides at zero temperature. with X in Eq. (3.767) Thus we merely have to employ the square-root trick (5.186) to the effective potential (3.767) and interchange X by Xo to obtain the variational approximation W1(xo) to the effective classical potantial to be varied. Explicitly, we replace, as in (5.188) , the one-loop contribtions, ground state energies nwT,L in (3.767) by

Vrl~,L + (W~,L(X) - rI~,L) "':; rlT,L + (W~,L(X) - rI~,d/2r1T,L' and exchange WL ,T(X) of the remaining two-loop terms by rlT,L. This yields for the D-dimensional rotated double-well potential (the Mexical hat potential in Fig. 3.13]: _ { W(xo) T='O v(X)

+ Ii

[rlL 4

_ + Wl,(X)] 4r1L + (D

D2 - 1 [VII(X) _ V1(X)]

+ rI~

li 2 - 6(2M)3

X2

{Ii

X3

+

1)1i

[rlT 4

~{...!.... (4) + Wf(X)] 4r1T + 8(2M)2 rli v (X)

2(D - 1) [VIII (X) _ 2V"(X) rlLrlT X X2

+

2V1(X)]} X3

2+ 2r1T 3(D-l) 1 [VII(X) VI(X)]2} + O(Ii)3} x~·xo(5.259) + rlL rlLrlf ---x- - --x'2

III

3 r1 [v (X) ]

This has to be extremized in rlT,L at fixed Xo .

5.21

Higher-Order Effective Classical Potential for Nonpolynomial Interactions

The systematic improvement of the Feynman-Kleinert approximation in Section 5.13 was based on Feynman diagrams and therefore applicable only to polynomial potentials. If we want to calculate higher-order effective classical potentials for nonpolynomial interactions such as the Coulomb interaction, we need a generalization of the smearing rule (5 .30) to the correlation functions of interaction potentials which occur in the expansion (5.180). The second-order term, for example, requires the calculation of

~oc = ,

l0

lif3 llif3 dT dT' (Vj~f(X(T))Vj~nX(T')))~Oc' 0

(5.260)

'

where 1 Vj~nx) = V(x) - ZMrl 2(xo)(x - XO)2 .

(5.261)

Thus we need an efficient smearing formula for local expectations of the form

(Fl(X(Tl))" X

f

.Fn(X(Tn)))~O =

z:'o n

'Dx(T)Fl(X(Tl))'" Fn(x(Tn))8(x - xo) exp { -~Ano [X(T)]} ,

(5.262)

where AnD [X(T)] and z~o are the local action and partition function of Eqs. (5.3) and (5.4). After rearranging the correlation functions to connected ones according to Eqs. (5.182) we find the cumulant expansion for the effective classical potential [see (5.180)]

1if3

Veff,cl(x o )=F.nxo

+~ )))XO lif3 JdT1 (v,xO(X(T mt 1 n ,c o

524

5 Variational Perturbation Theory h{3

- 2:2(3

h{3

J J (V;~~(X(T1))V;~~(X(T2)))~~c J J J (V;~~(X(T1))V;~~(X(T2))V;~~(xh)))~~c + ... . dT1

dT2

o

h{3

+ 6:3(3

(5 .263)

0

h{3

dT1

o

h{3

dT2

0

dT3

0

It differs from the previous expansion (5.180) for polynomial interactions by the potential V(x)

not being expanded around Xo. The first term on the right-hand side is the local free energy (5.6).

5.21.1

Evaluation of Path Integrals

The local pair correlation function was given in Eq. (5.19):

(8X(T)8x(T'))~O == ([X(T) -

XO][X(T') -

xo])~O

=

~G~)XO(T' T') =

a;T'(xo),

(5.264)

with [recall (5.19)- (5.24)]

n cosh[O(xo)(T-T' - n(3/2)] 2 aTT,(xo) = 2MO(xo) sinh[O(xo)n(3/2]

1 T

E (0, n(3) .

(5.265)

Higher correlation functions are expanded in products of these according to Wick's rule (3.302). For an even number of 8X(T)'S one has

(8X(T1)8x(T2) '" 8X(Tn))~O =

2: a;p(1)Tp

(2)

(5.266)

(xo) · ·· a;p(n _l)Tp(n) (XO),

pairs

where the sum runs over all (n - I)!! pair contractions. For an exponential, Wick's rule implies /

[

\ exp i

1

Inserting j(T) =

h{3

dTj(T) 8X(T)

2:::1 ki8(T -

]

) XO

n = exp

[

-~

1 1 h{3

dT

h{3

]

dT'j(T)a;T,(xo)j(T') .

(5.267)

Ti) , this gives for the expectation value of a sum of exponentials (5.268)

By Fourier-decomposing the functions F(X(T)) = J(dk/211')F(k) expik [xo obtain from (5.268) the new smearing formula

+ 8X(T)]

in (5.262), we

(5.269)

where a.;:;~,(xo) is the inverse of the n x n -matrix a~iT,(xo), This smearing formula determines the harmonic expectation values in the variational perturbation expansion (5.263) as convolutions with Gaussian functions. For n = 1 and only the diagonal elements a2(xo) = a;'T(xo) appear in the smearing formula (5.269), which reduces to the previous one in Eq. (5.30) [F(X(T)) = V(X(T)) ].

5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions

525

For polynomials F(X(T)) , we set X(T) = Xo + 8X(T) and expand in powers of 8X(T), and see that the smearing formula (5.269) reproduces the Wick expansion (5.266) . For two functions, the smearing formula (5.269) reads explicitly

(5.270) Specializing F2(x(T2)) to quadratic functions in X(T), we obtain from this

(F1 (x (T1 ))) flXO a2( Xo )

+

[1 -

a~'T2(XO)] a4 (XO)

IF ( ( )) [ ( ) l2)XO a~'T2(xO) 1 X T1 X T1 - Xo fl a4 (XO) ,

\

(5.271)

and

\I [x(Td - xol2 [xh) - xol 2)XO fl

5.21.2

=

a4 (Xo)

4 + 2aT1T2 (xo).

(5.272)

Higher-Order Smearing Formula in D Dimensions

The smearing formula can easily be generalized to D-dimensional systems, where the local pair correlation function (5.264) becomes a D x D -dimensional matrix: (5.273) For rotationally-invariant systems, the matrix can be decomposed in the same way as the trial frequency fI;j(xo) in (5.95) into longitudinal and transversal components with respect to Xo: (5.274) where PL;ij(XO) and PT;ij(XO) are longitudinal and transversal projection matrices introduced in (5.96). Denoting the matrix (5.274) by a; T' (xo), we can write the D-dimensional generalization of the smearing formula (5.275) as '

(5.275)

The inverse D x D -matrix a;k2,.k' (xo) is formed by simply inverting the n x n -matrices aL- ?Tk T k' (To), ";::?T T (ro) in the projection formula (5.274) with projection matrices Pdxo) and -1", k k' PT(xo): (5.276) a;k2,.k' (Xo) = aL;~kTk' (ro)pdxo) + aT~TkTk' (ro)PT(xo), 1

In D dimensions, the trial potential contains a D x D frequency matrix and reads

526

5 Variational Perturbation Theory

with the analogous decomposition 2 ) 0ij ( Xo

XOiXOJ'

2

= 0L(XO) - 2-

rO

+ 0T2 ( Xo )

(

8ij

-

XOiXOJ' ) -2-

.

(5.277)

XOj).

(5.278)

ro

The interaction potential (5.261) becomes

V;~~(x) =

V(x) -

~O~j(XO)(Xi -

XOi)(Xj -

To first order, the anisotropic smearing formula (5.275) reads (5.279) with the special cases =

2a},

(5.280)

Inserting this into formula (5.263) we obtain the first-order approximation for the effective classical potential 1i/3

Wl(xo)=F~o + n~

J

dTl

(V;~~(xh)))~~c'

(5.281)

o

in agreement with the earlier result (5.97) . To second-order, the smearing formula (5.275) yields [33]

(5.282) so that rule (5.271) for expectation values generalizes to

(5.283)

(5.284)

(5.285)

2a}ai ,

(5.286) (5.287)

5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions

5.21.3

527

Isotropic Second-Order Approximation to Coulomb Problem

To demonstrate the use of the higher-order smearing formula (5.275), we calculate the effective classical potential of the three-dimensional Coulomb potential e2 V(x) = -~

(5.288)

to second order in variational perturbation theory, thus going beyond the earlier results in Eq. (5.53) and Section 5.10. The interaction potential corresponding to (5.278) is

V;~~(x) =

~ f!2(XO)(x -

-1:1 -

XO)2 .

(5.289)

For simplicity, we consider only the isotropic approximation with only a single trial frequency. Then all formulas derived in the beginning of this section have a trivial extension to three dimensions. Better results will, of course be obtained with two trial frequencies f!i(ro) and f!~(ro) of Section 5.9. The Fourier transform 47re 2flkl2 of the Coulomb potential e2fixi is most conveniently written in a proper-time type of representation as V(k)

4~e21°O dae - ak2/2- ikx,

=

(5.290)

where a has the dimension length square. The lowest-order smeared potentials were calculated before in Section 5.10. For brevity, we consider here only the isotropic approximation in which longitudinal and transverse trial frequencies are identified [compare (5.112)] : \ [xh) - xof)

~o

=

3a2(xo) ,

(IX(~l)I):o

=

1:01 erf [

~] .

(5.291)

The first-order variational approximation to the effective classical potential (5.281) is then given by the earlier-calculated expression (5.115). To second order in variational perturbation theory we calculate expectation values

+ 6a;,T2 (xo) ,

(5.292)

2[3a (xo) - a;,T2 (xo)] .J7ra6 (xo)

(5.293)

9a4 (xo) 4

which follow from the obvious generalization of (5.271) , (5.272) to three dimensions. More involved is the Coulomb-Coulomb correlation function / 1 1 \ IX(Tl)llx(T2)1

)XO fl

(5.294)

Using these smearing results we calculate the second connected correlation functions of the interaction potential (5.289) appearing in (5.263) and find the effective classical potential to second order in variational perturbation theory

W1(xo)

+

[Me f!(Xo) 2

1i.J27ra6 (xo)

4

e Jr.f3

21i

o

_

3M f!3(XO)] 2

[4(xO)

41i

/ 1 1 dT \ IX(T) l lx(O)1

)XO fl

(5.295)

528

5 Variational Perturbation Theory

with the abbreviation

4(

) _ /i [4 + /i 2,B2f!2(xO) - 4 cosh /i,Bf!(xo)

+ /i,Bf!(xo) sinh /i,Bf!(xo) J 8,BM2f!3(xo) sinh[/i,Bf!(xo)j2] ,

I Xo =

(5.296)

the symbol indicating that this is a quantity of dimension length to the forth power. After an extremization of (5.295) with respect to the trial frequency f!(xo) , which has to be done numerically, we obtain the second-order approximation for the effective classical potential of the Coulomb system plotted in Fig. 5.25 for various temperatures. The curves lie all below the first-order ones, and the difference between the two decreases with increasing temperature and increasing distance from the origin.

:: oo\c:,cc~co:;>8fg0 ·0.6

/:;>"/,/ / .

/./:. ,/

/

-1

/

/~'/~

·0.8

/

-" .

.

.

V(ro)

,/ f

! 0.5

1.5

2.5

3.5

ro

Figure 5.25 Isotropic approximation to effective classical potential of Coulomb system in t he first (lines) and second order (dots). The temperatures are 10- 4 , 10- 3, 10- 2,10- 1, and 00 from top to bottom in atomic units. Compare also Fig. 5.9.

5.21.4

Anisotropic Second-Order Approximation to Coulomb Problem

The first-order effective classical potential W 1 (xo) was derived in Eqs. (5.97) and (5.117)- (5.121) . To obtain the second-order approximation W 2 (xo), we insert the Coulomb potential in the representation (5.290) into the second-order smearing formula (5.282) , and find

These are special cases of the more general expectation value

5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions

529

which furthermore leads to

(5.299)

From these smearing results we calculate the second-order approximation to the effective classical potential n{3

W2(xo) = F~o + h~

J

dTl

n{3

(Vi~~(X(Td))~~c -

2;2/3

o

n{3

JJ dTl

0

dT2

(Vi~~(X(Td)Vi~~(X(T2)))~~c . (5.300)

0

The result is

with the abbreviation

1

4 h [4 + h2/32n~,L - 4 cosh h/3nT,L + h/3nT,L sinh h/3nT,L IT,L = 8/3M2n~,L sinh[h/3nT,L/2] ,

(5.302)

which is a quantity of dimension (length)4. After an extremization of (5.115) and (5.301) with respect to the trial frequencies nT, n L which has to be done numerically, we obtain the second-order approximation for the effective classical potential of the Coulomb system plotted in Fig. 5.26 for various temperatures. The second order curves lie all below the first-order ones, and the difference between the two decreases with increasing temperature and increasing distance from the origin.

5.21.5

Zero-Temperature Limit

As a cross check of our result we take (5.301) to the limit T --t O. Just as in the lowest-order discussion in Sect. (5.4), the xo-integral can be evaluated in the saddle-point approximation which becomes exact in this limit, so that the minimum of WN(xo) in xo yields the nth approximation to the free energy at T = 0 and thus the nth approximations EYP the ground state energy E(O) of the Coulomb system. In this limit, the results should coincide with those derived from a direct variational treatment of the Rayleigh-Schrodinger perturbation expansion in Section 3.18. With the help of such a treatment, we shall also carry the approximation to the next order, thereby illustrating the convergence of the variational perturbation expansions. For symmetry reasons, the minimum of the effective classical potential occurs for all temperatures at the origin, such that we may restrict (5.115) and (5.301) to this point. Recalling the zero-temperature limit of the two-point correlations (5.19) from (3.246), lim a;T'(xo) =

{3~=

2

M~( Xo ) exp{ -n(xo) IT -

Til} ,

(5.303)

530

5 Variational Perturbation Theory

we immediately deduce for the first order approximation (5.115) with n = ncO) the limit

~nn - ~JMn e2 . 4 ,.fir n

E(O)(n) = lim Wfl(O) = {3-->oo

1

1

(5.304)

In the second-order expression (5.301) , the zero-temperature limit is more tedious to take. Performing the integrals over a1 and a2 , we obtain the connected correlation function (5.305) Inserting (5.303) , setting T1 = 0 and integrating over the imaginary times T = T2 E [0, n,8], we find

J/ 1i{3

1 1 )XO 4M { n2,82n2 dT \ IX(T) llx(O)1 fl c{3!";oo n2,8n eli{3fl - 1 - n,8n o '

x[e

li{3fl

2

-'Tr- - ;

arcsin

viI -

e-

21i{3fl +

~ In 0«,8) - ~ [In 0«,8)]2 - ~ J1 2

8

du In u

1}

l+u'

2

(5.306)

a({3)

with the abbreviation 10«,8) = 1 +

VI VI -

Inserting this into (5.301) and going to the limit ,8 (0)

.

fl

9

3

E2 (n)=J~~w2(0)=16nn-2,.fir -0.2

e 21i{3fl e 21i{3fl --> 00

JMn

(5.307)

we obtain

'Tr)

2 4 ( M 4 ----,;-e - ; 1+ln2-'2 n2e .

(5.308)

r------r-----.-----,------,---=~

-0.4

-0.6

-0.8

WfT=flL(ro) Wfn'flL(ro)

-1

V(ro) -1.2

L -_ _ _-'---'---_ _ _ _-'--_ _ _ _---"-_ _ _ _---'_ _ _ _- - - '

o

2

3

4

5

ro Figure 5.26 Isotropic and anisotropic approximations to effective classical potential of Coulomb system in first and second order at temperature 0.1 in atomic units. The lowest line represents the high temperature limit in which all isotropic and anisotropic approximations coincide.

5.21 Higher-Order Effective Classical Potential for Nonpolynomial Interactions

531

Postponing for a moment the extremization of (5.304) and (5.308) with respect to the trial frequency n, let us first rederive this result from a variational treatment of the ordinary RayleighSchrodinger perturbation expansion for the ground state energy in Section 3.18. According to the replacement rule (5.186), we must first calculate the ground state energy for a Coulomb potential in the presence of a harmonic potential of frequency w: M

2

e2

2

V(x) = 2W x - ~.

(5.309)

After this, we make the trivial replacement w -> v'n 2 + w2 - n2 and re-expand the energy in powers of w2 - n2, considering this quantity as being of the order e2 and truncating the reexpansion accordingly. At the end we go to w = 0, since the original Coulomb system contains no oscillator potential. The result of this treatment will be precisely the expansions (5.304) and (5.308). The Rayleigh-Schrodinger perturbation expansion of the ground state energy Er;p(w) for the potential (5.309) in Section 3.18 requires knowledge of the matrix elements of the Coulomb potential (5.288) with respect to the eigenfunctions of the harmonic oscillator with the frequency w : 271'

Vn,l,m ;n',l',m' =

7!'

00

JJ

Jdrr21j;~,I,m(r,19, T' > 0,

(5.384)

such that

Gr2(T T') = ~ (e-r(T-T') - 1- e-r(T+T') + ~e-2rT + ~e-2rT') ,

2r

2

2

'

for T > T' > O.

(5.385) In the limit r ---. 0, this becomes -!IT - T'I. We therefore obtain at zero temperature , n G-fl 'r (T, T) T~O 2M

[0r

{0

2 n -_ - - 1T-T 'I 2M

f2

2- ( , Go T, T)

2

_0

+r

r 2- -2 ( 1 +-

f3

0

2- 2 ')] f2 GO(T, T

(5 .386)

1 - 2rT - -e 1 - 2rT')} e- r IT- T'1+ e - r(T+T') - -e 2 2 '

to be inserted into (5.378) to get the expectation value (eik[X(T)-X(T')l) . The last three terms can be avoided by shifting the time interval under consideration and thus the Fourier expansion (5.352) from (0, n(3) to (-nf3 /2, nf3 /2) , which changes Green function (5.387) to

Gw2(T, T') =

L: sin Vn(T + nf3/2) sin Vn(T' + nf3/2) n=1 v; + w oo

2

sinhw(nf3/2 - T) sinhw(T' + nf3/2) £ T > T' > O. W sinh wnf3 ' or

(5.387)

541

5.22 Polarons

We have seen before at the end of Section 3.20 that such a shift is important when discussing the limit T ---> 0 which we want to do in the sequel. With the symmetric limits of integration, the Green function (5.384) looses its last term [compare with (3.147) for real times] and (5.386) simplifies to

n { -IT [22 Co,r(T T') ~ - - T'I ,

T=O

2M

r2

+r

2 - [22

f3

(1- e-r 1r - r , I) } .

At any temperature, we have the complicated expression for

Co,r(T, T') = -

2~ {~;

[T - T' -

n~ (T -

T

> T':

T')2]

r~::n~~;rf1 [sinhrf1(nN2-T) sinhrf1 (T' + n/3/2)-(T'

(5.388)

(5 .389) --->

T)-(T

--->

T')] }.

With the help of the Fourier integral (5.346) we find from this the expectation value of the interaction in (5.347) :

For zero temperature, this leads directly to the expectation value of the interaction in (5.347):

The expectation value of the harmonic trial interaction in (5.356), on the other hand, is simply found from the correlation function (5.376) [or equivalently from the second derivative of 1°,G (k, T, T') with respect to the momenta] : / \[X(T) - X(T')] 2)0,C = -6G- 0 'r (T, T').

(5.392)

At low temperatures, this leads to an integral (5.393) This expectation value contributes to the ground state energy a term (5.394)

542

5 Variational Perturbation Theory

Note that this term can be derived from the derivative of the ground state energy (5.368) as C8cE~,c. Together with -a/2V2 times the result of (5.391), the inequality (5.370) for the ground state energy becomes Eo

311,

0 W

2

< - (f - n) - liw--

10

y'7rW

- 4f

00

dt17

0

e-W~T

V-2Cfl,['(t17,0)

.

(5.395)

This has to be minimized in nand C, or equivalently, in nand f. Considering the low-temperature limit, we have taken the upper limit of integration to infinity (the frequency w corresponds usually to temperatures of the order of 1000 K). For small 0 , the optimal parameters nand f differ by terms of order o . We can therefore expand the integral in (5.395) and find that the minimum lies, in natural units with 11, = w = 1, at n = 3 and f = 3[1 + 20(1 - P)/3f, where P = 2[(1 - r)1/2 - 1]. From this we obtain the upper bound Eo :S

02 -0 -

81

+ ... r::::; - 0 -

0.01230 2

+ ...

(5.396)

This agrees well with the perturbative result [21] E!X

=

-0 -

0.01591962200 2

-

0.0008060700480 3

-

0(0 4 ).

(5.397)

The second term has the exact value (5.398)

In the strong-coupling region, the best parameters are n = 1, f = 40 2 /97T - [4 (log 2+ 1/2) - 1], where I r::::; 0.5773156649 is the Euler-Mascheroni constant (2.467). At these values, we obtain the upper bound 2

0 Eo :S - 37T - 3

(14 + log 2) + 0(0-

2 ) r::::;

-0.10610 - 2.8294 + 0(0- 2 ).

(5.399)

This agrees reasonably well with the precise strong-coupling expansion [23]. (5.400)

The numerical results for variational parameters and energy are shown in Table 5.12.

5.22.3

Effective Mass

By performing a shift in the velocity of the path integral (5.347) , Feynman calculated also an effective mass for the polaron. The result is (5.401)

543

5.22 Polarons

Table 5.12 Numerical results for variational parameters and energy.

a

r

n

1 3 5 7 9 11 15

3.110 3.421 4.034 5.S10 9.S50 15.41 30.0S

2.S71 2.560 2.140 1.604 1.2S2 1.162 1.076

~Eci2)

Eo -1.01 -3.13 -5.44 -S.l1 -11.5 -15.7 -26.7

E tot

-0.0035 -0.031 -0.OS3 -0.13 -0.17 -0.22 -0.39

-1.02 -3.16 -5.52 -S.24 -11.7 -15.9 -27.1

correction 0.35% 1.0% 1.5% 1.6% 1.4% 1.4% 1.5%

The reduced effective mass m == M eff / M has the weak-coupling expansion mw =

1+

a

"6 + 2.469136 x

1O- 2a 2 + 3.566719 x 1O-3a 3 +

...

(5.402)

and behaves for strong couplings like ms ~ ~

16 4 4 Sl7r2a - 31f (1

0.0201410: 4 -

2

+ log 4) a + 11.S5579 + .. . 1.0127750?+ 11.85579 + ... .

(5.403)

The exact expansions are [31]

a

1 +"6

+ 2.362763 x

0.0227019a4

5.22.4

1O- 2a 2 + O(a 4 ),

(5.404)

+ O(a 2 ).

(5.405)

Second-Order Correction

With some effort, also the second-order contribution to the variational energy has been calculated at zero temperature [32]. It gives a contribution to the ground state energy (2) _ ~Eo -

1 1 I - 21if31i2 \ (Ant

2)n,c -

Aint,harm)

c

(5.406)

.

Recall the definitions of the interactions in Eqs. (5.371) and (5.372). There are three terms

~E(2,1) = __1_~ o

21if31i2

{I\

A2 )n,c "'-'nt

_ [(A )n,C]2} mt

,

(5.407)

544

5 Variational Perturbation Theory

and (2,3)

6.Eo

_ -

1 1 - 2fi{3 fi2

{I\Aint,harm 2 )n,c -

[

(Aint,harm)

n,c]2}

.

(5.409)

The second term can be written as 1

(2,2) _

6.Eo

1 {

- - 2fi{3 fi2 -2Coc (Aint)

n,c}

the third as (2,3) _

6.Eo

1

1 {

- - 2fi{3 fi2 {I - Coel (Aint,harm)

,

n,c} .

(5.410)

(5.411)

The final expression is rather involved and given in Appendix 5C. The second-order correction leads to the second term (5.398) found in perturbation theory. In the strong coupling limit, it changes the leading term _0: 2 /37r ~ -0.1061 in (5.399) into 1 2 - 47r -;:

(2n)! 24n(n!)2n(2n + 1)

E 00

-17 +64 arcsin( \h;v'3) - 321og(4 Y2;v'3) ( ) 47r ' 5.412

which is approximately equal to -0.1078. The corrections are shown numerically in the previous Table 5.12.

5.22.5

Polaron in Magnetic Field, Bipolarons, Small Polarons, Polaronic Excitons, and More

Feynman's solution of the polaron problem has instigated a great deal of research on this subject [34]. There are many publications dealing with a polaron in a magnetic field. In particular, there was considerable discussion on the validity of the JensenPeierls inequality (5.10) in the presence of a magnetic field until it was shown by Larsen in 1985 that the variational energy does indeed lie below the exact energy for sufficiently strong magnetic fields . On the basis of this result he criticized the entire approach. The problem was, however, solved by Devreese and collaborators who determined the range of variational parameters for which the inequality remained valid. In the light of the systematic higher-order variational perturbation theory developed in this chapter we do not consider problems with the inequality any more as an obstacle to variational procedures. The optimization procedure introduced in Section 5.13 for even and odd approximations does not require an inequality. We have seen that for higher orders, the exact result will be approached rapidly with exponential convergence. The inequality is useful only in Feynman's original lowest-order variational approach where it is important to know the direction of the error. For higher orders, the importance of this information decreases rapidly since the convergence behavior allows us to estimate the limiting value quantitatively, whereas the inequality tells us merely the sign of the error which is often quite large in the lowest-order variational approach, for instance in the Coulomb system.

545

5.22 Polarons

There is also considerable interest in bound states of two polarons called bipolarons. Such investigations have become popular since the discovery of hightemperature superconductivity. 3

5.22.6

Variational Interpolation for Polaron Energy and Mass

Let us apply the method of variational interpolation developed in Section 5.18 to the polaron. Starting from the presently known weak-coupling expansions (5.397) and (5.404) we fix a few more expansion coefficients such that the curves fit also the strong-coupling expansions (5.400) and (5.405) . We find it convenient to make the series start out with a O by removing an overall factor -a from E and deal with the quantity -E!x/ a . Then we see from (5.400) that the correct leading power in the strong-coupling expansion requires taking p = 1, q = 1. The knowledge of bo and bl allows us to extend the known weak coupling expansion (5.397) by two further expansion terms. Their coefficients a3, a4 are solutions of the equations [recall (5.245)]

35 15 a2 128 aoc + al + 8-;;35ao 5a2 a3 - - - -- - 32 c 4 c3 c3 '

bo

2a3

35

15 a2 c2

8

4a3 -

~ -

(5.413) (5.414)

The constant c governing the growth of ON for a bo in c, which yields the equation 128 ao -

a4

+ --;} + c3'

---> 00

4a4

is obtained by extremizing

0

7 = .

(5.415)

The simultaneous solution of (5.413)- (5.415) renders

0.09819868, 6.43047343 x 10- 4 , -8.4505836 X 10-5 .

(5.416)

The re-expanded energy (5.238) reads explicitly as a function of a and 0 (for E including the earlier-removed factor -a)

(5.417) Extremizing this we find 0 4 as a function of a [it turns out to be quite well approximated by the simple function 0 4 ~ c4a+1/(1+0.07a)]. This is to be compared with the optimal frequency obtained from minimizing the lower approximation W 2 (a, 0):

O~ = 1 + 4a2x2 + 3ao

1(1

Y

+ 4a2x2 )2 -1, 3ao

(5.418)

546

5 Variational Perturbation Theory

which behaves like C2o+1+ . . . with C2 = V8a2/3ao ~ 0.120154. The resulting energy is shown in Fig. 5.28, where it is compared with the Feynman variational energy. For completeness, we have also plotted the weak-coupling expansion, the strongcoupling expansion, the lower approximation W2 (a) , and two Pade approximants given in Ref. [22] as upper and lower bounds to the energy.

-5

- 10 -15 -20 -25

-30

E

Figure 5.28 Variational interpolation of polaron energy (solid line) between the weakcoupling expansion (dashed) and the strong-coupling expansion (short-dashed) shown in comparison with Feynman's variational approximation (fat dots) , which is an upper bound to the energy. The dotted curves are upper and lower bounds coming from Pade approximants [22] . The dot-dashed curve shows the variational perturbation theory W 2 (a) which does not make use of the strong-coupling information.

Consider now the effective mass of the polaron, where the strong-coupling behavior (5.405) fixes p = 4, q = 1. The coefficient bo allows us to determine an approximate coefficient a3 and to calculate the variational perturbation expansion W3 (a) . From (5 .245) we find the equation (5.419)

whose minimum lies at C3 = V8a2/3ao where bo = V32aV27al' Equating bo of Eq. (5.419) with the leading coefficient in the strong-coupling expansion (5.405), we obtain a3 = [27a1b5 /32j1/3 ~ 0.0416929. The variational expression for the polaron mass is from (5.238) (5.420)

This is extremal at (5.421)

547

5.22 Polarons

2.5

2 1.5

\. 0. 5

~.

o

2

4

~











- .-.~ --.

6

8

10

12

14

Figure 5.29 Variational interpolation of polaron effective mass between the weak(dashed) and strong-coupling expansions (short-dashed). To see better the differences between the strongly rising functions, we have divided out the asymptotic behavior mas = 1 + bo0:4 before plotting the curves. The fat dots show Feynman's variational approximation. The dotted curves are upper and lower bounds coming from Pade approximants [22].

From this we may find once more c3 = V8a2/3ao . The approximation W3 (0:) = W3 (0: ,D3 ) for the polaron mass is shown in Fig. 5.29, where it is compared with the weak and strong-coupling expansions and with Feynman's variational result . To see better the differences between the curves which all grow fast with 0:, we have divided out the asymptotic behavior mas = 1 + b0 0: 4 before plotting the data. As for the energy, we have again displayed two Pade approximants given in Ref. [22] as upper and lower bounds to the energy. Note that our interpolation differs considerably from Feynman's and higher order expansion coefficients in the weak- or the strongcoupling expansions will be necessary to find out which is the true behavior of the model. Our curve has, incidentally, the strong-coupling expansion

m S = 0.02270190:4 + 0.1257220: 2 + 1.15304 + 0(0:- 2),

(5.422)

the second term ex: 0:2-term being in sharp contrast with Feynman's expression (5.403). On the weak-coupling side, a comparison of our expansion with Feynman's in Eq. (5.402) shows that our coefficient a3 ;:::; 0.0416929 is about 10 times larger than his. Both differences are the reason for our curve forming a positive arch in Fig. 2, whereas Feynman's has a valley. It will be interesting to find out how the polaron mass really behaves. This would be possible by calculating a few more terms in either the weak- or the strong-coupling expansion.

548

5 Variational Perturbation Theory

Note that our interpolation algorithm is much more powerful than Pade's. First, we can account for an arbitrary fractional leading power behavior o;P as 0; ---> 00. Second, the successive lower powers in the strong-coupling expansion can be spaced by an arbitrary 2/ q. Third, our functions have in general a cut in the complex 0;plane approximating the cuts in the function to be interpolated (see the discussion in Subsection 17.10.4). Pade approximants, in contrast, have always an integer power behavior in the strong-coupling limit, a unit spacing in the strong-coupling expansion, and poles to approximate cuts.

5.23

Density Matrices

In path integrals with fixed end points, the separate treatment of the path average Xo (l / n!3)Jon/3 dr x( r) looses its special virtues. Recall that the success of this separation in the variational approach was based on the fact that for fixed Xo, the fluctuation square width a2(xo) shrinks to zero for large temperatures like n2 / 12MkB T [recall (5.25)] . A similar shrinking occurs for paths whose endpoints held fixed, which is the case in path integrals for the density. Thus there is no need for a separate treatment of Xo , and one may develop a variational perturbation theory for fixed endpoints instead. These may, moreover, be taken to be different from one another Xb # Xa , thus allowing us to calculate directly density matrices.4 The density matrix is defined by the normalized expression

=

P(Xb, xa) =

~P(Xb' xa) ,

(5.423)

where P(Xb , xa) is the unnormalized transition amplitude given by the path integral

J

P(Xb' xa) = (Xb n!3lxaO) =

'Dx exp {-A[x]/n} ,

(5.424)

(Xa ,O),,-,,(x b,n/ kBT)

summing all paths with the fixed endpoints x(O) = Xa and x(n/kBT) = Xb. The diagonal matrix elements of the density matrix in the integrand yield, of course, the particle density (5.90) . The diagonal elements coincide with the partition function density z(x) introduced in Eq. (2.331). The partition function divided out in (5.423) is found from the trace 00

z=

Jdxp(x,x).

(5.425)

-00

5.23.1

Harmonic Oscillator

As usual in the variational approach, we shall base the approximations to be developed on the exactly solvable density matrix of the harmonic oscillator. For the sake of generality, this will be assumed to be centered around X m , with an action

An,xm [x]

=

r

Jo

n / k BT

dr

{ I} "2 Mi?(r)I + "2M02[x(r) - xm]2 .

(5.426)

4H. Kleinert, M. Bachmann, and A. Pelster, Phys. Rev. A 60 , 3429 (1999) (quant-ph/ 9812063).

549

5.23 Density Matrices

Its unnormalized density matrix is [see (2.409)]

with the abbreviation (5.428) At fixed endpoints Xb, Xa and oscillation center Xm , the quantum-mechanical correlation functions are given by the path integral 1 (Ol(x(rl)) 02(x(r2))'" )~~~x: = _ n,Xm ( ) Po Xb, Xa

J

x

VxOl(x(rl)) O2 (x(r2)) .. . exp{ -An,xm[xJlfi}. (5.429)

(Xa,O)""'(Xb,li/kBT)

The path x(r) at a fixed imaginary time r has a distribution p(x,r)

where

Xci (r)

_

n,Xm b, a

1 [ (x - XcI(r))2] expb2() , V 27rb2(r) 2 r

= (b(x-x(r)))x X = I

(5.430)

is the classical path of a particle in the harmonic potential Xci

( ) _ xbsinhnr+xasinhn(fi/kBT-r) r sinhfin/kBT '

(5.431)

and b2 (r) is the square width b2 (r)

= _fi_ {coth fin _ cosh[n(2r -fi/kBT)]} . kBT

2Mn

sinhfin/kBT

(5.432)

In contrast to the square width a2 (xo) in Eq. (5.24) this depends on the Euclidean time r, which makes calculations more cumbersome than before. Since the r lies in the interval 0 ::::: r ::::: fi/kBT, the width (5.432) is bounded by 2 fi fin b (r) ::::: 2Mn tanh 2k BT'

(5.433

)

thus sharing with a2 (xo) the property of remaining finite at all temperatures. The temporal average of (5.432) is -2 kBT b = T

rli/ kBT drb (r) =

io

2

fi ( fin kBT) 2Mn coth kBT - fin .

(5.434)

Just as a2 (xo), this goes to zero for T --+ 00. Note however, that the asymptotic behavior is b2 ---+ fin/6k BT, (5.435) T -->oo

which is twice as big as that of a 2 (xo) in Eq. (5.25) (see Fig. 5.30).

550

5 Variational Perturbation Theory

1.0

" " " " " I

I

a~ot

I

I

,,'

,,"

0.5

""

/

/

/

/

/

/

/

/

fluctuation width

/

/

//

II

2 atot,cl

/

\

/ \

/

0.0 1.0

0.0

2.0

Figure 5.30 Temperature dependence of fluctuation widths of any point X(T) on the path in a harmonic oscillator (l2 is the generic square length in units of 'hI MQ). The quantity a2 (dashed) is the quantum-mechanical width, whereas a2 (xo) (dash-dotted) shares the width after separating out the fluctuations around the path average Xo. The quantity a~l (long-dashed) is the width of the classical distribution, and b2 (solid curve) is the fluctuation width at fixed ends which is relevant for the calculation of the density matrix by variational perturbation theory (compare Fig. 3.14).

5.23.2

Variational Perturbation Theory for Density Matrices

To obtain a variational approximation for the density matrix, we separate the full action into the harmonic trial action and a remainder (5.436) with an interaction

Andx(T)] =

rn

io

{3

(5.437)

dTVint(X(T)) ,

where the interaction potential is the difference between the original one V(x) and the inserted displaced harmonic oscillator: (5.438) The path integral (5.424) is then expanded perturbatively around the harmonic expression (5.427) as

P(Xb,X a) = pofl,Xm(Xb,Xa) [1-

~(Aindx] )~,xxm + ~/ Arnt[x] )fl,Xm - ... J, 21i \ It

b, a

Xb,Xa

(5.439)

551

5.23 Density Matrices

with the harmonic expectation values defined in (5.429). The sum can be evaluated as an exponential of its connected parts, going over to the cumulant expansion:

P(Xb, xa) =pofl,xm (Xb, xa) exp

[-~( Aindx] )~,Xxma, + ~/ Afnt [x] \ fl ,x m 217, \ / C

b,

II,

-

. ..

Xb,Xa,C

J,(5.440)

where the cumulants are defined as usual [see (3.482), (3.483)]. The series (5.440) is truncated after the N-th term, resulting in the N-th order approximant for the quantum statistical density matrix ,. " f2,Xm

PN

_

,.., f2,Xm

(Xb, xa) - Po

(Xb, xa) exp

[

E n!lin N

(-

l)n

n

f2 ,Xm

]

(Aint[x] )Xb,Xa,C ,

(5.441)

which explicitly depends on the two variational parameters 0 and X m . By analogy with classical statistics, where the Boltzmann distribution in configuration space is controlled by the classical potential V (x) according to [recall (2.352)]

-fi T exp [-,8V(x)] , Y~

Pcb) =

(5.442)

we shall now work with the alternative type of effective classical potential Veff,cl(x a, Xb) introduced in Subsection 3.25.4. It governs the unnormalized density matrix [see Eq. (3.834)]

P(Xb, xa) =

J27rli~ ,8 exp [_,8Veff,cl(Xb, Xa)].

(5.443)

Variational approximations to Veff,cl(Xb, xa) of Nth order are obtained from (5.427) , (5.441), and (5.443) as a cumulant expansion

Wfl ,xm( ) 1 I sinh li,80 N Xb, Xa = 2,8 n li,80

MO + 2li,8 sinh li,80

{( -2

xb

-2) h li,80 + xa cos -

T - }

XbXa

(5.444) They have to be optimized in the variational parameters 0 and Xm for a pair of endpoints Xb, Xa . The result is denoted by W N(Xb, xa). The optimal values O(xa, Xb) and xm(x a, Xb) are denoted by ON(X a, Xb), x;;,(xa, Xb). The Nth-order approximation for the normalized density matrix is then given by

PN ( Xb,Xa) = ZN- 1-PNn~,X~( Xb,Xa) ,

(5.445)

where the corresponding partition function reads (5.446)

In principle, one could also optimize the entire ratio (5.445), but this would be harder to do in practice. Moreover, the optimization of the unnormalized density matrix is the only option, if the normalization diverges due to singularities of the potential, for example in the hydrogen atom

552

5.23.3

5 Variational Perturbation Theory

Smearing Formula for Density Matrices

In order to calculate the connected correlation functions in the variational perturbation expansion (5.441), we must find efficient formulas for evaluating expectation values (5.429) of any power of the interaction (5.437)

This can be done by an extension of the smearing formula (5.30) . For this we rewrite the interaction potential as

V'int(X(Tl)+X m )= !dZ1V'int(ZI+X m )! -00

~;ei'\IZlexp[_ fo hf3 dTiAI8(T-Tl)X(T)] '

-00

(5.448) and introduce a current n

J(T)

=

L

inA 18(T - Tl),

(5.449)

1=1

so that (5.447) becomes

The kernel Kfl,Xm bl represents the generating functional for all correlation functions of the displaced harmonic oscillator

For zero current j, this generating functional reduces to the Euclidean harmonic propagator (5.427): (5.452) and the solution of the functional integral (5.451) is given by (recall Section 3.1)

(5.453)

553

5.23 Density Matrices

where Xcl (T) denotes the classical path (5.431) and cgl (T, T') the harmonic Green function with Dirichlet boundary conditions (3.386), to be written here as

d 2l (

') _ _ n_coshO(IT - T'I- n(3) - coshO(T + T' n2 T, T - 2MO sinh nf30

-

n(3)

.

(5.454)

The expression (5.453) can be simplified by using the explicit expression (5.449) for the current j. This leads to a generating functional (5.455)

where we have introduced the n-dimensional vectors A = (AI, ... , An) and Xci = (Xcl(T1) , .. ' , Xcl(Tn))T with the superscript T denoting transposition, and the symmetric n x n-matrix C whose elements are C k1 = cgl(Tk,Tl)' Inserting (5.455) into (5.450), and performing the integrals with respect to AI, ... , An , we obtain the n-th order smearing formula for the density matrix

The integrand contains an n-dimensional Gaussian distribution describing both thermal and quantum fluctuations around the harmonic classical path Xcl(T) of Eq. (5.431) in a trial oscillator centered at X m , whose width is governed by the Green functions (5.454). For closed paths with coinciding endpoints (Xb = x a ), formula (5.456) leads to the n-th order smearing formula for particle densities

P(Xa) =

~P(Xa, xa) = ~

f DXO

(5.468)

The second term is determined by inserting the high-temperature limit of the total fluctuation width (5.26): 2 kBT (5.469) atot,cl = M02 ' and of the polynomials (5.464) into the expansion (5.461), leading to ) ~ Van( 2 Xa T~ ->00

~ _1_H ~ 2n I n

n=O

n.

(J

M02j3 ) Joo 2 Xa - 00

J

dz lJ; ( ) - Mn 2{3z2/ 2H (M02j3 ) Vint z e n 2 Z. M02 j3

21f /

(5.470)

Then we make use of the completeness relation for Hermite polynomials 1r:;;e _ x2 ~ 1 Hn () ~ ~ x Hn (') x n=O 2 n.

y1f

= is (x - x ') ,

(5.471)

which may be derived from Mehler's formula (2.295) in the limit b --+ 1-, to reduce the smeared interaction potential Va~(xa) to the pure interaction potential (5.438): (5.472)

Recalling (5.438) we see that the first-order effective classical potential (5.468) approaches the classical one: (5.473)

This is a consequence of the vanishing fluctuation width b2 of the paths around the classical orbits. This property is universal to all higher-order approximations to the effective classical potential (5.444). Thus all correction terms with n > 1 must disappear in the limit j3 --+ 0,

r

{3~

-1~(-I)n(An[l)n

---;3 f:2

n!1in

"'int X

Xa,Xa,C

0

= .

(5.474)

b) Zero-Temperature Limit

At low temperatures, the first-order effective classical potential (5.466) becomes

W- 1n,qm( Xa ) -_ -11,0 2

+

l'1m Va2n (Xa ) . {3->00

(5.475)

557

5.23 Density Matrices

The zero-temperature limit of the smeared potential in the second term defined in (5.467) follows from Eq. (5.461) by taking into account the limiting procedure for the polynomials C~n) in (5.465) and the zero-temperature limit of the total fluctuation width (5.26), which is equal to the zero-temperature limit of a2 (xo): a;ot ° = a;ot = , T=O

n/2Mn. Thus we obtain with Ho(x) = 1 and the inverse length K, == l/'xn = vMn/n [recall (2.301)]:

J~~ Va~(xa) =

1 ~Ho(K,Z)2 dz

exp{ _K,2 Z2} Vint(z).

(5.476)

- 00

Introducing the harmonic eigenvalues

E~=nn(n+D,

(5.477)

and the harmonic eigenfunctions [recall (2.299) and (2.300)] (5.478) we can re-express the zero-temperature limit of the first-order effective classical potential (5.475) with (5.476) by (5.479) This is recognized as the first-order Rayleigh-Schrodinger perturbative result for the ground state energy. For the discussion of the quantum-mechanical limit of the first-order normalized density, (5.480) we proceed as follows. First we expand (5.480) up to first order in the interaction, leading to

Inserting (5.428) and (5.461) into the third term in (5.481), and assuming n not to depend explicitly on X a , the xa-integral reduces to the orthonormality relation for Hermite polynomials

2nn~.j1r

JdXaHn(xa)Ho(xa)e-X~ 00

-00

=

8no ,

(5.482)

558

5 Variational Perturbation Theory

so that the third term in (5.481) eventually becomes

But this is just the n = 0 -term of (5.461) with an opposite sign, thus canceling the zeroth component of the second term in (5.481), which would have been divergent for j3 --+ 00. The resulting expression for the first-order normalized density is

The zero-temperature limit of C~n) is from (5.465) and (5.477) 1·

1m

(3->oo

j3C(n) _ {3

-

En n

2 -

(5.485)

En ' 0

so that we obtain from (5.484) the limit

(5.486)

Taking into account the harmonic eigenfunctions (5.478) , we can rewrite (5.486) as

. ( )1 2 = ["I.n( "I.n( ) ('t/J~ 1't/J~) PIn( Xa ) = 1"1 % Xa % Xa )]2 - 2"%I.n( Xa )"'"' ~ o/n Xa En1Vint _ En ' n>O

n

(5.487)

0

which is just equivalent to the harmonic first-order Rayleigh-Schrodinger result for particle densities. Summarizing the results of this section, we have shown that our method has properly reproduced the high- and low-temperature limits. Due to relation (5.487) , the variational approach for particle densities can be used to determine approximately the ground state wave function 't/JO (xa) for the system of interest.

5.23.5

Smearing Formula in Higher Spatial Dimensions

Most physical systems possess many degrees of freedom. This requires an extension of our method to higher spatial dimensions. In general, we must consider anisotropic harmonic trial systems, where the previous variational parameter n2 becomes a D x D -matrix n~v with j.t, v = 1, 2, ... , D.

559

5.23 Density Matrices

a) Isotropic Approximation An isotropic trial ansatz

(5.488) can give rough initial estimates for the properties of the system. In this case, the n-th order smearing formula (5.458) generalizes directly to

x

(5.489)

with the D-dimensional vectors Zl = (Zll' Z21, •.• , ZDl)T . Note, that Greek labels J.L, v, ... = 1, 2, . .. , D specify spatial indices and Latin labels k, I, ... = 0,1,2, ... , n refer to the different imaginary times. The vector Zo denotes r a , the matrix a2 is the same as in Subsection 5.23.3. The harmonic density reads

p~(r) =)27raoo 12

D

exp [-2 \

aoo

t x~] .

(5.490)

1'=1

b) Anisotropic Approximation In the discussion of the anisotropic approximation, we shall only consider radially-symmetric potentials V(r) = V(lr l) because of their simplicity and their major occurrence in physics. The trial frequencies decompose naturally into a radial frequency n£ and a transverse one nT as in (5.95) :

(5.491) with T a = 1r a I· For practical reasons we rotate the coordinate system by points along the first coordinate axis, Ta ,

(ra)1' == 21'0 = { 0, and

Xn

=

U Xn so that

J.L = 1, 2~J.L~D ,

ra

(5.492)

n2-matrix is diagonal: 0

nT2

n2 =

U

0 0

0

nT2

0

0

! )~un'u-' nT2

After this rotation, the anisotropic n-th order smearing formula in D dimensions reads

(5.493)

560

5 Variational Perturbation Theory

The components of the longit udinal and transversal matrices aL and a~ are aLkl = aL(Tk,Tz) ,

ahl = a~(Tk,Tz) ,

(5.495)

where the frequency n in (5.459) must be substituted by the new variational parameters nL , nT, respectively. For the harmonic density in the rotated system p~L ,T (1') which is used to normalize (5.494), we find Po{k ,T( r- ) -_

Appendix 5A

1 12- D-l exp -~~ - -221fa LOO

21fa TOO

[

- - 12-

2 a LOO

-2 X l

-

D ] " xp. -2 - 12- 'L.J . 2 a TOO 1"=2

(5.496)

Feynman Integrals for T=fo 0 without Zero Frequency

The Feynman integrals needed in variational perturbation theory of the anharmonic oscillator at nonzero temperature can be calculated in close analogy to those of ordinary perturbation theory in Section 3.20. The calculation proceeds as explained in Appendix 3D, except that the lines represent now the t hermal correlation function (5.19) with the zero-frequency subtracted from the spectral decomposition:

With the dimensionless variable x == li{3w , the results for the quantities a'fl defined of each Feynman diagram with L lines and V vertices as in (3.547) , but now without the zero Matsubara frequency, are [compare with the results (3D.3)-(3D.11)]

~!

wx

a~

(::coth:: -1) 2 2 '

(5A.1)

~ (4+ x2 -4 cosh x + x sinh x ) , ( ~)2 81xsinh2-

(5A.2)

W

2

~ (-3 x cosh:: + 2 x ( ~) 3 _1_ 64x sinh 2 3 _

W

3

cosh:: + 3 x cosh 3x 2 2

2

+48 sinh ~ + 6 x 2 sinh ~ - 16 sinh 3;) ,

a~

(5A.3)

~ (-864 + 18 X4 + 1152 cosh x + 32 x 2 cosh x ( ~) 4 ~ 768x sinh4W

2 - 288 cosh 2x - 32 x 2 cosh 2x - 288 x sinh x + 24 x 3 sinh x

+ 144 x sinh 2x + 3 x 3 sinh 2x) ,

(5A.4)

5 __1_ _ _1_ (672 xcosh:: _ 8 x 3 cosh:: + 24 x5cosh:: ( ~) w 4096x sinh5 :: 2 2 2 3

2

3x 3 3x 5x 3 5x - 1008 x cosh 2 + 3 x cosh 2 + 336 x cosh 2 + 5 x cosh 2 -7680 sinh

~-

352 x 2 sinh ~ + 72 x4 sinh

~ + 3840

sinh 3;

+ 224 x 2 sinh 3; + 12 x4 sinh 3; - 768 sinh 5; _ 64 x 2 sinh 5;) ,

(5A.5)

561

Appendix 5A Feynman Integrals for T =1= 0 without Zero Frequency

4~ (-107520 -7360 x 2 + 624 x4 + 96 x ( w~)6 49112 5 x smh

6

6 -

2

+ 161280 cosh x + 12000 x 2 cosh x - 777 X4 cosh x + 24 x 6 cosh x - 64512 cosh 2x - 5952 x 2 cosh 2x + 144 x4 cosh 2x + 10752 cosh 3x -28800 x sinh x + 1312 x 2 cosh 3x + 9 X4 cosh 3x + 1120 x 3 sinh x +324 x 5 sinh x

+ 23040 x sinh 2x -

320 x 3 sinh 2x - 5760 x sinh 3x

-160 x3 sinh3x),

(5A.6)

~ (- 24 - 4x2+24coshx+ x 2 cosh x- 9xsinhx) , ( w~) 3 24x1 sinh 2

2 -

2

(~) 28~X2 Si~3:: 4

(45 x cosh

~-

6 x 3 cosh

~-

45 x cosh 3;

2 -432 sinh ~ - 54 x 2 sinh ~ lO a 3'

(~

) 5

23014x 3

Si~

4::

+ 144

sinh 3;

+ 4 x 2 sinh 3;) ,

(-3456 - 414 x 2 - 6 x4

+ 4608

(5A.7)

cosh x+

2

496 x 2 cosh x - 1152 cosh 2x - 82 x 2 cosh 2x - 1008 x sinh x 16 x 3 sinh x

+ 504 x sinh 2x + 5 x 3 sinh2x) .

(5A.8)

Six of these integrals are the analogs of those in Eqs. (3.547). In addition there are the three integrals a~, a~, and ali? , corresponding to the three diagrams (5A.9)

respectively, which are needed in Subsection 5.14.2. They have been calculated with zero Matsubara frequency in Eqs. (3D.8)- (3D.11). In the low-temperature limit where x = nh(3 --t 00, the x-dependent factors in Eqs. (5A.l)(5A.8) converge towards the same constants (3D.13) as those with zero Matsubara frequency, and the same limiting relations hold as in Eqs. (3.550) and (3D.14). The high-temperature limits x --t 0, however, are quite different from those in Eq. (3D.16) . The present Feynman integrals all vanish rapidly for increasing temperatures. For L lines and V vertices, h(3(I/w) v -la?l goes to zero like (3v ((3/12) L. The first V factors are due to the V-integrals over T, the second are the consequence of the product of n/2 factors a2 Thus a?l behaves like (5A.1O) Indeed, the x-dependent factors in (5A.l)- (5A.8) vanish now like x/12, x 3 /720 , x 5 /30240, x 5 /241920, x 7 /11404800 , 193x8 /47551795200, x 4/30240, x 6 /1814400, x 7 /59875200,

(5A.11)

respectively. When expanding (5A.l)- (5A.8) into a power series, the lowest powers cancel each other. For the temperature behavior of these Feynman integrals see Fig. 5.32. We have plotted the reduced Feynman integrals a~L(x) in which the low-temperature behaviors (3.550) and (3D.14) have been divided out of a~L.

562

5 Variational Perturbation Theory

0.5

1

1.5

2

L/x Figure 5.32 Plot of the reduced Feynman integrals a~L(x) as functions of L/x = LkBTj'hw. The integrals (3D.4)- (3D.ll) are indicated by decreasing dash-Ienghts. Compare Fig. 3.16. The integrals (5A.2) and (5A.3) for a~ and ag can be obtained from the integral (5A.l) for a 2 via the operation 1 (5A.12) 2

'hn ( o)n 'hn ( n!

-

ow

= n!

-

o)n

2w ow

'

with n = 1 and n = 2, respectively. This is derived following the same steps as in Eqs. (3D.18)(3D.20). The absence of the zero Matsubara frequency does not change the argument. Also, as in Eqs. (5.195)- (3D.21), the same type of expansion allows us to derive the three integrals from the one-loop diagram (3.546) .

Appendix 5B

Proof of Scaling Relation for Extrema of W N

Here we prove the scaling relation (5.215), according to which the derivative of the Nth approximation W N to the ground state energy can be written as [14] (5B.l) where pN(a) is a polynomial of order N in the scaling variable a = fl(fl2 - 1)/g. For the sake of generality, we consider an anharmonic oscillator with a potential gx P whose power P is arbitrary. The ubiquitous factor 1/4 accompanying 9 is omitted, for convenience. The energy eigenvalue of the ground state (or any excited state) has an Nth order perturbation expansion N

EN(g) =

1

Wl:EI CC P!2)/J '

(5B .2)

1=0

where EI are rational numbers. After the replacement (5.188), the series is re-expanded at fixed r in powers of 9 up to order N, and we obtain (5B.3) with the re-expansion coefficients [compare (5.207)] (5B.4)

563

Appendix 5B Proof of Scaling Relation for the Extrema of W N

Here a is a scaling variable for the potential gx P generalizing (5.208) (note that it is four times as big as the previous a , due to the different normalization of g): _ n(P-2)/2(n 2 _ w2)

a=

(5B.5)

.

9

We now show that the derivative dWN(g,n)/dn has the following scaling form generalizing (5B.l): dWjJ ( 9 )N dn = n(P+2)/2 PN(a),

(5B.6)

where pN(a) is the following polynomial of order N in the scaling variable a: _2 dcN + 1 (a) da

2 ~ E ( (1 - If-j)/2 ) (N ~ J N+l-j

+ 1- j) (_a)N - j

j=O

The proof starts by differentiating (5B.3) with respect to dWjJ ~[ P+2 dn = ~ cl(a) - - 2 - lcl (a)

.

(5B.7)

n, yielding

dCl] (

+ n dn

9 )l n(p+2)/2 .

(5B.8)

Using the chain rule of differentiation we see from (5B.4) that del _ [n(P+2)/2 n dn - 2 9

+

P - 2 ] dCl 2

(5B.9)

ada'

and (5B.8) can be rewritten as dWjJ dn

N [(

~

P +2 ) 1- -2- I cl(a)

+

(n(p+2)/2 2 9

P - 2 ) del] ( 9 )l n(P+2)/2 . da

+ -2-a

(5B.IO) After rearranging the sum, this becomes

(5B.11) The first term vanishes trivially since co happens to be independent of a. The sum in the second line vanishes term by term: (5B.12) To see this we form the derivative

2 del + 1 = 2 ~ E- ( (1- Pt 2j)/2 ) ( . -I-I) (_ )l-j d

a

~

j=O

J

1+I-j

J

a

,

(5B.13)

564

5 Variational Perturbation Theory

and use the identity

2( (1-~j)/2) I+l-j

=

yj+21-1 ((I-~j)/2 ) j-l-l l-j

(5B.14)

to rewrite (5B.13) as

2 dct +1 da

=~E-( (1 - l-j ~j)/2) L..JJ

(P - 2 j +21_1) (-a)l-j

j=O

implying

P - 2 adct = ~ 2 da L..J

j=O

2

E- ( (1 - l-J Efj)/2 ) P-

'

2(1_ j)(-a)l-j.

2

J

(5B.15)

(5B.16)

By combining this with (5B.4), (5B.13) , we obtain Eq. (5B.12) which proves that the second line in (5B.ll) vanishes. Thus we are left with the last term on the right-hand side of Eq. (5B.ll). Using (5B .12) for 1= N leads to

dWjJ

=

9

-2 (

dO.

n(p+2)/2

) N dcN+1(a) da'

(5B.17)

When expressing d€N+l (a)/da with the help of (5B.4), we arrive at

dWjJ dO.

_

-

(

9

2 n(P+2)/2

{;EJ. (

)N N

(1 - ~j)/2 ) _. _ N-j N+l-j (N + l J)( a) . (5B.18)

This proves the scaling relation (5B.6) with the polynomial (5B.7) . The proof can easily be extended to physical quantities QN(9) with a different physical dimension n , which have an expansion (5B.19) rather than (5B.2). In this case the quantity [Q N(gW/" has again an expansion like (5B.2). By rewriting QN(g) as {[QN(gW /" }" and forming the derivative using the chain rule we see that the derivative vanishes whenever the polynomial PN(a) vanishes, which is formed from [QN(9W/a as in Eq. (5B.7).

Appendix 5C

Second-Order Shift of Polaron Energy

For brevity, we introduce the dimensionless variable p == wf!.r and

(5C.l) Going to natural units with 1i = M = w = 1, Feynman's variational energy (5.395) takes the form

Eo =

~(r 4r

0.)2 - _1_ r [= dpe- PF- 1/ 2(p). .,;:IT

Jo

(5C .2)

The second-order correction (5.406) reads

f!.E(2) = _ 412 r2J + _1_(r2 _ 0. 2 ) [= dpe-PF-1/2(p) _ _ 3_(r2 _ 0.2)2 o 7r 2r.,;:IT Jo 16r3 _ _ I_(r2 _ n2) [= dpe-PF-3/2(p) {(1+ 0. 2 ) (l -e-rp) + r2_n2 pe-rp}, 4.,;:IT Jo r r

(5C.3)

565

Notes and References where I denotes the integral

(5C.4) with

+ PI 2': 0 + PI 2': 0 P2 + PI < 0

Q = Ql for P3 - P2

and P3 - P2 2': 0,

Q = Q2 for P3 - P2

and P3 - P2 < 0,

Q = Q3 for P3 -

and P3 - P2 < 0,

(5C.5)

and Ql =

~F-l/2(pDF-l/2(p~) r2; 0 2e- rp3 (1 _ e- rp1 )(erp -

Q2 =

~F-l/2(pdF-l/2(P2)

x

Q3 = X

1),

(5C.6)

2

{r2; 2 0

[e - r (P2 - P3) _ e- rp3 (1

+ e- r (P' - P2) _

e- rP1 )] - 202(P2 - P3) } ,

(5C.7)

~F-l/2(pdF-l/2(P2)

{r2; 0 2

[ _e-rp3 (1

_ e- rp1 ) _

e- r (P2-P3) (e rP1 -

1)] - 20 2PI} .

(5C.8)

Notes and References The first-order variational approximation to the effective classical partition function V eff cl (xo) presented in this chapter was developed in 1983 by RP. Feynman and H. Kleinert, Phys. Rev. A 34, 5080 (1986) (http://www.physik.iu-berlin . derkleinert/159) . For further development see: H. Kleinert, Phys. Lett. A 118, 195 (1986) (ibid.http/148) ; B 181, 324 (1986) (ibid.http/151) ; W. Janke and B.K. Chang, Phys. Lett. B 129, 140 (1988); W . Janke, in Path Integrals from meV to MeV, ed. by V. Sa-yakanit et ai., World Scientific, Singapore, 1990. A detailed discussion of the accuracy of the approach in comparison with several other approximation schemes is given by S. Srivastava and Vishwamittar, Phys. Rev. A 44, 8006 (1991) . For a similar, independent development containing applications to simple quantum field theories, see R Giachetti and V. Tognetti, Phys. Rev. Lett. 55, 912 (1985); Int. J. Magn. Mater. 54-57, 861 (1986) ; R Giachetti, V. Tognetti, and R Vaia, Phys. Rev. B 33, 7647 (1986); Phys. Rev. A 37, 2165 (1988) ; Phys. Rev. A 38, 1521, 1638 (1988) ; Physica Scripta 40, 451 (1989). R Giachetti, V. Tognetti, A. Cuccoli, and R Vaia, lecture presented at the XXVI Karpacz School of Theoretical Physics, Karpacz, Poland, 1990. See also R Vaia and V. Tognetti, Int. J . Mod. Phys. B 4,2005 (1990); A. Cuccoli, V. Tognetti, and R Vaia, Phys. Rev. B 41 , 9588 (1990); A 44, 2743 (1991); A. Cuccoli, A. Maradudin, A.R McGurn, V. Tognetti, and R Vaia, Phys. Rev. D 46, 8839 (1992) .

566

5 Variational Perturbation Theory

The variational approach has solved some old problems in quantum crystals by extending in a simple way the classical methods into the quantum regime. See V.1. Yukalov, Mosc. Univ. Phys. Bull. 31, 10-15 (1976); S. Liu, G.K. Horton, and E .R. Cowley, Phys. Lett. A 152, 79 (1991) ; A. Cuccoli, A. Macchi, M. Neumann, V. Tognetti, and R. Vaia, Phys. Rev. B 45, 2088 (1992). The systematic extension of the variational approach was developed by H. Kleinert, Phys. Lett. A 173, 332 (1993) (quant-ph/9511020) . See also J. Jaenicke and H. Kleinert, Phys. Lett. A 176, 409 (1993) (ibid.http/217) ; H. Kleinert and H. Meyer, Phys. Lett. A 184, 319 (1994) (hep-th/9504048). A similar convergence mechanism was first observed within an order-dependent mapping technique in the seminal paper by R. Seznec and J. Zinn-Justin, J. Math. Phys. 20, 1398 (1979). For an introduction into various resummation procedures see C.M. Bender and S.A. Orszag, Advanced Mathematical Methods fo r Scientists and Engineers , McGraw-Hill, New York, 1978. The proof of the convergence of the variational perturbation expansion to be given in Subsection 17.10.5 went through the following stages: First a weak estimate was found for the anharmonic integral: I.R.C. Buckley, A. Duncan, H.F. Jones, Phys. Rev. D 47, 2554 (1993); C.M. Bender, A. Duncan, H.F. Jones, Phys. Rev. D 49, 4219 (1994). This was followed by a similar extension to the quantum-mechanical case: A. Duncan and H.F. Jones, Phys. Rev. D 47, 2560 (1993) ; C. Arvanitis, H.F. Jones, and C.S. Parker, Phys.Rev. D 52, 3704 (1995) (hep-ph/9502386); R. Guida, K. Konishi, and H. Suzuki, Ann. Phys. 241, 152 (1995) (hep-th/9407027). The exponentially fast convergence observed in the calculation of the strong-coupling coefficients of Table 5.9 was, however, not explained. The accuracy in the table was reached by working up to the order 251 with 200 digits in Ref. [13]. The analytic properties of the strong-coupling expansion were studied by C.M. Bender and T .T . Wu, Phys. Rev. 184, 1231 (1969); Phys. Rev. Lett. 27, 461 (1971) ; Phys. Rev. D 7, 1620 (1973) ; ibid. D 7, 1620 (1973); C.M. Bender, J . Math. Phys. 11, 796 (1970); T. Banks and C.M. Bender, J. Math. Phys. 13, 1320 (1972) ; J.J . Loeffel and A. Martin, Cargese Lectures on Physics (1970) ; D. Bessis ed., Gordon and Breach, New York 1972, Vol. 5, p.415; B. Simon, Ann. Phys. (N.Y.) 58, 76 (1970) ; Cargese Lectures on Physics (1970) , D. Bessis ed., Gordon and Breach, New York 1972, Vol. 5, p. 383. The problem of tunneling at low barriers (sliding) was solved by H. Kleinert , Phys. Lett. B 300, 261 (1993) (ibid.http/214) . See also Chapter 17. Some of the present results are contained in H. Kleinert, Pfadintegrale in Quantenmechanik, Statistik und Polymerphysik , B.-I. Wissenschaftsverlag, Mannheim, 1993. A variational approach to tunneling is also used in chemical physics: M.J. Gillan, J. Phys. C 20, 362 (1987) ; G.A. Voth, D. Chandler, and W.H. Miller, J. Chern. Phys. 91 , 7749 (1990); G.A. Voth and E.V. OGorman, J. Chern. Phys. 94, 7342 (1991) ; G.A. Voth, Phys. Rev. A 44,5302 (1991). Variational approaches without the separate treatment of Xo have been around in the literature for some time:

Notes and References

567

T. Barnes and G.1. Ghandour, Phys. Rev. D 22, 924 (1980) ; B.S. Shaverdyan and A.G. Usherveridze, Phys. Lett. B 123, 316 (1983) ; K. Yamazaki, J. Phys. A 17, 345 (1984) ; H. Mitter and K. Yamazaki, J. Phys. A 17, 1215 (1984); P.M. Stevenson, Phys. Rev. D 30,1712 (1985); D 32,1389 (1985); P .M. Stevenson and R Tarrach, Phys. Lett. B 176, 436 (1986); A. Okopinska, Phys. Rev. D 35, 1835 (1987); D 36, 2415 (1987) ; W. Namgung, P.M. Stevenson, and J.F. Reed, Z. Phys. C 45, 47 (1989) ; U. Ritschel, Phys. Lett. B 227, 44 (1989); Z. Phys. C 51 , 469 (1991) ; M.H. Thoma, Z. Phys. C 44,343 (1991); 1. Stancu and P.M. Stevenson, Phys. Rev. D 42, 2710 (1991); R Tarrach, Phys. Lett. B 262, 294 (1991); H. Haugerud and F. Raunda, Phys. Rev. D 43, 2736 (1991); A.N. Sissakian, 1.1. Solovtsov, and O.Y. Shevchenko, Phys. Lett. B 313, 367 (1993). Different applications of variational methods to density matrices are given in V.B. Magalinsky, M. Hayashi, and H.V. Mendoza, J. Phys. Soc. Jap. 63,2930 (1994); V.B. Magalinsky, M. Hayashi, G.M. Martinez Pena, and R Reyes Sanchez, Nuovo Cimento B 109, 1049 (1994). The particular citations in this chapter refer to the publications [1] H. Kleinert, Phys. Lett. A 173, 332 (1993) (quant-ph/9511020) . [2] The energy eigenvalues of the anharmonic oscillator are taken from F .T . Hioe, D. MacMillan, and E .W. Montroll, Phys. Rep. 43, 305 (1978); W. Caswell, Ann. Phys. (N.Y.) 123, 153 (1979) ; R1. Somorjai, and D.F. Hornig, J. Chern. Phys. 36, 1980 (1962) . See also K. Banerjee, Proc. Roy. Soc. A 364, 265 (1978) ; R Balsa, M. Plo, J.G. Esteve, A.F. Pacheco, Phys. Rev. D 28, 1945 (1983); and most accurately F. Vinette and J . Cizek, J. Math. Phys. 32, 3392 (1991) ; E.J . Weniger, J. Cizek, J . Math. Phys. 34, 571 (1993). [3] RP. Feynman, Statistical Mechanics , Benjamin, Reading, 1972, Section 3.5. [4] M. Hillary, RF. O'Connell, M.O . Scully, and E.P. Wigner, Phys. Rep. 106, 122 (1984). [5] H. Kleinert, Phys. Lett. A 118, 267 (1986) (ibid.http/145). [6] For a detailed discussion of the effective classical potential of the Coulomb system see W. Janke and H. Kleinert, Phys. Lett. A 118, 371 (1986) (ibid.http/153). [7] C. Kouveliotou et al., Nature 393, 235 (1998); Astroph. J. 510, L115 (1999) ; K. Hurley et al. , Astroph. J. 510, L111 (1999); V.M. Kaspi, D. Chakrabarty, and J. Steinberger, Astroph. J. 525, L33 (1999) ; B. Zhang and A.K. Harding, (astro-ph/0004067). [8] The perturbation expansion of the ground state energy in powers of the magnetic field B was driven to high orders in J.E. Avron, B.G. Adams, J. Cizek, M. Clay, M.L. Glasser, P. Otto, J. Paldus, and E. Vrscay, Phys. Rev. Lett. 43, 691 (1979); B.G. Adams, J.E. Avron, J. Cizek, P. Otto, J . Paldus, RK. Moats, and H.J. Silverstone, Phys. Rev. A 21, 1914 (1980). This was possible on the basis of the dynamical group 0(4,1) and the tilting operator (13.181) found by the author in his Ph.D. thesis. See H. Kleinert, Group Dynamics of Elementary Particles, Fortschr. Physik 6, 1 (1968)

568

5 Variational Perturbation Theory (ibid.http/i); H. Kleinert, Group Dynamics of the Hydrogen Atom, Lectures in Theoretical Physics, edited by W.E. Brittin and A.O. Barut, Gordon and Breach, N.Y. 1968, pp. 427-482 (ibid.http/4).

[9] Precise numeric calculations of the ground state energy of the hydrogen atom in a magnetic field were made by H. Ruder, G. Wunner, H. Herold, and F. Geyer, Atoms in Strong Magnetic Fields (SpringerVerlag, Berlin, 1994). [10] M. Bachmann, H. Kleinert, and A. Pelster, Phys. Rev. A 62, 52509 (2000) (quantph/0005074), Phys. Lett. A 279,23 (2001) (quant-ph/00051O). [11] L.D. Landau and E .M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965. [12] J .C. LeGuillou and J. Zinn-Justin, Ann. Phys. (N.Y.) 147, 57 (1983). [13] W. Janke and H. Kleinert, Phys. Rev. Lett. 75, 2787 (1995) (quant-ph/9502019) . [14] The high accuracy became possible due to a scaling relation found in W. Janke and H. Kleinert, Phys. Lett. A 199, 287 (1995) (quant-ph/9502018). [15] For the proof of the exponentially fast convergence see H. Kleinert and W. Janke, Phys. Lett. A 206, 283 (1995) (quant-ph/9502019); R. Guida, K. Konishi, and H. Suzuki, Ann. Phys. 249, 109 (1996) (hep-th/9505084) . The proof will be given in Subsection 17.10.5. [16] The oscillatory behavior around the exponential convergence shown in Fig. 5.22 was explained in H. Kleinert and W. Janke, Phys. Lett. A 206, 283 (1995) (quant-ph/9502019) in terms of the convergence radius of the strong-coupling expansion (see Section 5.15) . [17] H. Kleinert, Phys. Rev. D 60 , 085001 (1999) (hep-th/9812197); Phys. Lett. B 463, 69 (1999) (cond-mat/9906359). See also Chapters 19-20 in the textbook H. Kleinert and V. Schulte-Frohlinde, Critical Properties of ip4-Theories, World Scientific, Singapore 2001 (ibid.http/b8) [18] H. Kleinert, Phys. Lett. A 207, 133 (1995) . [19] See for example the textbooks H. Kleinert, Gauge Fields in Condensed Matter, Vol. I Superflow and Vortex Lines, Vol. II Stresses and Defects, World Scientific, Singapore, 1989 (ibid.http/bi). [20] R.P. Feynman, Phys. Rev. 97, 660 (1955). [21] S. Hohler and A. Miillensiefen, Z. Phys. 157, 159 (1959) ; M.A. Smondyrev, Theor. Math. Fiz. 68, 29 (1986); O.V. Selyugin and M.A. Smondyrev, Phys. Stat. Sol. (b) 155, 155 (1989) . [22] N.N. Bogoliubov (jun) and V.N. Plechko, Teor. Mat. Fiz. [Sov. Phys.-Theor. Math. Phys.]' 65, 423 (1985); illv. Nuovo Cimento 11, 1 (1988) . [23] S.J. Miyake, J. Phys. Soc. Japan, 38, 81 (1975). [24] J.E. Avron, LW. Herbst, B. Simon, Phys. Rev. A 20, 2287 (1979). [25] LD. Feranshuk and L.L Komarov, J . Phys. A: Math. Gen. 17, 3111 (1984) . [26] H. Kleinert, W. Kiirzinger, and A. Pelster, J. Phys. A: Math. Gen. 31, 8307 (1998) (quantph/9806016). [27] H. Kleinert, Phys. Lett. A 173, 332 (1993) (quant-ph/9511020).

Notes and References

569

[28] H. Kleinert, Phys. Rev. D 57, 2264 (1998) and Addendum: Phys. Rev. D 58, 107702 (1998). [29] F.J. Wegner, Phys. Rev. B 5, 4529 (1972); B 6, 1891 (1972). [30] H. Kleinert, Phys. Lett. A 207, 133 (1995) (quant-ph/9507005) . [31] J. Rossler, J. Phys. Stat. Sol. 25, 311 (1968). [32] J.T. Marshall and L.R. Mills, Phys. Rev. B 2, 3143 (1970). [33] Higher-order smearing formulas for nonpolynomial interactions were derived in H. Kleinert, W. Kiirzinger and A. Pelster, J. Phys. A 31 , 8307 (1998) (quant-ph/9806016). [34] The polaron problem is solved in detail in the textbook R.P. Feynman, Statistical Mechanics, Benjamin, New York, 1972, Chapter 8. Extensive numerical evaluations are found in T.D. Schultz, Phys. Rev. 116, 526 (1959); See also M. Dineykhan, G.V. Efimov, G. Ganbold, and S.N. Nedelko, Oscillator Representation in Quantum Physics, Springer, Berlin, 1995. An excellent review article is J.T. Devreese, Polarons, Review article in Encyclopedia of Applied Physics, 14, 383 (1996) (cond-mat /0004497). This article contains ample references on work concerning polarons in magnetic fields, for instance F .M. Peeters, J.T. Devreese, Phys. Stat. Sol. B 110, 631 (1982); Phys. Rev. B 25, 7281, 7302 (1982) ; Xiaoguang Wu, F.M. Peeters, J .T. Devreese, Phys. Rev. B 32, 7964 (1985); F. Brosens and J.T. Devreese, Phys. Stat. Sol. B 145, 517 (1988). For discussion of the validity of the Jensen-Peierls inequality (5.10) in the presence of a magnetic field, see J.T. Devreese and F . Brosens, Solid State Communs. 79, 819 (1991) ; Phys. Rev. B 45, 6459 (1992) ; Solid State Communs. 87, 593 (1993); D. Larsen in Landau Level Spectroscopy , Vol. 1, G. Landwehr and E. Rashba (eds.), North Holland, Amsterdam, 1991, p. 109. The paper D. Larsen, Phys. Rev. B 32, 2657 (1985) shows that the variational energy can lie lower than the exact energy. The review article by Devreese contains numerous references on bipolarons, small polarons, and polaronic excitations. For instance: J.T. Devreese, J . De Sitter, M.J. Goovaerts, Phys. Rev. B 5, 2367 (1972) ; L.F. Lemmens, J. De Sitter, J.T. Devreese, Phys. Rev. B 8, 2717 (1973) ; J.T. Devreese, L.F. Lemmens, J. Van Royen, Phys. Rev. B 15, 1212 (1977); J. Thomchick, L.F. Lemmens, J.T. Devreese, Phys. Rev. B 14, 1777 (1976) ; F .M. Peeters, Xiaoguang Wu, J .T . Devreese, Phys. Rev. B 34, 1160 (1986); F .M. Peeters, J .T. Devreese, Phys. Rev. B 34, 7246 (1986); B 35, 3745 (1987); J .T . Devreese, S.N. Klimin, V.M. Fomin, F . Brosens, Solid State Communs. 114, 305 (2000) . There exists also a broad collection of articles in E .K.H. Salje, A.S. Alexandrov, W .Y. Liang (eds.), Polarons and Bipolarons in High-Tc Superconductors and Related Materials , Cambridge University Press, Cambridge, 1995. A generalization of the harmonic trial path integral (5.356), in which the exponential function e-OIT-T' 1 at zero temperature is replaced by I( IT - T'I), has been proposed by M. Saitoh, J . Phys. Soc. Japan. 49, 878 (1980), and further studied by R. Rosenfelder and A.W. Schreiber, Phys. Lett. A 284, 63 (2001) (cond-mat/ 0011332).

570

5 Variational Perturbation Theory In spite of a much higher numerical effort, this generalization improves the ground state energy only by at most 0.1 % (the weak-coupling expansion coefficient -0.012346 in (5.396) is changed to -0.012598, while the strong-coupling coefficients in (5.399) remain unchanged at this level of accuracy. For the effective mass, the lowest nontrivial weak-coupling coefficient of 12 in (5.402) is changed by 0.0252 % while the strong-coupling coefficients in (5.403) remain unchanged at this level of accuracy.

Aevo rarissima nostro, simplicitas.

Simplicity, a very rare thing in our age. OVID, Ars Amatoria, Book 1, 241

6 Path Integra Is with T opologica I Constra ints The path integral representations of the time evolution amplitudes considered so far were derived for orbits x(t) fluctuating in Euclidean space with Cartesian coordinates. Each coordinate runs from minus infinity to plus infinity. In many physical systems, however, orbits are confined to a topologically restricted part of a Cartesian coordinate system. This changes the quantum-mechanical completeness relation and with it the derivation of the path integral from the time-sliced time evolution operator in Section 2.1. We shall consider here only a point particle moving on a circle, in a half-space, or in a box. The path integral treatment of these systems is the prototype for any extension to more general topologies.

6.1

Point Particle on Circle

For a point particle on a circle, the orbits are specified in terms of an angular variable 'P(t) E [0,271'] subject to the topological constraint that 'P = 0 and 'P = 271' be identical points. The initial step in the derivation of the path integral for such a system is the same as before: The time evolution operator is decomposed into a product .

('Pbtbl'Pata ) = ('Pb lexp

[-~(tb -

II

N+l

ta)iI] l'Pa) == ('Pb l

'

exp ( -~EiI) l'Pa }.

(6.1)

The restricted geometry shows up in the completeness relations to be inserted between the factors on the right-hand side for n = 1, . .. , N:

(6.2) If the integrand is singular at 'P = 0, the integrations must end at an infinitesimal piece below 271'. Otherwise there is the danger of double-counting the contributions from the identical points 'P = 0 and 'P = 271'. The orthogonality relations on these intervals are

(6.3) 571

572

6 Path Integrals with Topological Constraints

The O.

(6.26)

Given a free particle moving in such a geometry, we want to calculate N+l

(rbtblrata) =

.

(rbllI exp ( -~iiE)

Ira) .

(6.27)

576

6 Path Integrals with Topological Constraints

As usual, we insert N completeness relations between the N of a vanishing Hamiltonian, the amplitude (6.27) becomes

+ 1 factors.

In the case

(6.28) For each scalar product (rnlrn-l) = lS(rn - rn-l) , we substitute its spectral representation appropriate to the infinite-wall boundary at r = O. It consists of a superposition of the free-particle wave functions vanishing at r = 0:

(r lrl )

2 rOOdksinkrsinkr'

Jo

1 dk 00

-00

(6.29)

7r

27r

[expik(r - rl) - expik(r + r l)] = lS(r - rl) - lS(r + rl).

This Fourier representation does a bit more than what we need. In addition to the IS-function at r = r' , there is also a IS-function at the unphysical reflected point r = -r'. The reflected point plays a similar role as the periodically repeated points in the representation (6.11) . For the same reason as before, we retain the reflected points in the formula as though rl were permitted to become zero or negative. Thus we rewrite the Fourier representation (6.29) as

(r lrl)

=

x~r

1

00 -00

dp

[i

]

27r1i exp "h,p(x - Xl) + i7r(u(x) - U(XI)) x'=r"

where

u(x)

= 8(-x)

(6.30)

(6.31)

with the Heaviside function 8(x) of Eq. (1.309). For symmetry reasons, it is convenient to liberate both the initial and final positions rand rl from their physical half-space and to introduce the localized states Ix) whose scalar product exists on the entire x-axis:

(x lx l)

1

L 00 dP1i exp [~p(XIl - Xl) x"=±x -00 27r lS(x - Xl) - lS(x + Xl).

+ i7r(U(X") -

u(xl))] (6.32)

With these states, we write

(r lrl ) = (x lxl) lx=r,x'=r"

(6.33)

We now take the trivial transition amplitude with zero Hamiltonian (6.34) extend it with no harm by the reflected IS-function (6.35)

577

6.2 Infinite Wall

and factorize it into many time slices: (6.36)

(rb = r N+1, r a = ro), where the trivial amplitude of a single slice is (xnflxn-10)0

=

(XnIXn-l),

x E (-00,00).

(6.37)

With the help of (6.32), this can be written as

(6.38) The sum over the reflected points Xn = ±rn is now combined, at each n, with the integral Jooo drn to form an integral over the entire x-axis, including the unphysical half-space x < O. Only the last sum cannot be accommodated in this way, so that we obtain the path integral representation for the trivial amplitude

(6.39) The measure of this path integral is now of the conventional type, integrating over all paths which fluctuate through the entire space. The only special feature is the final symmetrization in Xb = ±rb. It is instructive to see in which way the final symmetrization together with the phase factor exp[i7ra(x)] = ±1 eliminates all the wrong paths in the extended space, i.e., those which cross the origin into the unphysical subspace. This is illustrated in Fig. 6.2. Note that having assumed Xa = ra > 0, the initial phase a(x a ) can be omitted. We have kept it merely for symmetry reasons. In the continuum limit, the exponent corresponds to an action (6.40)

The first term is the usual canonical expression in the absence of a Hamiltonian. The second t erm is new. It is a pure boundary term: (6.41)

which keeps track of the topology of the half space x > 0 embedded in the full space x E (-00,00). This is why the action carries the subscript "topol" .

578

6 Path Integrals with Topological Constraints

Xa

x

Figure 6.2 Illustration of path counting near reflecting wall. Each path touching the wall once is canceled by a corresponding path of equal action crossing the wall once into the unphysical regime (the path is mirror-reflected after the crossing). The phase factor exp[i11'0'(Xb) ] provides for the opposite sign in the path integral. Only paths not touching the wall at all cannot be canceled in the path integral.

The topological action (6.41) can be written formally as a local coupling of the velocity at the origin:

Afopol[xl = -11'n

l

tb

ta

dtx(t)8(x(t)).

(6.42)

This follows directly from (6.43) Consider now a free point particle in the right half-space with the usual Hamiltonian p2 H= 2M

(6.44)

The action reads

AlP, xl =

l

tb

dtlPx - p2/2M - n1fx(t)8(x(t))],

(6.45)

ta

and the time-sliced path integral looks like (6.39) , except for additional energy terms -p;'/2M in the action. Since the new topological term is a pure boundary term, all the extended integrals in (6.39) can be evaluated right away in the same way as for a free particle in the absence of an infinite wall. The result is

6.3 Point Particle in Box

579

with Xa = ra' This is indeed the correct result: Inserting the Fourier transform of the Gaussian (Fresnel) distribution we see that

(6.47) which is the usual spectral representation of the time evolution amplitude. Note that the first part of (6.46) may be written more symmetrically as

In this form, the phase factors e i7r 0 corresponds to a strong-coupling limit in the reduced coupling constant g. According to the general theory of variational perturbation theory and its strong-coupling limit in Sections 5.14 and 5.17, the Nth order approximation to the strong-coupling limit of p(g), to be denoted by p' , is found by replacing, in the series truncated after the 3The programs can be downloaded from www.physik.fu-berlin.derkleinert/b5/programs

586

6 Path Integrals with Topological Constraints

Nth term, FNCg/w), the frequency w by the identical expression where

Vn2 - gr /2M, 2

C6.81) For a moment , this is treated as an independent variable, whereas n is a dummy parameter. Then the square root is expanded binomially in powers of g, and FNCg/Vn 2 - gr 2 /2M) is re-expanded up to order gN. After that, r is replaced by its proper value. In this way we obtain a function FNCg, n) which depends on n, which thus becomes a variational parameter. The best approximation is obtained by extremizing FNCg , n) with respect to w. Setting w = 0, we go to the strong-coupling limit g -+ 00 . There the optimal n grows proportionally to g, so that gin = e- 1 is finite, and the variational expression FNCg, n) becomes a function of fNCe) . In this limit, the above re-expansion amounts simply to replacing each power w n in each expansion terms of FNCg) by the binomial expansion of C1 - 1) -n/2 truncated after the CN - n )th term, and replacing 9 by c 1 . The first nine variational functions fNCe) are listed in Table 6.1. The functions fNCe) are minimized starting from hCe) and searching the minimum of each successive hCe), hCe), .. . nearest to the previous one. The functions fNCe) together with their minima are plotted in Fig. 6.5. The minima lie at

Table 6.1

First eight variational functions fN(C),

hCe) = ~ + l~ c + ~~ hCe) = ~ + 3~C + ~ 1 1 15 35c f 4 C) e = 4: - 256c3 + 128c + 256 C) e C) J6 e f 7 C) e f C) J8 e f C) f

J5 f

J9

5 35 = 4:1 - 512c3 + 256c + 63c 512 1 1 35 315 231 c = 4: + 2048c5 - 2048c3 + 2048c + 2048 1 7 105 693 429c = 4: + 4096c5 - 4096c3 + 4096c + 4096 1 5 63 1155 3003 6435c = 4: - 65536c7 + 16384c5 - 32768c3 + 16384c + 65536 1 45 231 3003 6435 12155 c e = 4: - 131072 c7 + 32768 c5 - 65536 c 3 + 32768 c + 131072

c 0.8

0.494 0.492 0.49

Figure 6.5 Variational functions fN(C) for particle between walls up to N = 16 are shown

together with their minima whose y-coordinates approach rapidly the correct limiting value 1/2.

587

6.4 Strong-Coupling Theory for Particle in Box

(N, J'lJin ) = (2,0.466506), (3,0.492061), (4,0.497701), (5,0.499253), (6,0.499738), (7,0.499903), (8,0.499963), (9,0.499985), (10,0.499994), (11,0.499998), (12,0.499999), (13,0.5000), (14,0.50000), (15,0.50000), (16,0.5000).

(6.82)

They converge exponentially fast against the known result 1/2, as shown in Fig. 6.6.

6.4.4

Special Properties of Expansion

The alert reader will have noted that the expansion coefficients (6.80) possess two special properties: First, they lack the factorial growth at large orders which would be found for a single power [u 2 (r)]k+1 of the interaction potential, as mentioned in Eq.(3C.27) and will be proved in Eq. (17.323) . The factorial growth is canceled by the specific combination of the different powers in the interaction (6.72), making the series (6.71) convergent inside a certain circle. Still, since this circle has a finite radius (the ratio test shows that it is unity) , this convergent series cannot be evaluated in the limit of large 9 which we want to do, so that variational strong-coupling theory is not superfluous. However, there is a second remarkable property of the coefficients (6.80): They contain an infinite number of zeros in the sequence of coefficients for each odd number, except for the first one. We may take advantage of this property by separating off the irregular term a19 = 9/4 = n: 2/4d 2 , setting a = 9 2 / 4w 2 , and rewriting F(g) as N

F(a)

1 ] , 4"1 [ 1 + foh(a)

=

h(a) ==

L 22n+la2n a n .

(6.83)

n=O

Inserting the numbers (6.80), the expansion of h(a) reads

h(a) = 1

~ _ a2 a3 + 2 8 + 16

_

~ a4 128

+

..!....- a 5 _ ~a6 ~ a 7 _ 256

1024

+

2048

429

32768 a

8

+ .... (6.84)

We now realize that this is the binomial power series expansion of VI + a. Substituting this into (6.83), we find the exact ground state energy for the Euclidean action (6.67) (6.85)

0.5 0.499998 0.499996 0.499994

• •

0.00005



!'!Jin



0.499992

Figure 6.6 exact value.

Exponentially fast convergence of strong-coupling approximations towards

588

6 Path Integrals with Topological Constraints

Here we can go directly to the strong-coupling limit a -. 00 to recover the exact ground state energy E(O) = 7r 2 /2d 2 . The energy (6.85) can of course be obtained directly by solving the Schrodinger equation associated with the potential (6.72), 2

.\) -1 { -8-2 + [,\(1 - -22

8x

cos

d2 1] } 'l/J(x) = -E'l/J(x), 7r 2

X

where we have replaced u -. dX/7r and set w 2 d 4 /7r 4

(6.86)

== .\(.\ - 1), so that (6.87)

Equation (6.86) is of the Piischl-Teller type [see Subsection 14.4.5], and has the ground state wave function, to be derived in Eq. (14.163),

'l/JO(x) = const x cos'" x ,

(6.88)

with the eigenvalue 7r 2E(O) /d 2 = (.\2 - 1)/2, which agrees of course with Eq. (6.85). If we were to apply the variational procedure to the series h(a)/fo in F of Eq. (6.85), by replacing the factor l/w 2n contained in each power an by fl = y'fl2 - ra and re-expanding now in powers of a rather than g, we would find that all approximation hN(C) would possess a minimum with unit value, such that the corresponding extremal functions fN(C) yield the correct final energy in each order N .

6.4.5

Exponentially Fast Convergence

With the exact result being known, let us calculate the exponential approach of the variational approximations observed in Fig. (6.6). Let us write the exact energy (6.85) as 1

E(O) -- _(g 4 + ...)g2 + 4w2) .

(6.89)

After the replacement w -. ...)fl2 - pg, this becomes (6.90)

where g == g/fl2. The Nth-order approximant fN(9) of E(O) is obtained by expanding (6.91) in powers of g up to order N, N

(6.91) fN(g) = fl"Lhk(p)gk, o and substituting p by 2Mr2 = (1- C;P)/g [compare (6.81)], with 12>2 == w2/fl2 . The resulting function of g is then optimized. It is straightforward to find an integral representation for FN(g) . Setting rg == z, we have 1

FN =

-2· 7rZ

i

coz

dz N+1

1- zN+l 1 f(z),

- z

(6.92)

where the contour Co refers to small circle around the origin and

F(z)

=

~4 (=r + Jr2Z2 -4Z+4) (6.93)

Notes and References with branch points at

589

Zl,2

= 2r2 (1

±

VI - l/r2).

For z < 1, we rewrite

1 - zN+l = (1 - z)(1 + z + ... + zN) = (1 - z)(N + 1) -(1- z)2 [N + (N -1)z + .. . + zN- l] and estimate this for z

~

(6.94)

1 as (6.95)

Dividing the approximant (6.92) by 11, and indicating this by a hat, we use (6.94) to write FN as a sum over the discontinuities across the two branch cuts:

(N + 1) 27ri

(N + 1)

1 dzF(z) F(z) = leo zN+l

L1

00

2

i=l

Zi

(N + 1) F(N)(O) N!

d F(z). N:l

(6.96)

Z

The integrals yield a constant plus a product (6.97) which for large N can be approximated using Stirling's formula (5.204) by (6.98) In the strong-coupling limit of interest here, w2 = 0, and r = 1/9 = l1/g = c. In Fig. 6.5 we see that the optimal c-values tend to unity for N --+ 00, so that l!J.jN goes to zero like e- N , as observed in Fig. 6.6.

Notes and References There exists a large body of literature on this subject, for example L.S. Schulman, J. Math. Phys. 12, 304 (1971); M.G.G. Laidlaw and C. DeWitt-Morette, Phys. Rev. D 3, 1375 (1971); J.S. Dowker, J. Phys. A 5, 936 (1972); P.A. Horvathy, Phys. Lett. A 76, 11 (1980) and in Differential Geometric Methods in Math . Phys. , Lecture Notes in Mathematics 905, Springer, Berlin, 1982; J.J. Leinaas and J . Myrheim, Nuovo Cimento 37, 1, (1977). The latter paper is reprinted in the textbook F . Wilczek, Fractional Statistics and Anyon Superconductivity, World Scientific, 1990. See further P.A. Horvathy, G. Morandi, and E .C.G. Sudarshan, Nuovo Cimento D 11, 201 (1989) , and the textbook L.S. Schulman, Techniques and Applications of Path Integration, Wiley, New York, 1981. It is possible to account for the presence of hard walls using infinitely high 8-functions: C. Grosche, Phys. Rev. Lett. 71, 1 (1993); Ann. Phys. 2, 557 (1993); (hep-th/9308081); (hep-

th/9308082) ; (hep-th/9402110); M.J. Goovaerts, A. Babcenco, and J.T. Devreese, J . Math. Phys. 14, 554 (1973); C. Grosche, J . Phys. A Math. Gen. 17,375 (1984) .

590

6 Path Integrals with Topological Constraints

The physically important problem of membranes between walls has been discussed in W . Helfrich, Z. Naturforsch. A 33, 305 (1978); W. Helfrich and RM. Servuss, Nuovo Cimento D 3, 137 (1984); W . Janke and H. Kleinert, Phys. Lett. 58, 144 (1987) (http://www . physik. fu-berlin. del -kleinert/143); W. Janke, H. Kleinert, and H. Meinhardt, Phys. Lett. B 217, 525 (1989) (ibid.http/184) ; G. Gompper and D.M. Kroll, Europhys. Lett. 9, 58 (1989) ; RR Netz and R Lipowski, Europhys. Lett. 29. 345 (1995); F. David, J. de Phys. 51 , C7-115 (1990); H. Kleinert, Phys. Lett. A 257 ,269 (1999) (cond-mat/9811308); M. Bachmann, H. Kleinert, A. Pelster, Phys. Lett. A 261 , 127 (1999) (cond-mat/9905397). The problem has been solved with the help of the strong-coupling variational perturbation theory developed in Chapter 5 by H. Kleinert, Phys. Lett. A 257, 269 (1999) (cond-mat/9811308); M. Bachmann, H. Kleinert, and A. Pelster, Phys. Lett. A 261, 127 (1999) (cond-mat/9905397). The quantum-mechanical calculation presented in Section 6.4 is taken from H. Kleinert, A. Chervyakov, and B. Hamprecht, Phys. Lett. A 260, 182 (1999) (condmat/9906241).

Mirum, quod divina natura dedit agros. It's wonderful that divine nature has given us fields. VARRO, 82 B.C .

7 Many Particle Orbits Statistics and Second Quantization Realistic physical systems usually contain groups of identical particles such as specific atoms or electrons. Focusing on a single group, we shall label their orbits by x(v)(t) with // = 1,2, 3, ... , N. Their Hamiltonian is invariant under the group of all N! permutations of the orbital indices //. Their Schrodinger wave functions can then be classified according to the irreducible representations of the permutation group. Not all possible representations occur in nature. In more than two space dimensions, there exists a super-selection rule, whose origin is yet to be explained, which eliminates all complicated representations and allows only for the two simplest ones to be realized: those with complete symmetry and those with complete antisymmetry. Particles which appear always with symmetric wave functions are called bosons . They all carry an integer-valued spin. Particles with antisymmetric wave functions are called fermions 1 and carry a spin whose value is half-integer. The symmetric and antisymmetric wave functions give rise to the characteristic statistical behavior of fermions and bosons. Electrons, for example, being spin-1/2 particles, appear only in antisymmetric wave functions. The antisymmetry is the origin of the famous Pauli exclusion principle, allowing only a single particle of a definite spin orientation in a quantum state, which is the principal reason for the existence of the periodic system of elements, and thus of matter in general. The atoms in a gas of helium, on the other hand, have zero spin and are described by symmetric wave functions . These can accommodate an infinite number of particles in a single quantum state giving rise to the famous phenomenon of Bose-Einstein condensation. This phenomenon is observable in its purest form in the absence of interactions, where at zero temperature all particles condense in the ground state. In interacting systems, Bose-Einstein statistics can lead to the stunning quantum state of superfluidity. The particular association of symmetry and spin can be explained within relativistic quantum field theories in spaces with more than two dimensions where it is shown to be intimately linked with the locality and causality of the theory. 1 Had M. Born as editor of Zeitschrift fiir Physik not kept a paper by P. Jordan in his suitcase for half a year in 1925, they would be called jardanons . See the bibliographical notes by B. Schroer (hep-thj0303241) .

591

592

7 Many Particle Orbits - Statistics and Second Quantization

In two dimensions there can be particles with an exceptional statistical behavior. Their properties will be discussed in Section 7.5. In Chapter 16, such particles will serve to explain the fractional quantum Hall effect. The problem to be solved in this chapter is how to incorporate the statistical properties into a path integral description of the orbits of a many-particle system. Afterwards we describe the formalism of second quantization or field quantization in which the path integral of many identical particle orbits is abandoned in favor of a path integral over a single fluctuating field which is able to account for the statistical properties in a most natural way.

7.1

Ensembles of Bose and Fermi Particle Orbits

For bosons, the incorporation of the statistical properties into the orbital path integrals is quite easy. Consider, for the moment, distinguishable particles. Their many-particle time evolution amplitude is given by the path integral

II [J VDx(V)] N

t IX(l) X(N). t ) = (Xb(l) " . " X(Nl. b ,b a "." a ,a

eiA(N)

Iii ,

(7.1)

v=l

with an action of the typical form

(7.2) where V(x(v)) is some common background potential for all particles interacting via the pair potential Vint(x(v) - X(VI)). We shall ignore interactions involving more than two particles at the same time, for simplicity. If we want to apply the path integral (7.1) to indistinguishable particles of spin zero, we merely have to add to the sum over all paths x(v)(t) running to the final positions x~v) the sum of all paths running to the indistinguishable permuted final positions x~(v)) . The amplitude for n bosons reads therefore (1) (N).t)_"( (p(l)) , . .. ,X(p(N))·tl (1) (N) · t ) ( Xb(1) , ... , Xb(N).tl ,b xa , .. . , xa , a - L... Xb ,b Xa , ... , Xa , a , b

(73) .

p(v)

where p(v) denotes the N! permutations of the indices v. For bosons of higher spin, the same procedure applies to each subset of particles with equal spin orientation. A similar discussion holds for fermions. Their Schrodinger wave function requires complete antisymmetrization in the final positions. Correspondingly, the amplitude (7.1) has to be summed over all permuted final positions x~p(v)), with an extra minus sign for each odd permutation p(v). Thus, the path integral involves both sums and differences of paths. So far, the measure of path integration has always been a true sum over paths. For this reason it will be preferable to attribute the alternating sign to an interaction between the orbits, to be called a statistics interaction . This interaction will be derived in Section 7.4.

7.1 Ensembles of Bose and Fermi Particle Orbits

593

For the statistical mechanics of Bose- and Fermi systems consider the imaginarytime version of the amplitude (7.3): ( Xb(1), ... , Xb(N).,

"f31 Xa"' (1) "

IL

(N). Xa ,

0) -_ '" ((P(l)) (P(N)). "f31 (1) L..J fp(v) Xb , ... , Xb , IL Xa"'"

(N). Xa ,

p(v)

0) ,

(7.4)

where fp(v) = ±1 is the parity of even and odd permutations p(v) , to be used for Bosons and Fermions, respectively. Its spatial trace integral yields the partition function of N-particle orbits:

A factor liN! accounts for the indistinguishability of the permuted final configurations. For free particles, each term in the sum (7.4) factorizes:

where each factor has a path integral representation (xip(v))nf3lx~) 0)0 =

111i~ M ] VDx(v) exp - dT - X(v)2(T), l X(V)(Ii~)=x(P(V)) (O)=xl:') n0 2 [

b

(7.7)

X(V)

which is solved by the imaginary-time version of (2.130):

(7.8) The partition function can therefore be rewritten in the form Z oeN) --

l I ND

V27rn2f31M

N'

'

J

dD X (1)

...

d D X (N)", L..Jfp(v) p(v)

N II exp

{I

M [x(p(v)) _xCv)] 2 nf3

2}

----=-----:-::----''-

n

.

v=l

(7.9) This is a product of Gaussian convolution integrals which can easily be performed as before when deriving the time evolution amplitude (2.70) for free particles with the help of Formula (2.69). Each convolution integral simply extends the temporal length in the fluctuation factor by nf3 . Due to the indistinguishability of the particles, only a few paths will have their end points connected to their own initial points, i.e., they satisfy periodic boundary conditions in the interval (0, n(3) . The sum over permutations connects the final point of some paths to the initial point of a different path, as illustrated in Fig. 7.1. Such paths satisfy periodic boundary conditions on an interval (0, wn(3), where w is some integer number. This is seen most clearly by drawing the paths in Fig. 7.1 in an extended zone scheme shown in Fig. 7.2, which is reminiscent of Fig. 6.1. The extended zone scheme can, moreover , be placed on a

594

7 Many Particle Orbits - Statistics and Second Quantization

Ii~

T

Figure 7.1 Paths summed in partition function (7.9). Due to indistinguishability of particles, final points of one path may connect to initial points of another.

1

3 1

2

1

I-

-I

I-

-I

Ii~

,

,

31i_3

21if3

T

Figure 7.2 Periodic representation of paths summed in partition function (7.9), once in extended zone scheme, and once on D-dimensional hypercylinder embedded in D + 1 dimensions. The paths are shown in Fig. 7.1. There is now only one closed path on the cylinder. In general there are various disconnected parts of closed paths.

hypercylinder, illustrated in the right-hand part of Fig. 7.2. In this way, all paths decompose into mutually disconnected groups of closed paths winding around the cylinder, each with a different winding number w [1]. An example for a connected path which winds three times 3 around the D-dimensional cylinder contributes to the partition function a factor [using Formula (2.69)]: to!.Ztl3

=

1 3D jdDX(lldDX(2ldDX(3l exp .J21fn2(J/M

x exp

{-.!. M [X(3LX(2lt} n 2

l_X(1)t} {_.!.M[X(1)-X(3f}_ {_.!.M[X(2 n 2 n(J exp n 2 n(J -

n(J

VD D' (7.10) .J21fn23fJ/M

For cycles of length w the contribution is to!.zt lw

=

Zo(w(J) ,

(7.11)

where Zo(w(J) is the partition function of a free particle in a D-dimensional volume VD for an imaginary-time interval wn(J: (7.12)

595

7.1 Ensembles of Bose and Fermi Particle Orbits

In terms of the thermal de Broglie length leChj3) == V27rn 2 j3 / M associated with the temperature T = 1/kBj3 [recall (2.351)]' this can be written as

Zo(wj3)

=

Vv If (wnj3) .

(7.13)

There is an additional factor 1/w in Eq. (7.11) , since the number of connected windings of the total w! closed paths is (w - 1)!' In group theoretic language, it is the number of cycles of length w, usually denoted by (1,2 , 3, .. . , w), plus the (w -1)! permutations of the numbers 2, 3, . .. , w. They are illustrated in Fig. 7.3 for w = 2, 3, 4. In a decomposition of all N! permutations as products of cycles, the number of elements consisting of C1, C2 , C3 , ... cycles oflength 1, 2, 3, .. . contains (7.14) elements [2] . (12)

~= ~ !~2I (23)

(12 )

(31)

(123 )

(132 )

~= ~3 ~X~ ~X~3 ~~~ T~TiT~'f 333 3 3 3 ;.i! ........il.il3. ........3.; (12 ) 11 lxl 2 22 2 3 33 344 44 (132) (123 )

1;z:1

(13 )

(14 )

21~12 3 3 4-4 (134 )

1~(

l~f

(23 )

1-

1

(2 4) 11

2 2 2X2 3 3 3 3 4 4 4- 4 (14 3) (1 24)

32~23 4 4 (142 )

1~1 21~12

1 ~{

12 ~12 2 2 2 2 2 2 3 3 3 3 3 3 3 3 4 4- 4 4- 4 4 4 4 (1342) (13241 (1234 ) (124 3) '1 1

2 2 3 3 3 3 4 4 4 4 (1423) (14321

(34)

1-1 22 3X 3 4 4 (234)

1-1 32~23 4

4

(24 3) 1-1 2 2

3;Z:34 4

(12 )(34) lX l 2 2 3X 3 4 4

(13 )(24 ) (1 4)( 23) 12 3 4

~12 1)(1 2 2 3 3 4 4

3 4

I ~I! If~I! 2 ~2 I :~I! II§I! II~I! 3

3

.4

4

Figure 7.3 Among the w! permutations of the different windings around the cylinder, (w -I)! are connected. They are marked by dotted frames. In the cycle notation for permutation group elements, these are (12) for two elements, (123) , (132) for three elements, (1234) , (1243), (1324) , (1342), (1423), (1432) for four elements. The cycles are shown on top of each graph, with trivial cycles of unit length omitted. The graphs are ordered according to a decreasing number of cycles.

596

7 Many Particle Orbits - Statistics and Second Quantization

With the knowledge of these combinatorial factors we can immediately write down the canonical partition function (7.9) of N bosons or fermions as the sum of all orbits around the cylinder, decomposed into cycles: N

1 'L..Jtp(v)M(Cl, " Zo(N) (13) = N' . .. , C N ) • plY)

II

(7.15)

w=l N=~wwCw

The sum can be reordered as follows: N

Z(N)_~ o - N!

M(Cl, .. . , CN)tw ,C" .. . ,cn

II

[Zo(wJ3)tw

.

(7.16)

w=l

C" "" CN

N=~wwCw

The parity tw ,c"""cn of permutations is equal to (±1)~w (W+l)Cw . Inserting (7.14), the sum (7.16) can further be regrouped to

L

Cl,,,, ,CN N=~wwCw

(7.17)

For N = 0, this formula yields the trivial partition function Z~O) (13) = 1 of the noparticle state, the vacuum. For N = 1, i.e., a single particle, we find Z~l)(J3) = Zo(J3). The higher zt) can be written down most efficiently if we introduce a characteristic temperature T(O) C

[ ] 2/D N kBM VD((D /2) ,

= 27r1i2 -

(7.18)

and measure the temperature T in units of T~O), defining a reduced temperature t == T/T~O). Then we can rewrite Z~l)(J3) as t D/ 2VD. Introducing further the Ndependent variable N ]2/D [ (7.19) TN == ((D/2) t, we find Z~l) = Tf/2. A few low-N examples are for bosons and fermions: Z o(2)

±2-1-D/2 T2D/2 + T2D,

Z~3)

±3- 1- D/ 2Tf/ 2 + 2- 1- D/ 2Tf ± T12-1TtD/2, ±T2- DTf/ 2 + (T 3 - D + 3- 1- D/ 2) Tf±T2- ~T1D/2

Z~4)

(7.20)

+ 3- 1T 3Ti D.

From zt)(J3) we calculate the specific heat [recall (2.600)] of the free canonical ensemble: (N) _ d2 [ ~ [ (N)] (N)] _ (7.21 ) CO - T dT2 Tlog Zo - TN dTj, TN log Zo ,

597

7.1 Ensembles of Bose and Fermi Particle Orbits

and plot it [3] in Fig. 7.4 against t for increasing particle number N. In the limit ---> 00, the curves approach a limiting form with a phase transition at T = TJO) , which will be derived from a grand-canonical ensemble in Eqs. (7.67) and (7.70). The partition functions can most easily be calculated with the help of a recursion relation [4, 5], starting from Z~O) == 1:

N

zt)({)) =

~ fJ±1)n-1 Zo(n{))zt- n)({)) .

(7.22)

n=1

2 1.75 1.5

C~N)/NkB

1.25

1 0.75 0.5 '/

0.25

'/

Figure 7.4 Plot of the specific heat of free Bose gas with N = 10, 20, 50, 100, 500, 00 particles. The curve approaches for large T the Dulong-Petit limit 3kBN12 corresponding to the three harmonic kinetic degrees of freedom in the classical Hamiltonian p2/2M. There are no harmonic potential degrees of freedom.

This relation is proved with the help of the grand-canonical partition function which is obtained by forming the sum over all canonical partition functions zt) ({)) with a weight factor zN: 00

Zao({)) ==

The parameter z

L

zt)({) ) zN . (7.23) N=O is the Boltzmann factor of one particle with the chemical potential

f.1:

z = z ({)) == e(3Jl.

(7.24)

It is called the fugacity of the ensemble. Inserting the cycle decompositions (7.17),

the sum becomes

L C 1 , ... , C N

N=EwwCw

IT ~ [(±1)W_1 Z0 (W{))ew(3Jl]Cw . -1 w-

CWo

W

The right-hand side may be rearranged to

Zao({)) =

j}1C~O C~! [(±1)W_1Z0(w~eW(3Jlrw

(7.25)

598

7 Many Particle Orbits - Statistics and Second Quantization

(7.26) From this we read off the grand-canonical free energy [recall (1.542)] of noninteracting identical particles (7.27) This is simply the sum of the contributions (7.11) of connected paths to the canonical partition function which wind w = I, 2, 3, ... times around the cylinder [1 , 6] . Thus we encounter the same situation as observed before in Section 3.20: the free energy of any quantum-mechanical system can be obtained from the perturbation expansion of the partition function by keeping only the connected diagrams. The canonical partition function is obviously obtained from (7.27) by forming the derivative: (7.28) It is now easy to derive the recursion relation (7.22) . From the explicit form (7.27) , we see that

~Zao = - (~(3Fa) 8z 8z

(7.29)

Zao .

Applying to this N - 1 more derivatives yields

8 N- 1 [8 ] N- 1 (N _ I)! ( 81+1 ) 8 N- I- 1 8ZN- 1 8z Zao = - ~ l!(N -l-l)! 8z1+1(3Fa 8ZN- I- 1Zao . To obtain from this Ztl we must divide this equation by N! and evaluate the derivatives at z = O. From (7.27) we see that the l + 1st derivative of the grandcanonical free energy is

81+1 I 8 z 1+ 1(3Fa z=o = _(±l)ll! Zo((l + 1)(3).

(7.30)

Thus we obtain 1 8N I 1 N-1 N! 8z NZao z=o = N ~ (±1)1 Zo((l

1

+ 1)(3) (N -l-l)!

8 N- I- 1 I 8ZN- I- 1Zao z=o

Inserting here (7.28) and replacing l --+ n-1 we obtain directly the recursion relation (7.22). The grand-canonical free energy (7.27) may be simplified by using the property 1

Zo(w(3 ) = Zo((3) W D / 2

(7.31)

599

7.2 Bose-Einstein Condensation

of the free-particle partition function (7.12), to remove a factor 1/ ,juP from Zo(w(3). This brings (7.27) to the form 1

Fa

ew {31'

00

= - (3-Zo((3) L(±I)W-l w=l

D/2+1'

(7.32)

W

The average number of particles is found from the derivative with respect to the chemical potential2

fJ

N

00

= -7lFa = Zo((3) up,

ew {31'

L (±I)W- l W D/2' w=l

(7.33)

The sums over w converge certainly for negative or vanishing chemical potentialp" i.e. , for fugacities smaller than unity. In Section 7.3 we shall see that for fermions , the convergence extends also to positive p,. If the particles have a nonzero spin 8, the above expressions carry a multiplicity factor gs = 28 + 1, which has the value 2 for electrons. The grand-canonical free energy (7.32) will now be studied in detail thereby revealing the interesting properties of many-boson and many-fermion orbits, the ability of the former to undergo Bose-Einstein condensation, and of the latter to form a Fermi sphere in momentum space.

7.2

Bose-Einstein Condensation

We shall now discuss the most interesting phenomenon observable in systems containing a large number of bosons, the Bose-Einstein condensation process.

7.2.1

Free Bose Gas

For bosons, the above thermodynamic functions (7.32) and (7.33) contain the functions 00

(,,(z) ==

ZW

L-' w=l wI'

(7.34)

These start out for small z like z, and increase for z -+ 1 to ((v) , where ((z) is Riemann's zeta function (2.519). The functions (,,(z) are called Polylogarithmic functions in the mathematical literature [7] , where they are denoted by Li"(z). They are related to the Hurwitz zeta function ((v, a, z) == 'L'::=ozw/(w + a)" as (,,(z) = z((v, 1, z). The functions ¢(z , v, a) = ((v, a, z) are also known as Lerch functions . In terms of the functions (.,(z) , and the explicit form (7.12) of Zo((3), we may write Fa and N of Eqs. (7.32) and (7.33) simply as 2In grand-canonical ensembles, one always deals with the average particle number (N) for which one writes N in all thermodynamic equations [recall (1.548)]. This should be no lead to confusion.

7 Many Particle Orbits - Statistics and Second Quantization

600

Fa

(7.35)

N

(7.36)

The most interesting range where we want to know the functions (,..,(z) is for negative small chemical potential J.L . There the convergence is very slow and it is useful to find a faster-convergent representation. As in Subsection 2.15.6 we rewrite the sum over w for z = ef31-' as an integral plus a difference between sum and integral

(7.37) The integral yields f(l- v)( -f3J.LY-l, and the remainder may be expanded sloppily in powers of J.L to yield the Robinson expansion (2.579):

There exists a useful integral representation for the functions (v(z):

(7.39) where iv(a) denotes the integral v-I

00

iv(a)

= io(

dE

E

ee-a - 1

,

(7.40)

containing the Bose distribution function (3.93):

(7.41) Indeed, by expanding the denominator in the integrand in a power series

(7.42) and performing the integrals over E, we obtain directly the series (7.34). It is instructive to express the grand-canonical free energy Fa in terms of the functions i,..,(a). Combining Eqs. (7.35) with (7.39) and (7.40), we obtain

F. =-~Z ( )i D/ 2+l(f3J.L) =_ 1 VD {oodE ED/2 a f3 0 f3 f(D/2+1) f3r(D/2+1) V27rn2f3/MD io ee- f31-' _ l'

(

) 7.43

601

7.2 Bose-Einstein Condensation

The integral can be brought to another form by partial integration, using the fact that 1 0 --::--= -log(1 -

OE

e,,-,6JL - 1

e-,,+,6JL).

(7.44)

The boundary terms vanish, and we find immediately: (7.45)

This expression is obviously equal to the sum over momentum states of oscillators with energy fiwp == p2/2M, evaluated in the thermodynamic limit N --+ 00 with fixed particle density N /V, where the momentum states become continuous: (7.46) This is easily verified if we rewrite the sum with the help of formula (1.555) for the surface of a unit sphere in D dimensions and a change of variables to the reduced particle energy E = fJp 2/2M as an integral

"~

--+

VD

J

dDp 1 (27rfi)D - VDSD (27rfi)D

1

VD

f(D/2) V27rfi2fJ/MD

roo d

Jo

J

dpp

D- l (2M/fJ)D/21 - VDSD (27rfi)D 2

dEE

D/2- 1

(7.47)

D/2-1

EE

J

.

Another way of expressing this limit is

L p

--+

roo dE N",

Jo

(7.48)

where N" is the reduced density of states per unit energy interval: (7.49)

The free energy of each oscillator (7.46) differs from the usual harmonic oscillator expression (2.483) by a missing ground-state energy fiw p /2. The origin of this difference will be explained in Sections 7.7 and 7.14. The particle number corresponding to the integral representations (7.45) and (7.46) is

602

7 Many Particle Orbits - Statistics and Second Quantization

1 .__.__.__._.__.__.__.__._-_.__.__.

0.8 0.6 0.4 0.2 0.2

0.4

Figure 7.5 Plot of functions (y(z) for thermodynamics.

0.6

l/

0.8

1

z

= 3/2 and 5/2 appearing in Bose-Einstein

For a given particle number N, Eq. (7.36) allows us to calculate the fugacity as a function of the inverse temperature, z(f3) , and from this the chemical potential J.L(f3) = f3- 1 log z(f3) . This is most simply done by solving Eq. (7.36) for f3 as a function of z, and inverting the resulting function f3(z) . The required functions (y(z) are shown in Fig. 7.5. There exists a solution z(f3) only if the total particle number N is smaller than the characteristic function defined by the right-hand side of (7.36) at unit fugacity z(f3) = 1, or zero chemical potential J.L = 0:

VD

N < W(nf3) ((D/2).

(7.51)

Since le(nf3) decreases with increasing temperature, this condition certainly holds at sufficiently high T. For decreasing temperature, the solution exists only as long as the temperature is higher than the critical temperature Tc = l/kBf3c , determined by (7.52) This determines the critical density of the atoms. The de Broglie length at the critical temperature will appear so frequently that we shall abbreviate it by Pc: (7.53) The critical density is reached at the characteristic temperature T~O) introduced in Eq. (7.18). Note that for a two-dimensional system, Eq. (7.18) yields T~O) = 0, due to ((1) = 00, implying the nonexistence of a condensate. One can observe, however , definite experimental signals for the vicinity of a transition. In fact , we have neglected so far the interaction between the atoms, which is usually repulsive. This will give rise to a special type of phase transition called Kosterlitz-Thouless transition. For a discussion of this transition see other textbooks [8].

603

7.2 Bose-Einstein Condensation

By combining (7.18) with (7.36) we obtain an equation for the temperature dependence of z above Te: 1

= (T)D/2 (D/2(Z(T)) Te

((D/2) '

T > Te·

(7.54)

This is solved most easily by calculating T /Te as a function of z = ef3/L. If the temperature drops below Te , the system can no longer accommodate all particles N in a normal state. A certain fraction of them, say Neond(T), is forced to condense in the ground state of zero momentum, forming the so-called Bose-Einstein condensate. The condensate acts like a particle reservoir with a chemical potential zero. Both phases can be described by the single equation for the number of normal particles, i.e., those outside the condensate: (7.55) For T > Te , all particles are normal and the relation between fJ and the temperature is found from the equation Nn(T) = N, where (7.55) reduces to (7.36). For T < Te, however, the chemical potential vanishes so that z = 1 and (7.55) reduces to (7.56) which yields the temperature dependence of the number of normal particles: Nn(T) N

= (T)D/2 Te

, T < Te·

(7.57)

The density of particles in the condensate is therefore given by Neond(T) N

= 1- Nn(T) = 1- (T)D/2 Te

N

(7.58)

We now calculate the internal energy which is, according to the general thermodynamic relation (1.550), given by E

= Fa + TS + fJN = Fa - TfJrFa + fJN = 8f3(f3Fa) + fJN.

(7.59)

Expressing N as -8Fa/8fJ, we can also write (7.60) Inserting (7.35) we see that only the f3-derivative of the prefactor contributes since (f38f3 - fJ8/L) applied to any function of z = ef3/L vanishes. Thus we obtain directly D

E = --F.a

2

'

(7.61)

604

7 Many Particle Orbits - Statistics and Second Quantization

which becomes with (7.35) and (7.36):

D E = - 2{3Zo({3KD/2+l(Z) =

D (D/2+l(Z) (D/2(Z) NkBT.

(7.62)

-"2

The entropy is found using the thermodynamic relation (1.571): 1 (E - JlN - FG) = T 1 --2-FG ( D +- 2 S =T JlN) ,

(7.63)

or, more explicitly,

S = k [D + 2 Z ({3)1" ( )_ B 2 0 ,>D/2+1 Z

{3Jl N]

{3] Jl.

= k BN [D 2+ 2 (D/2+l(Z) (D/2(Z)

For T < Te , the entropy is given by (7.64) with Jl the number N n of normal particles of Eq. (7.57):

(7.64)

= 0, Z = 1 and N replaced by

(7.65) The particles in the condensate do not contribute since they are in a unique state. They do not contribute to E and FG either since they have zero energy and Jl = O. Similarly we find from (7.62):

E = D Nk T (T)D/2 (D/2+l(1) = ~TS < 2 B Te (D/2(1) D+2 T e , the chemical potential at fixed N satisfies the equation (7.68) This follows directly from the vanishing derivative (3o{3N = 0 implied by the fixed particle number N. Applying the derivative to Eq. (7.36) and using the relation zOz(v(z) = (v-l(Z), as well as (3o{3f(z) = zOzf(z) (30{3({3Jl) , we obtain

(7.69) thus proving (7.68).

605

7.2 Bose-Einstein Condensation

The specific heat C at a constant volume in units of kB is found from the derivative C = T8rSIN = -/32o(3EIN' using once more (7.68):

C = kBN [(D

+ 2)D (D/2+1 (Z) 4

(D/2(Z)

T> Te.

_ D2 (D/2(Z)] 4 (D/2-1(Z) ,

(7.70)

At high temperatures, C tends to the Dulong-Petit limit DkBN/2 since for small Z all (v(z) behave like z. Consider now the physical case D = 3, where the second denominator in (7.70) contains (1/2(Z). As the temperature approaches the critical point from above, Z tends to unity from below and (1/2(Z) diverges. Thus 1/(1/2(1) = 0 and the second term in (7.70) disappears, yielding a maximal value in three dimensions 15 (5/2(1) 4 1. Inserting (7.142) back into (7.162) , we may re-express the normal fraction as a function of the temperature as follows : (7.143) and the condensate fraction as (7.144) Including the next term in (7.162), the condensate fraction becomes (7.145) It is interesting to re-express the critical temperature (7.142) in terms of the particle density at the origin which is at the critical point, according to Eq. (7.134),

(7.146) where we have shortened the notation on an obvious way. From this we obtain

k

T(O) = B

e

27r1i 2

M

[

no(O) ] 2/D ((D/2)

(7.147)

Comparing this with the critical temperature without a trap in Eq. (7.18) we see that both expressions agree if we replace N /V by the uniform density no (0). As an

619

7.2 Bose-Einstein Condensation

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.2

0. 4

0.6

00

0.2 0.4

T ITc

0.6

0.8

1

1.2

1.4

1.6

1.8

T I Tc

Figure 7.9 Condensate fraction Ncond/N == 1 - Nn/N as function of temperature for total number of N particles. The long- and short-dashed curves on the left-hand show the zeroth and first-order approximations (7.144) and (7.163). The dotted curve displays the free-boson behavior (7.58). The right-hand figure shows experimental data for 40000 87Rb atoms near Tc by Ensher et. al., Phys. Rev. Lett. 77, 4984 (1996). The solid curve describes noninteracting Bose gas in a harmonic trap [ef. (7.144)]. The dotted curve corrects for finite-N effects [cf. (7.163)]. The dashed curve is the best fit to the data. The transition lies at Tc ,:::; 280 nK, about 3% below the dotted curve.

obvious generalization, we conclude that Bose condensation in any trap will set in when the density at the lowest point reaches the critical value determined by (7.147) [and (7.18)]. Note the different power D and argument in ((D) in Eq. (7.142) in comparison with the free-boson argument D/2 in Eq. (7.18). This has the consequence that in contrast to the free Bose gas, the gas in a trap can form a condensate in two dimensions. There is now, however, a problem in D = 1 dimension where (7.142) gives a vanishing transition temperature. The leading-order expression (7.142) for critical temperature can also be calculated from a simple statistical consideration. For small w , the density of states available to t he bosons is given by the classical expression (4.216):

- (-)

pcl(E) -

M 27rn?

D/2

1

f(D/2)

J

dD x [E

-

V(x)] D / 2-1 .

(7.148)

The number of normal particles is given by the equation Nn

=

roo

JEmin

dE

pcl(E)

eE / kBT

-

1'

(7.149)

where E min is the classical ground state energy. For a harmonic trap, the spatial integral in (7.148) can be done and is proportional to ED , after which the integral over E in (7.149) yields [kBT/nw ]D((D) = (T/TJol)DN , in agreement with (7.143).

620

7 Many Particle Orbits - Statistics and Second Quantization

Alternatively we may use the phase space formula dDp 1 _ ~ JdD dDp - n,B[p2/2M+V(xl ] X (27rn)D e,B[P2/2M+V(xl] _ 1 - ~ X (27rn)D e

N - JdD n -

=

f

1 D J dDxe-n,BV(x l , 2 n=1 V27rn nf3 / M

where the spatial integration produces a factor side becomes again (T /TJol)D N.

7.2.7

(7 .150)

V27r /M w nf3 so that the right-hand 2

More General Anisotropic Trap

The equation (7.150) for the particle number can be easily calculated for a more general trap where the potential has the anisotropic power behavior V(x) = M (;?a?

f (EJ)Pi ,

i=1 ai . f d _ . . average ah W IS some requency parameter an a IS t he geometnc were 2

(7.151)

== [rrDi=1 ai ]1/D.

Inserting (7.151) into (7.150) we encounter a product of integrals (7.152) so that the right-hand side of (7.150) becomes (T /TJol/j N , with the critical temperature (7.153) where jj is the dimensionless parameter _ D D 1 D== -+ 2 i=1 Pi

L-'

(7.154)

A harmonic which takes over the role of D in the harmonic formula (7.142). trap with different oscillator frequencies WI, . . . , WD along the D Cartesian axes, is a special case of (7.151) with Pi == 2, = (;;2a,2ja; and jj = D, and formula (7.153) reduces to (7.142) with W replaced by the geometric average of the frequencies (;; == (WI '" WD)I/D. The parameter ji disappears from the formula. A free Bose gas in a box of size VD = np:'I(2ai) = 2DjiD is described by (7.151) in the limit Pi ---> 00 where jj = D/2. Then Eq. (7.153) reduces to

w;

7rn2 [~] 2/D _ 27rn2 [ N ] 2/D - 2Mji2 ((D/2) - M VD((D/2) ,

(Ol _

kBTc

(7.155)

621

7.2 Bose-Einstein Condensation

in agreement with (7.53) and (7.18). Another interesting limiting case is that of a box of length L = 2al in the xdirection with PI = 00, and two different oscillators of frequency W2 and W3 in the other two directions. To find T~O) for such a Bose gas we identify c; 2 0,2 / a~ 3 = W~ 3 in the potential (7.151), so that C;4/ii 2 = w~wUai, and obtain " k B

T(O) c

7.2.8

=

nC; ( ~ )

1/5 [

2MC;

] 2/5 = n - ( 27r AWl AW2 ) N al((5/2) w L2

1/5 [ ~ ] 2/5

((5/2)

(7.156)

Rotating Bose-Einstein Gas

Another interesting potential can be prepared in the laboratory by rotating a Bose condensate [48] with an angular velocity 0 around the z-axis. The vertical trapping frequencies is W z ~ 27r x 11.0Hz ~ 0.58x nK , the horizontal one is W.l ~ 6 x W z . The centrifugal forces create an additional repulsive harmonic potential, bringing the rotating potential to the form V(x)

M w;

(

2

2

K

ri )

nwz ( :\2 Z2 rl K ri ) + 361]:\2 + 2'-:\4 '

= -2- z + 361]r .l + 2':\2 = 2 Wz

Wz

Wz

(7.157)

Wz

where rl = x 2 + y2, 1] == 1 - 0 2 / wi, K ~ 0.4, and Awz == 3.24511,ID ~ 1.42 x 10- 3 K. For 0 > W.l, 1] turns negative and the potential takes the form of a Mexican hat as shown in Fig. 3.13, with a circular minimum at r~ = -361]A~j K. For large rotation speed, the potential may be approximated by a circular harmonic well, so that we may apply formula (7.156) with al = 27rrm , to obtain the 1]-independent critical temperature k

T(O) B

c

~ n (K)I/5 [ N ]2/5 wz;;: ((5 / 2)

~

(7.158)

K = 0.4 and N = 300000, this yields Tc ~ 53nK. At the critical rotation speed 0 = W.l, the potential is purely quartic r.l /(X2+y2). To estimate TiO) we approximate it for a moment by the slightly different potential (7.151) with the powers PI = 2, P2 = 4, P3 = 4, al = A wz , a2 = a3 = Awz (K/2)1 /4, so that formula (7.153) becomes

For

(0) _

kBTc

7r 2 K

-nwz [ 16f4 (5/4)

] 1/5 [

N ] 2/5 ((5/2)

(7.159)

It is easy to change this result so that it holds for the potential ex r 4 = (x + y)4 rather than X4 + y4: we multiply the right-hand side of equation (7.149) for N by a factor 27r J rdrdxdy e- r4 7r 3/ 2 (7.160) xcy J dxdy e- ' r[5 /4]2 .

622

7 Many Particle Orbits - Statistics and Second Quantization

This factor arrives inversely in front of N in Eq. (7.161), so that we obtain the critical temperature in the critically rotating Bose gas 4 r;,

(0) _

kBTc

- nwz

1/5 [

( -:; )

N ] 2/5 ((5/2)

(7.161)

The critical temperature at 0 = W.L is therefore by a factor 41 / 5 ;:::; 1.32 larger than at infinite O. Actually, this limit is somewhat academic in a semiclassical approximation since the quantum nature of the oscillator should be accounted for.

7.2.9

Finite-Size Corrections

Experiments never take place in the thermodynamic limit. The particle number is finite and for comparison with the data we must calculate finite-size corrections coming from finite N where 11 w ;:::; 1/N1/D is small but nonzero. The transition is no longer sharp and the definition of the critical temperature is not precise. As in the thermodynamic limit, we shall identify it by the place where ZD = 1 in Eq. (7.139) for N n = N . For D > 3, the corrections are obtained by expanding the first term in the sum (7.138) in powers of wand performing the sums over wand subtracting ((0) for the sum E~=11: 1 [ Nn(Tc) = ((3nw)D ((D)

2

] (3nw D (wn(3) + -2((D-1) + ----u- D (3D-1)((D-2) + ...

((0). (7.162) The higher expansion terms contain logarithmically divergent expressions, for instance in one dimension the first term ((1), and in three dimensions ((D-2) = ((1) . These indicate that the expansion powers of (3nw is has been done improperly at a singular point. Only the terms whose (-function have a positive argument can be trusted. A careful discussion along the lines of Subsection 2.15.6 reveals that ((1) must be replaced by (reg(l) = -log((3nw) + const., similar to the replacement (2.588). The expansion is derived in Appendix 7A. For D > 1, the expansion (7.162) can be used up to the ((D - 1) terms and yields the finite-size correction to the number of normal particles Nn

N Setting N n

=

(T)D TJO)

D ((D - 1)

1

+ "2 (l - l/D(D) N1/D ·

(7.163)

= N, we obtain a shifted critical temperature by a relative amount bTc TJO) =

1 ((D - 1)

1

-"2 ((D- 1)/D(D) N1/D + ....

(7.164)

In three dimensions, the first correction shifts the critical temperature TJO) downwards by 2% for 40000 atoms. Note that correction (7.164) has no direct w-dependence whose size enters only implicitly via TJO) of Eq. (7.142). In an anisotropic harmonic trap, the temperature

623

7.2 Bose-Einstein Condensation

shift would carry a dimensionless factor w/w where w is the arithmetic mean w == (Lp:'l Wi) / D. The higher finite-size corrections for smaller particle numbers are all calculated in Appendix 7A. The result can quite simply be deduced by recalling that according to the Robinson expansion (2.581), the first term in the naive, wrong power series expansion of (v(e - b ) = L~=l e- wb /w v = Lk::o( -b)k((v - k)/k! is corrected by changing the leading term ((v) to r(1 - v)( _b)v- l + ((v) which remains finite for all positive integer v. Hence we may expect that the correct equation for the critical temperature is obtained by performing this change in Eq. (7.162) on each ((v). This expectation is confirmed in Appendix 7A. It yields for D = 3,2,1 the equations for the number of particles in the excited states 2

+]

=

Nn

=

(,8liw)2 ((2) -,8liw log,8nw - 'Y + 2

Nn

=

[ ,8liw 1 (,8liw)3 ((3)3 + -2-((2) - (,8liw) 2 ( log,8nw - 'Y + 1 24 9 +)...] , (7.167)

_ 1_

,8nw

[_ (log,8nw _ 'Y) _ ,8nw ((0) 2

+(,8liw)

Nn

1 [

(

12

(( -1)

... ,

1) - 7(,8nw? 12 ((0) + ... ] ,

(7.165) (7.166)

where'Y = 0.5772 . .. is the Euler-Mascheroni number (2.467). Note that all nonlogarithmic expansion terms coincide with those of the naive expansion (7.162). These equations may be solved for ,8 at N n = N to obtain the critical temperature of the would-be phase transition. Once we study the position of would-be transitions at finite size, it makes sense to include also the case D = 1 where the thermodynamic limit has no transition at all. There is a strong increase of number of particles in the ground state at a "critical temperature" determined by equating Eq. (7.165) with the total particle number N , which yields (0) _

1

~

1

kBTc - nwN (- l,8n N' og w + 'Y ) ~ liwN og1

D=l.

(7.168)

Note that this result can also be found also from the divergent naive expansion (7.162) by inserting for the divergent quantity ((1) the dimensionally regularized expression (reg (l) = -log(,8nw) of Eq. (2.588).

7.2.10

Entropy and Specific Heat

By comparing (7.126) with (7.125) we see that the grand-canonical free energy can be obtained from N n of Eq. (7.162) by a simple multiplication with -1/,8 and an increase of the arguments of the zeta-functions ((v) by one unit. Hence we have, up to first order corrections in w

Fa (,8,f.tc) = -

,8(,8~w)D

[ ((D + 1) + ,8liw~ ((D) + ...J .

(7.169)

624

7 Many Particle Orbits - Statistics and Second Quantization

From this we calculate immediately the entropy S = -OTFC = k B{320(3Fc as

S

=

-kB(D + l){3Fc

=

kB(D + 1) ({3~)D [ ((D + 1) + {3nw ~ ((D) + ... J . (7.170)

In terms of the lowest-order critical temperature (7.142), this becomes

T)D((D+1) D[((D)]l/D(T)D- l } S = kBN(D + 1) {( Ti O) ((D) + 2 ---y:{ Ti O) + ... . (7.171) From this we obtain the specific heat C = TOTS below Te:

T )D ((D+1) D(D-1) [((D)] l/D( T )D- l } C=kBN(D+1) { D ( Ti O) ((D) + 2 ---y:{ Ti O) + .... (7.172) At the critical temperature, this has the maximal value

((D+1) { [D-1 e+1/D(D) D ((D-1)] 1 } Cmax';::;jk BN(D+1)D ((D) 1+ - 2- ((D+1) -2 (l- l/D(D) Nl/D . (7.173) In three dimensions, the lowest two approximations have their maximum at (7.174) Above Te, we expand the total particle number (7.126) in powers of w as in (7.162). The fugacity of the ground state is now different from unity: (7.175) The grand-canonical free energy is (7.176) and the entropy

The specific heat C is found from the derivative -(30(3SIN as

C({3,f.1)

=kB({3n~)D +~

[D (D2

{(D+1)D(D+l(ZD)

+ 1) {3nw- D{3f.1- (30(3({3f.1)] (D(ZD)+ . . . }.

(7.178)

625

7.2 Bose-Einstein Condensation

The derivative f3o(3(f3fJ) is found as before from the condition: oN(f3,fJ(f3))/of3 = 0, implying

o=

( f3 :W)D

{- [D(D(ZD)

+ f317,w ~ (D - l)(D-l(ZD)]

+ [(D-l(ZD) + f31iw~ (D-2(ZD)] [f30(3(f3fJ)-f317,W~] + .. .}, so that we obtain

(7.179)

D

(D(ZD) + f317,W 2 (D- 2(ZD) + ... f30(3(f3fJ) = D D (D-l(ZD) + f317,2W(D-2(ZD) + .. .

(7.180)

(7.181)

so that from (7.60) (7.184) The chemical potential at fixed N satisfies the equation (7.185) The specific heat at a constant volume is (7.186) At high temperatures, C tends to the Dulong-Petit limit DNkB since for small

Z

all

(,,(z) behave like z. This is twice a big as the Dulong-Petit limit of the free Bose gas since there are twice as many harmonic modes. As the temperature approaches the critical point from above, from below and we obtain a maximal value in three dimensions

Z

tends to unity

C(O) = k N [12(4(1) - 9(3(1)] ::;:, k N 422785 max

B

(3(1)

(2(1)



.

(7.187)

The specific heat for a fixed large number N of particles in a trap has a much sharper peak than for the free Bose gas. The two curves are compared in Fig. 7.10, where we also show how the peak is rounded for different finite numbers N.3 3p.W. Courteille, V.S. Bagnato, and V.l. Yukalov, Laser Physics 2, 659 (2001).

626

7 Many Particle Orbits - Statistics and Second Quantization 10

10

\

1\\ N

harmonic trap

/ \t-N_=_1000 _

4

/f

'":;..

t

N= 100

/' ,...:,..,.,.,..,...1"''''

free Bose gas 0.2

0.4

0.6

0.8

I

1.2

= 10000

1.4

0.25

0.50

0.75

T /TJO)

1.00

1.25

1.50

T /TJO)

Figure 7.10 Peak of specific heat for infinite (left-hand plot) and various finite numbers 100, 1000, 10000 of particles N (right-hand plots) in harmonic trap. The large-N curve is compared with that of a free Bose gas.

7.2.11

Interactions in Harmonic Trap

Let us now study the effect of interactions on a Bose gas in an isotropic harmonic trap. This is most easily done by adding to the free part (7.132) the interaction (7.101) with n(x) taken from (7.131), to express the grand-canonical free energy by analogy with (7.102) as

Fa =

JdDx { -*Zw(/3) fo

ZD

d: (D({3nw; z; x)

+g

[Zw ({3)(D ({3nw; ZD; X)]2} . (7.188)

Using the relation (7.103), this takes a form more similar to (7.106):

Fa =

J {

dz -j31 Zw({3) dDx JorzD -;(D({3nw;z;x)-2fLl~(n{3)Zw({3) [(D({3nw;zD;X)]

2} .

(7.189) As in Eq. (7.113), we now construct the variational free energy to be extremized with respect to the local parameter a(x). Moreover, we shall find it convenient to express (( z) as foz (dz' / z')( (z') . This leads to the variational expression (7.190) where

ZD

=

e-(DIiw/2-J1.l/3

a ==

= e- D /3Iiw/2 z and

al~ (n{3) Zw ({3) = le(~{3) l~ (n{3)Zw({3).

The extremum lies at a(x) = ~(x)

~(x)

(7.191)

where by analogy with (7.114):

== -4a(D({3nw; zDe~(xl; x) .

(7.192)

627

7.2 Bose-Einstein Condensation

For a small trap frequency w, we use the function (v (f3nw ; Zv; x) in the approximate form (7.133) , written as

(df3nw; zv; x)

=L frf!-00

w 2 Mf3

w=l

21TW

v z'De- Mw {3w

2 2

x

/2.

(7.193)

In this approximation, Eq. (7.192) becomes

~(

u X ) wzo ~ -

4-a

~ ~

J

w=l

w2 Mf3 vzve w wI;(x) e -Mw{3w 2 x 2 /2 . -

(7.194)

21TW

For small ii, ~(x) is also small, so that the factor ewI;(x) on the right-hand side is close to unity and can be omitted. This will be inserted into the equation for the particle number above Tc: (7.195) Recall that in the thermodynamic limit for D > 1 where the phase transition properly exists, (v(f31iw; Zv; x) and c'v(f3nw; Zv; x) coincide, due to (7.139) and (7.141). From (7.195) we may derive the following equation for the critical temperature as a function of zv: 1

=

Zw(f3)

JdV

Zw(f3~O))

c'v(f3nw;zv eI;(x) ;x) (v(f3~O)1iw ; 1) ,

x-=--"::----'~----'~

(7.196)

where T~O) is the critical temperature in the trap without the repulsive interaction. The critical temperature Tc of the interacting system is reached if the second argument of c'df31iw; zVeI;(x); x) hits the boundary of the unit convergence radius of the expansion (7.130) for x = 0, i.e., if zVeI;(O) = 1. Thus we find the equation for Tc: 1=

Z w ((.I) fJc

J

I" (f3 1iw. eI;c(x)-I;c(O). x) dV x 0 dE N n~.

(7.222)

g

The Bose function contains a pole at 0: = 0 which prevents the existence of a solution for positive 0: . In the analytically continued fermionic function (7.215), on the other hand, the point 0: = 0 is completely regular. Consider now a Fermi gas close to zero temperature which is called the degenerate limit. Then the reduced variables E = E/kBT and 0: = l-£/kBT become very large and the distribution function (7.217) reduces to

nf

g

=

{I

0

for EO:

= e (E -

(7.223)

0:).

All states with energy E lower than the chemical potential 1-£ are filled, all higher states are empty. The chemical potential 1-£ at zero temperature is called Fermi energy E F : (7.224) 1-£1T=O =EF · The Fermi energy for a given particle number N in a volume VD is found by performing the integral (7.222) at T = 0: 1

D(PF)D r;: ,

gsV

f(D/2+1) J47TD

(7.225)

where (7.226) is the Fermi momentum associated with the Fermi energy. Equation (7.225) is solved for EF by

EF = 27rn? [r(D/2 + 1)]2/D (.!!.-)2/D, M

gs

VD

(7.227)

and for the Fermi momentum by

_

PF -

2V1in [f(D/2+1)]1/D(.!!.-)1/D v:D n. gs

(7.228)

632

7 Many Particle Orbits - Statistics and Second Quantization

Note that in terms of the particle number N, the density of states per unit energy interval and volume can be written as (7.229)

As the gas is heated slightly, the degeneracy in the particle distribution function in Eq. (7.223) softens. The degree to which this happens is governed by the size of the ratio kT / E F • It is useful to define a characteristic temperature, the so-called Fermi temperature (7.230)

For electrons in a metal, PF is of the order of 'Ii/IA. Inserting M = me = 9.109558 X 10- 28 g, further kB = 1.380622 X 10- 16 erg/K and 'Ii = 6.0545919 X 10- 27 erg sec, we see that TF has the order of magnitude TF

~

44000K.

(7.231)

Hence, even far above room temperatures the relation T /TF « 1 is quite well fulfilled and can be used as an expansion parameter in evaluating the thermodynamic properties of the electron gas at nonzero temperature. Let us calculate the finite-T effects in D = 3 dimensions. From Eq. (7.225) we obtain _ gsV 2 .f (f..L) (7.232) N = N(T, f..L) = 1~('lif3) y'1i 23 / 2 kBT . Expressing the particle number in terms of the Fermi energy with the help of Eq. (7.225), we obtain for the temperature dependence of the chemical potential an equation analogous to (7.54): (7.233)

To evaluate this equation, we write the integral representation (7.216) as i~(a)

=

1

00

- a

1 a

o

where E = x and obtain

+ a.

(a

+ x)n- 1

dx-'-----'--c-e X +l

dx

(a-x)n-1 e- X + 1

+

1

00

0

(a+x)n-1

dx-'-----'--c-eX

+1

In the first integral we substitute 1/(e- X

+ 1)

(7.234) = 1 - 1/(eX

+ 1),

(7.235)

633

7.3 Gas of Free Fermions

In the limit a--->oo, only the first term survives, whereas the last term is exponentially small, so that it can be ignored in a series expansion in powers of l/a. The second term is expanded in such a series:

f

roo dx-f-- = 2 L

(n-1)! a n- 1- k(1 - Tk)((k + 1). (n-1-k)! (7.236) In the last equation we have used the integral formula for Riemann's (-function4 2

k=odd

(n-1) a n- 1- k k Jo

e

+1

k=odd

(7.237)

At even positive and odd negative integer arguments, the zeta function is related to the Bernoulli numbers by Eq. (2.567). The lowest values of ((x) occurring in the expansion (7.236) are ((2) , ((4) , ... , whose values were given in (2.569), so that the expansion of (a) starts out like

i;

i~(a) =

.!.an + 2(n-1).!.((2)an- 2 + 2(n-1)(n-2)(n-3)~((4)an-4 + ... n 2 8

Inserting this into Eq. (7.233) where n 1 = (kBT)3/2 ~ [~ EF 2 3

.

(7.238)

= 3/2, we find the low-temperature expansion

(~)3/2 + 7r2(~)-1/2 + ~ (~)-5/2 ... J ' kBT

12

kBT

3 . 320

kBT

(7.239)

implying for fl the expansion (7.240)

These expansions are asymptotic. They have a zero radius of convergence, diverging for any T. They can, however, be used for calculations if T is sufficiently small or at all T after a variational resummation it la Section 5.18. We now turn to the grand-canonical free energy Fe. In terms of the function (7.216) , this reads 1 gsV 1 .f (7.241) Fe = -73l~(nf3) f(5/2) z5/2(a) . Using again (7.238) , this has the expansion Fc(T, fl , V)

1

57r2 (kBT)2 77r 4 (kBT)4 = Fc(O, fl , V) [1 + 8 ---;;- - 384 ---;;- + . .. ,

where Fc(O ,fl, V) 41.8. Gradshteyn and 1.M. Ryzhik, op. cit. , Formula 3.411.3.

(7.242)

634

7 Many Particle Orbits -

Statistics and Second Quantization

By differentiating Fe with respect to the temperature at fixed /-l, we obtain the low-temperature behavior of the entropy

B

7f2 kBT kB - E - N+ .. . 2 F

=

(7.243)

From this we find a specific heat at constant volume

C = TaB I aT V,N

T~O B = kB 7f2 kBT N + ...

(7.244)

2 EF

This grows linearly with increasing temperature and saturates at the constant value

3kB N /2 which obeys the Dulong-Petit law of Section 2.12 corresponding to three kinetic and no potential harmonic degrees of freedom in the classical Hamiltonian p2/2M. See Fig. 7.11 for the full temperature behavior. The linear behavior is C kEN 3

Dulong-Petit Law

2"

T

Figure 7.11 Temperature behavior of specific heat of free Fermi gas. As in the free Bose gas, the Dulong-Petit rule gives a high-temperature limit 3kBN/2 for the three harmonic kinetic degrees of freedom in the classical Hamiltonian p2/2M. There are no harmonic potential degrees of freedom. due to the progressive softening near the surface of the Fermi distribution which makes more and more electrons thermally excitable. It is det ected experimentally in metals at low temperature where the contribution of lattice vibrations freezes out as (T/TDt Here TD is the Debye temperature which characterizes the elastic stiffness of the crystal and ranges from T D ~ 90K in soft metals like lead over T D ~ 389K for aluminum to T D ~ 1890K for diamond. The experimental size of the slope is usually larger than the free electron gas value in (7.244) . This can be explained mainly by the interactions with the lattice which result in a large effective mass Meff > M . Note that the quantity Fc(O, /-l , V) is temperature dependent via the chemical potential/-l. Inserting (7.240) into (7.242) we find the complete T-dependence 57f2 (kBT)2 7f4 (kBT)4 Fc(T,/-l , V) = Fc(O, EF, V) [ 1 + 12 EF - 16 EF

+ .. .1,

(7.245)

7.4 Statistics Interaction

635

where (7.246) As in the boson gas, we have a relation (7.61) between energy and grandcanonical free energy: 3 E = --F.e (7.247) 2 ' such that equation (7.245) supplies us with the low-temperature behavior of the internal energy: (7.248) The first term is the energy of the zero-temperature Fermi sphere. Using the relation = aE/VaT, the second term yields once more the leading T -+ 0 -behavior (7.244) of specific heat. This behavior of the specific heat can be observed in metals where the conduction electrons behave like a free electron gas. Due to Bloch's theorem , a single electron in a perfect lattice behaves just like a free particle. For many electrons, this is still approximately true, if the mass of the electrons is replaced by an effective mass. Another important macroscopic system where (7.244) can be observed is a liquid consisting of the fermionic isotope 3He. There are two electron spins and an odd number of nucleon spins which make this atom a fermion. Also there the strong interactions in the liquid produce a screening effect which raises to an effective value of the mass to 8 times that of the atom.

Cv

7.4

Statistics Interaction

First, we consider only two identical particles; the generalization to n particles will be obvious. For simplicity, we ignore the one-body potentials V(x(v») in (7.2) since they cause only inessential complications. The total orbital action is then (7.249) The standard change of variables to center-of-mass and relative coordinates (7.250) respectively, separates the action into a free center-of-mass and a relative action

A = ACM+ Arel =

lb dt~ X? + l~b dt [~X2 - Vint(X)] ,

(7.251)

with a total mass M = M(1)+M(2) and a reduced mass f.t = M(l) M(2) /(M(1) +M(2»). Correspondingly, the time evolution amplitude of the two-body system factorizes

7 Many Particle Orbits - Statistics and Second Quantization

636

into that of an ordinary free particle of mass M, (XbtbIXata) , and a relative amplitude (Xbtblxata). The path integral for the center-of-mass motion is solved as in Chapter 2. Only the relative amplitude is influenced by the particle statistics and needs a separate treatment for bosons and fermions. First we work in one dimension only. Many of the formulas arising in this case are the same as those of Section 6.2, where we derived the path integral for a particle moving in a half-space x = r > 0; only the interpretation is different. We take care of the indistinguishability of the particles by restricting x to the positive semiaxis x = r 2 0; the opposite vector -x describes an identical configuration. The completeness relation of local states reads therefore

l)O drlr)(rl = 1.

(7.252)

To write down the orthogonality relation, we must specify the bosonic or fermionic nature of the wave functions. Since these are symmetric or antisymmetric, respectively, we express (rb lra) in terms of the complete set of functions with these symmetry properties:

(rb lra) = 2

roo dp {

Jo

7r'h

c?sprb/h c?spra/'h }.

sm prb/'h sm pra/'h

(7.253)

This may be rewritten as

(rb lra) =

1

00

- 00

dp

27r'h

(eiP(rb - ra)/1i

±

eiP(rb+ ra)/Ii)

= i5(rb - ra) ± i5(rb + ra) . (7 .254)

The infinitesimal time evolution amplitude of relative motion is then, in the canonical formulation,

(rn Elrn- l 0)

=

(rn le- i,k,el/lilrn_l)

(7.255) where H rel (p, x) is the Hamiltonian of relative motion associated with the action Arel in Eq. (7.251). By combining N + 1 factors, we find the time-sliced amplitude (7.256) x { exp

[~ }; Pn(rn -

rn- l)] ± exp

[* };: Pn(rn + rn- l)] } e-K' L,:~11

Hcel(Pn ,rn) ,

valid for bosons and fermions, respectively. By extending the radial integral over the entire space it is possible to remove the term after the ±sign by writing (7.257)

7.4 Statistics Interaction

637

where the function a-(x) vanishes identically for bosons while being equal to

a(x)

=

8(-x)

(7.258)

for fermions, where 8(x) is the Heaviside function (1.309) . As usual, we have identified Xb == XN+ l and Xa == Xo which is equal to ra' The final sum over Xb = ±rb accounts for the indistinguishability of the two orbits. The phase factors ei7r(7(xn) give the necessary minus signs when exchanging two fermion positions. Let us use this formula to calculate explicitly the path integral for a free twoparticle relative amplitude. In the bosonic case with a vanishing a-term, we simply obtain the free-particle amplitude summed over the final positions ±rb:

For fermions, the phases a(xn) in (7.257) cancel each other successively, except for the boundary term (7.260) When summing over Xb = ±rb in (7.257), this causes a sign change of the term with Xb = -rb and leads to the antisymmetric amplitude (7.261) Let us also write down the continuum limit of the time-sliced action (7.257). It reads (7.262) The last term is the desired Fermi statistics interaction. It can also be written as

Ar =

-n;r

l tb dti:(t)J(x(t)) = n;r ltb dt8 8(-x(t)). t

ta

(7.263)

ta

The right-hand expression shows clearly the pure boundary character of A f , which does not change the equations of motion. Such an interaction is called a topological

interaction. Since the integrals in (7.257) over x and p now cover the entire phase space and a(x) enters only at the boundaries of the time axis, it is possible to add to the action any potential \lint (r). As long as the ordinary path integral can be performed, also the path integral with the additional a-terms in (7.257) can be done immediately. It is easy to generalize this result to any number of fermion orbits x(v)(t), 1/ = 1, . . . ,n. The statistics interaction is then I:v .( )"'] 2 00 r J dx-Ixp)(xpl= dr _e-,m-n'Ye- r 211" 211" m,n=O

Jmf JnT

The angular integration enforces m = n, and the integrals over r2 cancel the factorials, as in (7.348), thus proving the resolution of the identity (7.342), which can also be written as J dZ:dZ Iz)(zi = 1. (7.355) This resolution of the identity can now be inserted into a product decomposition of a Boltzmann operator

(zble-.6iL IZa)

=

(zble-.6iL/(N+l)e-.6Hw/(N+l) . . . e-.6Hw /(N+l) IZa),

(7.356)

to arrive at a sliced path integral [compare (2.2)- (2.4)] (

)- rrN [J -11"dZ~dzn] Nrr+l( ) -(3/( N+1 ) . Zb 1e-.6Hwl Za-n=l n=lZn 1e -fHwl Zn-l,ZO-Za,ZN+l-Zb,f= (7.357)

We now calculate the matrix elements (Znle-fHw IZn- l) and find

(Znle-fHwIZn _l) ~ (znI 1 - titlzn- l) = (ZnIZn- l) - f(znljtlzn - l). Using (7.350) we find

(7.358)

7.8 Coherent States

653

The matrix elements of the operator Hamiltonian (7.298) is easily found. The coherent states (7.345) are eigenstates of the annihilation operator a with eigenvalue z: 00 zn 00 zn alz) = e- z' z/ 2 L -aln) = e- z' z/ 2 L In - 1) = zlz). n=O vInf n=l v(n - I)!

Thus we find immediately

(znIHwlzn-l) = 1iw(znl(at a + aat)IZn_l) = nw

(Z~Zn-l +

(7.360)

D.

Inserting this together with (7.359) into (7.358) , we obtain for small integral

(7.361) E

the path (7.362)

(7.363) The gradient terms can be regrouped using formula (2.35) , and rewriting its righthand side as PN+IXN+l - POXo + L.;:~l(Pn - Pn- l)Xn- l. This leads to

I)} .

n N+1{ z~V'zn+w (Z~Zn-l+2 A~[z*'Z]=2(-Z;Zb+Z:Za)+nE];

(7.364)

Except for the surface terms which disappear for periodic paths, this action agrees with the time-sliced Euclidean action (7.310) , except for a trivial change of variables

a ---; z. As a brief check of formula (7.362) we set N = 0 and find

A~[z*, z] = ~(-Z;Zb + z:za) + nz; (Zb -

za)

+ W (Z;Za + ~) ,

(7.365)

D] .

(7.366)

and the short-time amplitude (7.364) becomes

(zbl e- Elfw IZa) = exp

[-~(Zb'Zb + z:za) + zb'za -

Enw (Zb'Za

+

Applying the recovery operations (7.346) we find (Ole-ElfwIO) = [e(ZbZb+z~Za)/2(zbl e-Elfwlza)L.=o z=o = e- fnw / 2,

(1Ie - Elfw 11) = {Oz

[e(ZbZb+z~za)/2(zble-Elfw IZa) ] 8: L.=o,z=o = e- E3nw /2,

(Ole- EHw ll) = (1Ie- EHw IO) = O.

(7 .367) (7.368)

(7.369)

Thus we have shown that for fixed ends, the path integral gives the amplitude for an initial coherent state IZa) to go over to a final coherent state IZb)' The partition function (7.337) is obtained from this amplitude by forming the diagonal integral

Zw =

JdZ:dz (zle-.B Hw

Iz).

(7.370)

654

7 Many Particle Orbits -

7.9

Second-Quantized Fermi Fields

Statistics and Second Quantization

The existence of the periodic system of elements is based on the fact that electrons can occupy each orbital state only once (counting spin-up and -down states separately) . Particles with this statistics are called fermions. In the above Hilbert space in which n-particle states at a point x are represented by oscillator states In, x ), this implies that the particle occupation number n can take only the values n

° (no electron) ,

n

1 (one electron).

It is possible to construct such a restricted many-particle Hilbert space explicitly by subjecting the quantized fields ~t (x) , ~(x) or their Fourier components a~, ap to anticommutation relations, instead of the commutation relations (7.282) , i.e., by postulating

[~(x, t) , ~t(x' , t)l+ At At [1/J (x, t), 1/J (x', t)l+ [~(x, t), ~ (x' , t)l+

0,

(7.371)

0,

or for the Fourier components

[ap(t), a~, (t)l+ [a~(t) , a~, (t)l+ [ap(t) , ap (t)l+ l

Here

[A, Sl + denotes the

Jppl,

0,

0.

anticommutator of the operators

[A,Sl+ == As + sA.

(7.372)

A and S (7.373)

Apart from the anticommutation relations, the second-quantized description of Fermi fields is completely analogous to that of Bose fields in Section 7.6.

7.10

Fluctuating Fermi Fields

The question arises as to whether it is possible to find a path integral formulation which replaces the anticommuting operator structure. The answer is affirmative, but at the expense of a somewhat unconventional algebraic structure. The fluctuating paths can no longer be taken as c-numbers. Instead, they must be described by anticommuting variables.

7.10.1

Grassmann Variables

Mathematically, such objects are known under the name of Grassmann variables. They are defined by the algebraic property (7.374)

655

7.10 Fluctuating Fermi Fields

which makes them nilpotent: (7.375)

These variables have the curious consequence that an arbitrary function of them possesses only two Taylor coefficients, Fa and F1 , (7.376)

They are obtained from F(O) as follows :

F(O), F

, _a

(7.377)

= aoF.

The existence of only two parameters in F(O) is the reason why such functions naturally collect amplitudes of two local fermion states, Fa for zero occupation, Fl for a single occupation. It is now possible to define integrals over functions of these variables in such a way that the previous path integral formalism remains applicable without a change in the notation, leading to the same results as the second-quantized theory with anticommutators. Recall that for ordinary real functions, integrals are linear functionals. We postulate this property also for integrals with Grassmann variables. Since an arbitrary function of a Grassmann variable F(O) is at most linear in 0, its integral is completely determined by specifying only the two fundamental integrals J dO and J dO O. The values which render the correct physics with a conventional path integral notation are

dO J.j'h

0,

(7.378)

J ~O .j'h

1.

(7.379)

Using the linearity property, an arbitrary function F(O) is found to have the integral = Fl = F'. (7.380) J ~F(O) .j'h Thus, integration of F(O) coincides with differentiation. This must be remembered whenever Grassmann integration variables are to be changed: The integral is transformed with the inverse of the usual Jacobian. The obvious equation

dO ( dO' (') c· ) 0 = c· F , = c· J --F 0 J --F .j'h .j'h

(7.381)

for any complex number c implies the relation

dO' [dO] - lF(O')' J ~F(O'(O))=J .j'h .j'h dO'

(7.382)

656

7 Many Particle Orbits - Statistics and Second Quantization

For ordinary integration variables, the Jacobian dO/dO' would appear without the power -1. When integrating over a product of two functions F(O) and G(O), the rule of integration by parts holds with the opposite sign with respect to that for ordinary integrals:

dO G (O) oOF(O) a J V2ii

JdO[O] V2ii ooG(O) F(O).

=

(7.383)

There exists a simple generalization of the Dirac 8-function to Grassmann variables. We shall define this function by the integral identity

J

~8(0 -

O')F(O') == F(O).

(7.384)

Inserting the general form (7.376) for F(O), we see that the function

8(0-0')=0'-0

(7.385)

satisfies (7.384). Note that the 8-function is a Grassmann variable and, in contrast to Dirac's 8-function, antisymmetric. Its derivative has the property

8'(0 - 0') == 008(0 - 0') = -1.

(7.386)

It is interesting to see that 8' shares with Dirac's 8' the following property:

~8'(0 -

J

O')F(O')

=

-F'(O),

(7.387)

with the opposite sign of the Dirac case. This follows from the above rule of partial integration, or simpler, by inserting (7.386) and the explicit decomposition (7.376) for F(O).

The integration may be extended to complex Grassmann variables which are combinations of two real Grassmann variables 01, O2 :

a*

=

~(Ol -

i02 ),

(7.388)

The measure of integration is defined by

J

da*da == J d0 2 d0 1 == - J dada*. 7r

2m

(7.389)

7r

Using (7.378) and (7.379) we see that the integration rules for complex Grassmann variables are

J da:da

(7.390)

da* da * J --aa

(7.391 )

7r

657

7.10 Fluctuating Fermi Fields

Every function of a* a has at most two terms: F(a*a) = Fo

+ Fl

a*a.

(7.392)

In particular, the exponential exp{ -a* Aa} with a complex number A has the Taylor series expansion e - a'Aa = 1- a* Aa. (7.393) Thus we find the following formula for the Gaussian integral:

J

da: da e- a' Aa

= A.

(7.394)

The integration rule (7.390) can be used directly to calculate the Grassmann version of the product of integrals (7.315) . For a matrix A which can be diagonalized by a unitary transformation, we obtain directly (7.395)

Remarkably, the fermion integration yields precisely the inverse of the boson result (7.315).

7,10,2

Fermionic Functional Determinant

Consider now the time-sliced path integral of the partition function written like (7.309) but with fermionic anticommuting variables. In order to find the same results as in operator quantum mechanics it is necessary to require the anticommuting Grassmann fields a(T), a* (T) to be antiperiodic on the interval T E (0, Ti(3), i.e., a(Ti(3)

= -a(O) ,

(7.396)

or in the sliced form (7.397)

Then the exponent of (7.395) has the same form as in (7.315), except that the matrix A of Eq. (7.398) is replaced by 1 -l+EW f

A =10(1-

-

EW)~T

+ lOW =

0 1

0 0

-l+EW

0

0

0

0 0 1 -1

0 0 0 0

+ EW

0

. ..

-l+EW

1-

lOW

0 0 0 1 (7.398)

where the rows and columns are counted from 1 to N + 1. The element in the upper right corner is positive and thus has the opposite sign of the bosonic matrix

658

7 Many Particle Orbits - Statistics and Second Quantization

in (7.318). This makes an important difference: While for w gave

= 0 the bosonic matrix (7.399)

due to translational invariance in

T,

we now have (7.400)

The determinant of the fermionic matrix (7.398) can be calculated by a repeated expansion along the first row and is found to be (7.401) Hence we obtain the time-sliced fermion partition function (7.402) As in the boson case, we introduce the auxiliary frequency We

1

== --log(l - EW)

(7.403)

= 1 + e-(3liwe .

(7.404)

E

and write Z~N in the form Z:.:

This partition function displays the typical property of Fermi particles. There are only two terms, one for the zero-particle and one for the one-particle state at a point. Their energies are 0 and 'hw e , corresponding to the Hamiltonian operator (7.405) In the continuum limit into

E --+

0, where We

--+ W ,

the partition function Zw N goes over (7.406)

Let us generalize also the fermion partition function to a system with a timedependent frequency O(T), where it reads (7.407) with the sliced action N

A!: = 'h L [a~(an - an-i) n=i

+ EOna~an-il ,

(7.408)

or, expressed in terms of the difference operator V', N

A!:

=

'hE

L

n=i

a~ [(1- EOn)V' + On] an·

(7.409)

7.10 Fluctuating Fermi Fields

659

The result is N

Z~N

= detN+l[E(l - d2)\7 + EW] = 1 -

II (1 -

dl n ).

(7.410)

n=O

As in the bosonic case, it is useful to introduce the auxiliary frequency (7.411) and write

z5

N

in the form (7.412)

If we attempt to write down a path integral formula for fermions directly in the continuum limit, we meet the same phenomenon as in the bosonic case. The difference operator (7.398) turns into the corresponding differential operator

(7.413) which now acts upon periodic complex functions e-iw;"T with the odd Matsubara frequencies (7.414) = w(2m + l)kBT/n, m = 0, ±1, ±2, . . ..

w:"

The continuum partition function can be written as a path integral

f

Va*Va

[ { Iif3

-w-

exp -

Nwdet (aT

+ w),

Jo

dT (a*oTa + wa*a)

]

(7.415)

with some normalization constant N w determined by comparison with the timesliced result. To calculate Z~, we take the eigenvalues of the operator aT + w, which are now -iw;, + w, and evaluate the product of ratios 00

II m=-oo

_iwf +w ~ f

-twm

= cosh(nw;3/2).

(7.416)

This corresponds to the ratio of functional determinants det (aT + w) = h(" /.1/2) det (aT) cos nWfJ .

(7.417)

In contrast to the boson case (7.336), no prime is necessary on the determinant of aT since there is no zero frequency in the product of eigenvalues (7.416). Setting N w = 1/2det (aT) ' the ratio formula produces the correct partition function z~

= 2cosh(nw;3/2).

(7.418)

660

7 Many Particle Orbits - Statistics and Second Quantization

Thus we may write the free-fermion path integral in the continuum form explicitly as follows :

f Va*Va [_11i d

( *!l * )] _ 2 det (OT + w) a UTa + wa a d (!l) et U

i3

exp

7r

0

T

T

2 cosh(1iw,B/2).

(7.419)

The determinant of the operator OT

+ w)

det (OT

+w

= det (-OT

can again be replaced by

+ w)

= Jdet (-0;'

+ w2).

(7.420)

As in the bosonic case, this Fermi analog of the harmonic oscillator partition function agrees with the results of dimensional regularization in Subsection 2.15.4 which will ensure invariance of path integrals under a change of variables, as will be seen in Section 10.6. The proper fermionic time-sliced partition function corresponding to the dimensional regularization in Subsection 2.15.4 is obtained from a fermionic version of the time-sliced oscillator partition function by evaluating

(7.421 ) with a product over the odd Matsubara frequencies

w;". The result is (7.422)

with We given by (7.423)

sinh(we /2) = Ew/2 .

This follows from the Fermi analogs of the product formulas (2.398), (2.400):7

NIt (1 - _.- -,-:i-;: ~_:X" '+"'l)-1l")

cos(N + l)x cos x

N = even,

(7.424)

sin2 x ) . 2 (2m+1)1l" sm 2(N+l)

cos(N + l)x,

N=odd.

(7.425)

sm

m=O

(N - 1)/2 (

II m=O

1-

2(N+ 1)

For odd N , where all frequencies occur twice, we find from (7.425) that N

II m=O

(

sin 2 x ) 1- . 2~ sm 2(N+l)

1/2

=cos(N+1)x,

71.8. Gradshteyn and I.M. Ryzhik, op. cit. , Formulas 1.391.2, 1.391.4.

(7.426)

661

7.10 Fluctuating Fermi Fields

and thus, with (7.423), directly (7.422) . For even N, where the frequency with m = N/2 occurs only once, formula (7.424) gives once more the same answer, thus proving (7.426) for even and odd N . There exists no real fermionic oscillator action since x 2 and j;2 would vanish identically for fermions, due to the nilpotency (7.375) of Grassmann variables. The product of eigenvalues in Eq. (7.421) emerges naturally from a path integral in which the action (7.408) is replaced by a symmetrically sliced action. An important property of the partition function (7.418) of (7.422) is that the ground-state energy is negative: E(O)

= _ 1iw

(7.427)

2'

As discussed at the end of Section 7.7, such a fermionic vacuum energy is required for each bosonic vacuum energy to avoid an infinite vacuum energy of the world, which would produce an infinite cosmological constant, whose experimentally observed value is extremely small.

7.10.3

Coherent States for Fermions

For the bosonic path integral (7.304) we have studied in Section 7.8, the case that the endpoint values aa = a(Ta) and ab = a(Tb) of the paths a(T) are held fixed. The result was found to be the matrix element of the Boltzmann operator e- f3Hw between coherent states la ) = e- a*a/2eaa t 10) [recall (7.345)]. There exists a similar interpretation for the fermion path integral (7.415) if we hold the endpoint values aa = a(Ta) and ab = a(Tb) ofthe Grassmann paths fixed. By analogy with Eq. (7.345) we introduce coherent states [50]

(7.428) The corresponding adjoint states read

(7.429) Note that for consistency of the formalism, the Grassmann elements ( anticommute with the fermionic operators. The states 10) and (11 and their conjugates (0 1 and (1 1can be recovered from the coherent states I() and ((I by the operations:

In) = [1()e(*(/2

an] _1_ , ( (=0 v'nf

(7.430)

These formula simplify here to

10) = [1()e(*(/2L =0, 11) = [1 ()e(*(/2

a(L=o'

(7.431)

(0 1= [e(*(/2(( IL=0' (1 1= [o(*e(*(/2(( I]

(=0

.

(7.432)

662

7 Many Particle Orbits - Statistics and Second Quantization

For an operator agonal elements trO =

0 , the trace can be calculated from the integral over the antidi-

J

d(:d( (-(101()

=

J

d(:d( e-(*«((Ol- (*(11)0(10) - (11)).

(7.433)

Using the integration rules (7.390) and (7.391), this becomes trO = (01010)

+ (11011).

(7.434)

The states Ie) form an overcomplete set in the one-fermion Hilbert space. The scalar products are [compare (7.350)]:

e-

(i (, /2- (2 (2/2+ (i (2

(7.435)

e- (i«, - (2)/2+«i - (. We have shown that as in the Bose case the path integral with fixed ends gives the amplitude for an initial coherent state I(a) to go over to a final coherent state I(b). The fermion partition function (7.419) is obtained from this amplitude by forming the trace of the operator e- (3Hw , which by formula (7.434) is given by the integral over the antidiagonal matrix elements

(7.440) The antidiagonal matrix elements lead to antiperiodic boundary conditions of the fermionic path integral.

663

7.11 Hilbert Space of Quantized Grassmann Variable

7.11

Hilbert Space of Quantized Grassmann Variable

To understand the Hilbert space associated with a path integral over a Grassmann variable we recall that a path integral with zero Hamiltonian serves to define the Hilbert space via all its scalar products as shown in Eq. (2.18):

(Xbtblxata)

=

J1)x J~:n exp [i JdtP(t)±(t)]

=

(xblxa)

=

8(Xb - xa).

(7.441)

A momentum variable inside the integral corresponds to a derivative operator p == -ili8x outside the amplitude, and this operator satisfies with x = x the canonical commutation relation [p, x] = -iii [see (2.19)]. By complete analogy with this it is possible to create the Hilbert space of spinor indices with the help of a path integral over anticommuting Grassmann variables. In order to understand the Hilbert space, we shall consider three different cases.

7.11.1

Single Real Grassmann Variable

First we consider the path integral of a real Grassmann field with zero Hamiltonian

.] . J1)B27r exp [ih, Jdt 2iii B(t)B(t) From the Lagrangian

£(t)

i

.

(7.442)

21iB(t)B(t)

(7.443)

8£ iii Po=-. =--B.

(7.444)

=

we obtain a canonical momentum 8B

2

Note the minus sign in (7.444) arising from the fact that the derivative with resect to Banticommutes with the variable B on its left. The canonical momentum is proportional to the dynamical variable. The system is therefore subject to a constraint X =Po

iii

+ -2 B =

0.

(7.445)

In the Dirac classification this is a second-class constraint, in which case the quantization proceeds by forming the classical Dirac brackets rather than the Poisson brackets (1.21), and replacing them by ± i/Ii times commutation or anticommutation relations, respectively. For n dynamical variables qi and m constraints Xp the Dirac brackets are defined by (7.446) where C pq is the inverse of the matrix (7.447)

664

7 Many Particle Orbits - Statistics and Second Quantization

For Grassmann variables Pi, gi, the Poisson bracket (1.21) carries by definition an overall minus sign if A contains an odd product of Grassmann variables. Applying this rule to the present system we insert A = Po and B = () into the Poisson bracket (1.21) we see that it vanishes. The constraint (7.445), on the other hand, satisfies

{X,X}

=

{ Po + 22'n (),po + 22'n () }

-in{po,()}

=

=

-in.

(7.448)

Hence C = -in with an inverse i/n. The Dirac bracket is therefore (7.449) With the substitution rule {A, B}v ---; (-i/n)[A, B]+, we therefore obtain the canonical equal-time anticommutation relation for this constrained system: A

[P(t) , ()(t)]+ =

in

-2'

(7.450)

or, because of (7.444),

[O(t),O(t)]+ = 1.

(7.451)

The proportionality of Po and () has led to a factor 1/2 on the right-hand side with respect to the usual canonical anticommutation relation. Let 'Ij;(()) be an arbitrary wave function of the general form (7.376): (7.452) The scalar product in the space of all wave functions is defined by the integral (7.453) In the so-defined Hilbert space, the operator given by the differential operator

0 is diagonal,

while the operator p is

p = inoo,

(7.454)

to satisfy (7.450). The matrix elements of the operator pare

('Ij;'lpl'lj;) -=

J

d() 'Ij;'*(())in~'Ij;(()) 2'iT o()

= in'lj;'*'Ij; 1 1·

(7.455)

By calculating (7.456)

7.11 Hilbert Space of Quantized Grassmann Variable

665

we see that the operator fJ is anti-Hermitian, this being in accordance with the opposite sign in the rule (7.383) of integration by parts. Let 19) be the local eigenstates of which the operator 8 is diagonal: 819) = 919).

(7.457)

The operator 8 is Hermitian, such that (918 = (919 = 9(91 ·

(7.458)

The scalar products satisfy therefore the usual relation (9' - 9)(9'19)

=

o.

(7.459)

On the other hand, the general expansion rule (7.376) tells us that the scalar product S = (9'19) must be a linear combination of So + S19 + SW + S299'. Inserting this into (7.459), we find (7.460) (9'19) = -9' + 9 + S299', where the proportionality constants So and SI are fixed by the property (9'19) =

Jd9" (9'10")(9"19). 21f

(7.461)

The constant S2 is an arbitrary real number. Recalling (7.385) we see that Eq. (7.460) implies that the scalar product (9'19) is equal to a

J(27rn)D' dDpV

(7.558)

For infinitely thin time slices, t -> 0 and ,:;\(p) reduces to w(p) and the expressions (7.556), (7.557) turn into the usual free energies of bosons and fermions. They agree completely with the expressions (7.46) and (7.220) derived from the sum over orbits. Next we may introduce a fluctuating field in space and imaginary time

'IjJ(x, T)

=

~L

yV

eipXap(T) ,

(7.559)

p

and rewrite the action (7.553) in the local form

Ae['IjJ*''IjJ]

=

101i,6 dT

JdDx ['IjJ*(X, T)

n8T 'IjJ (x, T)

+ 2~ V 'IjJ*(x , T)V'IjJ(X, T)]

. (7.560)

677

7.14 Spatial Degrees of Freedom

The partition function is given by the functional integral

Z =

f V'l/lD'I/;*e- A.['I/J·,'l/Jl/li .

(7.561)

Thus we see that the functional integral over a fluctuating field yields precisely the same partition function as the sum over a grand-canonical ensemble of fluctuating orbits. Bose or Fermi statistics are naturally accounted for by using complex or Grassmann field variables with periodic or antiperiodic boundary conditions, respectively. The theory based on the action (7.560) is completely equivalent to the second-quantized theory of field operators. In order to distinguish the second-quantized or quantum field description of many particle systems from the former path integral description of many particle orbits, the former is referred to as the first-quantized approach, or also the world-line approach. The action (7.560) can be generalized further to include an external potential V(x, T), i.e. , it may contain a general Schrodinger operator if(T) = p2 + V(x , T) instead of the gradient t erm:

A" ['1/;* , '1/;] = foli/3dT J dDx {'I/;*(x, T)noT'I/; (x, T)+ 'I/;*(X, T) [if(T) -11] 'I/;(x, T)}. (7 .562) For the sake of generality we have also added a chemical potential to enable the study of grand-canonical ensembles. This action can be used for a second-quantized description of the free Bose gas in an external magnetic trap potential V(x) , which would, of course, lead to the same results as the first-quantized approach in Section 7.2.4. The free energy associated with this action 1

A

F = :eTrlog[noT + H(T) -11]

(7 .563)

was calculated in Eq. (3.139) as an expansion F = _l_Tr [rli/3 dT if(T)] 2n(3 Jo

-.!. f='!'Tr

{Te- nJoni3dT" [H(T"l - l-'l/li} ,

(3 n=l n

(7.564)

The sum can be evaluated in the semiclassical expansion developed in Section 4.9. For simplicity, we consider here only time-independent external potentials, where we must calculate Tr [e- n/3(H - l-'l]. Its semiclassical limit was given in Eq. (4.257) continued to imaginary time. From this we obtain F = ~Tr if _ 1 2 (3

1

V21fn2(3/M

D

~ JdD [z(x)]n ~ X D/2+1' n

(7.565)

n=l

where z(x) == e- /3[V(xl - I-'J is the local fugacity (7.24). The sum agrees with the previous first-quantized result in (7.132) and (7.133). The first term is due to the symmetric treatment of the fields in the action (7.562) [recall the discussion after Eq. (7.340)]. One may calculate quantum corrections to this expansion by including the higher gradient terms of the semiclassical expansion (4.257). If the potential is timedependent , the expansion (4.257) must be generalized accordingly.

678

7 Many Particle Orbits -

7.14.2

Statistics and Second Quantization

First versus Second Quantization

There exists a simple set of formulas which illustrates nicely the difference between first and second quantization, i.e., between path and field quantization. Both may be though of as being based on two different representations of the Dirac 8-function. The first-quantized representation is (7.566) the second-quantized representation

=

iM(D) (Xb - Xa)8(tb - ta)

f

V '¢V '¢* '¢(Xb , tb)'¢*(Xa, ta)e(i/Ii) J dDx

roo dt,p*(x,t),p(x,t).

(7.567) The first representation is turned into a transition amplitude by acting upon it with the time-evolution operator e-di(tb-t a), which yields (X

t Ix t )

bb a a

= edi (tb- t a)8(D) (xb - x a) =

JVDx

f

VDp (27rn)D

e(i/Ii)J,~dt(Px-H) .

(7.568)

By multiplying this with the Heaviside function 8(tb - t a), we obtain the solution of the inhomogeneous Schrodinger equation (7.569) This may be expressed as a path integral representation for the resolvent

in

\

Xbtb l . tnat -

1 A

H

Xata

)

-

-

-8(tb - ta)

l

X(t b)=Xb

V

D

x

f

x(ta)=Xa

VDp (i/Ii) J,'b dt(px- H) - ")D ( ea . (7 .570) 27rn

The same quantity is obtained from the second representation (7.567) by changing the integrand in the exponent from '¢*(x , t)'¢(x, t) to '¢*(x, t)(inat - H)'¢( x , t): / Xbtb I . in 1xata\ \ tnat - H / A

=

f

V '¢V'¢* '¢(Xb, tb)'¢*(Xa, ta)e(i/Ii) J dDx

roo dW(x,t)(ilio,-H),p(x,t).

(7.571)

This is the second-quantized functional integral representation of the resolvent.

7.14.3

Interacting Fields

The interaction between particle orbits in a grand-canonical ensemble can be accounted for by anharmonic terms in the particle fields. A pair interaction between orbits, for example, corresponds to a fourth-order self interaction. An example is the interaction in the Bose-Einstein condensate corresponding to the energy in Eq. (7.101) . Expressed in terms of the fields it reads

679

7.14 Spatial Degrees of Freedom

A~nt[7{i*' 7{i] = -l'f3dT flE = -~ l'f3 dT

Jd3x7{i*(x, T + TJ)7{i*(X, T + TJ)7{i(X , T)7{i(X, T) ,

(7.572) where TJ > 0 is an infinitesimal time shift. It is then possible to develop a perturbation theory in terms of Feynman diagrams by complete analogy with the treatment in Section 3.20 of the anharmonic oscillator with a fourth-order self interaction. The free correlation function is the momentum sum of oscillator correlation functions:

(7{i(x, T)7{i*(X', T')) =

L (a

p ( T)a~,

(T'))ei(px- p'x')

=

p,p'

L (a

p ( T)a~( T'))eiP(x-

x').

(7.573)

p

The small TJ > 0 in (7.572) is necessary to specify the side of the jump of the correlation functions (recall Fig. 3.2). The expectation value of flE is given by (7.101) , with a prefactor g rather than gj2 due to the two possible Wick contractions. Inserting the periodic correlation function (7.528), we obtain the Fourier integral

(7{i(x, T) 7{i* (X' , T'))

=

L(1 + nwp)e-Wp(T-T')+iP(X-X').

(7.574)

P

Recalling the representation (3.284) of the periodic Green function in terms of a sum over Matsubara frequencies , this can also be written as

(7{i(x, T) 7{i*(X', T')) = ~

L .

-1

nJ3 Wm ,p ZWm -

e-iWm(T-T')+ip(x-x') .

wp

(7.575)

The terms in the free energies (7.102) and (7.107) with the two parts (7.108) and (7.109) can then be shown to arise from the Feynman diagrams in the first line of Fig. 3.7. In a grand-canonical ensemble, the energy nwp in (7.575) is replaced by liwp - 1-£. The same replacement appears in wp of Eq. (7.574) which brings the distribution function nwp to [recall (7.529)] 1

(7.576)

ef3l' .

The expansion of the Feynman integrals in powers where z is the fugacity z = of z yields directly the expressions (7.102) and (7.107).

7.14.4

Effective Classical Field Theory

For the purpose of studying phase transitions, a functional integral over fields 7{i(x, T) with an interaction (7.572) must usually be performed at a finite temperature. Then is often advisable to introduce a direct three-dimensional extension of the effective classical potential Veffcl (xo) introduced in Section 3.25 and used efficiently in Chapter 5. In a field theory we can set up, by analogy, an effective classical action which is a functional of the three-dimensional field with zero Matsubara frequency ¢>(x) == 7{io(x). The advantages come from the reasons discussed in Section 3.25, that

680

7 Many Particle Orbits - Statistics and Second Quantization

the zero-frequency fluctuations have a linearly diverging fluctuation width at high temperature, following the Dulong-Petit law. Thus only the nonzero-modes can be treated efficiently by the perturbative methods explained in Subsection 3.25.6. By analogy with the splitting of the measure of path integration in Eq. (3.805) , we may factorize the functional integral (7.561) into zero- and nonzero-Matsubara frequency parts as follows: (7.577) and introduce the Boltzmann factor [compare (3.810)] contain the effective classical action (7.578) to express the partition function as afunctional integral over time-independent fields in three dimensions as: (7.579)

In Subsection 3.25.1 we have seen that the full effective classical potential Veffcl(XO) in Eq. (3.809) reduces in the high-temperature limit to the initial potential V(xo). For the same reason, the full effective classical action in the functional integral (7.577) can be approximated at high temperature by the bare effective classical action, which is simply the zero-frequency part of the initial action:

This follows directly from the fact that, at high temperature, the fluctuations in the functional integral (7.578) are strongly suppressed by the large Matsubara frequencies in the kinetic terms. Remarkably, the absence of a shift in the critical temperature in the first-order energy (7.113) deduced from Eq. (7.116) implies that the chemical potential in the effective classical action does not change at this order [34]. For this reason, the lowest-order shift in the critical temperature of a weakly interacting Bose-Einstein condensate can be calculated entirely from the three-dimensional effective classical field theory (7.577) with the bare effective classical action (7.580) in the Boltzmann factor. The action (7.580) may be brought to a more conventional form by introducing the differently normalized two-component fields

T > Ta (8.20)

8.2

Trouble with Feynman's Path Integral Formula in Radial Coordinates

In the above calculation we have shown that the expression (8.10) is certainly the correct radial path integral. It is, however, not of the Feynman type. In operator quantum mechanics we learn that the action of a particle moving in a potential V(r) at a fixed angular momentum L3 = mn contains a centrifugal barrier n2 (m 2 _ 1/4)/2Mr2 and reads (8.21) This is shown by separating the Hamiltonian operator into radial and azimuthal coordinates, over fixing the azimuthal angular momentum L 3 , and choosing for it the quantum-mechanical value nm. According to Feynman's rules, the radial amplitude therefore should simply be given by the path integral (8.22)

8.2 Trouble with Feynman's Path Integral Formula in Radial Coordinates

701

The reader may object to using the word "classical" in the presence of a term proportional to n? in the action. In this section, however, Tim is merely meant to be a parameter specifying the azimuthal momentum P

0, the number m would become infinitely large, leading to the correct high-temperature limit of the partition function. For a fixed finite m, however, the discrepancy is unavoidable.

704

8 Path Integrals in Polar and Spherical Coordinates

with the sliced action

(8.35) After integrating out the momenta, this becomes (8.36) with the action (8.37) The path integral formula (8.36) can in principle be used to find the amplitude for a fixed angular momentum of some solvable systems. An example is the radial harmonic oscillator at an angular momentum m, although it should be noted that this particularly simple example does not really require calculating the integrals in (8.36). The result can be found much more simply from a direct angular momentum decomposition of the amplitude (2.175). After a continuation to imaginary times t = -iT and an expansion of part of the exponent with the help of (8.5), it reads for D = 2

(8.38) By comparison with (8.9), we extract the radial amplitude (8.39) The limit w

----+

0 gives the free-particle result

M

y'rbra (rbTblraTa)m = -.;----Im n Tb - Ta

8.3

(M

rbra ) -.;---- . n Tb - Ta

(8.40)

Cautionary Remarks

It is important to emphasize that we obtained the correct amplitudes by performing the time slicing in Cartesian coordinates followed by the transformation to the polar coordinates in the time-sliced expression. Otherwise we would have easily missed

705

8.3 Cautionary Remarks

the factor -1/4 in the centrifugal barrier. To see what can go wrong let us proceed illegally and do the change of variables in the initial continuous action. Thus we try to calculate the path integral (XbTblxaTa) =

1 1

L

00

Vrr

l=- oo,oo 0

x exp {~ It

l

Tb

Ta

dT

00

Vip

-00

[M2 (f2 + r2rp2) + v(r)]} 1 00 limit of Enr = wn(2nr + l + D /2) ---> n2k 2/2M]' we obtain the spectral representation of the free-particle propagator (9.80)

For comparison, we derive the same results directly from the initial spectral representation (9.2) in one dimension (9.81)

Its "angular decomposition" is a decomposition with respect to even and odd wave functions

In D dimensions we use the expansion (8.100) for eikx to calculate the amplitude in the radial form

. J(21fdDk)D e,k(Xb-Xa) 1

(21f)D/2

Itk 2

e- 2M (Tb-Ta)

roo dkk2v (kRY 1 J (kR) -~~ h-Ta) v e ,

io

(9.83)

with v == D /2 - 1. With the help of the addition theorem for Bessel functions lO (8.187) we rewrite (9.84)

and expand further according to 1

(kR)VJv(kR)

(21f)D/2 00 * (Prbra)v ~ Jv+l(krb) J v+l (kra) ~Yim(Xb)Ylm(Xa), A

=

A

(9.85)

to obtain the radial amplitude (9.86)

l°I.S. Gradshteyn and I.M. Ryzhik, op. cit. , Formula 8.532.

763

9.4 Charged Particle in Uniform Magnetic Field

just as in (9.80). For D = 1, this reduces to (9.82) using the particular Bessel functions VzJ~1/2(Z) =

9.4

2 ~

v21l"

{COSZ} . . smz

(9.87)

Charged Particle in Uniform Magnetic Field

Let us also find the wave functions of a charged particle in a magnetic field . The amplitude was calculated in Section 2.18. Again we work with the imaginary-time version. Factorizing out the free motion along the direction of the magnetic field, we write (9.88) with (ZbTbIZaTa) =

1 exp {_ M (Zb - za)2} , V21l"n(Tb - Ta)/M 2n Tb - Ta

(9.89)

and have for the amplitude in the transverse direction (9.90) with the classical transverse action

At = ~w {! coth [wh -

Ta)/2](xt -

X~)2 + x~ X xt}.

(9.91)

This result is valid if the vector potential is chosen as 1

(9.92)

A="2B xx. In the other gauge with A

= (0, Bx, 0),

(9.93)

there is an extra surface term, and A~ is replaced by -1.

Acl

=

1.

Acl

Mw

+ -2-(XbYb -

xaYa).

(9.94)

The calculation of the wave functions is quite different in these two gauges. In the gauge (9.93) we merely recall the expressions (2.661) and (2.663) and write down the integral representation 1. ) ( xb1. Tb Ixa Ta

--

J

dpy eipy(Yb- Ya)/h (XbTb IXaTa ) xo=py/Mw, 21l"n

(9.95)

with the oscillator amplitude in the x-direction (XbTblxaTa)xo =

J21l"nSinh~~Tb _ Ta)] exp ( -~A~O '

(9.96)

764

9 Wave Functions

and the classical oscillator action centered around Xo

A~l

=

2'

h[~W

sm W Tb - Ta

)] {[(Xb - xo?

+ (xa -

xo?] cosh[w(Tb - Tb)]

-2(Xb - XO)(xa - xo)} .

(9.97)

The spectral representation of the amplitude (9.96) is then 00

(XaTblxaTa)xo =

L

'If;n(Xb - XO)'lf;n(xa - xo)e- (n+~)w(Tb-Tal,

(9.98)

n=O

where 'If;n(x) are the oscillator wave functions (2.300). This leads to the spectral representation of the full amplitude (9.88)

J dpz J dpYeiPz(Zb-Za)/1i 27rn

(9 .99)

27rn

00

xL 'If;n(Xb -

py/Mw)'lf;n(xa - py/Mw)e- (n+ ~)w(Tb -Ta).

n=O

The combination of a sum and two integrals exhibits the complete set of wave functions of a particle in a uniform magnetic field. Note that the energy (9.100) is highly degenerate; it does not depend on Py. In the gauge A = ~B x x, the spectral decomposition looks quite different. To derive it, the transverse Euclidean action is written down in radial coordinates [compare Eq. (2.667)] as

A~ = ~ {~coth[W(Tb -

Ta)/2]

b +Ta 2

2 -

2TbTaCOS(IPb - IPa)] -iWTbTasin(IPb-IPa)} .

(9.101)

This can be rearranged to

(9.102) We now expand

CAJ/Ii

into a series of Bessel functions using (8.5)

765

9.4 Charged Particle in Uniform Magnetic Field

The fluctuation factor is the same as before. Hence we obtain the angular decomposition of the transverse amplitude (9.104)

where

with (9.106)

To find the spectral representation we go to the fixed-energy amplitude

The integral is done with the help of formula (9.29) and yields

(9.108)

with

E

1/=

The Gamma function f(1/2 -

1/ -

(9.109)

Iml/2) has poles at _

1/

m

-wn + -. 2 1

Iml

= I/r = nr + "2 + ""2

(9.110)

of the form (9.111)

The poles lie at the energies (9.112)

766

9 Wave Functions

These are the well-known Landau levels of a particle in a uniform magnetic field. The Whittaker functions at the poles are (for m > 0) M- v,m/2(Z) W v,m/2(Z)

eZ / 2 Z ~ 2 M(-n r, l+m , -z) ,

(9.113)

( - ) nr (nr e- z /2 Z ~ 2

(9.114)

+,m)! M ( -nT) 1 + m, Z ) . m.

The fixed-energy amplitude near the poles is therefore (9.115) with the radial wave functions l l (9.116)

Using Eq. (9.53) , they can be expressed in terms of Laguerre polynomials

nr !

(n r M

X

(~ 211, r 2)

+ Iml)!

Mw exp ( ---r

L~(z):

2)

411,

M

Iml/2 Llml nr

(~ 211, r 2) .

(9.117)

The integral (9.54) ensures the orthonormality of the radial wave functions (9.118) A Laplace transformation of the fixed-energy amplitude (9.108) gives, via the residue theorem, the spectral representation of the radial time evolution amplitude (9.119) with the energies (9.112). The full wave functions in the transverse subspace are, of course, (9.120) Comparing the energies (9.112) with (9.100), we identify the principal quantum number n as

n=nr

Iml m +---· 2 2

(9.121)

l1Compare with L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965, p. 427.

767

9.4 Charged Particle in Uniform Magnetic Field

Note that the infinite degeneracy of the energy levels observed in (9.100) with respect to Py is now present with respect to m. This energy does not depend on m for m ~ o. The somewhat awkward m-dependence of the energy can be avoided by introducing, instead of m , another quantum number n' related to n , m by

m=n'-n.

(9.122)

The states are then labeled by n, n' with both nand n' taking the values 0, 1, 2, 3, .... For n' < n, one has n' = nr and m = n' - n < 0, whereas for n' ~ n one has n = nr and m = n' - n ~ O. There exists a natural way of generating the wave functions Wnrm (x) such that they appear immediately with the quantum numbers n , n' . For this we introduce the Landau radius

a=

VMw 21i ~ VeE =

(9.123)

as a length parameter and define the dimensionless transverse coordinates

z = (x

+ iy)/V'ia,

z* = (x - iy)/V'ia.

(9.124)

It is then possible to prove that the Wnrm's coincide with the wave functions

./, ( *) N n,n,e z'z ( - V'iuz* 1 '" 'Pn,n' Z, Z =

)n (- V'iuz 1 '" )n, e-2z'z .

(9.125)

The normalization constants are obtained by observing that the differential operators 1

ez'z ( - -1u'" . ) e-z'z

V'i( -Oz, + z),

ez· z ( - -1u"') e-z' z

~(-O V'i z + z* )

V'iz

V'iz

(9.126)

behave algebraically like two independent creation operators 1

V'i( -Oz, + z), bt

~(-Oz + z*),

(9.127)

whose conjugate annihilation operators are

a (9.128)

768

9 Wave Functions

The ground state wave function annihilated by these is

'l.jJo,o(z, z*) = (z, z*IO) ex e- Z • Z •

(9.129)

We can therefore write the complete set of wave functions as .o/n,n' 1. ( Z, Z

*) = N nn,a'tnb'tn'.0/0,0 I. (Z, Z *) .

(9.130)

Using the fact that at* = bt, bt* = at, and that partial integrations turn a, b, respectively, the normalization integral can be rewritten as

Jdx dy 'l.jJnl,n'l (z, Z*)'l.jJn2,n'2 (z, z*) Nnln'1Nn2n'2 Jdxdy [Cat)nl(bt)n'le[(a t t2(bt t'2 eNn1n'l N n2n'2 Jdxdye (anlbn'latn2btn'2) . Z' Z]

Z' Z ]

2z ' z

Here the commutation relations between in the last line to

bt ,at into

(9.131)

at, bt, a, bserve to reduce the parentheses (9.132)

The trivial integral (9.133) shows that the normalization constants are (9.134) Let us prove the equality of 'l.jJnr m and 'l.jJn,n' up to a possible overall phase. For this we first observe that z, Oz, and z*, Oz carry phase factors ei'P and e- i'P, respectively, so that the two wave functions have obviously the azimuthal quantum number m = nn'. Second, we make sure that the energies coincide by considering the Schrodinger equation corresponding to the action (2.641) - 1 ( -inV- -e A 2M c

)2 'I.jJ =E'I.jJ.

(9.135)

In the gauge where A

= (0, Bx, 0),

it reads

n - 2M 2

[

ax 2+ (Oy -

.eB

~--;;-x)

2+ Oz 2] 'I.jJ =

E 'I.jJ ,

(9.136)

and the wave functions can be taken from Eq. (9.99). In the gauge where

A = (-By/2, Bx/2, 0) ,

(9.137)

769

9.4 Charged Particle in Uniform Magnetic Field

on the other hand, the Schrodinger equation becomes, in cylindrical coordinates,

(9.138) Employing a reduced radial coordinate p = r / a and factorizing out a plane wave in the z-direction, eipzz/", this takes the form (9.139) The solutions are (9.140) where the confluent hypergeometric functions M ( -nr, for integer values of the radial quantum number

Iml + ~,p)

are polynomials

(9.141) as in (9.116). The energy is related to the principal quantum number by (9.142) Since 2Ma2 ~ the energy is

1

nw'

( 1)

p; E= n+- 'hw+-. 2 2M

(9.143)

(9.144)

We now observe that the Schrodinger equation (9.139) can be expressed in terms of the creation and annihilation operators (9.127), (9.128) as 4 [-(ata

+ 1/2) + n~

(E - :~)] ~(z, z*) = O.

(9.145)

This proves that the algebraically constructed wave functions ~n,n' in (9.130) coincide with the wave functions ~nrm of (9.116) and (9.140), up to an irrelevant phase. Note that the energy depends only on the number of a-quanta; it is independent of the number of b-quanta.

770

9.5

9 Wave Functions

Dirac 8-Function Potential

For a particle in a Dirac b-function potential, the fixed-energy amplitudes (xblxa)E can be calculated by performing a perturbation expansion around a free-particle amplitude and summing it up exactly. For any time-independent potential V(x), in addition to a harmonic potential Mw 2x2/2, the perturbation expansion in Eq. (3.475) can be Laplace-transformed in the imaginary time via (9.3) to find

(xblxa)E

J - ~2 J x2 JdDxl(Xblx2)w,EV(X2)(X2Ixl)w,EV(Xl)(Xllxa)w,E (xblxa)w ,E -

+ +

~ dDxl(Xblxl)w,EV(Xl)(Xllxa)w,E

dD

(9.146)

If the potential is a Dirac b-function centered around X,

V(X) = gb(D)(X - X),

n? 9 == Ml2-D '

(9.147)

this series simplifies to

ig g2 (xbIXa)E=(Xblxa)w,E- "hg(xbIX)w,E(Xlxa)w,E- 1i2 (xbIX)w ,E(XIX)w,E(Xlxa)w,E+ . .. , (9.148) and can be summed up to (9.149) This is, incidentally, true if a b-function potential is added to an arbitrary solvable fixed-energy amplitude, not just the harmonic one. If the b-function is the only potential, we use formula (9.149) with w = 0, so that (xblxa)o,E reduces to the fixed-energy amplitude (9.12) of a free particle, and obtain directly .M "P-2 K D / 2- 1 (KR) (xblxa)E = -2Zt; (27r)D/2 (KR)D/2- 1 .2M KD- 2 K D/ 2- 1(KR b) .2M KD- 2 K D/ 2- 1(KRa) ig Z 1i (27r)D/2 (KR b)D/2- 1 x zr;:- (27r)D/2 (KR a)D/2- 1 1i g2MKD 2KD/ 2 _ 1 (Kb) 1 - "h,r;:- 7rD/2 (Kb)D/2- 1

(9.150)

where R == IXb - xal and Ra,b == IXa,b - XI, and b is an infinitesimal distance regularizing a possible singularity at zero-distance. In D = 1 dimension, this reduces to (9.151)

771

9.5 Dirac t5-Function Potential

For an attractive potential with l < 0, the second term can be written as (9.152) exhibiting a pole at the bound-state energy EB = -Ti/2Ml 2 . In its neighborhood, the pole contribution reads (9.153) This has precisely the spectral form (1.321) with the normalized bound-state wave function

'If;B (X) =

VI~ e- Ix-X I/l .

(9.154)

In D = 3 dimensions, the amplitude (9.150) becomes (9.155) In the limit 15 ---* 0, the denominator requires renormalization. We introduce a renormalized coupling length scale 1 1 -=1+-

lr -

(9.156)

27rt5 '

and rewrite the last factor in (9.155) as 1

(9.157)

For lr < 0, this has a pole at the bound-state energy EB = -47r 2 Ti 2 /2Ml'k of the form 1

(9.158)

-lrEB E - EB The total pole term in (9.155) can therefore be written as

(9.159) with K,B = ..j2M EB/Ti = 27r /lr and the normalized bound-state wave functions

_ (K,~) 1/4 'If;B(X) - -42 7r

e- KB lx - XI

r

.

(9.160)

772

9 Wave Functions

In D = 2 dimensions, the situation is more subtle. It is useful to consider the amplitude (9.150) in D = 2 + E dimensions where one has

(xblxa)E

=

+

.M 1 K./2(IiR) -z-/ 1i 7r (27rIiR)€ 2 . M2 1 1 K€/2 (IiR b )K€/2 (liRa) Z1i 27r2 (27rIiR b )€/2 (27rIiRa)€/2 1i M 1 K ( ) -g + -1i7r (27rIiO)€/2 € /2 liE

(9.161)

Inserting here K€/2(IiO) ~ (1/2)r(E/2)(lio/2)-€/2, the denominator becomes

M f(E/2) -1ig + -21i7r (7rIiO)€/2

~

M -2 [ -1ig + -21i7r 1E - - log 7rIiO ] E 2 ( ).

(9.162)

Here we introduce a renormalized coupling constant 1 gr

1 g

M1 1i E'

-=-+-2

(9.163)

and rewrite the right-hand side as

1i M - - -log7r/i;o. gr 21i7r

(9.164)

This has a pole at (9.165) indicating a bound-state pole of energy EB = -1i21i~ /2M . We can now go to the limit of D = 2 dimensions and find that the pole term in (9.161) has the form

'ljJB(Xb)'ljJ~(Xa) E ~1iEB'

(9.166)

with the normalized bound-state wave function (9.167)

Notes and References The wave functions derived in this chapter from the time evolution amplitude should be compared with those given in standard textbooks on quantum mechanics, such as L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965. The charged particle in a magnetic field is treated in §Ill. The 8-function potential was studied via path integrals by C. Grosche, Phys. Rev. Letters, 71, 1 (1993) .

Make not my paths offensive to the Gods. AESCHYLOS,

Agamemnon, 891 b.c.

10 Spaces with Curvature and Torsion

The path integral of a free particle in spherical coordinates has taught us an important lesson: In a Euclidean space, we were able to obtain the correct time-sliced amplitude in curvilinear coordinates by setting up the sliced action in Cartesian coordinates Xi and transforming them to the spherical coordinates ql' = (r, ¢). It was crucial to do the transformation at the level of the finite coordinate differences, ~Xi ----+ ~ql'. This produced higher-order terms in the differences ~ql' which had to be included up to the order (~q)4/f. They all contributed to the relevant order f. It is obvious that as long as the space is Euclidean, the same procedure can be used to find the path integral in an arbitrary curvilinear coordinate system ql', if we ignore subtleties arising near coordinate singularities which are present in centrifugal barriers, angular barriers, or Coulomb potentials. For these, a special treatment will be developed in Chapters 12- 14. We are now going to develop an entirely nontrivial but quite natural extension of this procedure and define a path integral in an arbitrary metric-affine space with curvature and torsion. It must be emphasized that the quantum theory in such spaces is not uniquely defined by the formalism developed so far. The reason is that also the original Schrodinger theory which was used in Chapter 2 to justify the introduction of path integrals is not uniquely defined in such spaces. In classical physics, the equivalence principle postulated by Einstein is a powerful tool for deducing equations of motion in curved space from those in fiat space. At the quantum level, this principle becomes insufficient since it does not forbid the appearance of arbitrary coordinate-independent terms proportional to Planck's quantum 1i 2 and the scalar curvature R to appear in the Schodinger equation. We shall set up a simple extension of Einstein's equivalence principle which will allow us to carry quantum theories from fiat to curved spaces which are, moreover, permitted to carry certain classes of torsion. In such spaces, not only the time-sliced action but also the measure of path integration requires a special treatment. To be valid in general it will be necessary to find construction rules for the time evolution amplitude which do not involve the crutch of Cartesian coordinates. The final formula will be purely intrinsic to the general metric-affine space [1]. A crucial test of the validity of the resulting path integral formula will come from applications to systems whose correct operator quantum mechanics is known

e,

773

774

10 Spaces with Curvature and Torsion

on the basis of symmetries and group commutation rules rather than canonical commutation rules. In contrast to earlier approaches, our path integral formula will always yield the same quantum mechanics as operator quantum mechanics quantized via group commutation rules. Our formula can, of course, also be used for an alternative approach to the path integrals solved before in Chapter 8, where a Euclidean space was parametrized in terms of curvilinear coordinates. There it gives rise to a more satisfactory treatment than before, since it involves only the intrinsic variables of the coordinate systems.

10.1

Einstein's Equivalence Principle

To motivate the present study we invoke Einstein's equivalence principle, according to which gravitational forces upon a spinless mass point are indistinguishable from those felt in an accelerating local reference.l They are independent of the atomic composition of the particle and strictly proportional to the value of the mass, the same mass that appears in the relation between force and acceleration, in Newton's first law . The strict equality between the two masses, gravitational and inertial, is fundamental to Einstein's equivalence principle. Experimentally, the equality holds to an extremely high degree of accuracy. Any possible small deviation can presently be attributed to extra non-gravitational forces. Einstein realized that as a consequence of this equality, all spinless point particles move in a gravitational field along the same orbits which are independent of their composition and mass. This universality of orbital motion permits the gravitational field to be attributed to geometric properties of spacetime. In Newton's theory of gravity, the gravitational forces between mass points are inversely proportional to their distances in a Euclidean space. In Einstein's geometric theory the forces are explained entirely by a curvature of spacetime. In general the spacetime of general relativity may also carry another geometric property, called torsion. Torsion is supposed to be generated by the spin densities of material bodies. Quantitatively, this may have only extremely small effects, too small to be detected by present-day experiments. But this is only due to the small intrinsic spin of ordinary gravitational matter. In exceptional states of matter such as polarized neutron stars or black holes, torsion can become relevant. It is now generally accepted that spacetime should carry a nonvanishing torsion at least locally at those points which are occupied by spinning elementary particles [55]. This follows from rather general symmetry considerations. The precise equations of motion for the torsion field , on the other hand, are still a matter of speculation. Thus it is an open question whether or not the torsion field is able to propagate into the empty space away from spinning matter. lQuotation from his original paper Uber das Relativitiitsprinzip und die aus demselben gezogenen Folgerungen , Jahrbuch der Relativitat und Elektonik 4, 411 (1907): "Wir . .. wollen daher im folgenden die v611ige physikalische Gleichwertigkeit von Gravitationsfeld und entsprechender Beschleunigung des Bezugssystems annehmen" .

775

10.2 Classical Motion of Mass Point in General Metric-Affine Space

Even though the effects of torsion are small we shall keep the discussion as general as possible and study the motion of a particle in a metric-affine space with both curvature and torsion. To prepare the grounds let us first recapitulate a few basic facts about classical orbits of particles in a gravitational field. For simplicity, we assume here only the three-dimensional space to have a nontrivial geometry.2 Then there is a natural choice of a time variable t which is conveniently used to parametrize the particle orbits. Starting from the free-particle action we shall then introduce a path integral for the time evolution amplitude in any metric-affine space which determines the quantum mechanics via the quantum fluctuations of the particle orbits.

10.2

Classical Motion of Mass Point in General Metric-Affine Space

On the basis of the equivalence principle, Einstein formulated the rules for finding the classical laws of motion in a gravitational field as a consequence of the geometry of spacetime. Let us recapitulate his reasoning adapted to the present problem of a nonrelativistic point particle in a non-Euclidean geometry.

10.2.1

Equations of Motion

Consider first the action of the particle along the orbit x(t) in a flat space parametrized with rectilinear, Cartesian coordinates:

(10.1) It is transformed to curvilinear coordinates ql', p,

= 1,2,3, via some functions (10.2)

leading to

(10.3) where

(10.4) is the induced metric for the curvilinear coordinates. Repeated indices are understood to be summed over, as usual. The length of the orbit in the flat space is given by

l=

l

t

b

dtvgl'v(q)ql'qv.

(10.5)

ta

2The generalization to non-Euclidean spacetime will be obvious after the development in Chapter 19.

776

10 Spaces with Curvature and Torsion

Both the action (10.3) and the length (10.5) are invariant under arbitrary reparametrizations of space ql-' ---> q'l-'. Einstein's equivalence principle amounts to the postulate that the transformed action (10.3) describes directly the motion of the particle in the presence of a gravitational field caused by other masses. The forces caused by the field are all a result of the geometric properties of the metric tensor. The equations of motion are obtained by extremizing the action in Eq. (10.3) with the result (10.6) Here 1 f,\vl-' == 2(a,\gvl-'

+ avg,\1-' -

al-'g,\v)

(10.7)

is the Riemann connection or Christoffel symbol of the first kind [recall (1. 70)]. With the help of the Christoffel symbol of the second kind [recall (1.71)] (10.8) we can write (10.9) The solutions of these equations are the classical orbits. They coincide with the extrema of the length of a line 1 in (10.5). Thus, in a curved space, classical orbits are the shortest lines, the geodesics [recall (1.72)]. The same equations can also be obtained directly by transforming the equation of motion from (10.10) to curvilinear coordinates ql-', which gives (10.11) At this place it is again useful to employ the quantities defined in Eq. (1.357), the basis triads and their reciprocals (10.12) which satisfy the orthogonality and completeness relations (1.358): (10.13)

10.2 Classical Motion of Mass Point in General Metric-Affine Space

777

The induced metric can then be written as (10.14) Labeling Cartesian coordinates, upper and lower indices i are the same. The indices /-1, v of the curvilinear coordinates, on the other hand, can be lowered only by contraction with the metric gl'v or raised with the inverse metric gl'V == (gl'v)-l. Using the basis triads, Eq. (10.11) can be rewritten as d. -(e' dt I' q'l')

.

.

= e' I' q"l' + av e' I' q'l'q'v =

0,

or as (10.15) The quantity in front of q"'i/' is called the affine connection: (10.16) Due to (10.13), it can also be written as [compare (1.366)] (10.17) Thus we arrive at the transformed flat-space equation of motion (10.18) The solutions of this equation are called the straightest lines or autoparallels. If the coordinate transformation functions Xi(q) are smooth and single-valued, their derivatives commute as required by Schwarz's integrability condition (10.19) Then the triads satisfy the identity (10.20) implying that the connection fA",Jt is symmetric in the lower indices. In fact, it coincides with the Riemann connection, the Christoffel symbol f' A/' This follows immediately after inserting gl'v(q) = eil'(q)eiv(q) into (10.7) and working out all derivatives using (10.20). Thus, for a space with curvilinear coordinates ql' which can be reached by an integrable coordinate transformation from a flat space, the autoparallels coincide with the geodesics.

778

10 Spaces with Curvature and Torsion

10.2.2

Nonholonornic Mapping to Spaces with Torsion

It is possible to map the x-space locally into a q-space with torsion via an infinitesimal transformation

dx i = e\,(q)dq".

(10.21 )

We merely have to assume that the coefficient functions ei,,(q) do not satisfy the property (10.20) which follows from the Schwarz integrability condition (10.19) :

o)..e\(q) - ol.i 0 ",v e >. = ,

(10.40)

10.2 Classical Motion of Mass Point in General Metric-Affine Space

781

This will be of use later. From either of the two curvature tensors, Rp,VA I< and Rp,VA"', one can form the once-contracted tensors of rank two, the Ricci tensor (10.41)

and the curvature scalar (10.42)

The celebrated Einstein equation for the gravitational field postulates that the tensor (10.43)

the so-called Einstein tensor , is proportional to the symmetric energy-momentum tensor of all matter fields . This postulate was made only for spaces with no torsion, in which case Rp,v = Rp,v and Rp,v , Gp,v are both symmetric. As mentioned before, it is not yet clear how Einstein's field equations should be generalized in the presence of torsion since the experimental consequences are as yet too small to be observed. In this text, we are not concerned with the generation of curvature and torsion but only with their consequences upon the motion of point particles. It is useful to set up two simple examples for nonholonomic mappings which illustrate the way in which these are capable of generating curvature and torsion from a Euclidean space. The reader not familiar with this subject is advised to consult a textbook on the physics of defects [2] . where such mappings are standard and of great practical importance; every plastic deformation of a material can only be described in terms of such mappings. As a first example consider the transformation in two dimensions for i for i with an infinitesimal parameter

E

= 1, = 2,

(10.44)

and the multi-valued function (10.45)

The triads reduce to dyads, with the components

(10.46)

and the torsion tensor has the components (10.47)

782

10 Spaces with Curvature and Torsion

mapping.

dislocated

ideal

Figure 10.1 Edge dislocation in crystal associated with missing semi-infinite plane of atoms. The nonholonomic mapping from the ideal crystal to the crystal with the dislocation introduces a O-function type torsion in the image space.

If we differentiate (10.45) formally, we find (0J1.011 - 01l0J1.)¢J incorrect at the origin. Using Stokes' theorem we see that

== O. This, however, is (10.48)

for any closed circuit around the origin, implying that there is a o-function singularity at the origin with e2"S12 "

t (2) = -27rO (q) . 2

(10.49)

By a linear superposition of such mappings we can generate an arbitrary torsion in the q-space. The mapping introduces no curvature. In defect physics, the mapping (10.46) is associated with a dislocation caused by a missing or additional layer of atoms (see Fig. 10.1). When encircling a dislocation along a closed path C, its counter image C f in the ideal crystal does not form a closed path. The closure failure is called the Burgers vector (10.50)

It specifies the direction and thickness of the layer of additional atoms. With the help of Stokes' theorem, it is seen to measure the torsion contained in any surface S spanned by C: (10.51) where d2SJ1.11 = _d2SIlJ1. is the projection of an oriented infinitesimal area element onto the plane J.LV. The above example has the Burgers vector (10.52) A corresponding closure failure appears when mapping a closed contour C in the ideal crystal into a crystal containing a dislocation. This defines a Burgers vector: (10.53)

10.2 Classical Motion of Mass Point in General Metric-Affine Space

783

ideal

Figure 10.2 Edge disclination in crystal associated with missing semi-infinite section of atoms of angle fl. The nonholonomic mapping from the ideal crystal to the crystal with the disclination introduces a 8-function type curvature in the image space.

By Stokes' theorem, this becomes a surface integral If'

Is d sij 8ie/' = Is d sij ei" 8 e/, -Is d sij ei"e/, Sv).'" 2

2

v

2

(10.54)

the last step following from (10.17). As a second example for a nonholonomic mapping, we generate curvature by the transformation (10.55)

with the multi-valued function (10.45). The symbol Levi-Civita tensor. The transformed metric

E,w

denotes the antisymmetric (10.56)

is single-valued and has commuting derivatives. The torsion tensor vanishes since (81 8 2 - 8281 )X 1,2 are both proportional to q2,1b'(2) (q) , a distribution identical to zero. The local rotation field w(q) == H81 x 2 - 82 x 1 ), on the other hand, is equal to the multi-valued function -D.4>(q) , thus having the noncommuting derivatives: (10.57)

To lowest order in D., this determines the curvature tensor, which in two dimensions possesses only one independent component, for instance R 1212 . Using the fact that g"v has commuting derivatives, R1212 can be written as (10.58)

In defect physics, the mapping (10.55) is associated with a disclination which corresponds to an entire section of angle D. missing in an ideal atomic array (see Fig. 10.2). It is important to emphasize that our multivalued basis tetrads ea ,,(q) are not related to the standard tetrads or vierbein fields h ~(q) used in the theory of gravitation with spinning particles. The difference is explained in Appendix lOB.

784

10.2.3

10 Spaces with Curvature and Torsion

New Equivalence Principle

In classical mechanics, many dynamical problems are solved with the help of nonholonomic transformations. Equations of motion are differential equations which remain valid if transformed differentially to new coordinates, even if the transformation is not integrable in the Schwarz sense. Thus we postulate that the correct equations of motion of point particles in a space with curvature and torsion are the images of the equation of motion in a fiat space. The equations (10.18) for the autoparallels yield therefore the correct trajectories of spinless point particles in a space with curvature and torsion. This postulate is based on our knowledge of the motion of many physical systems. Important examples are the Coulomb system which will be discussed in detail in Chapter 13, and the spinning top in the body-fixed reference system [3]. Thus the postulate has a good chance of being true, and will henceforth be referred to as a new equivalence principle.

10.2.4

Classical Action Principle for Spaces with Curvature and Torsion

Before setting up a path integral for the time evolution amplitude we must find an action principle for the classical motion of a spinless point particle in a space with curvature and torsion, i.e., the movement along autoparallel trajectories. This is a nontrivial task since autoparallels must emerge as the extremals of an action (10.3) involving only the metric tensor gil'" The action is independent of the torsion and carries only information on the Riemann part of the space geometry. Torsion can therefore enter the equations of motion only via some novel feature of the variation procedure. Since we know how to perform variations of an action in the Euclidean x-space, we deduce the correct procedure in the general metric-affine space by transferring the variations c5Xi(t) under the nonholonomic mapping (10.59) into the qll-space. Their images are quite different from ordinary variations as illustrated in Fig. 10.3(a). The variations of the Cartesian coordinates c5Xi(t) are done at fixed endpoints of the paths. Thus they form closed paths in the x-space. Their images, however, lie in a space with defects and thus possess a closure failure indicating the amount of torsion introduced by the mapping. This property will be emphasized by writing the images c5Sqll(t) and calling them nonholonomic variations. The superscript indicates the special feature caused by torsion. Let us calculate them explicitly. The paths in the two spaces are related by the integral equation (10.60)

10.2 Classical Motion of Mass Point in General Metric-Affine Space

785

For two neighboring paths in x-space differing from each other by a variation bXi(t), equation (10.60) determines the nonholonomic variation b8 qJl(t) : (10.61) A comparison with (10.59) shows that the variation b8 and the time derivatives d/dt of ql"(t) commute with each other: (10.62) just as for ordinary variations bxi : .

d.

b:i:'(t) = dtbx'(t).

(10.63)

Let us also introduce auxiliary nonholonomic variations in q-space: (10.64)

In contrast to b8 ql"(t), these vanish at the endpoints, (10.65) just as the usual variations bXi(t) , i.e., they form closed paths with the unvaried orbits. Using (10.62), (10.63), and the fact that b8 x i (t) == bXi(t), by definition, we derive from (10.61) the relation

b8 et(q(t)W(t)

+ eiJl(q(t)) :tbXi(t)

b8 eil"(q(t)W(t)

+ et(q(t)) ![e\,(t)oqV(t)].

(10.66)

After inserting d i ( ) _ dt ev q -

r AV I"'A q ei 1'"

(10.67)

this becomes (10.68) It is useful to introduce the difference between the nonholonomic variation b8 ql" and an auxiliary closed nonholonomic variation bqJl:

(10.69)

786

10 Spaces with Curvature and Torsion

Then we can rewrite (10.68) as a first-order differential equation for 8s b":

!8Sb" = - f AV"8 Sb\,{ + 2SAv"il3 qv.

(10.70)

After introducing the matrices (10.71) and (10.72) equation (10.70) can be written as a vector differential equation: (10.73) Although not necessary for the further development , we solve this equation by

8Sb(t)

=

it

~(t') 3q(t'),

(10.74)

[-l dt"G(t")].

(10.75)

dt'U(t, t')

ta

with the matrix

U(t ,t') = Texp

In the absence of torsion, ~(t) vanishes identically and 8Sb(t) == 0, and the variations 8Sq"(t) coincide with the auxiliary closed nonholonomic variations 8q"(t) [see Fig. 1O.3(b)]. In a space with torsion, the variations 8Sq"(t) and 3q"(t) are different from each other [see Fig. 1O.3(c)]. Under an arbitrary nonholonomic variation 8Sq"(t) = 8q" + 8s b" , the action (10.3) changes by

8s A = M

lb

dt (g"vil8Sil

+ ~o"gA,,8Sq"qAq") .

(10.76)

After a partial integration of the 8q-term we use (10.65), (10.62), and the identity O"gvA == f"VA + f"AV' which follows directly form the definitions g"v == ei"ei v and f "V A == e/o"e i v, and obtain

lb

8SA = M

dt [-g"V (ij" + f\,," qAq") 3q" + (g"Vqv

!

8s b" + f "A,,8 Sb" qAq") ] . (10.77)

To derive the equation of motion we first vary the action in a space without torsion. Then 8S b"(t) == 0, and (10.77) becomes (10.78)

787

10.2 Classical Motion of Mass Point in General Metric-Affine Space

(c)

Figure 10.3 Images under holonomic and nonholonomic mapping of fundamental 8function path variation. In the holonomic case, the paths x(t) and x(t) +8x(t) in (a) turn into the paths q(t) and q(t) + oq(t) in (b). In the nonholonomic case with S~" =I- 0, they go over into q(t) and q(t) + 8S q(t) shown in (c) with a closure failure bI'- at tb analogous to the Burgers vector bit in a solid with dislocations. Thus, the action principle JS A = 0 produces the equation for the geodesics (10.9), which are the correct particle trajectories in the absence of torsion. In the presence of torsion, JSlJIL is nonzero, and the equation of motion receives a contribution from the second parentheses in (10.77). After inserting (10.70) , the nonlocal terms proportional to JSlJIL cancel and the total nonholonomic variation of the action becomes

(10.79)

r

The second line follows from the first after using the identity r A' so that the integration in each time slice (tn ' tn- I) with n = N + 1, N, . . . runs over dX~_ I . In a flat space parametrized with curvilinear coordinates, the transformation of the integrals over dDX~_ 1 into those over dDq~_1 is obvious:

rr:;=l

rr:;il

112 JdDx~_l = g {J dDq~_1 det [e~(qn-l)l}·

N+ I

N+ I

(10.122)

795

10.3 Path Integral in Metric-Affine Space

The determinant of ei ll is the square root of the determinant of the metric gilI': (10.123) and the measure may be rewritten as (10.124) This expression is not directly applicable. When trying to do the dD q~_l-integrations successively, starting from the final integration over dqiJv , the integration variable qn- l appears for each n in the argument of det [e~(qn- l)] or gllv(qn- l). To make this qn_l-dependence explicit, we expand in the measure (10.122) e~(qn- l) = eill(qntl.qn) around the postpoint qn into powers of tl.qn- This gives dx i =

e~(q -

tl.q)dqll =

e~dqll -

eill,vdqlltl.qV

+ ~eill,vAdqlltl.qV tl.qA + ...

(10.125)

omitting, as before, the subscripts of qn and tl.qn . Thus the Jacobian of the coordinate transformation from dx i to dqll is (10.126) giving the relation between the infinitesimal integration volumes dDxi and dDqll:

If JdDX~_l If {J dDq~_l =

n=2

Jon}.

(10.127)

n=2

The well-known expansion formula det (1 + B)

=

exptrlog(1 + B)

=

exptr(B - B2/2 + B 3 /3 - . . .)

(10.128)

allows us now to rewrite J o as (10.129) with the determinant det (e~) = Jg(q) evaluated at the postpoint. This equation defines an effective action associated with the Jacobian, for which we obtain the expansion

1ii AfJo

-_

1\ Il+ 1 [ei Il e/",VA i 1\ vI....l.q 1\ A + . .. . (10 .130) -ei I< ei 1:'j bJ.Lv + xJ.LXV /r2 and the Christoffel symbols (10.7) are fJ.L/' >:'j fJ.Lv'x >:'j bJ.L vx,X/r 2. Inserting this into (10.35) we obtain the curvature tensor for small xi-':

Since

(J-l

=

(10.161) This can be extended covariantly to the full surface of the sphere by replacing bJ.L,X by the metric gJ.L'x(x): -

1

RJ.Lv'x,,(x) = 2" [gJ.L,,(x)gv,X(x) - gJ.L,X(x)gv,,(x)] , r

(10.162)

so that Ricci tensor is [recall (10.41)] -

-

Rv,,(x) = RJ.Lv"i-'(x) =

D'-l

- - 2-gv,,(X)'

r

(10.163)

Contracting this with gV" [recall (10.42)] yields indeed the curvature scalar (10.156). The effective potential (10.152) is therefore

n? v"ff = - 6Mr 2 (D - 2)(D - 1) .

(10.164)

It supplies precisely the missing energy which changes the energy (8.224) near the sphere, corrected by the expectation of the quartic term 19! in the action, to the proper value El

n? = 2M r 2l(l + D - 2).

(10.165)

Astonishingly, this elementary result of Schrodinger quantum mechanics was found only a decade ago by path integration [5] . Other time-slicing procedures yield extra terms proportional to the scalar curvature R, which are absent in our theory. Here the scalar curvature is a trivial constant, so it would remain undetectable in atomic experiments which measure only energy differences. The same result will be derived in general in Eqs. (11.25) and in Section 11.3. An important property of this spectrum is that the ground state energy vanished for all dimensions D. This property would not have been found in the naive measure of path integration on the right-hand side of Eq. (10.147) which is used in most works on this subject. The correction term (10.151) coming from the nonholonomic mapping of the flat-space measure is essential for the correct result.

802

10 Spaces with Curvature and Torsion

More evidence for the correctness of the measure in (10.146) will be supplied in Chapter 13 where we solve the path integrals of the most important atomic system, the hydrogen atom. We remark that for t -+ t', the amplitude (10.153) shows the states Iq) to obey the covariant orthonormality relation (10.166) The completeness relation reads (10.167)

10.5

External Potentials and Vector Potentials

An important generalization of the above path integral formulas (10.146) , (10.149) , (10.153) of a point particle in a space with curvature and torsion includes the presence of an external potential and a vector potential. These allow us to describe, for instance, a particle in external electric and magnetic fields. The classical action is then (10.168) To find the time-sliced action we proceed as follows. First we set up the correct time-sliced expression in Euclidean space and Cartesian coordinates. For a single slice it reads, in the postpoint form,

A €=

M

.

-(~x')

2f

2

e

.

+ -Ai(X)~X' C

e 2c

-Ai

.

.

·(x)~x'~xJ

,J

- fV(X)

+ .. .

(10.169)

As usual, we have neglected terms which do not contribute in the continuum limit. The derivation of this time-sliced expression proceeds by calculating, as in (10.116), the action

A€ =

1~€ dtL(t)

(10.170)

along the classical trajectory in Euclidean space, where

L(t) = M x 2 (t) + ~A(x(t))x(t) - V(x(t)) 2

c

(10.171)

is the classical Lagrangian. In contrast to (10.116), however, the Lagrangian has now a nonzero time derivative (omitting the time arguments): (10.172)

10.5 External Potentials and Vector Potentials

803

For this reason we cannot simply write down an expression such as (10.117) but we have to expand the Lagrangian around the post point leading to the series

A< =

i

d

1

t

t-
1 taken at the end.

10.11.1

Diagrammatic Elements

The perturbation expansion for an evolution amplitude over a finite imaginary time proceeds as described in Section 10.6, except that the free energy in Eq. (10.201) becomes [recall (2.524)]

D

D

2(3 Tr log( -fP + w2 ) = 2(3

L log(w~ + w

2)

n

D . 7i log[2smh(n(3w/2)].

(10.389)

As before, the diagrams contain four types of lines representing the correlation functions (10.202)(10.191). Their explicit forms are, however, different. It will be convenient to let the frequency w in the free part of the action (10.187) go to zero. Then the free energy (10.389) diverges logarithmically in w. This divergence is, however, trivial. As explained in Section 2.9, the divergence is removed by replacing w by the length of the q-axis according to the rule (2.359) . For finite time intervals, the correlation functions are no longer given by (10.207) which would not have a finite limit for w -> O. Instead, they satisfy Dirichlet boundary conditions, where we can go to w = 0 without problem. The finiteness of the time interval removes a possible infrared divergence for w -> O. The Dirichlet boundary conditions fix the paths at the ends of the time interval (0, (3) making the fluctuations vanish, and thus also their correlation functions:

G~2J (0, 7') = G~2J ((3, 7') = 0,

G~2J (7,0) = G~2J (7, (3) = O.

(10.390)

The first correlation function corresponding to (10.206) is now

G~2J(7,7') = 8I'vLl(7,7') =

(10.391)

where

Ll(7,7') = Ll(7',7) =

~((3 -

7»7< =

~ [- «7 - 7')(7 - 7') +7+7'] -

7;',

(10.392)

abbreviates the Euclidean version of Go(t, t') in Eq. (3.39). Being a Green function of the free equation of motion (10.213) for w = 0, this satisfies the inhomogeneous differential equations '~(7,

7') = LX'(7, 7') = - 8(7 - 7'),

(10.393)

10.11 Perturbative Calculation on Finite Time Intervals

837

by analogy with Eq. (10.214) for w = O. In addition, there is now an independent equation in which the two derivatives act on the different time arguments: (lO.394) For a finite time interval, the correlation functions (lO.203) (10.204) differ by more than just a sign [recall (10.210)] . We therefore must distinguish the derivatives depending on whether the left or the right argument are differentiated. In the following, we shall denote the derivatives with respect to TorT' by a dot on the left or right, respectively, writing

.~ (T,T') == d~ ~(T,T'),

(10.395)

Differentiating (lO.392) we obtain explicitly

The discontinuity at T = T' which does not depend on the boundary condit ion is of course the same as before, The two correlation functions (lO.208) and (10.210) and their diagrammatic symbols are now

8TG~2J(T, T')

== (ti!'(T)qv(T') ) =8!,v ·~(T, T') 8T,G~2J(T, T') == (q!'(T)tiv(T') ) = 8!'v SeT, T')

(lO.397) (lO.398)

The fourth correlation function (lO.212) is now

8T8T,G~2J(T,T') = 8!'v·S(T,T') =

(10.399)

with ·S (T, T') being given by (lO.394). Note the close similarity but also the difference of this with respect to the equation of motion (10.393) .

10.11.2

Cumulant Expansion of D-Dimensional Free-Particle Amplitude in Curvilinear Coordinates

We shall now calculate the partition function of a point particle in curved space for a finite time interval. Starting point is the integral over the diagonal amplitude of a free point particle of unit mass (xa,BlxaO) in flat D-dimensional space (10.400) with the path integral representation

(xa,BlxaO)o =

J

VDxe-A(O)[Xi ,

(10.401)

r dTX2(T).

(lO.402)

where A(O) [x ] is the free-particle action A(O)[x] =

~

2

f3

Jo

Performing the Gaussian path integral leads to

(xa,BlxaO)o =

e-(D/2)'Irlog(-82 ) =

[21l",6] - D/2 ,

(10.403)

where the trace of the logarithm is evaluated with Dirichlet boundary conditions. The result is of course the D-dimensional imaginary-time version of the fluctuation factor (2.125) in natural units.

10 Spaces with Curvature and Torsion

838

A coordinate transformation Xi(T) = xi(ql'(T)) mapping Xa to q!:: brings the action (10.402) to the form (10.186) with V(q(T)) = 0:

_ I1f3 'I"V' _ 8Xi(q) 8Xi(q) A[q]- dTgl'v(q(T))q (T)q (T), with gl'v(q) = - 8 - - 8 - ' 2 0 ql' qV

(10.404)

In the formal notation (10.189), the measure transforms as follows:

J

VDX(T) ==

II JdDx(T) = J II JdDq(T) == J JVDq Jg(qa), T

(10.405)

T

where g(q) == detgl'v(q) and J is the Jacobian of the coordinate transformation generalizing (10.220) and (10.219)

8X

)=~_~(T, T) 4 /3 '

(10.435)

whose integrals yield - 1/3 2

1{3 0

dT~(T,T) =

1{3

/3 12

dT ·~2(T, T) = - .

0

(10.436)

Let us evaluate the second-order diagrams in /3f~i) , i = 1,2, 3,4. The sum of the local diagrams in (10.429) consists of the integrals by (10.437)

Replacing ·S(T, T) in Eq. (10.437) again by 8(0) - 1//3, on account of the equation of motion (10.394), and taking into account the right-hand equation (10.435),

/3f~l) = rl [38(0) 1{3 dT~2(T'T)] = 'T]2~~ 8(0) .

(10.438)

We now calculate the sum of bubble diagrams (10.430)- (10.432) , beginning with (10.430) whose analytic form is (10.439)

Inserting Eq. (10.394) into the last equal-time term, we obtain (10.440)

As we shall see below, the explicit evaluation of the integrals in this sum is not necessary. Just for completeness, we give the result:

(10.441)

We now turn to the three-bubbles diagrams (10.432) . Only three of these contain the correlation function /L~v(x, x') -+ ·S (T, T') for which Eq. (10.394) is not applicable: the second, fourth ,

843

10.11 Perturbative Calculation on Finite Time Intervals

and sixth diagram. The other three-bubble diagrams in (10.432) containing the generalization I'ill'(x, x) of the equal-time propagator 'S(r, r) can be calculated using Eq. (10.394) . Consider first a partial sum consisting of the first three three-bubble diagrams in the sum (10.432) . This has the analytic form

f3fJ3) I

=

1,2,3

+

_'!L21f31f3 drdr' {4il(r,r) 'il 2(r,r') 'S(r' , r') 2 0 0 2 'S(r, r)il 2(r, r')"S(r' , r')

(10.442)

+ 16'il (r, r)il(r, r') 'il (r, r') 'S(r', r')} .

Replacing 'S(r, r) and 'S(r',r') by 8(0) -1/13, according to of (10.394), we see that Eq. (10.442) contains, with opposite sign, precisely the previous sum (10.439) of all one-and two-bubble diagrams. Together they give

f3fJ2)

+ f3fJ3) I

= - 1722 rf3 rf3 drdr' { _ _ f34 il(r,r)"il 2(r,r')

1,2 ,3

+

;2

Jo Jo

il 2(r, r') -

~ 'il (r, r)il(r, r')'il (r, r')} ,

(10.443)

and can be evaluated directly to (10.444) By the same direct calculation, the Feynman integral in the fifth three-bubble diagram in (10.432) yields

aX): Is

f3 f3 Jro Jor drdr'·il(r,r)"il(r, r')S(r, r')S(r',r')=-~. 720

=

(10.445)

The explicit results (10.444) and (10.445) are again not needed, since the last term in Eq. (10.443) is equal, with opposite sign, to the partial sum of the fourth and fifth three-bubble diagrams in Eq. (10.432) . To see this, consider the Feynman integral associated with the sixth three-bubble diagram in Eq. (10.432): (10.446) whose d-dimensional extension is

1%

=

l f3 1f3 ddrddr ' "il(r, r)il(r, r')"ilf3(r, r')ilf3(r',r') .

(10.447)

Adding this to the fifth Feynman integral (10.445) and performing a partial integration, we find in one dimension

where we have used d['il(r, r)]/dr = - 1/13 obtained by differentiating (10.435) . Comparing (10.448) with (10.443), we find the sum of all bubbles diagrams, except for the sixth and seventh three-bubble diagrams in Eq. (10.432) , to be given by

13/2) + f3f(3) 2

2

I'

6,7

= 17

2

~f32.

2 15

(10.449)

844

10 Spaces with Curvature and Torsion

The prime on the sum denotes the exclusion of the diagrams indicated by subscripts. The correlation function 'S (T, T') in the two remaining diagrams of Eq. (10.432) , whose d-dimensional extension is ct ~,,(x, x') , cannot be replaced via Eq. (10.394), and the expression can only be simplified by applying partial integration to the seventh diagram in Eq. (10.432) , yielding

OX) : h -->

1" 1"dTdT'~(T,T)"~(T,T')"S(T, T')~' (T', T') 1"1" ~(T, T)ct~(T, T')ct~,,(T, T')~,,(T',T') ddT ddT'

~ 2

-->

r" r"ddT ddT' ~(T,T)~,,(T',T')8~ [ct~(T, T')l2

Jo Jo

~1" 1(3 2

1"dTdT' ~(T' T)"~(T"T')d~' ['~2(T, T')]

r" r"dTdT'~(T, T)"~2(T, T')

Jo Jo

= (32 .

(10.450)

90

The sixth diagram in the sum (10.432) diverges linearly. As before, we add and subtract t he divergence

0::0 : h

1" 1" 1"1" + 1"1"

dTdT'

~(T, T)"S2(T, T')~(T', T')

dTdT' ~(T, T) ['S2(T,T') - 82(T - T')] dTdT'

~(T', T')

~2(T, T)82(T-T').

(10.451)

In t he first , finite term we go to d dimensions and replace 8(T - T') --> 8(d)(T - T') = -~",, (T, T') using the field equation (10.418) . After this, we apply partial integration and find

If

-->

1"1" ~(T, T) [ct~~(T, T') ~;·k, T')] ~(T', T') 1"1" {-8ct[~(T, T)l~,,(T, T')ct~,,(T,T')~(T', T') 1"1"dTdT'2{-'~(T, T)S(T, T')"S(T, T')~(T', T')+ ddT ddT'

-

ddT ddT'

+ ~(T, T)~,, (T, T')~1'1' (T, T')8/3 [~(T', T'll)

-->

~(T, T)

S(T, T')"~ (T' , T') t1'(T, T')}.

(10.452)

In going to t he last line we have used d [~(T,T) l/dT = 2'~(T, T) following from (10.435) . By interchanging the order of integration T T' , the first term in Eq. (10.452) reduces to the integral (10.450). In the last term we replace t1'(T,T') using the field equation (10.393) and the trivial equation (10.376) . Thus we obtain

h =If + Itv

(10.453)

with

1R 6 16d iv

((32 (32) 1 ( 7(32) 2 - 90 - 120 = 2 -90

1"1"

dTdT'

~2(T, T) 82(T -

'

(10.454)

T') .

(10.455)

845

10.11 Perturbative Calculation on Finite Time Intervals

With the help of the identity for distributions (10.369), the divergent part is calculated to be

Ig iv

=

8(0)

[3

dT I~?(T, T) = 8(0)

~~.

(10.456)

Using Eqs. (10.450) and (10.453) yields the sum of the sixth and seventh three-bubble diagrams in Eq. (10.432) : (3) I

(3f2

=

6,7

1)2

1)2 [

-2 (2h + 1617) = -2

(33 28(0) 30

(32]

+ 10

.

(10.457)

Adding this to (10.449), we obtain the sum of all bubble diagrams

+ (3fJ3)

(3fJ2)

= -

~ [28(0)~~ + ~~] .

(10.458)

The contributions of the watermelon diagrams (10.432) correspond to the Feynman integrals

(3fJ4) = -21)2

+

1(31(3 dTdT' [~2(T,T')"S2(T, T')

4~(T, T')·~ (T,T') S(T,T')"S(T, T') + .~ 2(T, T') S2(T, T')J .

(10.459)

The third integral is unique and can be calculated directly:

() : ho

=

r(3 dT r(3 dT'

Jo

Jo

·~ 2(T, T') S2(T, T') =

(32 . 90

(10.460)

The second integral reads in d dimensions

0:

19

=

JJ

ddT ddT' ~(T, T')",~(T,T')~(3(T, T')",~(3(T, T').

(10.461)

This is integrated partially to yield, in one dimension,

19 =

- ~ho + 19' == - ~ho - ~

JJdTdT'~(T,T')S2(T, T')"lJ.(T, T').

(10.462)

The integral on the right-hand side is the one-dimensional version of 19,

=

-~

JJ

ddT ddT' ~(T,T')~~(T, T')

",,,,~(T,T').

(10.463)

Using the field equation (10.418) , going back to one dimension, and inserting ~(T,T'), S(T,T'), and .lJ. (T, T') from (10.392), (10.396), and (10.393), we perform all unique integrals and obtain

19' = (32

{4~

J

dT€2(T) 8(T)

+ 2~0} .

(10.464)

According to Eq. (10.375) , the integral over the product of distributions vanishes. Inserting the remainder and (10.460) into Eq. (10.462) gives:

19

(32 = - -.

720

(10.465)

We now evaluate the first integral in Eq. (10.459). Adding and subtracting the linear divergence yields

0 : Is = 1(31(3 dTdT' ~2(T,T')"S2(T,T') = 1(31(3 dTdT'~2(T,T)82(T-T') +

1(31(3 dTdT'~2(T,T') [·S2(T, T')-8 2(T-T')J.

(10.466)

846

10 Spaces with Curvature and Torsion

The finite second part of the integral (10.466) has the d-dimensional extension (10.467) which after partial integration and going back to one dimension reduces to a combination of integrals Eqs. (10.465) and (10.464) : R

18 = - 219

The divergent part of Is coincides with

+ 219,

(32

= - 72·

(10.468)

Ig iv in Eq. (10.455):

Ig iv = 1{31{3 dTdT'l:!>.2(T,T)82(T -

T') =

Ig iv = 8(0)~~ .

(10.469)

Inserting this together with (10.460) and (10.465) into Eq. (10.459), we obtain the sum of watermelon diagrams (10.470) For a flat space in curvilinear coordinates, the sum of the first-order diagrams vanish. To second order, the requirement of coordinate independence implies a vanishing sum of all connected diagrams (10.429)- (10.432) . By adding the sum of terms in Eqs. (10.438), (10.458), and (10.470), we find indeed zero, thus confirming coordinate independence. It is not surprising that the integration rules for products of distributions derived in an infinite time interval T E [0,00) are applicable for finite time intervals. The singularities in the distributions come in only at a single point of the time axis, so that its total length is irrelevant. The procedure can easily be continued to higher-loop diagrams to define integrals over higher singular products of E- and 8-functions. At the one-loop level, the cancellation of 8(0)s requires

J

dT l:!>.(T, T) 8(0) = 8(0)

J

dT l:!>.(T, T) .

(10.471)

The second-order gave, in addition, the rule (10.472) To n-order we can derive the equation (10.473) which reduces to (10.474) which is satisfied due to the integration rule (10.369). See Appendix lOC for a general derivation of (10.473).

10.11.5

Time Evolution Amplitude in Curved Space

The same Feynman diagrams which we calculated to verify coordinate independence appear also in the perturbation expansion of the time evolution amplitude in curved space if this is performed in normal or geodesic coordinates.

847

10.11 Perturbative Calculation on Finite Time Intervals

The path integral in curved space is derived by making the mapping from xi to q!:: in Subsection 10.11.2 nonholonomic, so that it can no longer be written as xi(r) = xi(q"'(r)) but only as dxi(r) = ei ,..(q)dql'(r). Then the q-space may contain curvature and torsion, and the result of the path integral will not longer be trivial one in Eq. (10.403) but depend on Rl'v>."'(qa) and S,..v>'(qa). For simplicity, we shall ignore torsion. Then the action becomes (10.404) with the metric 9,..v(q) = ei,..(q)eiv(q). It was shown in Subsection 10.3.2 that under nonholonomic coordinate transformations, the measure of a timesliced path integral transforms from the flat-space form I1ndDxn to I1ndDqV9nexp(~tRn/6) . This had the consequence, in Section 10.4, that the time evolution amplitude for a particle on the surface of a sphere has an energy (10.165) corresponding to the Hamiltonian (1.414) which governs the Schriidinger equation (1.420). It contains a pure Laplace-Beltrami operator in the kinetic part. There is no extra R-term, which would be allowed if only covariance under ordinary coordinate transformations is required. This issue will be discussed in more detail in Subsection 11.1.1. Below we shall see that for perturbatively defined path integrals, the nonholonomic transformation must carry the flat-space measure into curved space as follows: (10.475)

For a D-dimensional space with a general metric 9,..v(q) we can make use of the above proven coordinate invariance to bring the metric to the most convenient normal or geodesic coordinates (10.98) around some point qa. The advantage of these coordinates is that the derivatives and thus the affine connection vanish at this point. Its derivatives can directly be expressed in terms of the curvature tensor: (10.476)

Assuming qa to lie at the origin, we expand the metric and its determinant in powers of normal coordinates ~~I' around the origin and find, dropping the smallness symbols ~ in front of q and ~ in the transformation (10.98): (10.477) (10.478)

These expansions have obviously the same power content in ~,.. as the previous one-dimensional expansions (10.223) had in q. The interaction (10.410) becomes in normal coordinates, up to order T/ 2 :

rf3 dr {[T/ 61 R- ,..>'1'''''''' (:>'(:'" + 1 R0R(:>. (:'" (:U(:T] i:1'i:v T/ 45 >'1'''' UVTO.,.,.,., .,., 2

) 0

+T/

~8(0)Rl'v e~v + T/ 2 1~08(0)Rl'ov

U

R>.u", 0 e~ve~"'}.

(10.479)

This has again the same powers in ~I' as the one-dimensional interaction (10.425) , leading to the same Feynman diagrams, differing only by the factors associated with the vertices. In one dimension, with the trivial vertices of the interaction (10.425) , the sum of all diagrams vanishes. In curved space with the more complicated vertices proportional to Rl'v",>, and Rl'v , the result is nonzero but depends on contractions of the curvature tensor Rl'v",>,. The dependence is easily identified for each diagram. All bubble diagrams in (10.430)-(10.432) yield results proportional to R~v, while the watermelon-like diagrams (10.432) carry a factor R~v",>..

848

10 Spaces with Curvature and Torsion

When calculating the contributions of the first expectation value to the time evolut ion amplitude it is useful to reduce the D-dimensional expectation values of (10.479) to one-dimensional ones of (10.425) as follows using the contraction rules (8.63) and (8.64): 8fJ.V (~~),

(e'c)

(10.480)

. T) 2" >. T/1- q q c'

u

(10.538)

the diagrams (10.490) corresponding to the analytic expression [compare (10.491)]

(3jP) = -

~O>.f ,,{/1-V}l(3dT{g/1-Vg">'·S(T, T)b,(T , T)+2gl'''gv>' .b,2(T, T)-8(0)g/1- Vg">'b,(T, T)}

~g>''' (f>.v/1-f/1-"v + gTl'fT"Vf/1->'v) 8(0) l{3dTb,(T, T) .

(10.539)

Replacing ·b,·(T, T) by 8(0) -1/(3 according to (10.394), and using the integrals (10.436), the 8(0)terms in the first integral cancel and we obtain (10.540)

856

10 Spaces with Curvature and Torsion

In addition, there are contributions of order", from the second cumulant

- ~ lf3 dT lf3 dT' \ [f"/Lvq"(TW-'(T)qV(T) + 6(0)f /L,,/Lq"(T)] x [f"/Lvq"(T')q/L(T'W(T')

+ 6(0)f/L,,/Lq"(T')]

le'

(10.541)

These add to the free energy (2) - - '!!.gA" f /L f v [ :'-'~'-'" fJf1 -2 A/L"V .......... - 26(0) (.'~.'.""""" .

fJfi 3) =

-

",fA/L/L (i'" f "vV

+ gV" f V" A) [ (j-.(.:; -

fJfi 4 ) = - ~(g/LA g""f/LAvf""v

~(g/L" gVA f /LA" f "f3 + 3g/L"f /LA" f"" A)

fJf 1(6)

'" g A" (f /L f 2 AV /L" v + 9 /L" f "" v f /LAV ) ~ ~ .

-

6(0)

6 2 (0)

C;-.•

+ g/LVf/L,,"fvAA +2g/LVf/Lv"f"AA)

fJfi 5 ) = -

+

l,

u. () ,

(), (10.542)

The Feynman integrals associated with the diagrams in the first and second lines are 111 =

JJ

dTdT' fS(T , T)~(T,T')'S(T',T')-26(0) 'S(T, T)

~(T, T')+6 2(0) ~(T,T')} (10.543)

and

h2

=

JJ

dTdT'

r~(T, T)'~(T, T')'S(T' , T') -6(0)'~(T,T)'~(T, T')},

(10.544)

respectively. Replacing in Eqs. (10.543) and (10.544) '~'(T,T) and 'S(T',T') by 6(0) -l/fJ leads to cancellation of the infinite factors 6(0) and 6 2 (0) from the measure, such that we are left with 1 111 = fJ2

lf3 dT lf3 dT' ~(T,T') = o

0

-fJ

(10.545)

12

and

h2= _ _fJ1

rf3 dT Jorf3 dT'·~(T,T)'~(T,T')=- !!... .

Jo

12

(10.546)

The Feynman integral of the diagram in the third line of Eq. (10.542) has d-dimensional extension

h3

= ->

JJdTdT"~(T, T)S(T',T')'S(T,T') JJddXddX'/L~(x, x)~v(x',x')/L~v(x,x').

(10.547)

Integrating this partially yields

h3

=

~

JJ

dTdT'

S(T,T')'~(T', T') = ~ lf3 dT lf3 dT"~(T, T)'~(T,T') = ~, (10.548)

857

10.12 Effective Classical Potential in Curved Space

where we have interchanged the order of integration T .r T A I-' A A 0" 3Tei [T,O"jek T ek V,I-'L..l.q L..l.q L..l.q - 3Te >. O"T I-'V L..l.q L..l.q L..l.q , V

X

V

(11.58)

which vanishes if the q-space has no torsion. The Jacobian associated with (11.56) is

~x) J = 8( 8(~q)

=

det (i) I< i A V 1 I< i A V A >. ) el< det (SCI< u 1-'- ei e{l-',v}L..l.q + 2ei e{l-',v>.}L..l.q L..l.q +...

, ( 11.59)

and corresponds to the effective action

h,i A '} j ] A A >. + . . . . (11 .60) - ei I-' e{l.q ) -

#Q) V27riEh)M

D

[i () "

exp 1i,gJ1.V q L..:>.qJ1." L..:>.q v] ,

(11.68)

904

11 Schrodinger Equation in General Metric-Affine Spaces

the moment properties (11.64)- (11.66) are equivalent to (11.69)

0+ .. . ,

. nr

zf 2M Jt

Jtv

+ ... ,

(11. 70)

where the expectation values are taken with respect to the kernel KO(!::"q), the dots on the right-hand side indicating terms of the order f2. Note that the third of the three moment properties is trivially true since it receives only a contribution from the leading part of the kernel K.\(!::,.qv!::,.ql'« from (11.41), and working out the contractions. To calculate the V2 B -terms, it is useful to introduce the notation r 11'- == r I'-V v, r 21'- == rvl'-v, r 31'- == r VV I'- and similarly the matrices i\I'-~(rl'-h«, f'L~(r 1'-)«>" f'21'-~r>.I'-«' f'L~r «1'->" f'31'-~r>.«I'-' f'rl'-~r «>'1'-" For contractions such as r1l'-r11'- we write r1r1, and for r I'-v>.r I'-v>. we write f'1f'1 = f'2f'2 = f'3f'3, whichever is most convenient. Similarly, r I'-v>.r>.l'-v = f'1f'2 T = f'2f'3 T = f' 3f' 1. Then we work out 1 2 -T -T -S[(r1 + r 2) - r 3(r1 + r 1 + r 2 + r 2 )],

(l1A.8)

1 - -T S [r 3r 3 + r 3(r3 + r 3 )], 1

2

4" [(r1 +r2) +r3(r1 +r2)], 1 2 2 2 -S[r1 +r2 +r3 +2(r1r2+r2r3 +r3rd

+f'3(f'1 + f'f + f'2 + f'I + r3 + f'D]· It is easy to check that the sum of these V2 B-terms vanishes. Incidentally, if the symmetrizations in (11.49) following from our Jacobian action had been absent, we would find the additional terms

(l1A.9) (l1A.10) whose sum yields the additional contribution to the V3B-terms ~V3 =

12 2- -R - -a 81'- + -r381 6 3 I'3 '

(l1A.11)

916

11 Schrodinger Equation in General Metric-Affine Spaces

after having used the identity (l1A.12) The first term in (l1A.ll) is the R-term derived by K.S. Cheng6 as an effective potential in the Schr6dinger equation. For V2 B , we would find the extra terms 1

1

- -

1

2

-

-T

-

-

-T

boV2 = - 2(f1f1 - f3f2)+S [(f1+f2) - f3(f1+f1 + f 2 + f 2 )],

(l1A.13)

(l1A.14) which add up to 1 1 1- 1-boV2 = -28181 + 2f381 - Sf38 1 + 28183,

(l1A.15)

where we have written 8 1 for 8" and used some trivial identities such as

I'3I'2 T = I'3i\ .

(l1A.16)

Thus we would obtain an additional effective potential Veff = (1i 2/ M)v with

v=

12//

"6R -

sg" 8,,8// -

12

1--

I- -

2(81 - f 381) + 28183 - Sf381.

(l1A.17)

The second and the fourth term can be combined to

2

1

- -D 8" - - f 3 8 1 . 3" 6

(l1A.18)

Due to the presence off's in v, this is a noncovariant expression which cannot possibly be physically correct. In the absence of torsion, however, v happens to be reparametrization-invariant, and this is the reason why the resulting effective potential Veff = 1i2R/6M appeared acceptable in earlier works. A procedure which has no reparametrization-invariant extension to spaces with torsion cannot be correct.

Appendix lIB

DeWitt's Amplitude

Bryce DeWitt, in his frequently quoted paper,7 attempted to quantize the motion of a particle in a curved space using the naive measure of path integration as in Eq. (10.153), but with the short-time amplitude (with q == qn, q' == qn-d

Vg(q)g(q') 1/2(E/M)D/2V1/2

--'--':"':"O=-:c:....;~::::;:~:;D &----exp

v2niEIi/M

(i- A €) ,

(lIB.1)

Ii

where V is the curved-space analog of the Van Vleck-Pauli-Morette determinant [defined in flat space after Eq. (4.124)] : (l1B.2) After taking the Jacobian action Aio into account, this leads to an integral kernel K€(boq) which differs from our correct one in Eq. (11.4) by an extra factor 6K.S. Cheng, J. Math. Phys. 13,1723 (1972) . 7B.S. DeWitt, Rev. Mod. Phys. 29, 377 (1957) .

Vg(q)g(q')

1/2 (E/M)D / 2V1/2. This has

917

Notes and References

f2

the postpoint expansion 1 + RJ.'vf).qJ.' f).qv + . .. . When treated perturbatively, the extra term is equivalent to diR/12M, reducing the effective potential (11.106) by a factor 1/2. In order to obtain the correct amplitude, DeWitt had to add a nonclassical term to the action in the path integral. This term is proportional to Ii? and removes the unwanted term Veff, i.e., f).DwA€ = €Veff. Such a correction procedure must be rejected on the grounds that it runs contrary to the very essence of the entire path integral approach, in which the contribution of each path is controlled entirely by the phase eiA / 1i with the classical action in the exponent. The short-time kernel proposed by Cheng is the same as DeWitt's, except that it does not include the extra Van Vleck-Pauli-Morette determinant in (11B.l). In this case, the result is the full effective potential (11.106), which must be artificially subtracted from the classical action to obtain the correct amplitude.

Notes and References The first path integral in a curved space was written down by B.S. DeWitt, Rev. Mod. Phys. 29, 377 (1957), making use of previous work by C. DeWitt-Morette, Phys. Rev. 81 , 848 (1951). A modified ansatz is due to KS. Cheng, J . Math. Phys. 13, 1723 (1972). For recent discussions with results different from ours see H. Kamo and T. Kawai , Prog. Theor. Phys. 50, 680, (1973); T. Kawai, Found. Phys. 5, 143 (1975) ; H. Dekker, Physica A 103, 586 (1980); G.M. Gavazzi, Nuovo Cimento lOlA, 241 (1981) . A good survey of most attempts is given by M.S. Marinov, Phys. Rep. 60, 1 (1980) . In a space without torsion, C. DeWitt-Morette, working with stochastic differential equations, postulates a Hamiltonian operator without an extra R-term. See her lectures presented at the 1989 Erice Summer School on Quantum Mechanics in Curved Spacetime, ed. by V. de Sabbata, Plenum Press, 1990, and references quoted therein, in particular C. DeWitt-Morette, KD. Elworthy, B.L. Nelson, and G.S. Sammelman, Ann. Inst. H. Poincare A 32, 327 (1980). A path measure in a phase space path integral which does not produce any R-term was proposed by K Kuchar, J . Math. Phys. 24, 2122 (1983). The short derivation of the Schriidinger equation in Subsection 11.1.1 is due to P. Fiziev and H. Kleinert, J. Phys. A 29,7619 (1996) (hep-th/9604172).

To reach a depth profounder still, and still Profounder, in the fathomless abyss. WILLIAM COWPER (1731-1800), The Winter Morning Walk

12 New Path Integral Formula for Singular Potentials

In Chapter 8 we have seen that for systems with a centrifugal barrier, the Euclidean form of Feynman's original time-sliced path integral formula diverges for certain attractive barriers. This happens even if the quantum statistics of the systems is well defined. The same problem arises for a particle in an attractive Coulomb potential, and thus in any atomic system. In this chapter we set up a new and more flexible path integral formula which is free of this problem for any singular potential. This has recently turned out to be the key for a simple solution of many other path integrals which were earlier considered intractable.

12.1

Path Collapse in Feynman's formula for the Coulomb System

The attractive Coulomb potential V(r) = -e 2 /r has a singularity at the coordinate origin r = O. This singularity is weaker than that of the centrifugal barrier, but strong enough to cause a catastrophe in the Euclidean path integral. Recall that an attractive centrifugal barrier does not even possess a classical partition function. The same thing is true for the attractive Coulomb potential where formula (2.350) reads

The integral diverges near the origin. In addition, there is a divergence at large r . The leading part of the latter can be removed by subtracting the free-particle partition function and forming (12.1)

918

12.1 Path Collapse in Feynman 's formula for the Coulomb System

919

leaving only a quadratic divergence. In a realistic many-body system with an equal number of oppositely charged particles, this disappears by screening effects. Thus we shall not worry about it any further and concentrate only on the remaining small-r divergence. In a real atom, this singularity is not present since the nucleus is not a point particle but occupies a finite volume. However, this "physical regularization" of the singularity is not required for quantum-mechanical stability. The Schr6dinger equation is perfectly solvable for the singular pure -e 2 /r potential. We should therefore be able to recover the existing Schr6dinger results from the path integral formalism without any short-distance regularization. On the basis of Feynman's original time-sliced formula, this is impossible. If a path consists of a finite number of straight pieces, its Euclidean action (12.2) can be lowered indefinitely by a path with an almost stretched configuration which corresponds to a slowly moving particle sliding down into the -e 2 /r abyss. We call this phenomenon a path collapse. In nature, this catastrophe is prevented by quantum fluctuations . In order to understand how this happens, it is useful to reinterpret the paths in the Euclidean path integral as random lines parametrized by T E (Ta, Tb) . Their distribution is governed by the "Boltzmann factor" e- A/ Ii , whose effective "quantum" temperature is Teff == lilkB(Tb - Ta) .l The logarithm of the Euclidean amplitude (XbTblxaTa) multiplied by -Teff defines a free energy

F

=

E - kBTS

of the random line with fixed endpoints. The quantum fluctuations equip the path with a configurational entropy S. This must be sufficiently singular to produce a regular free energy bounded from below. Obviously, such a mechanism can only work if the exact path integral contains an infinite number of infinitesimally small sections. Only these can contain enough configurational entropy near the singularity to halt the collapse. The variational approach in Section 5.10 has shown an important effect of the configurational entropy of quantum fluctuations. It smoothes the singular Coulomb potential producing an effective classical potential that is finite at the origin. A path collapse was avoided by defining the path integral as an infinite product of integrals over all Fourier coefficients. The infinitely high-frequency components were integrated out and this produced the desired stability. These high-frequency components are absent in a finitely time-sliced path with a finite number of pieces, where frequencies Dm are bounded by twice the inverse slice thickness liE [recall Eqs. (2.106), (2.107)] . Unfortunately, the path measure used in the variational approach is unsuitable for exact calculations of nontrivial path integrals. Except for the free particle and IThis amounts to viewing the path as a polymer with configurational fluctuations in space, a possibility which is a major topic in Chapters 15 and 16.

920

12 New Path Integral Formula for Singular Potentials

the harmonic oscillator, these are all based on solving a finite number of ordinary integrals in a time-sliced formula. We therefore need a more powerful time-sliced path integral formula which avoids a collapse in singular potentials. For the Coulomb system, such a formula has been found in 1979 by Duru and Kleinert. 2 It has become the basis for solving the path integral of many other nontrivial systems. Here we describe the most general extension of this formula which will later be applied to a number of systems. For the attractive Coulomb potential and other singular potentials, such as attractive centrifugal barriers, it will not only halt the collapse, but also be the key to an analytic solution. The derived stabilization is achieved by introducing a path-independent width of the time slices. If the path approaches an abyss, the widths decrease and the number of slices increases. This enables the configurational entropy of Eq. (12.3) to grow large enough to cancel the singularity in the energy. To see the cancellation mechanism, consider a random line with n links which has, on a simple cubic lattice in D dimensions, (2D)n configurations with an entropy

S = nlog(2D) .

(12.3)

If the number of time slices n increases near the -e 2 /r singularity like const/r, then the entropy is proportional to l/r. A path section which slides down into the abyss must stretch itself to make the kinetic energy small. But then it gives up a certain entropy S, and this raises the free energy by kBTeffS according to (12.3). This compensates for the singularity in the potential and halts the collapse. The purpose of this chapter is to set up a path integral formula in which this stabilizing mechanism is at work. It should be pointed out that no instability problem would certainly arise if we were to define the imaginary-time path integral for the time evolution amplitude in the continuum

(12.4) without any time slicing as the solution of the Schrodinger differential equation (12.5) [compare Eq. (1.304)]. After solving the Schrodinger equation H'lj;n(x) = En'lj;n(x), the spectral representation (1.319) renders directly the amplitude (12.4). All subtleties described above are due to the finite number of time slices in the path integral. As explained at the end of Section 2.1, the explicit sum over all paths is an essential ingredient of Feynman's global approach to the phenomena of 21.H. Duru and H. Kleinert, Phys. Lett. B 84, 30 (1979) (http://www . physik. fu-berlin.derkleinert/65); Fortschr. Phys. 30, 401 (1982) (ibid.http/83) . See also the historical remarks in the preface.

12.2 Stable Path Integral with Singular Potentials

921

quantum fluctuations . Within this approach, the finite time slicing is essential for being able to perform this sum in any nontrivial system. We now present a general solution to the stability problem of time-sliced quantum-statistical path integrals.

12.2

Stable Path Integral with Singular Potentials

Consider the fixed-energy amplitude (9.1) which is the local matrix element

(xblxa)E = (xbIRlxa)

(12.6)

of the resolvent operator (1.315):

in R-----..,,--- E - H +i'r} A

(12.7)

Recall that the i'r}-prescription ensures the causality of the Fourier transform of (12.6), making it vanish for tb < ta [see the discussion after Eq. (1.323)] . The fixed-energy amplitude has poles of the form

at the bound-state energies, and a cut along the continuum part of the energy spectrum. The energy integral over the discontinuity across the singularities yields the completeness relation (1.326) . The new path integral formula is based on the following observation. If the system possesses a Feynman path integral for the time evolution amplitude, it does so also for the fixed-energy amplitude. This is seen after rewriting the latter as an integral (12.8) involving the modified time evolution operator

UE(t) == e-it(H-E)/\

(12.9)

that is associated with the modified Hamiltonian

HE == H - E .

(12.10)

Obviously, as long as the matrix elements of the ordinary time evolution operator U(t) = e- itH / 1i can be represented by a time-sliced Feynman path integral, the same is true for the matrix elements of the modified operator UE(t) = e-itHE/Ii . Its explicit form is obtained, as in Section 2.1 , by slicing the t-variable into N + 1 pieces, factorizing exp( -itHE/n) into the product of N + 1 factors, (12.11)

922

12 New Path Integral Formula for Singular Potentials

and inserting a sequence of N completeness relations

II JdDXnIXn) (xnl = 1 N

(12.12)

n=l

(omitting the continuum part of the spectrum) . In this way, we have arrived at the path integral for the time-sliced amplitude with tb - ta = E(N + 1) (12.13) where A~ is the sliced action N+l

A~

=

L

{Pn(xn - Xn-l ) - E[H(Pn, xn) -

n=l

In the limit of large N at fixed tb - ta

(xbIUE(t)lxa) =

=

J1JDx (t') J~;;.g2

E( N exp

En .

(12.14)

+ 1) , this defines the path integral

{~l dt'[px(t') -

HE(p(t'), x(t'))]}.

(12.15) It is easy to derive a finite-N approximation also for the fixed-energy amplitude (xblxa)E of Eq. (12.8). The additional integral over tb > ta can be approximated at the level of a finite N by an integral over the slice thickness E: (12.16) The resulting finite-N approximation to the fixed-energy amplitude, N

(XbIXa)E

= (N + 1) inroo dE(XbIUEN (E(N + l))l x a) = itroo dtb(XblUEN (tb A

A

ta)lxa) ,

o a (12.17) converges against the correct limit (xblxa)E. As an example, take the free-particle case where

(12.18) After a trivial change of the integration variable, this is the same integral as in (1.339) whose result was given in (1.344) and (1.351), depending on the sign of the energy E. The N -dependence happens to disappear completely as observed in Section 2.2.4. In the general case of an arbitrary smooth potential, the convergence is still assured by the dominance of the kinetic term in the integral measure. The time-sliced path integral formula for the fixed-energy amplitude (xblxa)E given by (12.17) , (12.13), (12.14) has apparently the same range of validity as the

923

12.2 Stable Path Integral with Singular Potentials

original Feynman path integral for the time evolution amplitude (Xbtblxata) . Thus, so far nothing has been gained. However, the new formula has an important advantage over Feynman's. Due to the additional time integration it possesses a new functional degree of freedom. This can be exploited to find a path-integral formula without collapse at imaginary times. The starting point is the observation that the resolvent operator R in Eq. (12.7) may be rewritten in the following three ways:

in

A

R=A fl(E - H A

A

+ i1])

!L,

(12.19)

or A

in

A

R = fr

A

(E - H

A

+ i1])fr

(12.20)

,

or, more generally, A

in

A

A

R = fr lz(E _ iI + i1])ir fl'

(12.21 )

lz,

where ir are arbitrary operators which may depend on f> and x. They are called regulating functions. In the subsequent discussion, we shall avoid operator-ordering subtleties by assuming iI, ir to depend only on X, although the general case can also be treated along similar lines. Moreover, in the specific application to follow in Chapters 13 and 14, the operators ir to be assumed consist of two different powers of one and the same operator i.e.,

lz,

i,

(12.22) whose product is

lzir = j.

(12.23)

Taking the local matrix elements of (12.21) renders the alternative representations for the fixed-energy amplitude (12.24) where UE(s) is the generalization of the modified time evolution operator (12.9), to be called the pseudo time evolution operator, (12.25) The operator in the exponent,

HE == fl(X)(iI -

E)fr(x),

(12.26)

924

12 New Path Integral Formula for Singular Potentials

may be considered as an auxiliary Hamiltonian which drives the state vectors Ix} of the system along a pseudotime s-axis, with the operator e-ishE(p,x)/h. Note that HE is in general not Hermitian, in which case UE(s) is not unitary. As usual, we convert the expression (12.24) into a path integral by slicing the pseudotime interval (0, s) into N +1 pieces, factorizing exp( -isHE/h) into a product of N + 1 factors, and inserting a sequence of N completeness relations. The result is the approximate integral representation for the fixed-energy amplitude,

(xblxa)E ::::; (N + 1)

lJO dEs (xbIU: (Es(N + 1)) Ix

a },

(12.27)

with the path integral for the pseudotime-sliced amplitude

whose time-sliced action reads N+l

A~

=

L

{Pn(xn - Xn-l) - Esft(Xn) [H(Pn' xn) - Elfr(Xn)} .

(12.29)

n=l

These equations constitute the desired generalization of the formulas (12.13)(12.17) . In the limit of large N, we can write the fixed-energy amplitude as an integral (12.30)

over the amplitude

The prime on x(s) denotes the derivative with respect to the pseudotime s. For a standard Hamiltonian of the form

H = T(p)

+ V(x) ,

(12.32)

with the kinetic energy

p2 T(p) = 2M'

(12.33)

the momenta Pn in (12.28) can be integrated out and we obtain the configuration space path integral (12.34)

12.2 Stable Path Integral with Singular Potentials

925

with the sliced action

In the limit of large N , this may be written as a path integral

with the slicing specification (12.35) . The path integral formula for the fixed-energy amplitude based on Eqs. (12.30) and (12.36) is independent of the particular choice of the functions fz(x), fr(x) , just like the most general operator expression for the resolvent (12.21). Feynman's original time-sliced formula is, of course, recovered with the special choice fz(x)

fr(x)

=1.

=

When comparing (12.25) with (12.9) , we see that for each infinitesimal pseudotime slice, the thickness of the true time slices has the space-dependent value (12.37)

The freedom in choosing f(x) amounts to an invariance under path-dependent time reparametrizations of the fixed-energy amplitude (12.30). Note that the invariance is exact in the general operator formula (12.21) for the resolvent and in the continuum path integral formula based on (12.30) and (12.36). However, the finite pseudotime slicing in (12.34), (12.35) used to define the path integral, destroys this invariance. At a finite value of N, different choices of f(x) produce different approximations to the matrix element of the operator UE(s) = fr(x)e-isHE(P,x)/lifz(x). Their quality can vary greatly. In fact , if the potential is singular and the regulating functions fr(x) , fz(x) are not suitably chosen, the Euclidean pseudotime-sliced expression may not exist at all. This is what happens in the Coulomb system if the functions fz(x) and fr(x) are both chosen to be unity as in Feynman's path integral formula. The new reparametrization freedom gained by the functions fz(x), fr(x) is therefore not just a luxury. It is essential for stabilizing the Euclidean time-sliced orbital fluctuations in singular potentials. In the case of the Coulomb system, any choice of the regulating functions fz(x), fr(x) with f(x) = r leads to a regular auxiliary Hamiltonian 'HE , and the path integral expressions (12.27)- (12.36) are all well defined. This was the important discovery of Duru and Kleinert in 1979, to be described in detail in Chapter 13, which has made a large class of previously non-existing Feynman path integrals solvable. By a similar Duru-Kleinert transformation with fz(x) , fr(x) with f(x) = fz(x)fr(x) = r2, the earlier difficulties with the centrifugal barrier are resolved, as will be seen in Chapter 14.

V

926

12.3

12 New Path Integral Formula for Singular Potentials

Time-Dependent Regularization

Before treating specific cases, let us note that there exists a further generalization of the above path integral formula which is useful in systems with a time-dependent Hamiltonian H(p, x, t). There we introduce an auxiliary Hamiltonian

it =

fz(x, t) [H(p, x, t) - Elfr(x, t),

(12.38)

where E is the differential operator for the energy which is canonically conjugate to the time t : (12.39) The auxiliary Hamiltonian acts on an extended Hilbert space, in which the states are localized in space and time. These states will be denoted by lx, t}. They satisfy the orthogonality and completeness relations

{x tlx' t'} =

J(D) (x

- x')J(t - t'),

(12.40)

and

JdDx Jdtlxt}{xtl =

1,

(12.41 )

respectively. By construction, the Hamiltonian 1t does not depend explicitly on the pseudotime s. The pseudotime evolution operator is therefore obtained by a simple exponentiation, as in (12.25),

71(s)

= fr(x , t)e-iS!t(x,t)(H-E)!r(x,t) fl(X, t).

(12.42)

The derivation of the path integral is then completely analogous to the timeindependent case. The operator (12.42) is sliced into N + 1 pieces, and N completeness relations (12.41) are inserted to obtain the path integral

(12.43) with the pseudotime-sliced action N+I

AN

=

L

{Pn(Xn -

Xn-I) -

En(tn - tn-I)

n=1

(12.44) where Xb = XN+ l, tb = tN+!; xa = Xu, ta = to. This describes orbital fluctuations in the phase space of spacetime which contains fluctuating worldlines x(s), t(s) and

927

12.3 Time-Dependent Regularization

their canonically conjugate spacetime p(s), E(s). In the limit N as

---> 00

we write this

(12.45) with the continuous action

A[p, x , E, t]

=

loS ds{p(s)x'(s) - E(s)t'(s)

-fl(P(S),x(s) , t(s)) [H(p(s) , x(s) , t(s)) - E(s)] fr(P(s),x(s), t(s))}. (12.46) In the pseudotime-sliced formula (12.43) , we can integrate out all intermediate energy variables En and obtain (12.47)

with the action N+I

A-N

" = '~

[Pn(Xn - Xn-I ) - fsfl(PnXn, tn)H(Pn, Xn, tn)fr(Pn-b Xn-I, tn-I )].

n=1

(12.48) This looks just like an ordinary time-sliced action with a time-dependent Hamiltonian. The constant width of the time slices E = (tb - ta)/(N + 1), however, has now become variable and depends on phase space and time: (12.49) The 8-function in (12.47) ensures the correct relation between the pseudotime sand the physical time t. In the continuum limit we may write (12.47) as

with the pseudotime action

A[p, x, t] = loS ds [px' - fl(X, t)H(p , x, t)fr(x, t)] ,

(12.51)

which is a functional of the s-dependent paths x(s) , p(s) , t(s). Note that in the continuum formula, the splitting of the regulating function f(x , t) into fl(X , t) and fr(x , t) according to the parameter A in Eq. (12.22) cannot be expressed properly since !l, H, and fr are commuting c-number functions. We have written them in a way indicating their order in the time-sliced expressions (12.47) , (12.48).

928

12 New Path Integral Formula for Singular Potentials

The integral over S yields the original time evolution amplitude (12.52) Indeed , by Fourier decomposing the scalar products {xbtblxata}, {X

t b b

Ix

t }

=

a a

J~ J dE eip(Xb- Xa)/h-iE(tb- ta)/h ' (27rn)D 27rn

(12.53)

we see that the right-hand side satisfies the same Schrodinger equation as the lefthand side: (12.54) [recall (1.304) and (12.5)]. If the o-function in (12.47) is written as a Fourier integral, we obtain a kind of spectral decomposition of the amplitude (12.52) , (12.55) with the pseudotime evolution amplitude:

UE(s) == fr(-P, x, t) e - isft(p,x ,t) (H- El!r(p,x,tl iz(p, x, t).

12.4

(12.56)

Relation to Schrodinger Theory. Wave Functions

For completeness, consider also the ordinary Schrodinger quantum mechanics described by the pseudo-Hamiltonian fl. This operator is the generator of translations of the system along the pseudotime axis s. Let ¢(x, t, s) be a solution of the pseudotime Schrodinger equation

H(p, x , E, t)¢(x, t , s) = inos¢(x, t , s) ,

(12.57)

written more explicitly as

fz(x , t) [H(p , x , t) - inot] fr(x, t)¢(x, t, s)

=

inos¢(x , t, s).

(12.58)

Since the left-hand side is independent of s, the s-dependence of ¢(x, t , s) can be factored out:

¢(x ,t , s)

=

¢s(x,t)e- iSs / h.

(12.59)

If H is independent of the time t, it is always possible to stabilize the path integral by a time-independent reparametrization function f(x) . Then we remove an oscillating factor e- iEt /h from ¢s(x, t) and factorize (12.60)

12.4 Relation to Schr8dinger Theory. Wave Functions

929

This leaves us with the time- and pseudotime-independent equation

1i(p, x, E) ' - 2 to 3>' i

"h,Af = 3>'

L

N+l

(iP)

log

(13.111)

;12 n Un-l

n=l

[compare (13.68)]. As in the two-dimensional case we shall at first ignore the subtleties due to the time slicing. Thus we set >. = 0 and apply the transformation formally to the continuum limit of the action Al¥, which has the form (13.10), except that x is replaced by x. Using the properties of the matrix (13.102)

4u2 A-I,

Vdet (AAT) = 16r

2,

(13.112)

we see that

4u2iP = 4rU'2 , 16r2 d4 u .

(13.113) (13.114)

In this way, we find the formal relation

eie2S/Ii~

(xbIUE(S)lx a) =

16

Jdx!ra (ubSluaO)

(13.115)

to the time evolution amplitude of the four-dimensional harmonic oscillator (13.116) with the action

Aos =

r

S

Jo

dS!!:.(U'2 _ w 2U2). 2

(13.117)

The parameters are, as in (13.27) and (13.28) , J1

= 4M,

W

= V-E/2M .

(13.118)

The relation (13.115) is the analog of (13.31). Instead of a sum over the two images of each point x in u-space, there is now an integral J dx!/r a over the infinitely many images in the four-dimensional u-space. This integral can be rewritten as an integral over the third Euler angle I using the relation (13.103). Since x and thus the polar angles B, 'P remain fixed during the integration, we have directly J dx!/ra = J d'a· As far as the range of integration is concerned, we observe that it may be restricted to a single period la E [0, 41l"]. The other periods can be included in the oscillator amplitude. By specifying a four-vector Ul" all paths are summed which run either to

947

13.4 Solution for the Three-Dimensional Coulomb System

the final Euler angle Ib or to all its periodic repetitions [which by (13.97) have the same Ub]. This was the lesson learned in Section 6.1. Equation (13.115) contains, instead, a sum over all initial periods which is completely equivalent to this. Thus the relation (13.115) reads, more specifically, (13.119) The reason why the other periods in (13.115) must be omitted can best be understood by comparison with the two-dimensional case. There we observed a two-fold degeneracy of contributions to the time-sliced path integral which cancel all factors 2 in the measure (13.71). Here the same thing happens except with an infinite degeneracy: When integrating over all images d4 u n of d4 xn in the oscillator path integral we cover the original x-space once for In E [0,47r] and repeat doing so for all periods In E [47rl , 47r(l + 1)]. This suggests that each volume element d4 u n must be divided by an infinite factor to remove this degeneracy. However, this is not necessary since the gradient term produces precisely the same infinite factor. Indeed, ~ ~ )2(Un ~ (Un + UnI

~

- Un-I

)2

(13.120)

is small for xn ~ Xn-I at infinitely many places of In -In-I, once for each periodic repetition of the interval [0, 47r]. The infinite degeneracy cancels the infinite factor in the denominator of the measure. The only place where this cancellation does not occur is in the integral J dx!fr a. Here the infinite factor in the denominator is still present, but it can be removed by restricting the integration over la in (13.119) to a single period [7]. Note that a shift of la by a half-period 27r changes U to -u and thus corresponds to the two-fold degeneracy in the previous two-dimensional system. The time-sliced path integral for the harmonic oscillator can, of course, be done immediately, the amplitude being the four-dimensional version of (13.33):

(ubSluaO) =

1

(27ri11,Es/ J.l)2

IT [1

n=1

4

[i f.

d b.un ] exp !!. (.!.b.U 2 27ri11,Es/ J.l 11, n=1 2 Es n

-

E W2

s

U2 )] n

{i

w S / )2 exp ~---:---S J.lW [( Ub ~ 2+U~2) S 2 ~ ~ ]} . ( 13.121 ) .f< . a COSW - UbUa (2 7rZnsmw J.l n smw 2

To find the fixed-energy amplitude we have to integrate this over S:

(xblxa)E =

roo dSeie2S/hJ:.... r47r dla(ubSluaO) . 16 io

io

(13.122)

Just like (13.35), the integral is written most conveniently in terms of the variables (13.38) , (13.40), so that we obtain the fixed-energy amplitude of the threedimensional Coulomb system

948

13 Path Integral of Coulomb System

In order to perform the integral over dxt, we now express UbUa in terms of the polar angles

UbUa = ylTbTa {cos(Bb/ 2) cos(Ba/2) cos[( 'Pb - 'Pa + 'Yb - 'Ya)/2] + sin(Bb/2) sin(Ba /2) cos[( 'Pb - 'Pa - 'Yb + 'Ya)/2]} .

(13.123)

A trigonometric rearrangement brings this to the form

UbUa = ylTbTa {COS [(Bb - Ba)/2] COS[('Pb - 'Pa)/2] COS[("(b - 'Ya)/2] - COS[(Bb + Ba)/2] sin[('Pb - 'Pa)/2] sin[("(b - 'Ya)/2]} ,

(13.124)

and further to

UbUa = V(TbTa + xbxa) /2cos [("(b - 'Ya + ,8)/2],

(13.125)

where ,8 is defined by (13.126)

or (13.127)

Ii

The integral 7r d'Ya can now be done at each fixed x. This gives the fixed-energy amplitude of the Coulomb system [1 , 8, 10, 9] .

(XbIXa)E

. MK,

11

dg (

r;r

=

-Z---.;-

x

exp [-K, 1 + g (Tb

1fn

0

l-g

ll

1- g

(2yfQ

Flo 2K,--V(TbTa+ XbXa)/2

)

1- g

+ Ta)]

,

(13.128)

where K, and 1I are the same parameters as in Eqs. (13.40). The integral converges only for 1I < 1, as in the two-dimensional case. It is again possible to write down another integral representation which converges for all 1I "I 1,2,3,.. . by changing the variables of integration to ( == (1 + g)/(1 - g) and transforming the integral over ( into a contour integral encircling the cut from ( = 1 to 00 in the clockwise sense. Since the cut is now of the type (( - 1) - 11, the replacement rule is

1 d(((-I) 00

- II

...

1

This leads to the representation

(XbIXa)E

=

1fei7r1l

-->-.-

1

sm 1f1l c

d( - II -.((-1) ... 21fz

(13.129)

1

. M K, 1fei7r1l d( -II II -z- - - .- . (( - 1) (( + 1) 1fn. 2 sm 1f1l C 21fz xlo(2K,V(2 - IV'(T-b-Ta-+-X-b-X-a)-/-2)e-I«(rb+ral.

(13.130)

13.5 Absence of Time Slicing Corrections for D

13.5

=3

949

Absence of Time Slicing Corrections for D

=

3

Let us now prove that for the three-dimensional Coulomb system also, the finite time-slicing procedure does not change the formal result of the last section. The reader not interested in the details is again referred to the brief argument in Section 13.6. The action Al¥ in the time-sliced path integral has to be supplemented, in each slice, by the Jacobian action [as in (13.62)] i AE

h

- el'e i {I',v} ~uv - eil'e i {I.,,« 2 fv«"f.x,,« - 2(- Dv>.U + 2uvu.x)/it +uvu.x/it . fv«"f.x,,« - 2f v«"S>.,,«

(13.142)

Collecting all terms, the Jacobian action (13.131) becomes (13.143)

A,.

In contrast to the two-dimensional equation (13.66), this cannot be incorporated into Although the two expressions contain the same t erms, their coefficients are different [see (13.111)] : i

,

n, A f

3>' log (:;;, )

(13.144)

un _ 1

2 n2 + 2 (Ut!.Un) 3'1 t!.Un _ t!.u /\ og [2 Un ;;'2 ;;'2 ;;'2 + ...]

=

un

Un

Un

.

It is then convenient to rewrite (omitting the subscripts n)

U2

- 2log [ (u _ t!.U)2

]

+

ut!.u U2

(13.145)

t!.U2 3 (Ut!.U)2 - U2 +2 U2 + ... , and absorb the first term into Aj, which changes 3>' to (3)' - 2). Thus, we obtain altogether the additional action [to be compared with (13.76)]

n,i t!. corr A '

!..4M t!.U2 [(2)' _ 1) ut!.u Ii

2E

+(3)' - 2) [2

U2

+ (!4 _ >.) t!.U2 + 2>.2 (Ut!.U) 2] U2 U2

U~U _ ~~-'2 + 2 ( u~ur]

ut!.u (t!.U) 2 3 (Ut!.U) 2 + U2 - ~+2 U2 + . . ..

(13.146)

951

13.6 Geometric Argument for Absence of Time Slicing Corrections Using this we now show that the expansion of the correction term

(13.147)

has the vanishing expectations

(C)o (C (pt::.i1))0

0, (13.148)

0,

i.e.,

* (t::.corrA')o +

~ (*) 2 ((t::.corrA')2)0

0,

(13.149)

*(t::.corrA (pt::.i1))O

0,

(13.150)

as in (13.87) , (13.88). In fact , using formula (13.86) the expectation (13.150) is immediately found to be proportional to

2+ 2(3A - 2) -1+ -1]

D + i [ -2(2A -1) - -

16

4

4'

(13.151)

which vanishes identically in A for D = 4. Similarly, using formula (13.85) , the first term in (13.149) has an expectation proportional to i [- 2

(!4 _A) (D +162)D _ 4A2D16+ 2 _ (3A - 2) (E.4 _ ~)4 _(E.4 _~)] 8 '

(13.152)

i.e. , for D = 4, (13.153)

to which the second term adds (13.154)

which cancels (13.154) for D = 4. Thus the sum of all time slicing corrections vanishes also in the three-dimensional case.

13.6

Geometric Argument for Absence of Time Slicing Corrections

As mentioned before, the basic reason for the absence of the time slicing corrections can be shown to be the property of the connection (13.155)

which, in turn, follows from the basic identity 81'e~ = 0 satisfied by the basis tetrad, and from the diagonality of the metric gl'l/ ex (jl'//. Indeed, it is possible to apply the techniques of Sections 10.1, 10.2 to the general pseudotime evolution amplitude (12.28) with the regulating functions

fl = f(x),

fr == 1.

(13.156)

952

13 Path Integral of Coulomb System

Since this regularization affects only the postpoints at each time slice, it is straightforward to repeat the derivation of an equivalent short-time amplitude given in Section 11.3. The result can be expressed in the form (dropping subscripts n)

K'(~q) =

y'ij(Qj

J27riEnf 1M

D

exp [i(A€ n

+ Aj)] ,

(13.157)

where f abbreviates the postpoint value f(xn) and A€ is the short-time action

A€

2~gI'V(q)~ql'~qv .

=

(13.158)

There exists now a simple expression for the Jacobian action. Using formula (11.75), it becomes simply

inA J €

-

-

~r 2 I' I' v ~ qv _

.

tE

!!:L(r 8M I' I' v )2 .

(13.159)

In the postpoint formulation, the measure needs no further transformation. This can be seen directly from the time-sliced expression (13.8) for .>.. = 0 or, more explicitly, from the vanishing of the extra action Aj in (13.69) for D = 2 and in (13.144) for D = 3. As a result , the vanishing contracted connection appearing in (13.159) makes all time-slicing corrections vanish. Only the basic short-time action (13.158) survives:

~A€

=

4M (~u)2 .

(13.160)

2Es

Thanks to this fortunate circumstance, the formal solution found in 1979 by Duru and Kleinert happens to be correct.

13.7

Comparison with Schrodinger Theory

For completeness, let us also show the significance of the geometric property r I'll>' = Consider the Schrodinger equation of the Coulomb system

o within the Schrodinger theory.

2

2 2)

e 1 ( - 2Mli V - E 1/J(X) = --;:1/J(x) ,

(13.161)

to be transformed to that of a harmonic oscillator. The postpoint regularization of the path integral with the functions (13.156) corresponds to multiplying the Schrodinger equation with il = r from the left. This gives

(-

2~li2rV2 -

Er) 1/J(x)

=

e21/J(x).

(13.162)

We now go over to the square root coordinates u l1 transforming - Er into the harmonic potential - E( U I1 )2 and the Laplacian V2 into gill' 8J3v - r 1111>'8>. . The geometric property r 1111>' = 0 ensures now the absence of the second term and the result is simply g11V 811 8 v . Since gill' = Jill' /4r, the Schrodinger equation (13.162) takes the simple form (13.163)

953

13.7 Comparison with Schrodinger Theory

Due to the factor (uJL? accompanying the energy E, the physical scalar product in which the states of different energies are orthogonal to each other is given by (13.164)

This corresponds precisely to the scalar product given in Eq. (11.95) with the purpose of making the Laplace operator (here ~ = (1/4u2)8~) hermitian in the uJL-space with torsion. Indeed, the once-contracted torsion tensor SJL = SJLV v can be written as a gradient of a scalar function: (13.165)

Quite generally, we have shown in Eq. (11.104) that if SJL(q) is a partial derivative of a scalar field a(q), the physical scalar product is given by (11.95):

(1j!2 11flt)phys ==

J

dD qvg(q)e- 2U (Q)1j!;(q)1j!1 (q).

(13.166)

yg = 16~ ,

(13.167)

From (13.132), we have

so that the physical scalar product is

8i

The Laplace operator obtained from by the nonholonomic KustaanheimoStiefel transformation is ~ = (1/4it2)8;. This is Hermitian in the physical scalar product (13.168), but not in the naive one (11.90) with the integral measure

J d4 u16u4 . In two dimensions, the torsion vanishes and the physical scalar product reduces to the naive one: (13.169)

With Jl = 4M and - E harmonic oscillator:

=

Jlw2 /2, Eq. (13.163) is the Schrodinger equation of a

+ !!.W2(UJL?] 1j!(uJL) = E1j!(uJL). [-~n,282 2Jl JL 2

(13.170)

The eigenvalues of the pseudoenergy E are (13.171)

where Du = 4 is the dimension of uJL-space, (13.172)

954

13 Path Integral of Coulomb System

sums up the integer principal quantum numbers of the factorized wave functions in each direction of the ull-space. The multivaluedness of the mapping from x to u ll allows only symmetric wave functions to be associated with Coulomb states. Hence N must be even and can be written as N = 2(n - 1). The pseudoenergy spectrum is therefore En

= fiw2(n + Dul4 -1),

n

= 1,2,3, ....

(13.173)

According to (13.170), the Coulomb wave functions must all have a pseudoenergy

(13.174) The two equations are fulfilled if the oscillator frequency has the discrete values

W

With w 2

= Wn ==

2(n + Dul 4 - 1)'

n = 1,2,3 ....

(13.175)

= - E 12M and Du = 4, this yields the Coulomb,energies En

= -2Mw~ =

Me 4 1 -Y2n2 '

(13.176)

showing that the number N 12 = n - 1 corresponds to the usual principal quantum number of the Coulomb wave functions. Let us now focus our attention upon the three-dimensional Coulomb system where Du = 4. In this case, not all even oscillator wave functions correspond to Coulomb bound-state wave functions. This follows from the fact that the Coulomb wave functions do not depend on the dummy fourth coordinate X4 (or the dummy angle ')'). Thus they satisfy the constraint ox4'l/J = 0, implying in ull-space [recall (13.132)]

-ire41l01l'l/J(UIl ) =

-i~

[(U 2 01 - U1(2) + (U403 - u3(4)] 'l/J(u ll )

-iO"j'l/J(ull ) = O.

(13.177)

The explicit construction of the oscillator and Coulomb bound-state wave functions is most conveniently done in terms of the complex coordinates (13.96). In terms of these, the constraint (13.177) reads

~ [20.. -

ZOz] 'l/J(z, z*) = O.

(13.178)

This will be used below to select the Coulomb states. To solve the Schrodinger equation (13.170) we simplify the notation by going over to atomic natural units, where fi = 1, M = 1, e2 = 1, 11 = 4M = 4. All lengths are measured in units of the Bohr radius (4.344), whose numerical value is aH = fi2 1Me2 = 5.2917 x 10- 9 cm, all energies in units of EH == e2 I aH =

955

13.7 Comparison with Schrodinger Theory

Me 4 /n2 = 4.359 x 10- 11 erg = 27.210 eV, and all frequencies w in units of WH == Me 4 /n 3 = 4.133 x 1016 / sec(= 41l'X Rydberg frequency VR)' Then (13.170) reads (after multiplication by 4M/n2 )

(13.179) The spectrum of the operator it is obviously 4w(N + 2) = 8wn. To satisfy the equation, the frequency w has to be equal to Wn = 1/ 2n. We now observe that the operator it can be brought to the standard form (13.180) with the help of the w-dependent transformation (13.181) in which the operator b is an infinitesimal dilation operator which in this context is called tilt operator [11, 12] : (13.182) and {) is the tilt angle {) = log(2w).

(13.183)

The Coulomb wave functions are therefore given by the rescaled solutions of the standardized Schrodinger equation (13.180): (13.184) Note that for a solution with a principal quantum number n the scale parameter ffw depends on n: (13.185) The standardized wave functions 'lj;~ (uP,) are constructed most conveniently by means of four sets of creation and annihilation operators at, at bt, bt and al, a2 , bl, b2 . They are combinations of Zl , Z2 , their complex-conjugates, and the associated differential operators OZl ' OZ2 ' oZi' OZ2' The combinations are the same as in (9.127), (9.128), written down once for Zl and once for Z2. In addition, we choose the indices so that ai and bi transform by the same spinor representation of the rotation group. If Ci j is the 2 x 2 matrix . 2 = C=UJ'

(

01)

-1 0

'

(13.186)

13 Path Integral of Coulomb System

956

then

C;jZj

transforms like

We therefore define the creation operators

Z;.

at1 == -~(-o. + Z2) J2 z2

'

at2 == ~(-o J2 zl• + Zl) '

(13.187)

and the annihilation operators

(13.188) Note that

01 = -oz ..

The standardized oscillator Hamiltonian is then

h

8

=

2(at a + btb + 2),

(13.189)

where we have used the same spinor notation as in (13.94). The ground state of the four-dimensional oscillator is annihilated by a!, 0,2 and bI, b2 • It has therefore the wave function 1 • • = -1e - (U,,)2 (z z*) = _e-ZjZl-Z2Z2 (z , z*IO) = ./, '1'8,0000, Vir Vir·

(13.190)

The complete set of oscillator wave functions is obtained, as usual, by applying the creation operators to the ground state, Atn~ a2 Atn~bAtn~bAtn~ 10) , 1n a1 , n 2a , n b1 , n 2b) _- N n al' n 2'a n bl' n b2 al 1 2

(13.191)

with the normalization factor (13.192) The eigenvalues of as

h8

are obtained from the sum of the number of a- and b-quanta

2(n~

+ n~ + n~ + n~ + 2) = 2(N + 2) = 4n.

(13.193)

The Coulomb bound-state wave functions are in one-to-one correspondence with those oscillator wave functions which satisfy the constraint (13.178), which may now be written as A L05

1

= -"2(0,

t

0,-

AtA b b)'lj;8

= O.

(13.194)

These states carry an equal number of a- and b-quanta. They diagonalize the (mutually commuting) a- and b-spins (13.195)

13.8 Angular Decomposition of Amplitude, and Radial Wave Functions

957

with the quantum numbers la = (n~ + n~)/2, lb = (n~ + n~)/2,

where l,m are the eigenvalues of i},

m a = (n~ - n~)/2, m b = (n~ - n~)/2,

(13.196)

L3 . By defining

n~ == nl + m, n~ == n2, n~ = n2 + m, n~ = nl, n~ == nl , n~ == n2 - m , n~ = n2 , n~ = nl - m,

for m 2: 0, for m ::; 0,

(13.197)

we establish contact with the eigenstates InlJ n2, m) which arise naturally when diagonalizing the Coulomb Hamiltonian in parabolic coordinates. The relation between these states and the usual Coulomb wave function of a given angular momentum Inlm) is obvious since the angular momentum operator Li is equal to the sum of aand b-spins. The rediagonalization is achieved by the usual vector coupling coefficients (see the last equation in Appendix 13A). Note that after the tilt transformation (13.184), the exponential behavior of the oscillator wave functions 'ljJ~(u/J.) ex: polynomial(u/J.) x e - (u,,)2 goes correctly over into the exponential r-dependence of the Coulomb wave functions 'ljJ(x) ex: polynomial(x) x e- r / n . It is important to realize that although the dilation operator b is Hermitian and the operator eWD at a fixed angle f) is unitary, the Coulomb bound states 'ljJn, arising from the complete set of oscillator states 'ljJ~ by applying eWD , do not span the Hilbert space. Due to the n-dependence of the tilt angle f)n = log(l/n), a section of the Hilbert space is not reached. The continuum states of the Coulomb system, which are obtained by tilting another complete set of states, precisely fill this section. Intuitively we can understand this incompleteness simply as follows. The wave functions 'ljJ~(u/J.) have for increasing n spatial oscillations with shorter and shorter wavelength. These allow the completeness sum 2:n 'ljJ~(u/J.)'ljJ~*(u/J.) to build up a 8-function which is necessary to span the Hilbert space. In contrast, when forming the sum of the dilated wave functions

L 'ljJ~(u/J. /y'n)'ljJ~*(u/J. /y'n), n

the terms of larger n have increasingly stretched spatial oscillations which are not sufficient to build up an infinitely narrow distribution. A few more algebraic properties of the creation and annihilation operator representation of the Coulomb wave functions are collected in Appendix 13A and 13.10.

13.8

Angular Decomposition of Amplitude, and Radial Wave Functions

Let us also give an angular decomposition of the fixed-energy amplitude. This serves as a convenient starting point for extracting the radial wave functions of the

958

13 Path Integral of Coulomb System

Coulomb system which will, in Chapter 14, enable us to find the Coulomb amplitude to D dimensions. We begin with the expression (13.128),

(13.198) and rewrite the Bessel function as Io(zcos(Bj2)), where B is the relative angle between Xa and Xb, and

z

_

2..fl

= 2Kyfrbra--.

(13.199)

1-Q

Now we make use of the expansion1 1 ( 2kz

)1-'-1/

00 1 f(l + J.l) IAkz) = kl-' ~ IT f(l + v) (2l + J.l)F( -l , l + J.li 1 + vi k 2)( - )II2l+I-'(z).

(13.200) Setting k = cos(Bj2), v = 2q > 0, J.l = 1 + 2q, and using formulas (1.442) , (1.443) for the rotation functions, this becomes (13.201) reducing for q

= 0 to 2

00

Io(zcos(Bj2)) = - 2)2l + 1)Pt(cosB)I21+1(Z). z 1=0 After inserting this into (13.128) and substituting y

z

=

=

-~

(13.202)

log Q, so that

1 sm y

2Kyfrbra--:-h '

(13.203)

we expand the fixed-energy amplitude into spherical harmonics 1 00 2l + 1 2:(rblra)E,I--Pt(cosB) rbra 1=0 41l'

t

_ 1_ f)rblra)E I l'im (Xb)Yz;"(xa) , rb r a 1=0 ' m=- l

(13.204)

IG.N. Watson, Theory of Bessel FUnctions , Cambridge University Press, London, 1966, 2nd ed., p.140, formula (3).

13.8 Angular Decomposition of Amplitude, and Radial Wave Functions

959

with the radial amplitude

2M -iyfrbra1i

1 dy-.--e 1

00

smhy

0

2" Y

x exp [-I\; coth y(rb + ra)] hl+1

(13.205)

(2I\;yfrbra~h ). sm y

Now we apply the integral formula (9 .29) and find

(rblra)E,1 =

. M r( -v + l + 1) (2l + I)! W",I+1/2 (2I\; r b) M",I+1/2 (2I\;ra).

-2 1i 1\;

(13.206)

V-2M

On the right-hand side the energy E is contained in the parameters I\; = E /1i2 2 2 4 and v =e /2w1i=V -e M /21i E. The Gamma function has poles at v = n with n = l + I, l + 2, l + 3, . .. . These correspond to the bound states of the Coulomb system. Writing 1 1

(13.207)

1\;=--

aHv'

with the Bohr radius (13.208)

(for the electron, aH :::::; 0.529 x 1O- 8 cm), we have the approximations near the poles at v:::::; n,

r( -v + l

+ 1) 1

v- n

:::::;

:::::;

(- )nr

1 n r ! v - n' 21i21\;2 1

------

n 2M E - En ' 1 1

where nr

=n- l-

(13.209)

1. Hence

(13.210)

Let us expand the pole parts of the spectral representation of the radial fixed-energy amplitude in the form (13.211)

The radial wave functions defined by this expansion correspond to the normalized bound-state wave functions

'l/Jnlm(X)

=

~RnI(r)Yzm(X) . r '

(13.212)

960

13 Path Integral of Coulomb System

By comparing the pole terms of (13.206) and (13.211) [using (13.210) and formula (9.48) for the Whittaker functions, together with (9.50)], we identify the radial wave functions as 1 (n + l)! + I)! (n - l - I)! (2r /naH )1+le- r / naH M( -n + l + I , 2l + 2, 2r /naH) 1

Rnl(r)

a~2n

x

1

(2l

(n-l-1)!e- r / naH (2r/na )l+lL21+1 (2r/na) (13.213) (n + l)! H n- l- l H .

To obtain the last expression we have used formula (9.53).2 It must be noted that the normalization integrals of the wave functions Rnl(r) differ by a factor z/2n = (2r/naH)/2n from those of the harmonic oscillator (9.54) , which are contained in integral tables. However, due to the recursion relation for the Laguerre polynomials zL~(z)

= (2n + /1 + l)L~(z) - (n + /1)L~_l(Z) - (n + l)L~+1(z) ,

(13.214)

the factor z/2n leaves the values of the normalization integrals unchanged. The orthogonality of the wave functions with different n is much harder to verify since the two Laguerre polynomials in the integrals have different arguments. Here the group-theoretic treatment of Appendix 13A provides the simplest solution. The orthogonality is shown in Eq. (13A.28) . We now turn to the continuous wave functions. The fixed-energy amplitude has a cut in the energy plane for positive energy where", = -ik and v = i/aHk are imaginary. In this case we write v = iv'. From the discontinuity we can extract the scattering wave functions. The discontinuity is given by

In the second term, we replace (13.216) and use the relation, valid for argz E (-7r/2,37r/2) , 2/1

i=

-I , -2, -3, . . . ,

2Compare L.D. Landau and E .M. Lifshitz, Quantum Mechanics , Pergamon, London, 1965, p. 119. Note the different definition of our Laguerre polynomials L~ =[( _ )1' j(n + j.t)!]Ln+I'I'IL.L ..

13.9 Remarks on Geometry of Four-Dimensional up,-Space

961

to find

Mlf(-iv'+l+1)1 2 7rv' . . (2l + 1)!2 e Miv',l+1/2 (- 2zkrb) M _iv',1+1/2 (2zkra). (13.218) The continuum states enter the completeness relation as .

dISC (rblra)E,1 = nk

(13.219) [compare (1.326)]. Inserting (13.215) and replacing the continuum integral by the momentum integral 1':'00 dkkn/27r M , the continuum part of the completeness relation becomes

1000 dE /27rn

(13.220) with the radial wave functions

_ 0I f (-iv'+l+1)1 7rl//2 . V2; (2l + I)! e MiV',I+1/2( -2zkr) .

Rkl(r) -

(13.221)

By expressing the Whittaker function M)..,p,(z) in terms of the confluent hypergeometric functions, the Kummer functions M(a, b, z), as (13.222) we recover the well-known result of Schrodinger quantum mechanics: 3

R (r) = kl

01f( -iv' + l + 1)1 e7rv'/2eikr(_2ikr)l+1 M(-iv' + l + 1 2l + 2 -2ikr) . V2; (2l+1)! ' , (13.223)

13.9

Remarks on Geometry of Four-Dimensional uJL-Space

A few remarks are in order on the Riemann geometry of the it-space in four dimensions with the metric gp,v = 4U2op,v' As in two dimensions, the Cartan curvature tensor Rp,v)..1< vanishes trivially since ei 1'(it) is linear in it: (13.224)

In contrast to two dimensions, however, the Riemann curvature tensor Rp,v).. I< is nonzero. The associated Ricci tensor [see (10.41)] , has the matrix elements

(13.225) 3L.D. Landau and E.M. Lifshitz, op. cit., p. 120.

962

13 Path Integral of Coulomb System

yielding the scalar curvature

v>. 9 R = g Rv>. = - 2'114 .

(13.226)

In general, a diagonal metric of the form

(13.227) is called conformally fiat since it can be obtained from a fiat space with a unit metric conformal transformation a la Weyl

g/l-V = b/l-v by a

(13.228) Under such a transformation, the Christoffel symbol changes as follows:

(13.229) the subscript separated by a comma indicating a differentiation, i.e., n ,/lIn D dimensions, the Ricci tensor changes according to

R/l-V

--->

== 8/l-n.

n- 2 R/l- v - (D - 2)(n- 3 n;/l-V - 2n- 4 n,/l-n,v) -9/l-v9>'1< [(D - 3)n- 4 n,>.n,1
.",]

n- 2 R/l- v + (D - 2)n- 1 (n- 1 bv - 9/l-v(D - 2)-ln- D (n D - 2 );>'1'1

il[! dD~qngl/2(qn~]

f(qa) D{)OdS V27ritsf(qa)n/M 0 n=2

eiAtot/n,

(14.214)

V27ritsMn/M

with the total time-sliced action N+l

A tot =

L

~ot·

(14.215)

n=l

Each slice contains three terms ~ot

= Af + A:J + A;ot·

(14.216)

In the postpoint form, the first two terms were given in (13.158) and (13.159). They are equal to M g () A !' A .nr!' A f n2 (r!'!' v)2 . A f + AfJ -_ 2tf !'11 q I....l.q I....l.q - 22" !' vl....l.q - ts 8M II

II

(14.217)

The third term contains the effect of a potential and a vector potential as derived in (10.183). After the DK transformation, it reads (14.218)

14.6

Path Integral of the Dionium Atom

We now apply the generalized D-dimensional Duru-Kleinert transformation to the path integral of a dionium atom in three dimensions. This is a system of two particles with both electric and magnetic charges (el?gl) and (e2,g2) [12]. Its Lagrangian for the relative motion reads

L=

~ *2 + A(x)* -

V(x),

(14.219)

where x is the distance vector pointing from the first to the second article, M the reduced mass, V(x) a Coulomb potential e2 V(x) = - -, r

(14.220)

14.6 Path Integral of the Dionium Atom

1005

and A(x) the vector potential

A(x) = n/' x x r

(_1 ___ 1_) = nq (xy - yx)z. r- z r+z r(x 2 + y2)

(14.221)

The coupling constants are q == -(elg2 - e2g1)/nc and e2 == -(ele2 + glg2) . The vector potential (14.221) implies an obvious generalization of the magnetic monopole interaction (8.299) with an electric charge [recall Appendix lOA.3] The potenial If we take the coupling as and e2 == -ele2 - glg2 in (14.221) we allow for the two particles to carry both electric and magnetic charges of the two particles, if we take for V(x) the potential V(x)

e2

= - -.

(14.222) r The hydrogen atom is a special case of the dionium atom with el = -e2 = e and q = 0, lo = O. An electron around a pure magnetic monopole has el = e, g2 = g, e2 = gl = O. In the vector potential (14.221) we have made use of the gauge freedom A ----> A(x) + V' A(x) to enforce the transverse gauge V' A(x) = O. In addition, we have taken advantage of the extra monopole gauge invariance which allows us to choose the shape of the Dirac string that imports the magnetic flux to the monopoles. The field A(x) in (14.221) has two strings of equal strength importing the flux, one along the positive x 3 -axis from minus infinity to the origin, the other along the negative x 3 -axis from plus infinity to the origin. It is the average of the vector potentials (10A.59) and (lOA.60). For the sake of generality, we shall assume the potential V(x) to contain an extra 1/r2-potential: (14.223) The extra potential is parametrized as a centrifugal barrier with an effective angular momentum nlo. At the formal level, i.e., without worrying about path collapse and time slicing corrections, the amplitude has been derived in Ref. [13] . Here we reproduce the derivation and demonstrate, in addition, that the time slicing produces no corrections.

14.6.1

Formal Solution

We extend the action of the type (14.11) by a dummy fourth coordinate as in the Coulomb system and go over to it-coordinates depending on the radial coordinate u= and the Euler angles 0, r.p, r as given in Eq. (13.97). Then the action reads

vr

J{

M 2v?+2u M 4 [.02 + (/+'':?+2 (1'+ nMu q ) I.{; cos 0] - ue2 - 2M~4 n 2l 2 +E } . A= dt 24u 2 4

(14.224)

1006

14 Solution of Further Path Integrals by Duru-Kleinert Method

By performing the Duru-Kleinert time reparametrization dt = dsr(s) and changing the mass to fJ = 4M, the action takes the form

This can be rewritten in a canonical form

A

=

foS ds(puu' + Pe() + P..2i , where the sum over i is limited by power of >..2 up to which we want to carry the perturbation series; also [' is restricted to a finite number of terms only, because of the banddiagonal structure of the Il\l~. Extracting the coefficients of the power expansion in >.. from (15.179) we obtain all desired moments of the end-to-end distribution, in particular the second and fourth moments (15.155) and (15.164). Higher even moments are easily found with the help of a MATHEMATICA program, which is available for download in notebook form [5]. The expressions are too lengthy to be written down here. We may, however, expand the even moments (Rn) in powers of L /~ to find a general large-stiffness expansion valid for all even and odd n :

where

(15.199) with

Do = 3 (-561O+2921n-822n 2 +67n3 ) ,D1 = 8490+ 12103n-3426n2 +461n3 , D 2 =45

(-2-187n-46n 2 +7n3 )

,

D4

= 35 (-6+31n+30n 2 +5n3 )

.

(15 .200)

The lowest odd moments are, up to order [4,

Y!:l

= 1_ ~

L

(R3 J = 1- ~ £3 2

15.9.3

+

5D-7 Z2 _ 33 - 43D + 14D2 Z3 _ 861 - 1469D + 855D 2 - 175D3 Z4 180(D-1) 3780(D _1)2 453600(D _1) 3

+

5D-4 Z2 _ 195-484D+329D2 Z3 _ 609-2201D+2955D 2 -1435D 3 Z4 30 (D-1) 7560(-1+d)2 151200 (D_1)3 ....

6

From Moments to End-to-End Distribution for D=3

We now use the recursively calculated moments to calculate the end-to-end distribution itself. It can be parameterized by an analytic function of r = R/ L [4]: (15.201)

1048

15 Path Integrals in Polymer Physics

whose moments are exactly calculable:

(15.202)

We now adjust the three parameters k, /3, and m to fit the three most important moments of this distribution to the exact values, ignoring all others. If the distances were distributed uniformly over the interval r E [0,1], the moments would be (r 21 )unif = 1/(2l + 2) . Comparing our exact moments (r21) (~) with those of the uniform distribution we find that (r21)(~)/(r21)unif has a maximum for n close to nmax(~) == 4~/ L. We identify the most important moments as those with n = nmax(~) and n = nmax(~) ± 1. If nmax(~) :-=:; 1, we choose the lowest even moments (r2), (r4), and (r 6). In particular, we have fitted (r2), (r 4) and (r 6) for small persistence length ~ < L/2. For ~ = L/2, we have started with (r4), for ~ = L with (r 8) and for ~ = 2L with (r 16 ), including always the following two higher even moments. After these adjustments, whose results are shown in Fig. 15.4, we obtain the distributions shown in Fig.15.6 for various persistence lengths~. They are in excellent agreement with the Monte Carlo data (symbols) and better than the one-loop perturbative results (thin curves) of Ref. [6], which are good only for very stiff polymers. The MATHEMATICA program to do these fits are available from the internet address given in Footnote 5.

. ' 1

k

17.5 15 12.5 10 7.5 5 2.5

8

f3 ... . .. .

ID

40

20 0.5

2

•. . ... . .

....

.....

0.5

22. 20 18

m

16 . 14 12

2

....

0.5

2

Figure 15.4 Paramters k, (3, and m for a best fit of end-to-end distribution (15.201). For small persistence lengths ~/ L = 1/400, 1/100, 1/30, the curves are well approximated by Gaussian random chain distributions on a lattice with lattice constant aeff = 2~, i.e., PL(R) ---+ e- 3R2/4L{ [recall (15.75)]. This ensures that the lowest moment (R2) = aeffL is properly fitted. In fact , we can easily check that our fitting program yields for the parameters k, /3, m in the end-to-end distribution (15.201) the ~ ---+ behavior: k ---+ -~ , /3 ---+ 2 + 2~, m ---+ 3/4~, so that (15.201) tends to the correct Gaussian behavior. In the opposite limit of large~, we find that k ---+ 1O~ -7/2, /3 ---+ 40~ +5, m ---+ 10, which has no obvious analytic approach to the exact limiting behavior PL(R) ---+ (1r )- 5/2 e- l/4{(I - r) , although the distribution at ~ = 2 is fitted numerically extremely well.

°

1049

15.9 Schrodinger Equation and Recursive Solution for Moments

The distribution functions can be inserted into Eq. (15.89) to calculate the structure factors shown in Fig. 15.5. They interpolate smoothly between the Debye limit (15.90) and the stiff limit (15.105).

~/L= 2

S(q)

~/L= 1 ~/L = 1/2

0.6 0.4

~/ L =

1/5, ... ,1/400

0.2

+--~5---'1~0---'1~ 5 --::'20:--25:---3:'"::0 q../l, Figure 15.5 Structure functions for different persistence lengths ~/ L = 1/400,1/100, 1/30,1/10, 1/5, 1/2, 1, 2, (from bottom to top) following from the end-to-end distributions in Fig. 15.6. The curves with low ~ almost coincide in this plot over the ~-dependent absissa. The very stiff curves fall off like l/q, the soft ones like 1/q2 [see Eqs. (15.105) and (15.90)].

15.9.4

Large-Stiffness Approximation to End-to-End Distribution

The full end-to-end distribution (15.132) cannot be calculated exactly. It is, however, quite easy to find a satisfactory approximation for large stiffness [6]. We start with the expression (15.170) for the end-to-end distribution PL(R). In Eq. (3.230) we have shown that a harmonic path integral including the integrals over the end points can be found, up to a trivial factor, by summing over all paths with Neumann boundary conditions. These are satisfied if we expand the fields u( s) into a Fourier series of the form (2.450): 00

u(s) = Uo +'TJ(s) = Uo + L

UnCOSllnS,

lin

= n7f/L.

(15.203)

n=l

Let us parametrize the unit vectors U in D dimensions in terms of the first D - 1 -dimensional coordinates ul-' == ql-' with f1 = 1, .. . , D - 1. The Dth component is then given by a power series (15.204) The we approximate the action harmonically as follows: -

A

A(O) -

+ A int =!5'.

L

f ds [u'(sW

2 io

L

1

1

+ -8(0) log(1 _ 2

L

~!5'.f ds[q'(sW- - 8(0)f dsq2. 2

io

2

q2)

io

(15.205)

1050

15 Path Integrals in Polymer Physics

The last term comes from the invariant measure of integration dD- 1q/ ~ [recall (10.636) and (10.641)]. Assuming, as before, that R points into the z, or Dth, direction we factorize

8(D)

(R-

foL dSU(S)) = x

where R

8

8(D-l)

{~q2(S) + ~[q2(SW + ... })

+ foL ds

(R - L

(foL dsq(s)) ,

(15.206)

== IRI. The second 8-function on the right-hand side enforces q=L-1 fo L dsql"(s) =0,

jl=1, . .. ,d-1,

(15.207)

and thus the vanishing of the zero-frequency parts ql{ in the first D - 1 components of the Fourier decomposition (15.203). It was shown in Eqs. (10.632) and (10.642) that the last 8-function has a distorting effect upon the measure of path integration which must be compensated by a Faddeev-Popov action FP

Ae =

D -1 r 2 ----u;io ds q , L

(15.208)

where the number D of dimensions of ql"-space (10.642) has been replaced by the present number D - 1. In the large-stiffness limit we have to take only the first harmonic term in the action (15.205) into account, so that the path integral (15 .170) becomes simply

PL(R) ex

r

iNBC

V 1D - 1q 8(R-L+

rL ds~q2(S))

io

e-(K./2) iOL ds [q'(s) ]2 .

2

(15.209)

The subscript of the integral emphasizes the Neumann boundary conditions. The prime on the measure of the path integral indicates the absence of the zero-frequency component of ql"(s) in the Fourier decomposition due to (15.207). Representing the remaining 8-function in (15.209) by a Fourier integral, we obtain

PL(R) ex K

1

ioo

-ioo

dw2 - 2 (L- R) _.el-, and calculating Z(K,). An integral J dK,e-im>'Z(A) will then select any specific linking number m. But a phase factor eim >. is simply produced in the partition function (16.178) by attaching to one of the current couplings in the interaction (16.172), say to that of C2 , a factor A, thus changing (16.172) to (16.179) The A-dependent partition function is then

(16.180) Ultimately, we want to find the probability distribution of the linking numbers m as a function of the lengths of C 1 and C2 . The solution of this two-polymer problem may be considered as an approximation to a more interesting physical problem in which a particular polymer is linked to any number N of polymers, which are effectively replaced by a single long "effective" polymer [9]. Unfortunately, the full distribution of m is very hard to calculate. Only a calculation of the second topological moment is possible with limited effort. This quantity is given by the expectation value (m 2 ) of the square of the linking number m . Let PL"L2(xl , x2 ;m) be the configurational probability to find the polymer C 1 of length L1 with fixed coinciding endpoints at Xl and the polymer C 2 of length L2 with fixed coinciding

1129

16.8 Entangled Pair of Polymers

endpoints at X2, entangled with a Gaussian linking number m. The second moment (m 2 ) is given by the ratio of integrals 2

(m) =

J d3Xld3X2 r~::: dm m 2P L1 ,L2(Xl , X2 ; m) +00 ' J d3Xld3X2 1-00 dmPL" L2(Xl,X2; m)

(16.181)

performed for either of the two probabilities. The integrations in d3xld3x2 covers all positions of the endpoints. The denominator plays the role of a partition function of the system:

z ==

J

d3Xld3X21:00 dmPL1,L2(xl,x2 ;m).

(16.182)

Due to translational invariance of the system, the probabilities depend only on the differences between the endpoint coordinates: (16.183)

Thus, after the shift of variables, the spatial double integrals in (16.181) can be rewritten as

J

d3xld3x2PL l, L2(xl,x2 ;m) = V

J

d 3xPL" L2(x;m) ,

(16.184)

where V denotes the total volume of the system.

16.8.1

Polymer Field Theory for Probabilities

The calculation of the path integral over all line configurations is conveniently done within the polymer field theory developed in Section 15.12. It permits us to rewrite the partition function (16.180) as a functional integral over two 7f>fl(Xl) and 7f>~2(X2) with nl and n2 replica (al = 1, ... , nl, a2 = 1, ... , n2)' At the end we shall take nl, n2 - t 0 to ensure that these fields describe only one polymer each, es explained in Section 15.12. For these fields we define an auxiliary probability Pz(Xl, X2 ; i\) to find the polymer C l with open ends at Xl, x~ and the polymer C2 with open ends at X2 ' X~, The double vectors Xl == (xl,xD and X2 == (X2'X~) collect initial and final endpoints of the two polymers C l and C 2 • The auxiliary probability Pz(Xl, X2; i\) is given by a functional integral (16.185)

where V(fields) indicates the measure of functional integration, and A the total action (16.180) governing the fluctuations. The expectation value is calculated for any fixed pair (aI , (2) of replica labels, i.e., replica labels are not subject to Einstein's summation convention of repeated indices. The action A consists of kinetic terms for the fields , a quartic interaction of the fields to account for the fact that two monomers of the polymers cannot occupy the same point, the so-called excludedvolume effect, and a Chern-Simons field to describe the linking number m . Neglecting at first the excluded-volume effect and focusing attention on the linking problem only, the action reads

A = ACS12 +

Ae,curr

+ Apol + AGF,

(16.186)

with a polymer field action 2

~J d3 X [Apol = L.. ID i7f>il 2 + m i2 l\[Jil 2] .

(16.187)

i=l

They are coupled to the polymer fields by the covariant derivatives Di=V+i"YiAi,

(16.188)

1130

16 Polymers and Particle Orbits in Multiply Connected Spaces

with the coupling constants 11,2 given by 11 = 1,

(16.189)

12 = >..

The square masses of the polymer fields are given by m~ = 2Mzi .

(16.190)

where M = 3/a, with a being the length of the polymer links [recall (15.79)]' and Zi the chemical potentials of the polymers, measured in units of the temperature. The chemical potentials are conjugate variables to the length parameters L1 and L 2 , respectively. The symbols Wi collect the replica of the two polymer fields (16.191) and their absolute squares contain the sums over the replica ni

IDi~i12 =

L Cti=l

IDi1jJf' 12,

IWil2 =

ni

L

l1jJf' 12.

(16.192)

Cti=l

Having specified the fields , we can now write down the measure of functional integration in Eq. (16.185): (16.193) By Eq. (16.180) , the parameter>' is conjugate to the linking number m . We can therefore calculate the probability P L" L2(:X\' X2 ; m) in which the two polymers are open with different endpoints from the auxiliary one Pz (X1 , X2 ; >.) by the following Laplace integral over = (Zl , Zl) :

z

(16.194)

16.8.2

Calculation of Partition Function

Let us use the polymer field theory to calculate the partition function (16.182). By Eq. (16.194) , it is given by the integral over the auxiliary probabilities

Z= (16.195) The integration over m is trivial and gives 27r8(>'), enforcing >. = 0, so that (16.196)

To calculate Pz (Xl, X2 ; 0) , we observe that the action A in Eq. (16.186) depends on >. only via the polymer part (16.187), and is quadratic in >.. Let us expand A as (16.197)

1131

16.8 Entangled Pair of Polymers with the ,X-independent part

(16.198)

a linear coefficient (16.199)

containing a pseudo-current of the second polymer field (16.200)

and a quadratic coefficient (16.201)

With these definitions we write with the help of (16.198) : Pz(Xl, X2 ; 0) =

J

V(fields) e-Ao¢fl (Xl)¢~"l (XD¢~2(X2)¢~2(X').

(16.202)

In the action (16.198), the fields \[12, \[12 are obviously free , whereas the fields \[11, \[Ii are apparently not because of the couplings with the Chern-Simons fields in the covariant derivative Dl . This coupling is, however, without physical consequences. Indeed, by integrating out A~ in (16.202), we find from ACS12 the flatness condition:

v

Al

X

=

O.

(16.203)

On a flat space with vanishing boundary conditions at infinity this implies Al = O. As a consequence, the functional integral (16.202) factorizes as follows [compare (15.370)] (16.204)

where Go(X; - x; ; Zi) are the free correlation functions of the polymer fields: (16.205)

In momentum space, the correlation functions are (16.206)

such that G ( ') o Xi - Xi ; Zi

=

J

d3 k ik-x (2'71-)3 e

k; +1 m;'

(16.207)

and

1 C

+iOO Mdz· -2-.' ez,L'Go(X' - x'·t' z·) t ~

c - ioo

~ 2

(

~) 411o£i

7T't

3/2

e- M (x , -x;l/2L, .

(16.208)

1132

16 Polymers and Particle Orbits in Multiply Connected Spaces

The partition function (16.196) is then given by the integral

Z=27rJd3Xld3X2 lim

X~ - Xl

Co(xl - x~;LI)CO(X2 - X~;L2) '

(16.209)

X; - X2

The integrals at coinciding endpoints can easily be performed, yielding (16.210) It is important to realize that in Eq. (16.195) the limits of coinciding endpoints x; --t Xi and the inverse Laplace transformations do not commute unless a proper renormalization scheme is chosen to eliminate the divergences caused by the insertion of the composite operators 1'¢"(x)12. This can be seen for a single polymer. If we were to commuting the limit of coinciding endpoints with the Laplace transform, we would obtain

1

C+ioo

c - ioo

dz -2 e zL ~im Co(x-x';z) = 1r

x

1

C+ioo

c - ioo

--+ X

dz -2.ezLCo(O,z),

(16.211)

1n

where

CoCo; z)

=

(16.212)

(1'¢(xW)·

This expectation value, however, is linearly divergent: (16.213)

16.8.3

Calculation of Numerator in Second Moment

Let us now turn to the numerator in Eq. (16.181):

N == J d3xld3x2

i:

(16.214)

dm m 2 PL"L 2 (xI , x2;m).

We shall set up a functional integral for N in terms of the auxiliary probability Pz(XI, X2 ; 0) analogous to Eq. (16.195) . First we observe that N

-_

J d3Xl d3r2

1

00

i:

-00

X

eZlLl+ZzLz

dm m 2 I'1m x~ -

Xl

1

CTi oo

c-ioo

i 2 M-dZ M-dZ ..27ft

27r1,

X~ -X2

d)..e- im >'PZ(XI , X2 ; )..).

(16.215)

The integration in m is easily performed after noting that (16.216) After two integrations by parts in ).., and an integration in m , we obtain

(16.217)

1133

16.8 Entangled Pair of Polymers Performing the now the trivial integration over A yields

(16.218)

To compute the term in brackets, we use again (16.197) and Eqs. (16.198)- (16.223) , to find

N

=

x x

J [(J

D(fields) exp( - Ao)I1f>r' (xl)1 211f>g2 (X2W

d3 xA2 · \I!;V\I!2r

+

~ Jd3xA~ 1\I!212].

(16.219)

In this equation we have taken the limits of coinciding endpoint inside the Laplace integral over Zl , Z2 . This will be justified later on the grounds that the potentially dangerous Feynman diagrams containing the insertions of operations like l\I!il 2 vanish in the limit nl , n2 --> O. In order to calculate (16.219) , we decompose the action into a free part (16.220)

and interacting parts (16.221)

with a "current" of the first polymer field (16.222)

and (16.223)

Expanding the exponential eAo =

eAg+A~+A~

= eAo

[1- A6+ (~)2 -A5+ ... ] ,

(16.224)

and keeping only the relevant terms, the functional integral (16.219) can be rewritten as a purely Gaussian expectation value

N = x x x

J [(J [(J

D(fields) exp( - A8)I1f>r'(XIWI 1f>g2(X2W

d3xA 1



\I!~V\I!lr+ ~ Jd3xAi 1\I! 12]

d3xA 2 · \I!;V\I!2r+

1

~ Jd3xA~ 1\I! 212] .

(16.225)

1134

16 Polymers and Particle Orbits in Multiply Connected Spaces

+

§

+

+

Figure 16.23 Four diagrams contributing in Eq. (16.225) . The lines indicate correlation functions of wi-fields. The crossed circles with label i denote the insertion of IWi(Xi)12 . Note that the initially asymmetric treatment of polymers C 1 and C 2 in the action (16.187) has led to a completely symmetric expression for the second moment. Only four diagrams shown in Fig. 16.23 contribute in Eq. (16.225). The first diagram is divergent due to the divergence of the loop formed by two vector correlation functions. This infinity may be absorbed in the four-w interaction accounting for the excluded volume effect which we do not consider at the moment. We now calculate the four diagrams separately.

16.8.4

First Diagram in Fig. 16.23

From Eq. (16.225) one has to evaluate the following integral (16.226)

As mentioned before, there is an ultraviolet-divergent contribution which must be regularized. The system has, of course, a microscopic scale, which is the size of the monomers. This, however, is not the appropriate short-distance scale to be uses here. The model treats the polymers as random chains. However, the monomers of a polymer in the laboratory are usually not freely movable, so that polymers have a certain stiffness. This gives rise to a certain persistence length ~o over which a polymer is stiff. This length scale is increased to ~ > ~o by the excluded-volume effects. This is the length scale which should be used as a proper physical short-distance cutoff. We may impose this cutoff by imagining the model as being defined on a simple cubic lattice of spacing~. This would, of course, make analytical calculations quite difficult. Still, as we shall see, it is possible to estimate the dependence of the integral N1 and the others in t he physically relevant limit in which the lengths of the polymers are much larger than the persistence length ~ . An alternative and simpler regularization is based on cutting off all ultraviolet-divergent continuum integrals at distances smaller than~. After such a regularization, the calculation of N1 is rather straightforward. Replacing the expectation values by the Wick contractions corresponding to the first diagram in Fig. 16.23, and performing the integrals as shown in Appendix 16A, we obtain N1

:

x

11

(~;6 (L 1L 2)- ! dt [(1 -

t)tj- ~

11

J

ds [(1 - s)sj - !

d 3ye- My2 /2t(1-t)

J Jd3x~ Ix~14 '

d3xe - MX2/2s(1 -s)

(16.227)

The variables x and y have been rescaled with respect to the original ones in order to extract the behavior of N1 in L1 and L 2. As a consequence, the lattices where x and yare defined have now spacings ~ /.;r;; and ~ /,;r;; respectively.

1135

16.8 Entangled Pair of Polymers

The x, y integrals may be explicitly computed in the physical limit L 1, L2 » ~ , in which the above spacings become small. Moreover, it is possible to approximate the integral in x~ with an integral over a continuous variable I and a cutoff in the ultraviolet region:

J

3

d

/I

Xl

1

Ix~ 14

(16.228)

After these approximations, we finally obtain

N = V7l'1/2~(L L )- 1/2 c - 1 1

16.8.5

(47l')3

1

2

(16.229)

.".

Second and Third Diagrams in Fig. 16.23

Here we have to calculate

(16.230) The above amplitude has no ultraviolet divergence, so that no regularization is required. The Wick contractions pictured in the second Feynman diagrams of Fig. 16.23 lead to the integral

N 2 = -4 v:t. '2VL-l/2L-lM3l1 dtlt dt'C(t ' t') , 2 1 6 7l'

0

(16.231)

0

where C(t, t') is a function independent of Ll and L 2:

[(1- t)t'(t - t,)] - 3/2

C(t, t') x

_ My' /2t') ( ~j v ye

(~;

J

v xe

d3xd3yd3ze- M (y - X)' /2(1-t)

- Mx' /2(t -t'))

[8;jz, (z + x) - (z

+ X); Zj ]

IzI31z + xl 3

.

(16.232)

As in the previous section, the variables x , y, Z have been rescaled with respect to the original ones in order to extract the behavior in L 1 . If the polymer lengths are much larger than the persistence length one can ignore the fact that the monomers have a finite size and it is possible to compute C(t, t') analytically, leading to (16.233) where K is the constant

K

1 (5 1) 1) 1 (9 1) = "61 B (32' 21) + 2B 2' 2 - B (72' 2 + SB 2' 2

=

1971' 384 "" 0.154,

(16.234)

and B(a, b) = r(a)r(b)jr(a + b) is the Beta function. For large h ---t 00, this diagram gives a negligible contribution with respect to N 1 . The third diagram in Fig. 16.23 give the same as the second, except that Ll and L2 are interchanged. (16.235)

1136

16 Polymers and Particle Orbits in Multiply Connected Spaces

16.8.6

Fourth Diagram in Fig. 16.23

Here we have the integral

_41i:2~

2

lim

j C+ioo

_0 n2- 0

c - 't O.

(16A.3)

Contracting the fields in Eq. (16.226) , and keeping only the contributions which do not vanish in the limit of zero replica indices, we find with the help of Eqs. (16A.l) and (16A.2):

J

Nl

d3Xl,d3X21Ll ds 1

GO(X2 -

X

L2 dt

X~ ; t)Go(x~ -

Jd3x~d3x~ GO(XI -X~; s)Go(x~

X2; L2 - t)

I ,

Xl -X 2 4

1

'

1

- Xl ; Ll - s)

(16A.4)

'

Performing the changes of variables

,

,

s

s

=

Ll

t

t

'

=

(16A.5)

L2 '

and setting x~ == x~ - x~ , we easily derive (16.227) . For small ~/,;r;; and ~/../L2 , we use the approximation (16.228). The space integrals can be done using the formula (16A.3) . After some work we obtain the result (16.240) . For the amplitude N2 in Eq. (16.230) we obtain likewise the integral N2 = X

X

J Jd3x~d3x~d3x~ [l 1 ds'Go(x~ Dik(X~ x2)Djk(X~ X~) [l d3xld3x2 L1

8

ds

- Xl ; Ll - s)

-

-

L2

V~;,GO(X1 - X~; s')V~; Go(x~ - X~; s X~; L2 - t)Go(X~ -

dtG O(X2 -

x2; t)] ,

s')]

(16A.6)

where Dij(X, x') are the correlation functions (16.174) and (16.175) ofthe vector potentials. Setting X2 == ../L2u + X2 and supposing that ~/../L2 is small, the integral over u can be easily evaluated with the help of the Gaussian integral (16A.3). After the substitutions x~ = ,;r;;y + Xl X~ = ,;r;;(y - X) + Xl , X2 = ,;r;;(y - X - z) +Xl and a rescaling of the variables s, s' by a factor Ll1, we derive Eq. (16.231) with (16.232). For small ~/,;r;;, ~ /../L2, the spatial integrals are easily evaluated leading to: N2

=

- v'2VL- 1/2L- lM- 1/21l 2

(4n)6

After the change of variable t' the type c( n, m) =

-t

1

1

1t

dt't'(I-t)

0

V

t - t' . I - t + t'

(16A.7)

til = t - t', the double integral is reduced to a sum of integrals

1 1t o

dt

0

dttm

dt't ln

0

{6' 1- t

--,,

m, n = integers .

These can be simplified by replacing t m by dt m+1/dt(m + 1) , and doing the integrals by parts. In this way, we end up with a linear combination of integrals of the form:

1 1

o

dt t"+!

vr=t

= B

(/\;+ ~)). 22

The calculations of N3 and N4 are very similar, and are therefore omitted.

(16A.8)

Appendix 16B Kauffman and BLM/Ho polynomials

Appendix 16B

1153

Kauffman and BLM/Ho Polynomials

The Kauffman polynomials are given by F(a,x) = a- WA(a,x), where w is the writhe and A(a,x) satisfies the skein relation (16B.1) The subscripts refer to the same loop configurations as in Figs. 16.10 and 16.12. The trivial loop has

a+ a-I A(a,x) = - - - l . z

(16B.2)

While the Kauffman polynomial is a knot invariant, the A-polynomial is only a ribbon invariant.12 If a winding LT+ or L T - is removed from a loop with the help of a Reidemeister move of type I in Fig. 16.6 (which for infinitely thin lines would be trivial while changing the writhe of a ribbon by one unit) then A(a, x) receives a factor a or a- I, respectively (see Fig. 16.25).

>

a x

) L TO

.-'x) L TO

Figure 16.25 Trivial windings LT+ and LT- . Their removal by means of Reidemeister move of type I decreases or increases writhe w by one unit. The Kauffman polynomials arise from Wilson loop integrals of a nonabelian Chern-Simons theory, if the action (16.314) is SO(N)- rather than SU(N)-symmetric. A list of these polynomials can be found in papers by Lickorish and Millet and by Doll and Hoste quoted at the end of the chapter. The BLMHo polynomials are special cases of the Kauffman polynomials. The relation between them is Q(x) == F(l, x).

Appendix 16C

Skein Relation between Wilson Loop Integrals

Here we sketch the derivation of the skein relation (16.323) for the expectation values of Wilson's loop integrals (16.319). Let us decompose Ai in terms of the N 2 - 1 generators Ta of the group

SO(N) : (16C.1) They satisfy the commutation rules

(16C.2) I2More precisely, F( a, x) is invariant under the three Reidemeister moves which, in the projected picture of the knot in Fig. 16.6, define the ambient isotopy, whereas A changes under the first Reidemeister move, associated only with regular isotopy. whereas

1154

16 Polymers and Particle Orbits in Multiply Connected Spaces

For simplicity, we assume k to be very large so that we can restrict the treatment to the lowest order in 11k. To avoid inessential factors of the constants e, c, n, we set these equal to 1. Under a small variation of the fields one has (16C.3) where the path-ordering operator P arranges the expression to its right in such a way that Ta is situated in WL at the correct path-ordered place. To emphasize this, we have recorded the position of Ta by means of an x-argument. More precisely, if we discretize the loop integral and write (16C.4) where xn are the midpoints of the intervals ~xn , a differentiation with respect to one of the Ai(xn)-fields replaces the associated factor eiA,(xn)L'.x~ by iTaeiA,(Xn)L'.x;' . With the ,,-function on a line L defined in Eq. (10A.8), we write (16C.3) as (16C.5) For simplicity, we assume x to be only once traversed by the loop L. If the shape of the loop is deformed infinitesimally by dSi = Eijkdxid'Xj, then W L changes by (16C.6)

"WL = idxid'xj PFiJ(x)Ta(x)WL ,

where FiJ are the N 2 - 1 components of the nonabelian field strengths (16C.7)

Fij = OiAj - OjAi - i[Ai , Aj]

and x the midpoints of the parallelograms spanned by dx and d'x. The derivation of Eq. (16C.6) is based on the observation that a change of the path by a small parallelogram adds to the line integral WL a factor Wo' which is a Wilson loop integral around the small parallelogram. The latter is evaluated as follows: W

0

eiA, (x - d'X/2)dx, eiAj (x+dx/2)d' Xj e - iA, (x+d'X/2)dx, e - iAj (x - dx /2)d' Xj ei [A, (x)dx , -OJ A, (x)dx ,d' Xj +... ]ei[Aj (x)d' Xj +o,A j (x)dx ,d' Xj +... ]

x

e - i [A,(X)dx,+oj A,(X)dx,d'xj+"']e-i[Aj(X)d'xj - o,Aj(X)dx, d'x j+ ... ] eiFij( x)dxid'xj.

(16C.8)

The last line is found with the help of the Baker-Hausdorff formula eAe B = e A+ B+ [A,BI! 2+ ... (recall Appendix 2A).

W

Let us denote the Chern-Simons functional integral over WdA ] by L. Their N x N-traces are WdA] and WL . The latter differs from the expectation (WdA]) in (16.320) by not containing the normalizing denominator, i.e. , (16C.9) This changes under the loop deformation by

J

DA"Wd A]e- Ae.cs

idxid'Xj

J

DAFiJ(x)Ta(x)WdA] e-Ae,Cs ,

(16C.1O)

1155

Appendix 160 Skein Relation between Wilson Loop Integrals

with the tacit agreement that a generator Ta(x) written in front of the trace has to be evaluated within the trace at the correct path-ordered position. Now we observe that F;'j can also be obtained by applying a functional derivative to the Chern-Simons action (16.314) : .47r

8A."cs

ZTfi jk

Mk(x)

a

(16C.11)

= Fij(X) ,

This allows us to rewrite (16C.1O) as

-~ J J ~J VA

and further as

dSi :~(~ Ta(x)WdA]e-Ae,es

VAdSiTa(x)WL M:(X) e-Ae,es.

A partial functional integration produces

which brings the variation to the form

-

8W L =

~J VAdSi8i(x,L)Ta(x)Ta(x)WdA]e - A .,es, -T

(16C.12)

The expectation of Wilson's loop integral W L changes under a deformation only if the loop crosses another line element. This property makes W L a ribbon invariant, i.e., an invariant of regular isotopy. For a finite deformation, the right-hand side has to be integrated over the area S across which the line has swept. Using the integral formula

is

dSi8i(x , L) = {

~

} if the line L {

:~:~:: ~

},

(16C.13)

(16C.14) The two generators Ta(x) lie path-ordered on the different line pieces of the crossing. To establish contact with the knot polynomials, the left-hand sides have been labeled by the loop subscripts L+ and L _ appearing in the skein relations of Fig. 16.114. The product of the generators on the right-hand side is the Casimir operator of the N x N -representation of SO(N):

(Ta)"/3(Ta ),),o =

1

'2 8,,0 8/3')' -

1 2N 8"/38,),0 '

When inserted into Eq. (16C.14) , we obtain the graphical relation: a

I

X"'(3

D

The second graph on the right-hand side can be decomposed into

(16C.15)

1156

16 Polymers and Particle Orbits in Multiply Connected Spaces

Taking these two terms to the left-hand side of (16C.14) , we obtain the skein relation 7ri ) ( 7ri ) 27ri( 1- Nk WL+ - 1+ Nk WL _ = -TWLo .

(16C.16)

We now apply this relation to the windings displayed in Fig. 16.25. They decompose into a line and a circle. Due to the trace operation in W L o ' the circle contributes a factor N . Thus we obtain the relation (16C.17) Now we remove on the left-hand side the windings according to the graphical rules of Fig. 16.25. Under this operation, the Wilson loop integral is not invariant. Like BLMHo polynomials, it acquires a factor a or a - 1: (16C.18) To be compatible with (16C.17), the parameter a must satisfy 7ri( a=I - N2 - 1)

Nk

'

7ri (2 ) a -1 = 1+ Nk N - 1.

(16C.19)

Due to (16C.18) , the Wilson loop integral is only a ribbon invariant exhibiting regular isotopy. A proper knot invariant which distinguishes ambient isotopy classes arises when multiplying W L by a- w . The polynomials HL == e-wW L satisfy the skein relation (16C.20) The prefactors on the left-hand side can be written for large k as 1 - 27riN/k ~ qN/2 and 1 + 27riN/k ~ q- N/2 with q = 1 - 27ri7r /k. The prefactor on the right-hand side is equal to q1/2 q-1/2 . To leading order in l/k, we have t hus derived the skein relation (16.323) for the HOMFLY polynomials H L .

Appendix 16D

London Equations

Consider an ideal fluid of charged particles. By definition, it is non-viscous and incompressible, satisfying v· v = o. If the charge of the particles is e (which we take to be negative for electrons) , the electric current density is j = pev,

(16D.l)

where p is the particle density. The current is obviously conserved. The equation of motion of the particles in an electric and magnetic field is governed by the Lorentz force and reads (16D.2) Using the kinematic identity

(1 2) - vx(Vxv)

dv= -{)v ov + ( v · V)v=-+V -v dt 8t 8t 2

'

(16D.3)

1157

Appendix 16D London Equations this leads to the partial differential equation for the velocity field v(x, t) M

C;;; + V ( ~ v 2 )

=

eE

+M v

.

(V x

v

+;

c B) .

(16D.4)

Consider the time dependence of the vector field on the right-hand side (16D.5) Using Maxwell's equation {)

8tB =

-cV x E ,

(16D.6)

V x (V x X) .

(16D.7)

we derive {) {)tX =

Suppose now that there is initially no B-field in the ideal fluid at rest which therefore has X == 0 everywhere. If a magnetic field is turned on, Eq. (16D.7) guarantees that X remains zero at all times. This implies that



pe 2 B Me '

(16D.8)

VXJ=- -

which is the first London equation. By inserting the first London equation into Eq. (16D.4) , we find the second London equation (16D.9) If the vector potential is taken in the transverse gauge V· A = 0 (which in this context is also called London gauge), then the first London equation can be solved and yields 2

J'= - ~A Me .

(16D.1O)

By inserting this equation into the Maxwell equation with no electric field E , we obtain

V x B = 47r j = _ p47re 2 A . c Mc2

(16D.ll)

When rewritten in the form

V x (V x A)

+ ),- 2 A

=

0,

(16D.12)

with ), - 2 =

p47re 2 Mc2'

(16D.13)

the equation exhibits directly the finite penetration depth ), of a magnetic field into the fluid, the celebrated Meissner effect. It is the ideal manifestation of the Lenz rule, according to which an incoming magnetic field induces currents reducing the magnetic field - in the present case to extinction.

1158

Appendix 16E

16 Polymers and Particle Orbits in Multiply Connected Spaces

Hall Effect in Electron Gas

A gas of electrons with a density p carries an electric current (16E.1)

j = pev .

In a magnetic field, the particle velocities change due to the Lorentz force by Mil = e (

E+ ~v B) . x

(16E.2)

If lTo denotes the conductivity of the system without a magnetic field, the electric current is obviously given by j

E+ ~v x B) (E+ _l_j B).

lTo ( lTo

x

pec

(16E.3)

The second term shows the classical Hall resistance (16.286) .

Notes and References For the Aharonov-Bohm effect, see the original work by Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959). For a review see S. Ruijsenaars, Ann. Phys. (N.Y.) 146, 1 (1983). See also the papers A. Inomata and V.A. Singh, J . Math. Phys. 19, 2318 (1978); E. Corinaldesi and F . Rafeli, Am. J. Phys. 46, 1185 (1978); M.V. Berry, Eur. J . Phys. 1 , 240 (1980) ; S. Ruijsenaars, Ann. Phys. 146, 1 (1983); G. Morandi and E . Menossi, Eur. J. Phys. 5, 49 (1984) ; R. Jackiw, Ann. Phys. 201 , 83 (1990); and in "M. A. B. Beg Memorial Volume" (A. Ali and P. Hoodbhoy, Eds.), World Scientific, Singapore, 1991; G. Amelino-Camelia, Phys. Lett. B 326, 282 (1994) ; Phys. Rev. D 51, 2000 (1995) ; C. Manuel and R. Tarrach, Phys. Lett. B 328, 113 (1994); S. Ouvry, Phys. Rev. D 50, 5296 (1994); C.R. Hagen, Phys. Rev. D 31, 848 (1985); D 52 2466 (1995); P. Giacconi, F . Maltoni, and R. Soldati, Phys. Rev. D 53, 952 (1996) ; R. Jackiw and S.-Y. Pi, Phys. Rev. D 42, 3500 (1990) ; O. Bergman and G. Lozano, Ann. Phys. (NY) 229, 416 (1994); M. Boz, V. Fainberg, and N.K. Pak, Phys. Lett. A 207,1 (1995) ; Ann. Phys (N.Y.) 246, 347 (1996) ; M. Gomes, J.M.C. Malbouisson, and A.J. da Silva, Phys. Lett. A 236, 373 (1997) ; Int. J . Mod. Phys. A 13, 3157 (1998); (hep-th/ 0007080) . Path integrals in multiply connected spaces and their history are discussed in the textbook L.S. Schulman, Techniques and Applications of Path Integration , Wiley, New York, 1981. Details on Lippmann-Schwinger equation is found in most standard textbooks, say S.S. Schweber, Relativistic Quantum Field Theory , Harper and Row, New York, 1961, Section lIb. In chemistry, the properties of self-entangled polymer rings, called catenanes , were first investigated by

Notes and References

1159

H.L. Frisch and E. Wasserman, J. Am. Chern. Soc. 83, 3789 (1961). Their existence was proved mass-spectroscopically by R. Wolovsky, J. Am. Chern. Soc. 92, 2132 (1961) ; D.A. Ben-Efraim, C. Batich, and E . Wasserman, J . Am. Chern. Soc. 92, 2133 (1970). In optics, the Kirchhoff diffraction formula can be rewritten as a path integral formula with linking terms: J .H. Hannay, Proc. Roy. Soc. Lond. A 450,51 (1995), In biophysics, J.C. Wang, Accounts Chern. Res. D 10, 2455 (1974) showed that DNA molecules can get entangled and must be disentangled during replication. The path integral approach to the entanglement problem in polymer systems was pioneered by S.F. Edwards, Proc. Phys. Soc. 91, 513 (1967) ; S.F . Edwards, J . Phys. A 1, 15 (1968) . See also M.G. Brereton and S. Shaw, J . Phys. A 13, 2751 (1980) and later works of these authors. Investigations via Monte Carlo simulations were made by A.V. Vologodskii, A.V. Lukashin, M.D. Frank-Kamenetskii, and V.V. Anshelevin, Sov. Phys. JETP 39, lO95 (1974); A.V. Vologodskii, A.V. Lukashin, and M.D. Frank-Kamenetskii, Sov. Phys.-JETP 40, 932 (1975) . See also the review article by M.D. Frank-Kamenetskii and A.V. Vologodskii, Sov. Phys. Usp. 24, 679 (1982) . This article also discussed ribbons. For further computer work on knot distributions see J.P.J. Michels and F.W. Wiegel, Phys. Lett. A 9 , 381 (1982); Proc. Roy. Soc. A 403,269 (1986), and references therein. The work is summarized in the textbook by F .W . Wiegel, Introduction to Path-Integral Methods in Physics and Polymer Science, World Scientific, Singapore, 1986. See also S. Windwer, J. Chern. Phys. 93, 765 (1990). The parameter C at the end of Section 6.4 was found by A. Kholodenko, Phys. Lett. A 159, 437 (1991), who mapped the problem onto a q-state Potts model with q = 4. This mapping gives Q = 0 and C = 2e- 7r / 6 ~ 1.18477. For the Gauss integral as a topological invariant of links see the original paper by G.F. Gauss, Koenig. Ges. Wiss. Goettingen 5, 602 (1877) . The writhing number Wr was introduced by F.B. Fuller, Proc. Nat. Acad. Sci. USA 68, 815 (1971), who applied the mathematical relation to DNA. See also F .H.C. Crick, Proc. Nat. Acad. Sci. USA 68,2639 (1976) . The relation Lk = Tw + Wr was first written down by G. Calagareau, Rev. Math. Pur. et Appl. 4 , 58 (1959); Czech. Math. J. 4 , 588 (1961), and extended by J .H. White, Am. J. Math. 90, 1321 (1968). In particle physics, ribbons are used to construct path integrals over fluctuating fermion orbits: A.M. Polyakov, Mod. Phys. Lett. A 3, 325 (1988). For more details see

1160

16 Polymers and Particle Orbits in Multiply Connected Spaces

C.H. Tze, Int. J. Mod. Phys. A 3 , 1959 (1988). The construct ion of t he Alexander polynomial of links is described in A.V. Vologodskii, A.V. Lukashin, and M.D. Frank-Kamenetskii, JETP 40, 932 (1974) . Their derivation from the skein relations is shown in J.H. Conway, An Enumeration of Knots and Links , Pergamon, London, 1970, pp. 329-358; L.H. Kauffman, Topology 20, 101 (1981). In the mathematical literature, the various knot polynomials are discussed by L.H. Kauffman, Topology 26, 395 (1987) ; Contemporary Mathematics AMS 78, 283 (1988) ; Trans. Amer. Math. Soc. 318, 417 (1990) ; On Knots, Princeton University Press, Princeton, 1987; Knots and Physics , World Scientific, Singapore, 1991; J . Math. Phys. 36, 2402 (1995). V. Jones, Bull. Am. Math. Soc. 12, 103 (1985); Ann. Math. 126, 335 (1987); P. Freyd , D. Yet ter, J. Hoste, W . B. R. Lickorish, K.C. Millet , and A. Ocneanu, Bull. Am. Math. Soc. 12, 239 (1985) ; W . B. R. Lickorish and K.C. Millet , Math. Magazine 61 , 3 (1987). The lower HOMFLY polynomials are tabulated in the text. For the higher ones see the microfilm accompanying the article H. Doll and J. Hoste, Math. of Computation 57, 747 (1991) and the unpublished tables by M.B. Thistlethwaite, University of Knoxville, Tennessee. The author is grateful for a copy of them. A collection of many relevant articles is found in T . Kuhno (ed.), New Developments in the Th eory of Knots , World Scientific, Singapore, 1990. A short introduction to the classificat ion problem of knots is given in the popular articles W .F.R. Jones, Scientific American, November 1990, p. 52, I. Stewart , Spektrum der Wissenschaft, August 1990, p. 12. The Chern-Simons actions have in recent years received increasing attention due to their relevance for explaining the fractional quantum Hall effect and a possible presence in high-temperature superconductivity. Actions of this type were first observed in four-dimensional quantum field theories in the form of so-called anomalies by J. Wess and B. Zumino, Phys. Lett. B 36, 95 (1971) . The action (16.253) in three spacetime dimensions was first analyzed by S. Deser, R. Jackiw, and S. Templeton, Ann. Phys. 140, 372 (1982), who pointed out the connection with the Chern classes of differential geometry described by S. Chern, Complex Manifolds without Potential Theory, Springer, Berlin, 1979. In particular they found the mass of the electromagnetic field which was conjectured to be the origin of the Meissner effect in high-temperature superconductors. See A.L. Fetter, C. Hanna, and R.B. Laughlin, Phys. Rev. B 39, 9679 (1989) ; Y.-H. Chen and F. Wilczek, Int. J. Mod. Phys. B 3, 117 (1989); Y.-H. Chen, F . Wilczek, E . Witten, and B.I. Halperin, Int. J . Mod. Phys. B 3, 1001 (1989) ; A. Schakel, Phys. Rev. D 44, 1198 (1992). However, the recent finding of dissipation in anyonic systems by D.V. Khveshchenko and 1.1. Kogan, Int. J. Mod. Phys. B 5, 2355 (1991) speaks against an anyon mechanism of this phenomenon. A Chern-Simons type of action appeared when integrating out fermions in H. Kleinert , Fortschr. Phys. 26, 565 (1978) (http://www.physik.fu-berlin.de;-kleinert/55). The relation with Chern classes was recognized by M.V. Berry, Proc. Roy. Soc. A 392, 45 (1984); B. Simon, Phys. Rev. Lett. 51, 2167 (1983). The Chern-Simons action in the text was derived for a degenerate electron liquid in two dimensions by

1161

Notes and References

T. Banks and J.D. Lykken, Nuc!. Phys. B 336, 500 (1990) ; S. Randjbar-Daemi, A. Salam, and J. Strathdee, Nuc!. Phys. B 340, 403 (1990), P.K. Panigrahi, R Ray, and B. Sakita, Phys. Rev. B 42, 4036 (1990). There is related work by M. Stone, Phys. Rev. D 33, 1191 (1986); I.J.R Aitchison, Acta Physica Polonica B 18, 207 (1987). See also the reprints of many papers on this subject in A. Shapere and F. Wilczek, Geometric Phases in Physics, World Scientific, Singapore, 1989. F. Wilczek, Fractional Statistics and Anyon Superconductivity, World Scientific, Singapore, 1990, which itself provides a clear introduction to the subject and contains many important reprints. A good review is also contained in the lectures J.J. Leinaas, Topological Charges in Gauge Theories, Nordita Preprint, 79/43, ISSN 0106-2646. Textbooks on this subject are A. Lerda, Anyons-Quantum Mechanics of Particles with Fractional Statistics, Lecture Notes in Physics, m14, Springer, Berlin 1992; A. Khare, Fractional Statistics and Quantum Theory , World Scientific, Singapore, 1997. The Lerda book contains many useful examples and explains the origin of difficulties in treating interacting anyons. The Khare book provides a well-motivated treatment and includes a brief introduction to the Braid group. Both include discussions of the Quantum Hall Effect and Anyon Superconductivity. For the relation between the Chern-Simons theory and knot polynomials see E . Witten, Comm. Math. Phys. 121, 351 (1989), Nuc!. Phys. B 330, 225 (1990). See also P. Cotta-Ramusino, E . Guadagnini, M. Martellini, and M. Mintchev, Nuc!. Phys. B 330, 557 (1990); G.V. Dunne, R Jackiw, and C. Trugenburger, Ann. Phys. 194, 197 (1989) ; A. Polychronakos, Ann. Phys. 203, 231 (1990) ; E . Guadagnini, I. J. Mod. Phys. A 7, 877 (1992) . The integer quantum Hall effect was found by K. vonKlitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 (1980) ; the fractional one by D.C. Tsui, H.L. Stormer, and A.C. Gossard, Phys. Rev. Lett. 48, 1559 (1980). Theoretical explanations are given in RB. Laughlin, Phys. Rev. Lett. 50, 1395 (1983) , Phys. Rev. B 23, 3383 (1983); F .D.M. Haldane, Phys. Rev. Lett. 51, 605 (1983); B.1. Halperin, Phys. Rev. Lett. 52, 1583 (1984) ; D.P. Arovas, J .R Schrieffer, F . Wilczek, Phys. Rev. Lett. 53, 722 (1984); D.P. Arovas, J.R Schrieffer, F. Wilczek, and A. Zee, Nuc!. Phys. B 251, 117 (1985) ; J .K. Jain, Phys. Rev. Lett. 63, 199 (1989). The exceptional filling factor v = ~ is discussed in R Willet et al., Phys. Rev. Lett. 59, 17765 (1987); S. Kivelson, C. Kallin, D.P. Arovas, and J .R Schrieffer, Phys. Rev. Lett. 56, 873 (1986). The Chern-Simons path integral is treated semiclassically in D.H. Adams, Phys. Lett. B 417, 53 (1998) (hep-th/9709147) . For a simple discussion of the change from Bose to Fermi statistics at the level of creation and annihilation operators via a topological interaction see E . Fradkin, Phys. Rev. Lett. 63, 322 (1989) ; Field Theories of Condensed Matter Physics, Addison-Wesley, 1991. The lattice form of the action

Lx El"v),AI"(x)V'vA),(x)

used by that author is not correct since

1162

16 Polymers and Particle Orbits in Multiply Connected Spaces

The lattice form of the action ExEJtVAAJt(x)V'vAA(X) used by that author is not correct since it violates gauge invariance. This can, however, easily be restored without destroying the results by replacing the first AJt(x)-field by AJt(x - eJt), where eJt is the unit vector in the jt-direction. See the general discussion of lattice gauge transformations in H. Kleinert, Gauge Fields in Condensed Matter, Vol. I, World Scientific, Singapore, 1989, Chapter 8 (http://www.physik.iu-berlin.de;-kleinert/bl). See furthermore D. Eliezer and G.W. Semenoff, Anyonization of Lattice Chern-Simons Theory, Ann. Phys.217, 66 (1992). For the London equations see the original paper by F. London and H. London, Proc. Roy. Soc. A 147, 71 (1935) and the extension thereof: A.B. Pippard, ibid., A 216, 547 (1953). The individual citations refer to [1] P.A.M. Dirac, Proc. Roy. Soc. A 133, 60 (1931); Phys. Rev. 74, 817 (1948), Phys. Rev. 74, 817 (1948). See also J. Schwinger, Particles, Sources and Fields, Vols. 1 and 2, Addison Wesley, Reading, Mass., 1970 and 1973. [2] For a review, see G. Giacomelli, in Monopoles in Quantum Field Theory, World Scientific, Singapore, 1982, edited by N.S. Craigie, P. Goddard, and W. Nahm, p. 377. [3] B. Cabrera, Phys. Rev. Lett. 48, 1378 (1982). [4] For a detailed discussion of the physics of vortex lines in superconductors, see H. Kleinert, Gauge Fields in Condensed Matter, World Scientific, Singapore, 1989, Vol. I, p. 331 (http://www.physik.iu-berlin.de/-kleinert/bl). [5] See the reprint collection A. Shapere and F. Wilczek, Geometric Phases in Physics, World Scientific, Singapore, 1989. In particular the paper by D.P. Arovas, Topics in Fractional Statistics, p.284. [6] The Tait number or writhe must not be confused with the writhing number Wr introduced in Section 16.6 which is in general noninteger. See P.G. Tait, On Knots I, II, and III, Scientific Papers, Vol. 1. Cambridge, England: University Press, pp. 273-347, 1898. [7] See Ref. [7] of Chapter 19. There exists, unfortunately, no obvious extension to four spacetime dimensions. [8] The development in this Section follows F. Ferrari, H. Kleinert, and 1. Lazzizzera, Phys. Lett. A 276,1 (2000) (cond-mat/0002049); Eur. Phys. J. B 18, 645 (2000) (cond-mat/0003355); nt. Jour. Mod. Phys. B 14, 3881 (2000) (cond-mat/0005300); Topological Polymers: An Application of Chern-Simons Field Theories, in K. Lederer and N. Aust (eds.), Chemical and Physical Aspects of Polymer Science and Engineering, 5-th Oesterreichische Polymertage, Leoben 2001, Macromolecular Symposia, 1st Edition, ISBN 3-527-30471-1, Wiley-VCH, Weinheim, 2002. [9] M.G. Brereton and S. Shah, J. Phys. A: Math. Gen. 15,989 (1982).

Notes and References

1163

[10] R. Jackiw, in Current Algebra and Anomalies, ed. by S.B. Treiman, R. Jackiw, B. Zumino, and E. Witten, World Scientific, Singapore, 1986, p. 21l. [11] L.D. Faddeev and V.N. Popov, Phys. Lett. B 25, 29 (1967).

So many paths, that wind and wind, While just the art of being kind Is all the sad world needs. ELLA WILCOX (1855-1919), The World's Needs

17 Tunneling

Tunneling processes govern the decay of metastable atomic and nuclear states, as well as the transition of overheated or undercooled thermodynamic phases to a stable equilibrium phase. Path integrals are an important tool for describing these processes theoretically. For high tunneling barriers, the decay proceeds slowly and its properties can usually be explained by a semiclassical expansion of a simple model path integral. By combining this expansion with the variational methods of Chapter 5, it is possible to extend the range of applications far into the regime of low barriers. In this chapter we present a novel theory of tunneling through high and low barriers and discuss several typical examples in detail. A useful fundamental application arises in the context of perturbation theory since the large-order behavior of perturbation expansions is governed by semiclassical tunneling processes. Here the new theory is used to calculate perturbation coefficients to any order with a high degree of accuracy.

17.1

Double-Well Potential

A simple model system for tunneling processes is the symmetric double-well potential of Eq. (5.78). It may be rewritten in the form

V(x)

=

w2

2

2

(17.1)

8a2 (x-a) (x+a) ,

which exhibits the two degenerate symmetric minima at x The coupling strength is

= ±a (see Fig. 17.1). (17.2)

Near the minima, the potential looks approximately like a harmonic oscillator potential V±(x) = w 2 (x T a? /2: 2 x T a + . .. ) == V±(x) + V(x) = w 2(x T a) 2 ( 1 ± -a-

1164

~V±(x)

+ ...

(17.3)

1165

17.1 Double-Well Potential

v (x)

Figure 17.1 Plot of symmetric double-well potential V(x) = (x - a?(x

+ a)2w 2/8a 2 for

w=1anda=1. The height of the potential barrier at the center is

v.

_ (wa)2

max-

8

(17.4)

.

In the limit a

---+ 00 at a fixed frequency w, the barrier height becomes infinite and the system decomposes into a sum of two independent harmonic-oscillator potentials widely separated from each other. Correspondingly, the wave functions of the system should tend to two separate sets of oscillator wave functions

(17.5)

where the quantities ~x±

== x±a

(17.6)

measure the distances of the point x from the respective minima. A similar separation occurs in the time evolution amplitude which decomposes into the sum of the amplitudes of the individual oscillators a-->oo

---+

(XbtblxataL + (Xbtblxata)+

JDx(t) exp {~l:b dt~[i;2 -

(17.7) w 2(x

+ a)2]} + (a ---+ -a) .

For convenience, we have assumed a unit particle mass M = 1 in the Lagrangian of the system: L=

i;2

2 -

V(x) .

(17.8)

1166

17 Tunneling

If a is no longer infinite, a particle in either of the two oscillator wells has a nonvanishing amplitude for tunneling through the barrier to the other well, and the wave functions of the right- and left-hand oscillators are mixed with each other. Since the action is symmetric under the mirror reflection x --+ -x , the solutions of the Schrodinger equation

H'l/J(x, t) = H( -i8x, x)'l/J(x, t) = inBt'l/J(x, t),

(17.9)

with the Hamiltonian p2

= 2 + V(x) ,

H(p,x)

(17.10)

can be separated into symmetric and antisymmetric wave functions. As usual, the symmetric states have a lower energy than the antisymmetric ones since a smaller number of nodes implies less kinetic energy for the particles. If the distance parameter a is very large, then, to leading order in a --+ 00, the lowest two wave functions coincide approximately with the symmetric and antisymmetric combinations of the harmonic-oscillator wave functions (17.11) Due to tunneling, the lowest two energies show some deviation from the harmonic ground state value

~nw + 6.E(O). 2 s,a

E(O) = s,a

(17.12)

At a large distance parameter a, this deviation is very small. In quantum mechanics, the level shifts 6.Es ,a can be calculated in lowest-order perturbation theory by inserting the approximate wave functions (17.11) into the formula 6.Es,a =

JdX'l/Js,aH'l/Js,a.

(17.13)

Since the wave functions 'l/Jo(x ± a) of the individual potential wells fall off exponentially like e- x2/ 2 at large x, the level shifts 6.Es ,a are exponentially small in the square distance a2 . In this chapter we derive the level shifts 6.Es ,6.Ea and the related tunneling amplitudes from the path integral of the system. For large a, this will be relatively simple since we can have recourse to the semiclassical approximation developed in Chapter 4 which becomes exact in the limit a --+ 00. As long as we are interested only in the lowest two states, the problem can immediately be simplified. We take the spectral representation of the amplitude

JVx(t)e(i/Ii)

Jt~ dt[j;2/2-V(x))

L'l/Jn(Xb)'l/Jn(Xa)e-iEn(tb- ta)/1i n

(17.14)

1167

17.2 Classical Solutions - Kinks and Antikinks

to imaginary times ta,b

--+

Ta,b

= ~iL/2,

where it becomes

JDx(T)e

(Xb L/2lxa - L/2)

-(lin) J~D2 dr[x,2/2+V(x)]

(17.15)

L'l/Jn(Xb)'l/Jn(Xa)e- EnLln, n

with the notation X'(T) == dX(T)/dT. In the limit of large L, the spectral sum (17.15) is obviously most sensitive to the lowest energies, the contributions of the higher energies En being suppressed exponentially. Thus, to calculate the small level shifts of the two lowest states, 6..Es ,a, we have only to find the leading and subleading exponential behaviors. Since the wave functions are largest close to the bottoms of the double well at x '" ±a, we may consider the amplitudes with the initial and final positions Xa and Xb lying precisely at the bottoms, once on the same side of the potential barrier,

(a L/21a - L/2)

(-a L/21-a -L/2),

(17.16)

(a L/21-a -L/2) = (-a L/21a -L/2).

(17.17)

=

and once on the opposite sides

For these amplitudes we now calculate the semiclassical approximation in the limit L --+ 00. The results will lead to level shift formula in Section 17.7.

17.2

Classical Solutions -

Kinks and Antikinks

According to Chapter 4, the leading exponential behavior of the semiclassical approximation is obtained from the classical solutions to the path integral. The fluctuation factor requires the calculation of the quadratic fluctuation correction. The result has the form exp{ -AcI/li} x F ,

(17.18)

class. solutions

where

Act denotes the action of each classical solution and F the fluctuation factor.

The amplitude (17.16), which contains the bottom of the same well on either side, is dominated by a trivial classical solution which remains all the time at the same bottom:

X(T) == ±a.

(17.19)

Classical solutions exist also for the other amplitudes (17.17) which connect the different bottoms at -a and a. These solutions cross the barrier and read, in the limit L --+ 00,

X(T)

= X~(T)

== ±atanh[w(T - To)/2],

(17.20)

1168

17 Tunneling T

4

2

-2

---

2

x

-2

-4 lOV(x)

Figure 17.2 Classical kink solution (solid curve) in double-well potential (short-dashed curve with units marked on the lower half of the vertical axis). The solution connects the two degenerate maxima in the reversed potential. The long-dashed curve shows a solution which starts out at a maximum and slides down into the adjacent abyss.

with an arbitrary parameter 70 specifying the point on the imaginary time axis where the crossing takes place. The crossing takes place within a time of the order of 2/w. For large positive and negative 7, the solution approaches ±a exponentially (see Fig. 17.2). Alluding to their shape, the solutions X~(7) are called kink and antikink solutions, respectively.1 To derive these solutions, consider the equation of motion in real time,

x(t) = -V'(x(t)), where V'(x)

(17.21)

== dV(x)/dx. In the Euclidean version with 7 = -it, this reads X"(7) = V'(X(7)).

(17.22)

Since the differential equation is of second order, there is merely a sign change in front of the potential with respect to the real-time differential equation (17.21). The Euclidean equation of motion corresponds therefore to a usual equation of motion of a point particle in real time, whose potential is turned upside down with respect to Fig. 17.1. This is illustrated in Fig. 17.3. The reversed potential allows obviously for a classical solution which starts out at x = -a for 7 --+ -00 and arrives at x = a for 7 --+ +00. The particle needs an infinite time to leave the initial potential mountain and to climb up to the top of the final one. The movement through the central valley proceeds within the finite time ~ 2/w. If the particle does not start 1 In field-theoretic literature, such solutions are also referred to as instanton or anti-instanton solutions, since the valley is crossed within a short time interval. See the references quoted at the end of the chapter.

17.2 Classical Solutions - Kinks and Antikinks

1169

Figure 17.3 Reversed double-well potential governing motion of position x as function of imaginary time T.

its movement exactly at the top but slightly displaced towards the valley, say at x = -a + E, it will reach x = a - E after a finite time, then return to x = -a + E, and oscillate back and forth forever. In the limit E -+ 0, the period of oscillation goes to infinity and only a single crossing of the valley remains. To calculate this movement, the differential equation (17.22) is integrated once after multiplying it by x' = dx / dT and rewriting it as 1 d'2

"2 dT X =

d

(17.23)

dT V(X(T)) .

The integration gives X'2

2 + [-V(X(T))]

= const.

(17.24)

T is reinterpreted as the physical time, this is the law of energy conservation for the motion in the reversed potential - V (x) . Thus we identify the integration constant in (17.24) as the total energy E in the reversed potential:

If

const == E.

(17.25)

Integrating (17.24) further gives (17.26) A look at the potential in Fig. 17.3 shows that an orbit starting out with the particle at rest for T -+ -00 must have E = o. Inserting the explicit potential (17.1) into (17.26) , we obtain for Ixl < a

T-TO

=

2a ±-

lX

dx' (a - x') (x' 2 x ±-arctanh-. w a W

0

+ a)

1 a+x =±-log-w a- x

(17.27)

1170

17 Tunneling

Thus we find the kink and antikink solutions crossing the barrier:

Xcl(T) = ±atanh[(T - To)w/2].

(17.28)

The Euclidean action of such a classical object can be calculated as follows [using (17.24) and (17.25)]:

(17.29) The kink has E

=

0, so that (17.30)

and the classical action becomes

Acl = -W

2a

Note that for E

=

ja dx(a - a

2

2

2

- x ) = -a 3

2 W

3

= -w . 3g

(17.31)

0, the classical action is also given by the integral (17.32)

There are also solutions starting out at the top of either mountain and sliding down into the adjacent exterior abyss, for instance (see again Fig. 17.3) T -

TO

2a

=fW

1 (x' - a)(x' dx' = + a) 00

x

1 x +a ±-log-W x- a

(17.33)

2 x ±-arccoth-. w a However, these solutions cannot connect the bottoms of the double well with each other and will not be considered further. Being in the possession of the classical solutions (17.19) and (17.28) with a finite action, we are now ready to write down the classical contributions to the amplitudes (17.16) and (17.17). According to the semiclassical formula (17.18), they are

(a L/21a - L/2) = 1 x Fw(L)

(17.34)

and (17.35) The factor 1 in (17.34) emphasizes the vanishing action of the trivial classical solution (17.19). The exponential e- Acl / h contains the action of the kink solutions (17.28).

1171

17.3 Quadratic Fluctuations

The degeneracy of the solutions in TO is accounted for by the fluctuation factor Fel (L ), as will be shown below. Actually, the classical kink and antikink solutions (17.28) do not occur exactly in (17.35) since they reach the well bottoms at x = ±a only at infinite Euclidean times T ---. ±oo. For the amplitude to be calculated we need solutions for which x is equal to ±a at large but finite values T = ±L/2. Fortunately, the error can be ignored since for large L the kink and antikink solutions approach ±a exponentially fast . As a consequence the action of a proper solution which would reach ±a at a finite L differs from the action Acl only by terms which tend to zero like e- wL . Since we shall ultimately be interested only in the large- L limit we can neglect such exponentially small deviations. In the following section we determine the fluctuation factors F el .

17.3

Quadratic Fluctuations

The semiclassical limit includes the effects of the quadratic fluctuations. These are obtained after approximating the potential around each minimum by a harmonic potential and keeping only the lowest term in the expansion (17.3). The fluctuation factor of a pure harmonic oscillator of frequency wand unit mass has been calculated in Section 2.3 with the result

Fw(L)

=

I

~

V27l'nsmhwL

'"'" fW e- wL/2 + O(e- 3wL /2). V;r;,

(17.36)

The leading exponential at large L displays the ground state energy w /2, while the corrections contain all information on the exited states whose energy is (n + 1/2)w with n = 1, 2,3, .... Note that according to the spectral representation of the amplitude (17.15), the factor JW/7l'n in (17.36) must be equal to the square of the ground state wave function wo(Llx±) at the potential minimum. This agrees with (17.5) . Consider now the fluctuation factor of a single kink contribution. It is given by the path integral over the fluctuations Y(T) == 8X(T)

Fel(L) =

JDY(T)e

JL/2

[ /2

-(lin) _L/2 dr (1/2) y

+V

/I

(xcl(r))y

21 ,

(17.37)

where Xel(T) is the kink solution, and Y(T) vanishes at the endpoints:

y(L/2) = y( -L/2) = O.

(17.38)

Suppose for the moment that L = 00. Then the kink solution is given by (17.28) and we obtain the fluctuation potential

1172

17 Tunneling

i--"-~~~~~-

2

3

4

5

6

T

7

Figure 17.4 Potential (17.41) for quadratic fluctuations around kink solution (17.28) in Schrodinger equation. The dashed lines indicate the bound states at energies 0 and 3w 2 / 4.

Thus, the quadratic fluctuations are governed by the Euclidean action

Al( =

1 1[ L/2

-L/2

(3

dT- y,2 + w 2 1 _ 2

1

2 cosh2 [w(T - To)/2]

)]

y2 .

(17.40)

The rules for doing a functional integral with a quadratic exponent were explained in Chapter 2. The paths y (T) are expanded in terms of eigenfunctions of the differential equation (17.41) with An being the eigenvalues. This is a Schrodinger equation for a particle moving along the T-axis in an attractive potential well of the Rosen-Morse type [compare (14.157) and see Fig. 17.4]:

(17.42) The eigenfunctions Yn(T) satisfy the usual orthonormality condition (17.43) Given a complete set of these solutions Yn(T) with n = 0,1,2, .. . , we now perform the normal-mode expansion 00

y~O,6 " "(T)

=

L

n=O

~nYn(T).

(17.44)

1173

17.3 Quadratic Fluctuations

After inserting this into (17.40), we perform a partial integration in the kinetic term and the T-integrals with the help of (17.43), and the Euclidean action of the quadratic fluctuations takes the simple form (17.45) With this, the fluctuation factor (17.37) reduces to a product of Gaussian integrals over the normal modes Fcl (L)

= N

d~n] - en [100-00 -,.}27rn °o

",,00

.

L...m=o'n n

/21i

n=O

= N - -1- . ,.}IInAn

(17.46)

The normalization constant N, to be calculated below, accounts for the Jacobian which relates the time-sliced measure to the normal mode measure. First we shall calculate the eigenvalues An. For this we use the amplitude of the Rosen-Morse potential obtained via the Duru-Kleinert transformation (14.160) in Eq. (14.161). If the potential is written in the form

V(T) = w 2

_

Va

(17.47)

cosh 2 [m(T - TO)]'

there are bound states for n = 0, 1,2, ... ,nmax < s, where

8

1[

== '2

-1

8

is defined by [1]

.~ +V 1 + 4:n12 J .

(17.48)

Their wave functions are, according to (14.163), and (14.165),

Yn(T) =

~ -, . fr(8 - n)f(l + 28 - n) r( n. V xF( -n, 1 +

28 -

n ;8

-

2n- s

1 +8

n + 1; HI

-

- n

) coshn- S[m(T - TO)]

tanh[m(T - TO)])),

where F(a, b; C; z) are the hypergeometric functions (1.450). parameter Va becomes

(17.49)

In terms of 8, the (17.50)

The bound-state energies are A~

= w 2 - m2 (8

-

nf

(17.51)

In the Schrodinger equation (17.41), we have m = w/2, Va = 3w 2 /2 , so that and there are exactly two bound-state solutions. These are

8

=2

(17.52)

1174

17 Tunneling

and



w 1 h[ ( )/ {(-1,4; 2;H1-tanh[w(T-To)/2])) 4 cos W T - TO 2

Yl(T)

{&.;

V4

sinh[w(T - To)/2] cosh 2[w(T - To)/2]'

(17.53)

The negative sign in (17.52) is a matter of convention, to give Yl(T) the same sign as Xcl(T - TO) in Eq. (17.89) below. The normalization factors can be checked using the formula

rOO

io

sinh" x = ~B(11 + 1 cosh v x 2 2'

l/

-11) 2 '

(17.54)

with B(x,y) = r(x)r(y)/r(x + y) being the Beta function. The corresponding eigenvalues are (17.55) The existence of a zero-eigenvalue mode is a general property of fluctuations around localized classical solutions in a system which is translationally invariant along the T-axis. It prevents an immediate application of the quadratic approximation, since a zero-eigenvalue mode is not controlled by a Gaussian integral as the others are in Eq. (17.46). This difficulty and its solution will be discussed in Subsection 17.3.l. In addition to the two bound states, there are continuum wave functions with An ~ w 2 . For an energy Ak = w 2

+ k2 ,

(17.56)

they are given by a linear combination of Yk(T) ex AeikT F(s + 1, -s; 1 - ik/m; !(1- tanh[m(T - TO)]))

(17.57)

and its complex-conjugate (see the end of Subsection 14.4.4). Using the identity for the hypergeometric function 2 F(a, b, c; z)

r(c)r(c - a - b)

= r(c _ a)r(c _ b) F(a, b; a + b - c + 1; 1- z)

+(l-z)

c_a_br(C)r( -c + a + b) r(a)r(b)

F(c-a,c-b;c-a-b+1;1-z), (17.58)

and F(a, b; c; 0) = 1, we find the asymptotic behavior F

T -> OO

------> T ~- OO

F

------>

1,

(17.59)

r( -ik/m)r(l - ik/m) _2ikTr(ik/m)r(1 - ik/m) r(-s-ik/m)r(s+l-ik/m) +e r(-s)r(l+s) ' (17.60)

2M. Abramowitz and I. Stegun, op. cit., formula 15.3.6.

1175

17.3 Quadratic Fluctuations

These limits determine the asymptotic behavior of the wave functions (17.57). With an appropriate choice of the normalization factor in (17.57) we fulfill the standard scattering boundary conditions T ---- -00,

(17.61)

T ---- 00 .

These define the transmission and reflection amplitudes. From (17.59) and (17.60) we calculate directly T: _ f( -s - ik/m)f(s + 1 - ik/m) k r( -ik/m)f(l - ik/m) , Rk

(17.62)

= f( -s - ik/m)r(s + 1- ik/m) f(ik/m) = Tk f(ik/m)f(l - ik/m). (17.63) f(-s)f(l+s)

f(-ik/m)

f(-s)f(l+s)

Using the relation f(z) = 1l' / sin(1l'z)r(1 - z), this can be written as f(s + 1 - ik/m) f(l + ik/m) sin(ik/m) f(s + 1 + ik/m) f(l - ik/m) sin(s + ik/m) , T: sin( 1l's) ksin(ik/m) .

(17.64) (17.65)

The scattering matrix (17.66) is unitary since (17.67) It is diagonal on the state vectors

(17.68) which, as we shall prove below, correspond to odd and even partial waves. The respective eigenvalues A%'!O = e2iO~'· define the phase shifts 5~'!O, in terms of which

~(e2iO~ + e2iO~),

(17.69)

~2 (e 2iOk _ e2iO~) .

(17.70)

Let us verify the association of the eigenvectors (17.68) with the even and odd partial waves. For this we add to the wave function (17.61) the mirror-reflected solution T ---- -00, T ---- 00,

(17.71)

1176

17 Tunneling

and obtain T -+ -00, T -+ 00.

(17.72)

Inserting (17.69), this can be rewritten as T -+ -00, T -+ 00.

(17.73)

The odd combination, on the other hand, gives T -+ -00, T -+ 00,

(17.74)

and becomes with (17.69):

From Eqs. (17.69) , (17.70) we see that ITkl2

= COS2 (O(T) +

AO 1T dT' w - L/2

[~(T)1](T') -1](T)~(T')] if>O(T').

(17.163)

The limits of integration ensure that if>~(T) vanishes at T = -L/2. The eigenvalue AO is determined by enforcing the vanishing also at T = L/2. Taking for if>O(T) the zero-eigenvalue solution D( T)

AO = -D(L/2)W

[~(L/2) 1L/2 dT 1](T)D(T) -1](L/2) 1L/2 dT ~(T)D(T)] - L/2 - L/2

-.1

(17.164)

Inserting (17.154) and using the orthogonality of ~(T) and 1](T) (following from the fact that the first is symmetric and the second antisymmetric) , this becomes

Ao=-D(L/2)w2 [~(-L/2)~(L/2) 1

L/2

-~

dT1]2(T)

+ 1](-L/2)1](L/2)

1 L / 2 ] -1 dTe(T) . -~

(17.165)

1188

17 Tunneling

Invoking once more the symmetry of (17.152) and (17.153), we obtain

Ao = -D(L/ 2)W 2 [e- wL l

~(T)

and TJ(T) and the asymptotic behavior

lL/2 ] - 1 L/2 dTTJ 2(T) - ewL dTe(T) -L/2 -L/2

(17.166)

The first integral diverges like ewL ; the second is finite. The prefactor makes the second integral much larger than the first , so that we find for large L the would-be zero eigenvalue (17.167) This eigenvalue is exponentially small and positive, as expected. Inserting it into (17.162) and using (17.156) and (17.160), we find the eigenvalue ratio

TIn A~

~

nn

An

l'

1

= L ---+ 1m 2w Joo d t2( ) . oo T ~ T

(17.168)

- 00

The determinant D(L/ 2) has disappeared and the only nontrivial quantity to be evaluated is the normalization integral over the translational eigenfunction ~ (T) . The normalization inte&ral requires the knowledge of the full T-behavior of the zero-eigenvalue solution ¢>01) (T) ; the asymptotic behavior used up to this point is insufficient. Fortunately, the classical solution Xci (T) also supplies this information. The normalized solution is Yo = aX~I(T) behaving asymptotically like 2aawe- W T • Imposing the normalization convention (17.152) for ¢>~1)(T) , we identify (17.169) Using the relation (17.32), the normalization integral is simply

1

00

dT~(T)

- 00

2

Al = -4a2W 2'

(17.170)

With it the eigenvalue ratio (17.168) becomes

TIn A~ 4a 2 w A . TI n An = 2w .I"lcl l

By inserting the value of the classical action

Al = 2a2w/ 3 from (17.31), we obtain

TIn AA~ = 12W 2 , TI n n l

(17.171)

(17.172)

just as in (17.146) and (17.129). It is remarkable that the calculation of the ratio of the fluctuation determinants with this method requires only the knowledge of the classical solution XcI(T).

17.6 Wave Functions of Double-Well

17.6

1189

Wave Functions of Double-Well

The semiclassical result for the amplitudes

(a L/21a -L/2),

(a L/21-a -L/2),

(17.173)

with the endpoints situated at the bottoms of the potential wells can easily be extended to variable endpoints Xb i= a, Xa i= ±a, as long as these are situated near the bottoms. The extended amplitudes lead to approximate particle wave functions for the lowest two states. The extension is trivial for the formula (17.34) without a kink solution. We simply multiply the fluctuation factor by the exponential exp( -Ad/n) containing the classical action of the path from Xa to Xb. If Xa and Xb are both near one of the bottoms of the well, the entire classical orbit remains near this bottom. If the distance of the orbit from the bottom is less than 1/avw, the potential can be approximated by the harmonic potential W 2X2. Thus near the bottom at x = a, we have the simple approximation to the action

Acl

w { [(xa - a) 2 ~ 2nsinhwL

+ (Xb - a) 2] coshwL - 2(Xb - a)(xa - a) } . (17.174)

For a very long Euclidean time L, this tends to W

2

2

Act ~ 2n [(Xb - a) + (xa - a) ].

(17.175)

The amplitude (17.34) can therefore be generalized to

(Xb L/2lxa L/2) ~ fWe-(w/21i)[(xb-a)2+(xa-a)21e-wL/21i.

V;h

This can also be written in terms of the bound-state wave functions (17.5) for n = 0 as (17.176) For the amplitude (17.35) with the path running from one potential valley to the other, the construction is more subtle. The approximate solution is obtained by combining a harmonic classical path running from (xa, -L/2) to (a, -L/4), a kink solution running from (a, -L/4) to (a , L/4), and a third harmonic classical path running from (a, L/4) to (Xb, L/2). This yields the amplitude

~e-(W/21i)(Xb-a)2 K Le- AcJ / li e-(w/21i)(xa-a)2.

(17.177)

Note that by patching the three pieces together, it is impossible to obtain a true classical solution. For this we would have to solve the equations of motion containing a kink with the modified boundary conditions x(-L/2) = Xa, x(L/2) = Xb. From the exponential convergence of X(T) --+ ±a (like e- wL ) it is, however, obvious that the

1190

17 Tunneling

true classical action differs from the action of the patched path only by exponentially small terms. As before, the prefactor in (17.177) can be attributed to the ground state wave functions 'l/Jo(x ), and we find the amplitude for Xb close to -a and Xa close to a: (17.178)

17.7

Gas of Kinks and Antikinks and Level Splitting Formula

The above semiclassical treatment is correct to leading order in e- wL . This accuracy is not sufficient to calculate the degree of level splitting between the two lowest states of the double well caused by tunneling. Further semiclassical contributions to the path integral must be included. These can be found without further effort. For very large L, it is quite easy to accommodate many kinks and antikinks along the T-axis without a significant deviation of the path from the equation of motion. Due to the fast approach to the potential bottoms x = ±a near each kink or antikink solution, an approximate solution can be constructed by smoothly combining a number of individual solutions as long as they are widely separated from each other. The deviations from a true classical solution are all exponentially small if the separation distance !::!.T on the T-axis is much larger than the size of an individual kink (i.e., !::!.T» l/w). The combined solution may be thought of as a very dilute gas of kinks and antikinks on the T-axis. This situation is referred to as the dilute-gas limit . Consider such an "almost-classical solution" consisting of N kink-antikink solutions Xcl(T) = ±atanh[w(T-Ti)/2] in alternating order positioned at, say, Tl » T2 » T3 » ... » TN and smoothly connected at some intermediate points 1'1 , . .. , TN-I. In the dilute-gas approximation, the combined action is given by the sum of the individual actions. For the amplitude (17.34) in which the paths connect the same potential valleys, the number of kinks must be equal to the number of antikinks. The action combined is then an even multiple of the single kink action: (17.179) For the amplitude (17.35) , where the total number is odd, the combined action is

A 2n+1

~ (2n

+ l)Acl.

(17.180)

As the kinks and antikinks are localized objects of size 2/w , it does not matter how they are distributed on the large-T interval [-L/2 , L/2]' as long as their distances are large compared with their size. In the dilute-gas limit, we can neglect the sizes. In the path integral, the translational degree of freedom of widely spaced N kinks and antikinks leads, via the zero-eigenvalue modes, to the multiple integral

1

L/2

- L/2

dTN

1TN - L/2

dTN- l...

1T1 - L/2

dTI

LN

= N' . .

(17.181)

1191

17.7 Gas of Kinks and Antikinks and Level Splitting Formula

The Jacobian associated with these N integrals is [see (17.112)]

(17.182) The fluctuations around the combined solution yield a product of the individual fluctuation factors. For a given set of connection points we have

1

1

1

vn~An l- VIl~Anl LN

LN - 1

x ... x VIl~Anl - '

(17.183)

L1

where Li == Ti - Ti-l are the patches on the T-axis in which the individual solutions are exact. Their total sum is

(17.184) now include the effect of the fluctuations at the intermediate times Ti where individual solutions are connected. Remembering the amplitudes (17.176), we that the fluctuation factor for arbitrary endpoints Xi, Xi-l near the bottom of potential valley must be multiplied at each end with a wave function ratio 'l/Jo(x ± a)/'¢o(O). Thus we have to replace We the see the

1

v'Iln An

-+

'l/JO(Xi ± a) 'l/Jo(O)

1

v'Iln An

'l/J~(Xi-l ± a) t 'l/Jo(O)

(17.185)

The adjacent xi-values of all fluctuation factors are set equal and integrated out, giving

Due to the unit normalization of the ground state wave functions, the integrals are trivial. Only the l'l/Jo(O)l2-denominators survive. They yield a factor 1

l'l/Jo(O)1 2 (N-l)

r w- (N - l)

=

Vrli

.

(17.187)

It is convenient to multiply and divide the result by the square root of the product of eigenvalues of the harmonic kink-free fluctuations , whose total fluctuation factor is known to be

(17.188)

1192

17 Tunneling

Then we obtain the total corrected fluctuation factor

[W- (N- 2) -wL/2h

V;h,

e

VIIn An LVO~AnILl VO~AnIL2 \0

I

1

1

1

X ... X

VO~An I LN

(

) 17.189

We now observe that the harmonic fluctuation factor (17.188) for the entire interval VOn A~IL = VW/'rrnexp( -wL/2n) can be factorized into a product of such factors for each interval Ti, Ti - 1 as follows:

~I =~ -(N-1) ~I V~n~' on L 7rn V~n~"n L,

... V~Ion LN. ~n~'

(17.190)

The total corrected fluctuation factor can therefore be rewritten as

~e -WL/2h 7rn

On' A~ I On An L,

X ... X

(17.191)

Each eigenvalue ratio gives the Li-independent result

K'=

(17.192)

with K' of Eq. (17.131) . Expressing K' in terms of K via (17.193)

the factors vAcl/27rn 1 remove the Jacobian factors (17.182) arising from the positional integrals (17.181). Altogether, the total fluctuation factor of N kink-antikink solutions with all possible distributions on the 'T-axis is (17.194)

Summing over all even and odd kink-antikink configurations, we thus obtain

(a L/21 ± a - L/2) = [We- wL / 2h

V;h,

L ~(KLe-Acl/h(. N.

(17.195)

even odd

This can be summed up to

(aL/21±a

(17.196)

1193

17.7 Gas of Kinks and Antikinks and Level Splitting Formula

As in the previous section, we generalize this result to positions Xb, Xa near the potential minima (with a maximal distance of the order of Vn/w) . Using the classical action (17.175) and expressing it in terms of ground state wave functions, we can now add the contribution of the amplitudes for all possible configurations, arriving at

(Xb L/2lxa - L/2)

=

e- wL /2n

(17.197)

x {'l/JO(Xb - a)'l/JO(xa -

a)~

[exp(K e-AcJ/n L) + exp( - K e-AcJ/n L)]

+'l/JO(Xb - a)'l/JO(xa -

a)~

[exp(Ke- Acl/nL) - exp(-Ke-Ac1/nL)]

+(Xb -+ -Xb) + (xa

-+

-xa) + (Xb

-+

-Xb , Xa

-+

-xa) }.

The right-hand side is recombined to 1

J2['l/JO(Xb - a)

+ 'l/JO(Xb + a)] x exp [-

1 x J2['l/Jo(xa - a)

+ 'l/JO(xa + a)]

(~- Ke - AcJ/n) L]

1

1

+ J2['l/JO(Xb - a) - 'l/JO(Xb + a)] x J2['l/Jo(xa - a) - 'l/JO(xa + a)] x exp [-

(~ + K e- AcJ/n) L] .

(17.198)

Here we identify the ground state wave function as the symmetric combination of the ground state wave functions of the individual wells (17.199) Its energy is (17.200) The first excited state has the antisymmetric wave function (17.201) and the slightly higher energy [(1)

=

L:l.E(O)

E(O)

+ - 2 - = (w/2 + Ke-AcJ/n) n.

(17.202)

The level splitting is therefore (17.203)

1194

17 Tunneling

Inserting K from (17.131), we obtain the formula D.E

=

4v'3V ;;~ fiwe - .AcIl'"',

(17.204)

with Al = (2/3)a 2 w. When expressing the action in terms of the height of the potential barrier Vrnax = a 2 w 2 /8 = 3wAI/16, the formula reads

(17.205) The level splitting decreases exponentially with increasing barrier height. Note that Vrnax is related to the coupling constant of the x 4-interaction by Vrnax = w4 /16g. To ensure the consistency of the approximation we have to check that the assumption of a low density gas of kinks and antikinks is self-consistent. When looking at the series (17.195) for the exponential (17.196), we see that the average number of contributing terms is given by

(17.206) The associated average separation between kinks and antikinks is

(17.207) If we compare this with their size

2/w, we find the ratio

fiw distance size ;:::, D.E·

(17.208)

For increasing barrier height, the level splitting decreases and the dilution increases exponentially. Thus the dilute-gas approximation becomes exact in the limit of infinite barrier height.

17.8

Fluctuation Correction to Level Splitting

Let us calculate the first fluctuation correction to the level splitting formula (17.204). For this we write the potential (17.1) as in (5.78) : w2

V(x) = - - x 4

2

1 + 9-4 x 4 + -49'

(17.209)

with the interaction strength (17.210) Expanding the action around the classical solution, we obtain the action of the fluctuations y(T) = X(T) - Xci (T). Its quadratic part was given in Eq. (17.40) which we write as (17.211)

1195

17.8 Fluctuation Correction to Level Splitting with the functional matrix

OW(T,T')

=[-~+W2(1-~ 1 )]'8(T-T') dT2 2 cosh 2[w(T - To)/2]

(17.212)

associated with the Schriidinger operator for a particle in a Rosen-Morse potential (14.158). The prime indicates the absence of the zero eigenvalue in the spectral decomposition of Ow (T, T') . Since the associated mode does not perform Gaussian fluctuations, it must be removed from Y(T) and treated separately. At the semiclassical level, this was done in Subsection 17.3.1, and the zero eigenvalue appeared in the level splitting formula (17.204) as a factor (17.112). The removal gave rise to an additional effective interaction (17.110):

A:

ff =

-!Hog [1 + A;;;I

J

dT X~I(T)Y'(T)] .

(17.213)

With (17.88)-(17.91), this can be rewritten after a partial integration as (17.214)

The interaction between the fluctuations is (17.215) In the path integral, we now perform a Taylor series expansion of the exponential e-(A}f' + A:ff)/1i in powers of the coupling strength g. A perturbative evaluation of the correlation functions of the fluctuations Y(T) according to the rules of Section 3.20 produces a correction factor to the path integral

(17.216)

where h, 12 , and

h are the dimensionless integrals running over the entire T-axis:

::2 J ;~; J

h

dT (y4(T))Ow'

h

h

=

dTdT' Xcl(T)(y 3(T)y 3(T'))Ow Xcl (T'),

-~: Ifj

J

dTdT'

(17.217)

Y~(T)(Y(T)y3(T'))OwXcl(T').

In order to check the dimensions we observe that the classical solution (17.28) can be written with (17.210) as Xcl(T) = ylw 2/2gtanh[W(T - To)/2], while y(T) and T have the dimensions and l/w, respectively. The Dirac brackets (.. ,)ow denote the expectation with respect to the quadratic fluctuations controlled by the action (17.211) . Due to the absence of a zero eigenvalue, the fluctuations are harmonic. The expectation values of the various powers of Y(T) can therefore be expanded according to the Wick rule of Section 3.17 into a sum of pair contractions involving products of Green functions

Vh7W

G~JT,T')

=

(Y (T)Y(T') )ow = 1i0;:/(T,T') ,

(17.218)

where O;;;-l(T, T') denotes the inverse of the functional matrix (17.212) . The first term in (17.217) gives rise to three Wick contractions and becomes (17.219)

1196

17 Tunneling

The integrand contains an asymptotically constant term which produces a linear divergence for large L . This divergence is subtracted out as follows:

h

= L3w -

16

3 + -3w 4h2

J

['2 ( )

2

d7 Go 7 7 - -h 2 ] . w' 4w

(17.220)

The first term is part of the first-order fluctuation correction without the classical solution, i.e., it contributes to the constant background energy of the classical solution. It is obtained by replacing

G,2 (77') Ow

'

hG (7 - 7') =

---t

w

~e-WIT-T' 1 2w

(17.221)

[recall (3.301) and (3.246)] . In the amplitudes (17.195), the background energy changes only the exponential prefactor e- wL / 21i to e-(l+3gli/16w 3 )wL/21i and does not contribute to the level splitting. The level splitting formula receives a correction factor C,

=

[1 -

gh3 CI w

+ . .. ] =

[ 1 - (' II

+ 12, + 13') wgh3 + 0 (92)]

,

(17.222)

in which all contributions proportional to L are removed. Thus h is replaced by its subtracted part 1~ == h - L3w /16. The integral 12 has 15 Wick contractions which decompose into two classes:

Each of the two subintegrals hi and h2 contains a divergence with L which can again be found via the replacement (17.221). The subtracted integrals in (17.222) are 1~1 = 121 + wL/8 and 1~2 = 122 +3wL/16. Thus, altogether, the exponential prefactor e- wL / 21i in the amplitudes (17.195) is changed to e-[1/2+(3/16-1/8-9/16)gli/w 3 )wL/21i = e -(1 /2-91i/2w 3 )wL/21i, in agreement with (5.258). To compare the two expressions, we have to set w = v'2 since the present w is the frequency at the bottom of the potential wells whereas the w in Chapter 5 [which is set equal to 1 in (5.258)] parametrized the negative curvature at x = O. The Wick contractions of the third term lead to the finite integral

Is =

1~ = -3~:

Ifi J

d7d7'

yb(7)G~j7, 7')G~j7', 7')Xcl(7').

(17.224)

The correction factor (17.216) can be pictured by means of Feynman diagrams as

C = I-3 CD +~(6 ~ +9 ~ ) + 3 ~ +0(g2),

(17 .225)

where the vertices and lines represent the analytic expressions shown in Fig. 17.5. For the evaluation of the integrals we need an explicit expression for G~w (7,7') . This is easily found from the results of Section 14.4.4. In Eq. (14.161), we gave the fixed-energy amplitude (xb lxa)ERM ,EpT solving the Schriidinger equation

(17.226)

1197

17.8 Fluctuation Correction to Level Splitting Inserting EpT =

(n? /2J-L)s(s + 1), the amplitude reads for Xb > Xa

-iJ-L (xb lxa)ERM ,EpT = Tf(m(EnM) - s)f(s + m(EnM)

+ 1)

xps-m(ERMl (tanh Xb)Ps-m(ERMl (- tanhx a ),

(17.227)

with (17.228) After a variable change x = WT /2 and h 2/ J-L = w2/2 , we set s = 2 and insert the energy EnM = -3w 2/4 . Then the operator in Eq. (17.226) coincides with Ow(T, T') of Eq. (17.212), and we obtain the desired Green function for T > T'

h Go (T, T') = -f(m - 2)f(m + 3) P'2w

m

W

WT WT' (tanh - ) P'2- m ( - tanh - ), 2 2

(17.229)

with m = 2. Due to translational invariance along the T-axis, this Green function has a pole at EnM = - 3w 2/4 which must be removed before going to this energy. The result is the subtracted Green function G ow (T, T') which we need for the perturbation expansion. The subtraction procedure is most easily performed using the formula Gow = (d/dEnM)EnMGow IERM=-3w2/4' In terms of the parameter m , this amounts to

GOw (T, T') =

2~ d~ (m 2 -

4)Gow(T,

T')lm=2'

(17.230)

Inserting into (17.227) the Legendre polynomials from (14.165) ,

p;m(z) =

f(l~m) c~:rm/2

[1-I:m(I-Z)+

(l+m)~2+m)(I-Z)2] ,

(17.231)

the Green function (17.230) can be written as (17.232) where T> and T< are the greater and the smaller of the two times T and T' , respectively, and YO(T) , YO(T) are the wave functions

~ -2 ( -tanhWT) = YO(T) = -2v6wP2 2

~

X >VV\.IV'o---

fi- - - 2 - ' 1 8 cosh

(17.233)

W2T

9

/iXcI(T) 9

4h

!fiY~(T) GOw(T,T')

Figure 17.5 Eq. (17.225).

Vertices and lines of Feynman diagrams for correction factor C in

1198

17 Tunneling

(17.234) From (17.231) we see that

d~ P2where "1

~

m

(tanh

~r) Im=2=

yo(r)[6(3 - 2"1 + wr) - e- WT (8 + e- WT )],

:

(17.235)

0.5773156649 is the Euler-Mascheroni constant (2.467) . Hence

Yo(r)

1

12w2Yo(r)[e- WT(e - WT

=

+ 8) -

2(2 + 3wr)] .

(17.236)

For r = r', the Green function is

, (r, r) Go ~

n

=

1

---2wcosh4~T

( cosh4 -wr 2

+ cosh2 -wr 2

- -11 ) . 8

(17.237)

Note that an application of the Schriidinger operator (17.212) to the wave functions Yo(r) and Yo(r) produces - yo(r) and 0, respectively. These properties can be used to construct the Green function Go~ (r, r') by a slight modification of the Wronski method of Chapter 3. Instead of the differential equation OwG( r , r') = n8( r - r'), we must solve the projected equation O~GoJr, r') =

n[8(r - r') - yo(r)yo(r')] ,

(17.238)

where the right-hand side is the completeness relation without the zero-eigenvalue solution:

LYn(r)Yn(r') = 8(r - r') - yo(r)yo(r') .

(17.239)

n#O

The solution of the projected equation (17.238) is precisely given by the combination (17.232) of the solutions Yo(r) and yo(r) with the above-stated properties. The evaluation of the Feynman integrals h,I21,I22,I3 is somewhat tedious and is therefore described in Appendix 17A. The result is

I'-~ 1 - 560'

I'

-~ 420'

21 -

I' _ 117 22 - 560'

I _ 49 3 - 20'

(17.240)

These constants yield for the correction factor (17.222)

C'

=

[1 _ 71 gn 24w 3

+ O(9 2)] ,

(17.241)

modifying the level splitting formula (17.204) for the ground state energy to l:l.E(O) =

4V3V w27rn/3g 3

liwe - w3 /3gli - 71gli/24w 3 + ....

(17.242)

This expression can be compared with the known energy eigenvalues of the lowest two double-well states. In Section 5.15, we have calculated the variational approximation W 3 (xo) to the effective classical potential of the double well and obtained for small 9 an energy (see Fig. 5.24) which did not yet incorporate the effects of tunneling. We now add to this the level shifts ±l:l.E(O) /2 from Eq. (17.242) and obtain the curves also shown in Fig. 5.24. They agree reasonably well with the Schriidinger energies.

1199

17.9 Tunneling and Decay

'--v

x ..

\r

v-

\'t,\.

~

",ill

",ill

\11\11

(1]aJr

" I

'\J min

V

'v0 \.,r-l

Figure 17.6 Positions of extrema Xex in asymmetric double-well potential, plotted as function of asymmetry parameter E. If rotated by 90°, the plot shows the typical cubic shape. Between E> and E 0, this slightly depresses the left minimum at x of the extrema are found from the cubic equation

E _]

V'(x ex ) = w 2a [(X ex )3 _ Xex _ _ = O. 2 a a w 2 a2

= -a. The positions

(17.244)

They are shown in Fig. 17.6. For large E, there is only one extremum, and this is always a minimum. In the region where Xex has three solutions, say X_ , Xo, X+, the branches denoted by "rei min." in Fig. 17.6 correspond to relative minima which lie higher than the absolute minimum. The central branch corresponds to a maximum. As E decreases from large positive to large negative values, a classical particle at rest at the minimum follows the upper branch of the curve and drops to the lower branch as E becomes smaller than E 00, this becomes an expansion for the ground state energy E(O)(g). In the limit L -> 00, Z (g) behaves like

Z( 9 )

->

e

_ E(O) (g)L

(17.281 )

,

exhibiting directly the ground state energy. Since the path integral can be done exactly at the point 9 = 0, it is suggestive to expand the exponential in powers of 9 and to calculate the perturbation series

Z(g) =

L

Zk ( g3)k

k=O

W

(17.282)

As shown in Section 3.20, the expansion coefficients are given by the path integrals

(17.283) By selecting the connected Feynman diagrams in Fig. 3.7 contributing to this path integral, we obtain the perturbation expansion in powers of 9 for the free energy F. In the limit L -> 00, this becomes an expansion for the ground state energy E(O)(g), in accordance with (17.281). By following the method in Section 3.18, we find similar expansions for all excited energies E(n)(g) in powers of g. For 9 = 0, the energies are, of course, those of a harmonic oscillator, n ) = w(n + 1/2). In general, we find the series

Ea

(17.284)

1209

17.10 Large-Order Behavior of Perturbation Expansions

Most perturbation expansions have the grave deficiency observed in Eq. (3C.27) . Their coefficients grow for large order k like a factorial k! causing a vanishing radius of convergence. They can yield approximate results only for very small values of g. Then the expansion terms Ekn ) (g/4)k decrease at least for an initial sequence of kvalues, say for k = 0, .. . ,N. For large k-values, the factorial growth prevails. Such series are called asymptotic. Their optimal evaluation requires a truncation after the smallest correction term. In general, the large-order behavior of perturbation expansions may be parametrized as E k -- I'P1J+1kf3(_4 a )k(Pk)'.

[1 + 1'1k + 1'2k + ...] ,

(17.285)

2

where the leading term (pk)! grows like (pk)! = (k!)p(pp)kk(1 - p)/2

~

(21l')(p-1)/2

[1 + O(l/k)] .

(17.286)

This behavior is found by approximating n! via Stirling's formula (5.204). It is easy to see that the kth term of the series (17.284) is minimal at k ~ kmin

1

(17.287)

== p(algI)1/p'

This is found by applying Stirling's formula once more to (k!)P and by minimizing l'(k!)Pkf3' (ppalgl)k with (3' = (3 + (1 - p)/2, which yields the equation p log k + log(ppalgl)

+ ((3 + p/2)/k + ... = O.

(17.288)

An equivalent way of writing (17.285) is Ek = I'P( -4a) k r(pk

+ (3 + 1)

[1 c 1+ (3 + pk

C2 + (pk + (3)(pk + (3 _

1)

+ ...] .(17.289)

The simplest example for a function with such strongly growing expansion coefficients can be constructed with the help of the exponential integral (17.290) Defining (17.291) this has the diverging expansion E(g) = 1 - 9 + 2!g2 - 3!g3 + ... + (_l)N N!gN

+ ...

(17.292)

At a small value of g, such as 9 = 0.05, the series can nevertheless be evaluated quite accurately if truncated at an appropriate value of N. The minimal correction

1210

17 Tunneling

is reached at N = 1/9 = 20 where the relative error with respect to the true value E ;:::; 0.9543709099 is equal to b.E/E ;:::; 1.14.10-8 . At a somewhat larger value 9 = 0.2, on the other hand, the optimal evaluation up to N = 5 yields the much larger relative error;:::; 1.8%, the true value being E ;:::; 0.852110880. The integrand on the right-hand side of (17.291), the function

B(t) __1_ - 1 +t'

(17.293)

is the so-called Borel transform ofthe function E(g). It has a power series expansion which can be obtained from the divergent series (17.292) for E(g) by removing in each term the catastrophically growing factor kL This produces the convergent series

B(t) = 1 - t + t 2 - t 3 + ... ,

(17.294)

which sums up to (17.293). The integral

F(g) =

loo

OO

dt _e- t / g B(t) 9

(17.295)

restores the original function by reinstalling, in each term tk, the removed k!-factor. Functions F(g) of this type are called Borel-resummable . They possess a convergent Borel transform B(t) from which F(g) can be recovered with the help ofthe integral (17.295). The resummability is ensured by the fact that B(t) has no singularities on the integration path t E [0, 00) , including a wedge-like neighborhood around it. In the above example, B(t) contains only a pole at t = -1 , and the function E(g) is Borel-resummable. Alternating signs of the expansion coefficients of F(g) are a typical signal for the resummability. The best-known quantum field theory, quantum electrodynamics , has divergent perturbation expansions, as was first pointed out by Dyson [5]. The expansion parameter 9 in that theory is the fine-structure constant

0: =

1/137.035963(15) ;:::; 0.0073.

(17.296)

Fortunately, this is so small that an evaluation of observable quantities, such as the anomalous magnetic moment of the electron

10:

(0:)2 +1.1765(13) (0:)3 + . ..

ae =b.f.t - = - - - 0.3284789657 f.t

21f

1f

,

(17.297)

1f

gives an extremely accurate result: a!heor =

(1159652478 ± 140) . 10- 12 .

(17.298)

The experimental value differs from this only in the last three digits, which are 200 ± 40. The divergence of the series sets in only after the 137th order.

1211

17.10 Large-Order Behavior of Perturbation Expansions

A function E(g) with factorially growing expansion coefficients cannot be analytic at the origin. We shall demonstrate below that it has a left-hand cut in the complex g-plane. Thus it satisfies a dispersion relation E( ) = _1 9

27l'2'

loo d ,discE(-g') 0

g, 9 +g

'

(17.299)

where disc E(g') denotes the discontinuity across the left-hand cut discE(g)

== E(g - iTJ) - E(g + iTJ).

(17.300)

It is then easy to see that the above large-order behavior (17.289) is in one-to-one

correspondence with a discontinuity which has an expansion, around the tip of the cut,

The parameters are the same as in (17.289). The one-to-one correspondence is proved by expanding the dispersion relation (17.299) in powers of g/4, giving

Ek = (-4)

kl

OO

o

dg' 1 . , - . -dlscE(-g). 27l'2 g'k+1

(17.302)

The expansion coefficients are given by moment integrals of the discontinuity with respect to the inverse coupling constant 1/g. Inserting (17.301) and using the integral formula 5

roo d9 _1_ - l/(alg ll(l/p) '" r( ) Igl"'+1 e - a p pa,

io

(17.303)

we indeed recover (17.289). From the strong-coupling limit of the ground state energy of the anharmonic oscillator Eq. (5.168) we see that the discontinuity grow for large 9 like gl/3. In this case, the dispersion relation (17.304) needs a subtraction and reads

E(g) = E(O)

+~ 27l'2

roo dg' discE(-g').

io

g'

g'

+9

(17.304)

This does not influence the moment formula (17.302) for the expansion coefficients, except that the lowest coefficient is no longer calculable from the discontinuity. Since the lowest coefficient is known, there is no essential restriction.

17.10.2

Semiclassical Large-Order Behavior

The large-order behavior of many divergent perturbation expansions can be determined with the help of the tunneling theory developed above. Consider the potential 51.8. Gradshteyn and 1.M. Ryzhik, op. cit., Formula 3.478.

1212

17 Tunneling

of the anharmonic oscillator at a small negative coupling constant 9 (see Fig. 17.11). The minimum at the origin is obviously metastable so that the ground state has only a finite lifetime. There are barriers to the right and left of the metastable minimum, which are very high for very small negative coupling constants. In this limit, the lifetime can be calculated accurately with the semiclassical methods of the last section. The fluctuation determinant yields an imaginary part of Z(g) of the form (17.270) , which determines the imaginary part of the ground state energy via (17.276) , which is accurate near the tip of the left-hand cut in the complex g-plane. From this imaginary part, the dispersion relation (17.302) determines the large-order behavior of the perturbation coefficients. The classical equation of motion as a function of 7 is

X"(7) - V'(X(7))

O.

=

(17.305)

The differential equation is integrated as in (17.26) , using the first integral of motion

1

1'2 - -w22 -x x - -g4 x = E = const 2 2 4 ' from which we find the solutions for E = 0 7 - 70 = ± -1

w

J

dx

1 = =f -1 arcosh xVI - (lgl /2w 2 )x2 w

(17.306)

(~W21) -Igl

x

(17.307)

'

or

X(7) = Xcl(7) == ±

W2 ~ -I 9 1cosh[

1

(17.308)

( )]" W 7 - 70

They represent excursions towards the abysses outside the barriers and correspond precisely to the bubble solutions of the tunneling discussion in the last section. The excursion towards the abyss on the right-hand side is illustrated in Fig. 17.11. The associated action is calculated as in (17.29) :

1 d7 [1 L/2

-L~

2

where

Xm

-X~~(7)

2

]

+ V(Xcl(7)) = 2 1L/2 d7 [X~(7) 0

foxmdxV2(E + V) -

EL,

(17.309)

is the maximum of the solution. The bubble solution has E

(Xm

Act = 2 Jo

E]

dxv'2V =

4w 3

3jgf.

= 0, so that (17.310)

Inserting the fluctuating path x( 7) = Xcl (7) + y( 7) into the action (17.278) and expanding it in powers of Y(7), we find an action for the quadratic fluctuations of the same form as in Eq. (17.211) , but with a functional matrix

Ow(7, 7') =

d2 ]' < V2w2/lgl . Its action is approximately given by

A 'cl In comparison with the limit L is replaced by

~ ---7

4w3( 1 - 12e-wL) . 3jgf 00,

(17.320)

the Boltzmann-like factor e- Acl of this solution

(17.321) The exponentials in the sum raise the reference energies in the imaginary part of Z( -lgl-i1]) in (17.316) from w/2 to w(n+1/2). The imaginary parts for the energies to the nth excited states become n . 12w ImE( n )(-Igl- Z1])= - -

n!

(f6 ~ -4w3 Y;: 31g1

1 +2n

e

_ 4w 3 3

/Igl

(17.322)

implying an asymptotic behavior of the perturbation coefficients: n

Wff12 3 k Ek(n) = -- - I (-3/w ) r(k + n + 1/2). 7f 7f n.

(17.323)

It is worth mentioning that within the semiclassical approximation, the dispersion integrals for the energies can be derived directly from the path integral (17.279). This can obviously be rewritten as

1 Jo JVX(T) iOO

Z(g)

dA.

roo

-ioo 27fZ

x

da e-(gaHa)/4 4

exp { -

i~~2 dT [~X~ + ~2 X2 - ~X4]} .

(17.324)

The integration over A generates a 8-function 48(JdTX4(T) - a) which eliminates the additionally introduced a-integration. The integral over a is easily performed. It yields a factor l/(A + g), so that we obtain the integral formula

Z(g) =

1

ioo

dA. _l_Z( _A).

-ioo 27fZ

A+ g

(17.325)

The integrand has a pole at A = -g and a cut on the positive real A-axis. We now deform the contour of integration in A until it encloses the cut tightly in the clockwise sense. In the semiclassical approximation, the discontinuity across the cut is given by Eq. (17.316) , i.e., with the present variable A: (17.326)

1216

17 Tunneling

On the upper branch of the cut, v').. is positive, on the lower negative. Thus we arrive at a simple dispersion integral from A = 0 to A = 00:

Z(g)=2w ('.JO dA_1_ f§.J4W 3 e - 4w3/3Ae - WL/2 . Jo 27r A + g Y-; 3A

(17.327)

For the ground state energy, this implies

E(O)(g)= _ 2w {'Xl dA_1_ f§.J4w 3 e - 4w3/3A . Jo 27r A + g Y-; 3A

(17.328)

Of course, this expression is just an approximation, since the integrand is valid only at small A. In fact, the integral converges only in this approximation. If the full imaginary part is inserted, the integral diverges. We shall see below that for large A, the imaginary part grows like A1/3. Thus a subtraction is necessary. A convergent integral representation exists for E(O) (g) - E(O)(O). With E(O)(O) being equal to w/2, we find the convergent dispersion integral

E(O)(

g

)=~ 2

+

roo dA

2 wg Jo

1 f§.J4w 3 - 4w 3 /3A 27rA(A+g)Y-; 3A e .

(17.329)

The subtraction is advantageous also if the initial integral converges since it suppresses the influence of the large-A regime on which the semiclassical tunneling calculation contains no information. After substituting A ~ 4g/3tw 3 , the integral (17.329) is seen to become a Borel integral of the form (17.295). By expanding 1/ (A + g) in a power series in g _ 1_ A+g

= f(-l)kl A- k- t, k=O

(17.330)

we obtain the expansion coefficients as the moment integrals of the imaginary part as a function of 1/g: (17.331) This leads again to the large-k behavior (17.319). The direct treatment of the path integral has the virtue that it can be generalized also to systems which do not possess a Borel-resummable perturbation series. As an example, one may derive and study the integral representation for the level splitting formula in Section 17.7.

17.10.3

Fluctuation Correction to the Imaginary Part and Large-Order Behavior

It is instructive to calculate the first nonleading term clalgl in the imaginary part (17.301), which gives rise to a correction factor 1 + cdk in the large-order behavior

1217

17.10 Large-Order Behavior of Perturbation Expansions

(17.289). As in Section 17.8, we expand the action around the classical solution. The interaction between the fluctuations Y(7) is the same as before in (17.215). The quadratic fluctuations are now governed by the differential operator Ow(7,7') =

[-dd?2+W2(1~7 7 cosh 2

W

70

))]'oo

(17.393)

1230

17 Tunneling

At the extremum, fk(r) has the value

fk -----; k log(3k/4eo-) - 20".

(17.394)

k-->oo

The constant -20" in this limiting expression arises when expanding the second term of Eq. (17.391) into a Taylor series, (40"/3')')(1- ,),)3/2 = 40"/3')'k - 20" + .... Only the first two terms survive the large-k limit. Thus, to leading order in k, the kth term of the re-expanded series becomes (17.395) The corresponding re-expansion coefficients are (0)

ck

ex: e

-2uE(0)

(17.396)

k.

They have the remarkable property of ~rowing in precisely the same manner with k as the initial expansion coefficients E kO), except for an overall suppression factor e- 2u . This property was found empirically in Fig. 5.20b. In order to estimate the convergence of the variational perturbation expansion, we note that with (17.397) and fj from (5.213), we have All O"g = - ~V.

(17.398)

For large n, this expression is smaller than unity. Hence the powers (O"fj)k alone yield a convergent series. An optimal re-expansion of the energy can be achieved by choosing, for a given large maximal order N of the expansion, a parameter 0" proportional to N: (17.399) Inserting this into (17.391), we obtain for large k

=N (17.400)

The extremum of this function lies at (17.401)

1231

17.10 Large-Order Behavior of Perturbation Expansions

The constant e is now chosen in such a way that the large exponent proportional to N in the exponential function efN(-y) due to the first term in (17.400) is canceled by an equally large contribution from the second term, i.e. , we require at the extremum (17.402) The two equations (17.401) and (17.402) are solved by

"( = -0.242964029973520 ... ,

e = 0.186047272 987 975 .. .

(17.403)

In contrast to the extremal "( in Eq. (17.393) which dominates the large-k limit, the extremal "( of the present limit, in which k is also large but of the order of N, remains finite (the previous estimate holds for k » N). Accordingly, the second term (4e/3"()(1 - ,,()3/2 in fN("() contributes in full, not merely via the first two Taylor expansion terms of (1- "()3/2, as it did in (17.394). Since fN("() vanishes at the extremum, the Nth term in the re-expansion has the order of magnitude SN ex (IJNgN)N

=

(1 - o\J

(17.404)

N

According to (17.397) and (17.399), the frequency ON grows for large N like (17.405) As a consequence, the last term of the series decreases for large N like

S (C) N

1

[1 _ 1 ]

N

ex

(IJNg)2/3

~

~ e

_N/(ug)2/3

~

~ e

_N 1/3/(cg)2/3

.

(17.406)

This estimate does not yet explain the convergence of the variational perturbation expansion in the strong-coupling limit observed in Figs. 5.21 and 5.22. For the contribution of the cut C 1 to SN, the derivation of such a behavior requires including a little more information into the estimate. This information is supplied by the empirically observed property, that the best ON-values lie for finite N on a curve [recall Eq. (5.211)]: IJN '" eN

6.85) ( 1 + N2/3 .

(17.407)

Thus the asymptotic behavior (17.399) receives, at a finite N, a rather large correction. By inserting this IJN into f N ("() of (17.400), we find an extra exponential factor 4e (1 - ,,()3/2 6.85] ~ exp [N 3 "( N2/3 97N1/3 6.85] exp [-Nlog(-"() N2/3 ~ e- ' .

(17.408)

1232

17 Tunneling

This reduces the size of the last term due to the cut G1 in (17.406) to _[9.7+(cg) - 2/3]Nl/3 SN (G) 1 cx:e ,

(17.409)

which agrees with the convergence seen in Figs. 5.21 and 5.22. There is no need to evaluate the effect of the shift in the extremal value of I caused by the correction term in (17.407), since this would be of second order in

I/N 2/3. How about the contributions of the other cuts? For GI, the integrals in (17.384) run from g = -2/1J to -00 and decrease like (_2/1J) - k. The associated last term SN(GI) is of the negligible order e-NlogN. For the cuts G 2 ,2,3, the integrals (17.384) start at g = 1/1J and have therefore the leading behavior

s k(o)(C2,2,3 )

'"

IJ k .

(17.410)

This implies a contribution to the Nth term in the re-expansion of the order of (17.411) which decreases merely like (17.406) and does not explain the empirically observed convergence in the strong-coupling limit. As before, an additional information produces a better estimate. The cuts in Fig. 17.16 do not really reach the point IJg = 1. There exists a small circle of radius b.f} > 0 in which E(O)(f}) has no singularities at all. This is a consequence of the fact unused up to this point that the strong-coupling expansion (5.231) converges for g > gs. For the reduced energy, this expansion reads:

E(O)(g) =

A I ] -2/3 [A 1 ] -4/3 } ( ~A) 1/3 { ao + a1 [ 4~3 (1 _ 1JfJ)3/2 + a2 4~3 (1 _ IJg)3/2 + ... . (17.412)

The convergence of (5.231) for g > gs implies that (17.412) converges for alllJg in a neighborhood of the point IJg = 1 with a radius (17.413) where gs == gs/w 3. For large N , b.(lJg) goes to zero like 1/(Nlgsle)2/3. Thus the integration contours of the moment integrals (17.384) for the contributions sLO) (Gi ) of the other cuts do not begin at the point IJg = 1, but a little distance b.(lJg) away from it. This generates an additional suppression factor (17.414) Let us set -gs = Igslexp(i'Ps) and Xs == (-g/gs?/3 = -l xs lexp(iB) , and introduce the parameter a == 1/[igsleF/3 . Since there are two complex conjugate contributions we obtain, for large N a last term of the re-expanded series the order of (17.415)

1233

17.10 Large-Order Behavior of Perturbation Expansions exp(6.41 - 9.42N 1 / 3 ) 3

4

5

-10

-20

-30

-40

Figure 17.17 Theoretically obtained convergence behavior of Nth approximants for to be compared with the empirically found behavior in Fig. 5.21.

0::0,

Figure 17.18 Theoretically obtained oscillatory behavior around exponentially fast asymptotic approach of 0::0 to its exact value as a function of the order N of the approximant, to be compared with the empirically found behavior in Fig. 5.22, averaged between even and odd orders. By choosing

Igsl '"" 0.160,

B '"" -0.467,

(17.416)

we obtain the curves shown in Figs. 17.17 and 17.18 which agree very well with the observed Figs. 5.21 and 5.22. Their envelope has the asymptotic falloff e-9.23Nl/3. Let us see how the positions of the leading Bender-Wu singularities determined by (17.416) compare with what we can extract directly from the strong-coupling series (5.231) up to order 22. For a pair of square root singularities at Xs = -Ixsl exp(±iB), the coefficients of a power series L:: Q:nxn have the asymptotic ratios Rn == Q:n+d Q:n '""

1234

17 Tunneling

0.2

+

.

0.1

+

5

-0.1 -0.2

••• ~ t +

10

"

20 "

" 15

n

**

* *"

+

+

-0.3 -0.4

Figure 17.19 Comparison of ratios Rn between successive expansion coefficients of strong-coupling expansion (dots) with ratios R':;' of expansion of superposition of two singularities at 9 = 0.156 x exp(±0.69) (crosses) .

0.62 0.6 0.58

3rd order PT /

0.56

O.M 0.52

/

_ ./

::;.-

.6-~_~ \ exact

0.1

0.2

0.3

0.4

0.5

9

Figure 17.20 Strong-coupling expansion of ground state energy in comparison with exact values and perturbative results of 2nd and 3rd order. The convergence radius in 1/9 is larger than 1/ 0.2.

17.11 Decay of Supercurrent in Thin Closed Wire

1235

R': == - cos[(n + 1)0 + J]/Ixsl cos(nO + (5). In Fig. 17.19 we have plotted these ratios against the ratios R,. obtained from the coefficients an of Table 5.9 . For large n, the agreement is good if we choose Ixsl

= 1/0.117, 0 = -0.467,

(17.417)

with an irrelevant phase angle 15 = -0.15. The angle 0 is in excellent agreement with the value found in (17.416). From Ixsl we obtain Igsl = 411/xs I3 / 2 = 0.160, again in excellent agreement with (17.416). This convergence radius is compatible with the heuristic convergence of the strong-coupling series up to order 22, as can be seen in Fig. 17.19 by comparing the curves resulting from the series with the exact curve. It is possible to extend the convergence proof to the more general divergent power series discussed in Section 5.17, whose strong-coupling expansions have the more general growth parameters p and q [14]. The convergence is assured for 1/2 < 2/q < 1 [15]. If the interaction of the anharmonic oscillator is J dT Xn(T) with n =J 4, the dimensionless expansion parameter for the energies is g/ wn/2+l rather than g/ w 3 . Then q = n/2 + 1, such that for n ~ 6 the convergence is lost. This can be verified by trying to resum the expansions for the ground state energies of n = 6 and n = 8, for example. For n = 6, the cut in Fig. 17.16 becomes circular such that there is no more shaded circle C3 in which the strong-coupling series converges.

17.11

Decay of Supercurrent in Thin Closed Wire

An important physical application of the above tunneling theory explains the temperature behavior of the resistance of a thin7 superconducting wire. The superconducting state is described by a complex order parameter 'lj;(z) depending on the spatial variable z along the wire. We then speak of an order field. The variable z plays the role of the Euclidean time T in the previous sections. We shall consider a closed wire where 'lj;(z) satisfies the periodic boundary condition

'lj;(z)

=

'lj;(z

+ L) .

(17.418)

The energy density of the system is described approximately by a Ginzburg-Landau expansion in powers of 'lj; and its gradients containing only the terms (17.419)

The total fluctuating energy is given by the functional

E['lj;*,'lj;] =

1

L/2

dzc:(z),

(17.420)

-L/2

7 A superconducting wire is called thin if it is much smaller than the coherence length to be defined in Eq. (17.425).

1236

17 Tunneling

and the probability of each fluctuation is determined by the Boltzmann factor exp{ -E['I,b*, 'l,bl/kBT}. The parameter m 2 in front of 1'I,b(zW is called the mass term of the field. It vanishes at the critical temperature Te and behaves near Te like m2

~ m~ (~ -1).

(17.421)

Below Te , the square mass is negative and the wire becomes superconducting. One can easily estimate, that each term in the Landau expansion is of the order of 11 - T/TeI 2 and any higher expansion term in (17.419) would be smaller than that by at least a power 11- T/TeI 1 / 2 . The partition function of the system is given by the path integral (17.422)

If T does not lie too close to Te [although close enough to justify the Landau expansion, i.e., the neglect of higher expansion terms in (17.419) suppressed by a factor 11 - T /TeI 1 / 2 ], this path integral can be treated semiclassically in the way described earlier in this chapter [16]. The basic microscopic mechanism responsible for the phenomenon of superconductivity will be irrelevant for the subsequent discussion. Let us only recall the following facts: A superconductor is a metal at low temperatures whose electrons near the surface of the Fermi sea overcome their Coulomb repulsion due to phonon exchange. This enables them to form bound states between two electrons of opposite spin orientations in a relative s-wave, the celebrated Cooper pairs. 8 The attraction which binds the Cooper pairs is extremely weak. This is why the temperature has to be very small to keep the pairs from being destroyed by thermal fluctuations. The critical temperature Te , where the pairs break up, is related to the binding energy of the Cooper pairs by E pair = kBTe . The field-theoretic process called phonon exchange is a way of describing the accumulation of positive ions along the path of an electron which acts as an attractive potential wake upon another electron while screening the Coulomb repulsion. The attraction is very weak and leads to a bound state only in the s-wave (the centrifugal barrier ex: l(l + 1)/r2 preventing the formation of a bound state in higher partial waves). The potential between the electrons may well be approximated by a 5-function potential V(x) ~ -g5(r). The critical temperature Te , usually a few degrees Kelvin, is found to satisfy the characteristic exponential relation (17.423)

The parameter J.l denotes the upper energy cutoff of the phonon spectrum TDkB' where TD is the Debye temperature of the lattice vibration. 8We consider here only with old-fashioned superconductivity which sets in below a very small critical temperature of a few-degree Kelvin. The physics of the recently discovered hightemperature superconductors is at present not sufficiently understood to be discussed along the same lines.

17.11 Decay of Supercurrent in Thin Closed Wire

1237

9

Figure 17.21 Renormalization group trajectories in the g , {t plane of superconducting electrons (g=attractive coupling constant, {t=Debye temperature, kB = 1). Curves with same Te imply identical superconducting properties. The renormalization group determines the reparametrizations of a fixed superconductive system along any of these curves. An important result of the theory, confirmed by experiment, is that all Tdependent characteristic equilibrium properties of the superconductor near Te depend only on the single parameter Te. Thus, many quite different systems with different microscopic parameters J-l == T D and g will have the same superconducting properties (see Fig. 17.21) . The critical temperature is an important prototype for the understanding of the so-called dimensionally transmuted coupling constant in quantum field theories, which plays a completely analogous role in specifying the system. In quantum field theory, an arbitrary mass parameter J-l is needed to define the coupling strength of a renormalized theory and physical quantities depend only on the combination9

(17.424) The set of all changes J-l which are accompanied by a simultaneous change of g(J-l) such as to stay on a fixed curve with Me from the renormalization group. The curve J-l, g(J-l) is called the renormalization group trajectory [15]. If one works in natural units with ti = kB = M = 1, the critical temperature corresponds to a length of;::;:J lOOOA. This length sets the scale for the spatial correlations of the Cooper pairs near the critical point via the relation ~(T)

const. =-

Te

(T) -1/2 (T) -1/2 1- Te ;::;:J lOooA 1 -Te-

(17.425)

9In quantum chromodynamics, this dimensionally transmuted coupling constant is of the order of the pion mass and usually denoted by A.

1238

17 Tunneling

The Cooper pairs are much larger than the lattice spacing, which is of the order of 1A. Their size is determined by the ratio ti2kF/me1fkBTe, where kF is the wave number of electrons of mass me on the surface of the Fermi sphere. The temperature Te in conventional superconductors of the order of 1 K corresponds to 1/11604.447 eV. Thus the thermal energy kBTe is smaller than the atomic energy EH = 27.21OeV (recall the atomic units defined on p. 13.7) by a factor 2.345 x 10- 3, and we find that ti2/me1fkBTe is of the order of 102a1. Since aH is of the order of l/kF ' we estimate the size of the Cooper pairs as being roughly 100 times larger than the lattice spacing. This justifies a posteriori the 8-function approximation for the attractive potential, whose range is just a few lattice spacings, i.e., much smaller than ~(T). The presence of such large bound states causes the superconductor to be coherent over the large distance ~(T). For this reason, ~(T) is called the coherence length. Similar Cooper pairs exist in other low-temperature fermion systems such as 3He, where they give rise to the phenomenon of superfiuidity. There, the interatomic potential contains a hard repulsive core for r < 2.7A. This prevents the formation of an s-wave bound state. In addition it produces a strong spin-spin correlation in the almost fully degenerate Fermi liquid, with a preference of parallel spin configurations. Because of the necessary antisymmetry of the pair wave function of the electrons, this amounts to a repulsion in any even partial wave. For this reason, Cooper pairs can only exist in the p-wave spin triplet state. The binding energy is much weaker than in a superconductor, suppressing the critical temperature by roughly a factor thousand. Experimentally, one finds Te = 27mK at a pressure of p = 35 bar. Since the masses of the 3He atoms are larger than those of the electrons by about the same factor thousand, the coherence length ~ has the same order of magnitude in both systems, i.e., l/Te has the same length when measured in units of A. The theoretical description of the behavior of the condensate is greatly simplified by re-expressing the fundamental Euclidean action in terms of a Cooper pair field which is the composite field (17.426) Such a change of field variables can easily be performed in a path integral formulation of the field theory. The method is very similar to the introduction of the auxiliary field tp(x) in the polymer field theory of Section 15.12. Since this subject has been treated extensively elsewhere lO we shall not go into details. The partition function of the system reads (17.427) lOThe way to describe the pair formation by means of path integrals is explained in H. Kleinert, Collective Quantum Fields, Fortschr. Phys. 26,565 (1978) (http://www.physik.fu-berlin. derkleinert/55).

1239

17.11 Decay of Supercurrent in Thin Closed Wire

By going from integration variables 'l/Je to partition function

z=

'l/Jpair> we can derive the alternative pair

JV_I.- . (X)V_I. . (x)e o/pau

o/palr

-A(.p;ai".ppai') ,

(17.428)

where 'l/Jpair is the Cooper pair field (17.426). In general, the new action is very complicated. For temperatures close to Te , however, it can be expanded in powers of the field 'l/Jpair and its derivatives, leading to a Landau expansion of the type (17.419) . For static fields the Euclidean field action is A ['I/J;air, 'l/Jpair] = E /kBT =

=

k~T

Jd x c(x) 3

(17.429)

k~T Jd x [(-log If + :2) l'l/Jpairl2 + 2~; l'l/Jpairl 3

4

+ ~; 1V'l/Jpair 12 + .. .J,

where the dots denote the omitted higher powers of 'l/Jpair and of their derivatives, each accompanied by an additional factor l/Te . Let us discuss the path integral (17.429) first in the classical limit. We observe that with the critical temperature (17.423), the mass term in the energy can be written as - log Te T

T) I'l/Jpairl· 2 I'l/Jpairl 2 rv - ( 1 - Te

(17.430)

It has the "wrong sign" for T < Te , so that the field has no stable minimum at 'l/Jpair = O. It fluctuates around one of the infinitely many nonzero values with the fixed absolute value

(17.431) It is then useful to take a factor Te (1- T/Te)1/2 out of the field 1

'I/J(x) == 'l/Jpair(X) Te(1 _ T /Te)1/2'

'l/Jpair, define (17.432)

and write the renormalized energy density as (17.433)

Here we have made use ofthe coherence length (17.425) to introduce a dimensionless space variable x, replacing x ----t x~. We also have dropped an overall energy density factor proportional to (1 - T /Te)2 T;. In the rescaled form (17.433), the minimum of the energy lies at I'l/Jol = 1, where it has the density

c=ce=-1/2.

(17.434)

1240

17 Tunneling

The negative energy accounts for the binding of the Cooper pairs in the condensate (in the present natural units) and is therefore called the condensation energy. In terms of (17.433) , the partition function in equilibrium can be written as (17.435) We are now prepared to discuss the flow properties of an electric current of the system carried by the Cooper pairs. It is carried by the divergenceless pair current [compare (1.102)] 1 0 at one place z, with an appreciable measure in the functional integral (17.435), it must start from a nontrivial solution of the equations of motion which carries p(z) as closely as possible to zero. From our experience with the mechanical

17.11 Decay of Supercurrent in Thin Closed Wire

1243

Figure 17.24 Order parameter ~(z) = p(z)ei-y(z) of superconducting thin circular wire neglecting fluctuations. The order parameter is pictured as a spiral of radius Po and pitch 8"'((z)/8z = 21fn/ L winding around the wire. At T = 0, the supercurrent is absolutely stable since the winding number n is fixed topologically.

motion of a mass point in a potential such as - V(p) of Eq. (17.441), it is easily realized that there exists such a solution. It carries p(z) from Po = viI - k 2 at z = -00 across the potential barrier to the small value PI = v'2k and back once more across the barrier to Po at z = 00 (see Fig. 17.25). Using the first integral of motion of the differential equation (17.440), the law of energy conservation 1 2 1 1 1 -p - -V(p) = E = --V(Po) = -Po(Po + 2pI) 2 z 2 2 4

(17.448)

leads to the equation

pz =

J2E + V(p).

(17.449)

z. V VIp)

I

I

I I I P, = --12 k 2 I I I

P

I I I I IP, =~ I I I

Figure 17.25 Extremal excursion of order parameter in superconducting wire. It corresponds to a mass point starting out at Po, rolling under the influence of "negative gravity" up the mountain unto the point PI = ,j2k, and returning back to Po , with the variable z playing the role of a time variable.

1244

17 Tunneling

This is solved by the integral

(17.450) yielding (17.451) Inverting this, we find the bubble solution 2 ( ) Z

Pel

w 2 /2

2

= 1 - k - cosh2[w(Z

-

Zl

(17.452)

)/2] '

where (17.453) is the curvature of V(p) close to Po , i.e., (17.454) The extra energy of the bubble solution is (17.455) The explicit solution (17.452) reaches the point of smallest p at value is

Zl,

where its

(17.456) This value is still nonzero and does not yet permit a phase slip. However, we shall now demonstrate that quadratic fluctuations around the solution (17.456) do, in fact, to reduce the current. For this, we insert the fluctuating order field

p(Z) = Pel(z)

+ Jp(z)

(17.457)

into the free energy. With Pel being extremal, the lowest variation of E is of second order in Jp(z) (17.458)

17.11 Decay of Supercurrent in Thin Closed Wire

1245

V"(p)

w'

Figure 17.26 Infinitesimal translation of critical bubble yields antisymmetric wave function of zero energy P~l solving differential equation (17.509). Since this wave function has a node, there must be a negative-energy bound state.

This expression is not positive definite as can be verified by studying the eigenvalue problem

The potential V"(Pcl(Z)) has asymptotically the value w 2. When approaching z = ZI from the right, it develops a minimum at a negative value (see Fig. 17.26). After that it goes again against w 2 • The energy eigenvalues Ao and A- I lie as indicated in the figure. The fact that there is precisely one negative eigenvalue A- I can be proved without an explicit solution by the same physical argument that was used to show the instability of the fluctuation problem (17.253): A small temporal translation of the classical solution corresponds to a wave function which has no energy and a zero implying the existence of precisely one lower wave function with A- I < 0 and no zero. The negative eigenvalue makes the critical bubble solution unstable against contraction or expansion. The former makes the fluctuation return to the spiral classical solution (17.452) of Fig. 17.24, the second removes one unit from the winding number of the spiral and reduces the supercurrent. For the precise calculation of the decay rate, the reader is referred to the references quoted at the end of the chapter. Here we only give the final result which is [18] rate = const x Lw(k)e- Ecl/kBT, with the k-dependent prefactor

(17.460)

1246

17 Tunneling

,

.......

\

-1

-2

-3

log R/RN

-4

-5

1.5

1.0

0.5

o

T - Tc (lO-3K)

Figure 17.27 Logarithmic plot of resistance of thin superconducting wire as function of temperature at current O.2ilA in comparison with experimental data (vertical axis is normalized by the Ohmic resistance Rn = O.5f! measured at T > Tc , see papers quoted at end of chapter). I (1-3k2)1/4 [3J2k (VI-3k 2)] w(k) = 21A_ll (1 _ k2)1/2 exp - VI _ 3k2arctan J2k '

(17.461)

where (17.462) is the negative eigenvalue of the fluctuations in the complex field 'lj;(z) [which is not directly related to A_I of Eq. (17.459) and requires a separate discussion of the initial path integral (17.435)]. This complicated-looking expression has a simple quite accurate approximation which had previously been deduced from a numerical evaluation of the fluctuation determinant [19]: (17.463) Both expressions vanish at the critical value k = kc = 1/ v'3. The resistance of a thin superconducting wire following from this calculation is compared with experimental data in Fig. 17.27.

17.12

Decay of Metastable Thermodynamic Phases

A generalization of this decay mechanism can be found in the first-order phase transitions of many-particle systems. These possess some order parameter with an

17.12 Decay of Metastable Thermodynamic Phases

1247

effective potential which has two minima corresponding to two different thermodynamic phases. Take, for instance, water near the boiling point. At the boiling temperature, the liquid and gas phases have the same energy. This situation corresponds to the symmetric potential. At a slightly higher temperature, the liquid phase is overheated and becomes metastable. The potential is now slightly asymmetric. The decay of the overheated phase proceeds by the formation of critical bubbles [3]. Their outside consists of the metastable water phase, their inside is filled with vapor lying close to the stable minimum of the potential. The radius of the critical bubble is determined by the equilibrium between the gain in volume energy and the cost in surface energy. If (J is the surface tension and E the difference in energy density, the energy of bubble solution depends on the radius as follows: (17.464) A plot of this energy in Fig. 17.28 looks just like that of the action A(~) in Fig. 17.8. Thus the role of the deformation parameter ~ is played here by the bubble radius R .

-------r~---L--~_4_R

Figure 17.28 Bubble energy as function of its radius R. At the critical bubble, the energy has a maximum. The fluctuations of the critical bubble must therefore have a negative eigenvalue. This negative eigenvalue mode accounts for the fact that the critical bubble is unstable against expansion and contraction. When expanding, the bubble transforms the entire liquid into the stable gas phase. When contracting, the bubble disappears and the liquid remains in the overheated phase. Only the first half of the fluctuations have to be counted when calculating the lifetime of the overheated phase. It is instructive to take a comparative look at the instability of a critical bubble to see how the different spatial dimensions modify the properties of the solution. We shall discuss first the case of three space dimensions. As in the case of superconductivity, the description of the liquid-vapor phase transition makes use of a space-dependent order parameter, the real order field rp(x). The two minima of the potential V (rp) describe the two phases of the system. The kinetic term X'2 in the path integral is now a field gradient term [(8x rp(X))2] which ensures finite correla-

1248

17 Tunneling

tions between neighboring field configurations. The Euclidean action controlling the fluctuations is therefore of the form (17.465) where V(tp) is the same potential as in Eq. (17.1) , but it is extended by the asymmetric energy (17.243). Within classical statistics, the thermal fluctuations are controlled by the path integral for the partition function (17.466) Here T is the temperature measured in multiples of the Boltzmann constant kB . The path integral J Vtp(x) is defined by cutting the three-dimensional space into small cubes of size f and performing one field integration at each point. The critical bubble extremizes the action. Assuming spherical symmetry, the bubble satisfies in D dimensions the classical Euler-Lagrange field equation

rf2 D-1d) ( - dr2 - - r - dr

tpel

,

+ V (tpel(r)) = O.

(17.467)

This differs from the equation (17.305) for the one-dimensional bubble solution by the extra gradient term -[(D-1)l r]ortpel(r). Such a term is an obstacle to an exact solution of the equation via the energy conservation law (17.306). The relevant qualitative properties of the solution can nevertheless be seen in a similar way as for the bubble solution. As in Fig. 17.11 we plot the reversed potential and imagine the solution tp(r) to describe the motion of a mass point in this potential with -r playing the role of a "time" . Setting tpel(r) = x(-t), the field equation (17.467) takes the form

x (t) - D - 1 x(t) - V'(x(t)) t

=

O.

(17.468)

In this notation, the second term, i.e., the term -[(D - l) l r]ortpel(r) in (17.467) plays the role of a negative "friction" accelerating motion of the particle along x(t). This effect decreases with time like l i t. With our everyday experience of mechanical systems, the qualitative behavior of the solution can immediately be plotted qualitatively as shown in Fig. 17.29. For D = 1, the energy conservation makes the particle reach the right-hand zero of the potential. For D > 1, the "antifriction" makes the trajectory overshoot. At r = 0, the solution is closest to the stable minimum (the maximum of the reversed potential) on the left-hand side. In the superheated water system, this corresponds to the inside of the bubble being filled with vapor. As r moves outward in the bubble, the state moves closer to the metastable state, i.e., it becomes more and more liquid. The antifriction term has the effect that the point of departure on the left-hand side lies energetically below the final value of the metastable state.

1249

17.12 Decay of Metastable Thermodynamic Phases r

---------'------+-1--

1>cl(r)

r=O

\

2

_'!!!:.-1>2 _ 2..1>' 2

4!

Figure 17.29 Qualitative behavior of critical bubble solution as function of its radius.

Consider now the fluctuations of such a critical bubble in D = 3 dimensions. Suppose that the field deviates from the solution of the field equation (17.467) by 8cp(x). The deviations satisfy the differential equation

- ~~ + j} + VII(cpcl(r))] 8cp(x) = [ -~ dr2 r dr r2

A8cp(x) ,

(17.469)

where j} is the differential operator of orbital angular momentum (in units 1i = 1). Taking advantage of rotational invariance, we expand 8cp(x) into eigenfunctions of angular momentum CPnlm , the spherical harmonics Yim(x): (17.470)

nlm The coefficients CPnlm satisfy the radial differential equation ~ 2 d [ - dr2 - ;: dr

l(l + 1) II ] + -r-2- + V (cpcl(r)) CPnlm(r) = AnICPnlm(r),

(17.471)

with (17.472)

1250

17 Tunneling

One set of solutions is easily found, namely those associated with the translational motion of the classical solution. Indeed, if we take the bubble at the origin, (17.473) to another place x

+ a,

we find, to lowest order in a,

+ aOx'Pcl(X) 'Pcl(r) + aXor'Pcl(r) . 'Pcl(X)

(17.474)

x

But is just the Cartesian way of writing the three components of the spherical harmonics Ylm(X). If we introduce the spherical components of a vector as follows

(17.475)

we see that (17.476) Thus, b'P(X) = aXor'P(x) must be a solution of Eq. (17.469) with zero eigenvalue A. This can easily be verified directly: The factor x causes i} to have the eigenvalue 2, and the accompanying radial derivative b'P(X) = Or'Pcl(r) is a solution of Eq. (17.471) for l = 1 and Anl = 0, as is seen by differentiating the Euler-Lagrange equation (17.467) with respect to r. Choosing the principal quantum number of these translational modes to be n = 1, we assign the three components of xOr'Pcl(r) to represent the eigenmodes 'Pl,l,m. As long as the bubble radius is large compared to the thickness of the wall, which is of the order l /w, the 1/r 2 -terms will be very small. There exists then an entire family of solutions 'Pllm(X) with all possible values of l which all have approximately the same radial wave function Or'Pcl(r). Their eigenvalues are found by a perturbation expansion. The perturbation consists in the centrifugal barrier but with the l = 1 barrier subtracted since it is already contained in the derivative Or'Pcl(r), i.e. , Vpert = [l(l

+ 1) -

2J / 2r2 .

(17.477)

The bound-state wave functions 'Pllm are normalizable and differ appreciably from zero only in the neighborhood of the bubble wall. To lowest approximation, the perturbation expansion produces therefore an energy A ~l(l+1)-2 nl ~

2

rc

'

(17.478)

1251

17.12 Decay of Metastable Thermodynamic Phases

where rc is the radius of the critical bubble. As a consequence, the lowest l eigenstate has a negative energy

= 0

(17.479) Physically, this single l = 0 -mode corresponds to an infinitesimal radial vibration of the bubble. As already explained above it is not astonishing that a radial vibration has a negative eigenvalue. The critical bubble lies at a maximum of the action. Expansion or contraction is energetically favorable . Since Yoo(x) is a constant, the wave function is proportional to (d/dr)'Pcl(r) itself without an angular factor. This is seen directly by performing an infinitesimal radial contraction (17.480) The variation rOr'Pcl(r) is almost zero except in the vicinity of the critical radius r c, so that rOr'Pcl(r) : : : : rcOr'Pcl(r) which is the above wave function. Being the ground state of the Schrodinger equation (17.469), it should be denoted by 'Pooo(r). Since it solves approximately the Schrodinger equation (17.471) with l = I , it also solves this equation approximately with l = 0 and the energy (17.479). Finally let us point out that in D > 1 dimensions, the value of the negative eigenvalue can be calculated very simply from a phenomenological consideration of the bubble action. Since the inside of the bubble is very close to the true ground state of the system whose energy density lies lower than that of the metastable one by E, the volume energy of a bubble of an arbitrary radius R is RD Ev = -SDDE,

(17.481)

where SDR D- 1 is the surface of the bubble and SDR D/ D its volume. The surface energy can be parametrized as (17.482) where u is a constant proportional to the surface tension. Adding the two terms and differentiating with respect to R, we obtain a critical bubble radius at (17.483) with a critical bubble energy SD

D- l

E C =J5Rc

u=

SD D SD DD(D_1)Rc E=J5(D-1)

l

uD ED - 1 '

(17.484)

The second derivative with respect to the radius R is, at the critical radius,

EI

d2 D-1 _ = -DEc -- ' dR 2 R-rc rc2

(17.485)

1252

17 Tunneling

Identifying the critical bubble energy Ec with the classical Euclidean action find the variation of the bubble action as - 1 J2AI::::; --1 ()2 JR D ADI -'

r;

2

Al we

(17.486)

We now express the dilational variation of the bubble radius in terms of the normal coordinate of (17.470) . The normalized wave function is obviously (17.487) But the expression under the square root is exactly D times the action of the critical bubble (17.488) To prove this we introduce a scale factor evaluate the action

into the solution of the bubble and

8

Ael (17.489) Since

Ael is extremal at 8 =

1, it has to satisfy

OAeII 08

_0 8=1 -

,

(17.490)

or (17.491) Hence (17.492) implying that 1(2

D-2)J dD X [Ortpcl(r)] 2 2D

1 D

DX

Jd

[Ortpcl(r)] 2 .

(17.493)

With (17.487) , the tpooo contribution to Jtp(x) reads Jtp(x) =

~oootpooo(r) = ~ooo v'~~I '

(17.494)

1253

17.12 Decay of Metastable Thermodynamic Phases

and we arrive at (17.495) Inserting this into (17.486) shows that the second variation of the Euclidean action 82Acl can be written in terms of the normal coordinates associated with the normalized fluctuation wave function 'Pooo as (17.496) From this relation, we read off the negative eigenvalue D-1

(17.497)

AOO=---' 2r~

For D = 3, this is in agreement with the D = 3 value (17.479). For general D, the eigenvalue corresponding to (17.479) would have been derived with the arguments employed there from the derivative term -[(D - l)/r]d/dr in the Lagrangian and would also have resulted in (17.497). All other multipole modes 'Pnlm have a positive energy. Close to the bubble wall (as compared with the radius), the classical solutions (l/r)'Pnlm(r) can be taken approximately from the solvable one-dimensional equation

The wave functions with n = 0 are (17.499) and have the eigenvalues A ~l(l+1)-2 01 ~ 2 2 . rc

(17.500)

The n = 1 -bound states are

'Pllm>::;



w sinh[w(r - r c )/2] -4 cosh 2[w (r - r c )/ 2]'

(17.501)

with eigenvalues \

All>::;

3

2

SW +

l(l

+ 2) 2r~

2

.

(17.502)

1254

17 Tunneling

Xi

Figure 17.30 Decay of metastable false vacuum in Minkowski space. It proceeds as a shock wave which after some time traverses the world almost with light velocity, converting the false into the true vacuum.

17.13

Decay of Metastable Vacuum State in Quantum Field Theory

The theory of decay presented in the last section has an interesting quantum fieldtheoretic application. Consider a metastable scalar field system in aD-dimensional Euclidean spacetime at temperature zero. At a fixed time, there will be a certain average number of bubbles, regulated by the "quantum Boltzmann factor" exp( -Actin). If the bubble gas is sufficiently dilute (i.e., if the distances between bubbles is much larger than the radii) , each bubble is described quite accurately by the classical solution. In Minkowski space, a Euclidean radius r = v'X2 + c2r2 corresponds to r = v'X2 - c2 t 2 , where c is the light velocity. The critical bubble has therefore the spacetime behavior (17.503) From the above discussion in Euclidean space we know that cp will be equal to the metastable false vacuum in the outer region r > r e , i.e., for (17.504) The inside region (17.505) contains the true vacuum state with the lower energy. Thus a critical bubble in spacetime has the hyperbolic structure drawn in Fig. 17.30. Therefore, the Euclidean critical bubble describes in Minkowski space the growth of a bubble as a function of time. The bubble starts life at some time t = rei c and expands almost instantly to a radius of order re. The position of the shock wave is described by (17.506)

17.14 Crossover from Quantum Tunneling to Thermally Driven Decay

1255

This implies that a shock wave that runs through space with a velocity

Ixi

VI -

c

v--------;===== - t r~/c2t2

(17.507)

and converts the metastable into the stable vacuum - a global catastrophe. A Euclidean bubble centered at another place Xb , Tb would correspond to the same process starting at Xb and a time (17.508) A finite time after the creation of a bubble, of the order rc/c, the velocity of the shock wave approaches the speed of light (in many-body systems the speed of sound). Thus, we would hardly be able to see precursors of such a catastrophe warning us ahead of time. We would be annihilated with the present universe before we could even notice.

17.14

Crossover from Quantum Tunneling to Thermally Driven Decay

For completeness, we discuss here the difference between a decay caused by a quantum-mechanical tunneling process at T = 0, and a pure thermally driven decay at large temperatures. Consider a one-dimensional system possessing, at some place X" a high potential barrier, much higher than the thermal energy kBT, with a shape similar to Fig. 17.10. Let the well to the left of the barrier be filled with a grand-canonical ensemble of noninteracting particles of mass M in a nearly perfect equilibrium. Their distribution of momenta and positions in phase space is governed by the Boltzmann factor e-,6[p2/2M+V(x)l . The rate, at which the particles escape across the barrier, is given by the classical statistical integral rcl

where

Zcl

=

Zci 1 J dx J 2;n e -,6[p2/2M +V(X)]6(x - x.) ~e(p) ,

(17.509)

is the classical partition function

Z = JdxJ dp

27rn

cl

e - ,6[p2/2M+V(x)l.

(17.510)

The step function e (p) selects the particles running to the right across the top of the potential barrier. Performing the phase space path integral in (17.509) yields Z- l r cl -- _c_l_e - V(x.) 27rnj3 .

(17.511)

If the metastable minimum of the potential is smooth, V(x) can be replaced ap-

proximately in the neighborhood of Xo by the harmonic expression

V(x)

~

M

2

2

2wo(x - xo) .

(17.512)

1256

17 Tunneling

The classical partition function is then given approximately by 1

Zcl

~ li{3wo'

(17.513)

and the decay rate follows the simple formula

r

cl

~ Wo e - t3V(x.) ~

27r

.

(17.514)

Let us compare this result with the decay rate due to pure quantum tunneling. In the limit of small temperatures, the decay proceeds from the ground state, and the partition function is approximately equal to (17.515) The decay rate is given by the small imaginary part of the partition function:

r---+~lmZ T-->O li{3 Re Z·

(17.516)

In contrast to this, the thermal rate formula (17.511) implies for the hightemperature regime, where r becomes equal to r cJ, the relation: r ---+ w. 1m Zcl . T-->oo 7r Re Zcl

(17.517)

The frequency w. is determined by the curvature of the potential at the top of the barrier, where it behaves like (17.518) The relation (17.517) follows immediately by calculating in the integral (17.510) the contribution of the neighborhood of the top of the barrier in the saddle point approximation. As in the integral, this is done (17.261) by rotating the contour of integration which starts at x = x. into the upper complex half-plane. Writing x = x. + iy this leads to the following integral: 1m Zcl ~

roo

Jo

dy e - t3 [V(x.l+¥W;y2 ] ~ _l_ e - t3v(x.). J27rli 2 {3/M 2li{3w.

(17.519)

Since the real part is given by (17.513) we find the ratio 1m Zcl ~ Wo - t3V(x.) ReZcl 2w. '

--~- e

(17.520)

so that (17.511) is equivalent to (17.517). The two formulas (17.516) and (17.517) are derived for the two extreme regimes T » To and T « To, respectively, where To denotes the characteristic temperature

1257

Appendix 17A Feynman Integrals for Fluctuation Correction

associated with the curvature of the potential at the metastable minimum To

tiwo/kB. Numerical studies have shown that the applicability extends into the close neighborhood of To on each side of the temperature axis. The crossover regime is quite small, of the order O(ti3 / 2 ) . Note that given the knowledge of the imaginary parts of all excited states, which can be obtained as in Section 17.9, it will be possible to calculate the average lifetime of a metastable state at all temperatures without the restrictions of the semiclassical approximation. This remains to be done.

Appendix 17A

Feynman Integrals for Fluctuation Correction

For the integral (17.219) we obtain with (17.237) immediately the result stated in Eq. (17.240)

I'1 = ~ 560'

(17A.1)

To calculate the remaining three double integrals hi,I22,h in (17.223) and (17.224) we observe that because of the symmetry of the Green function G ow (T, T') in T, T' the measure of integration can be rewritten as 2 dT J~= dT'. We further introduce the dimensionless classical functions

J::'=

(17A.2) use natural units with w = 1, 9 = 1, since these quantities cancel in all integrals, and define

Xcl(T)GOw(T,T)

[1.~ GOw(T,T')XG(T')dTl

[1.~ Gbw(T, T')XG (T')dt'] A '

(17A.3)

where the subscript A denotes the antisymmetric part in T. Because of the antisymmetry of the symmetric part gives no contribution to the integrals which can be written as

hi h2 h

61= 91= 241=

Xc],

dTXcl(T)XK3G(T) ,

(17A.4)

dTXG(T)XKG(T),

(17A.5)

dTyh(T)XKG(T).

(17A.6)

When evaluating the integrals in (17A.3) the antisymmetry of YO(T) is useful. We easily find _ 112T+tanh(T/2) 2 4 12cosh (T /2)

XKG (T) - Inserting this into h2 and

h we encounter integrals of two types

(17A.7)

1258

17 Tunneling

The former can be performed with the help of formula (17.54) , the latter require integrations by parts of the type

1= o

f( T/ 2) tanh ~ dT = 2

-1= f' 0

(T /2) In(2 cosh ~) dT,

(17A.8)

2

which lead to a finite sum of integrals of the first type plus integrals of the typel l

1=

logcosh(T/2) sinhm(T /2) / coshn(T/2) .

-ov Jo=

dT sinhm (T/2) / coshv+ 1 (T/2) so that it is evaluated again via These, in turn, are equal to (17.54). After performing the subtraction in 122 we obtain the values given in Eq. (17.240) . The evaluation of the integral 121 is more tedious since we must integrate over the third power of the Green function , which is itself a lengthy function of T . It is once more useful to exploit the symmetry properties of the integrand. We introduce the abbreviations

~ [H~(T) ± H~(-T)]Y3-n (T) ,

1"' f~/ (T')Xcl (T')dT' ,

(17A.9)

A

and find XK3G in the form

XK3G(T)

=

f3'\(T)[No - Fg(T)] + 3f:i"-(T) [N 1 - Fr(T)] + 3fi'(T)[N2 +3fi(T)Fi'(T) + 3fr(T)F:i"-(T) + fg(T)Ft (T) ,

-

Fi(T)] (17A.1O)

with

3 No = 32 '

N _ 203 _ log2 2 - 512 2 '

(17A.ll)

Explicitly:

sech7~ 5T) -- [ 3T(58 cosh -T - 27 cosh -3T - 3 cosh3 . 29 2 2 2 (753 sinh ~

+ 36 cosh ~ - 108 cosh

+ 48 sinh 32T + 22 sinh 52T + sinh 72T)

In(2 cosh ~) (6T

~ (I

T

dT' T' tanh

+ 8 sinh T + sinh2T)

~) ] .

(17A.12)

The integrals to be performed are of the same types as before, except for the one involving the last term which requires one further partial integration:

~1= dT d~ [sech6~]lT dT'T'tanh~ 6T T 16 -:311= 0 dT sech "2 T tanh"2 = - 135 . (17 A.13) IlLS. Gradshteyn and I.M. Ryzhik, op. cit., Formulas 2.417.

1259

Notes and References

After the necessary subtraction of the divergent term we find the value of hI given in Eq. (17.240). The final result for the first coefficient of the Taylor expansion of the subtracted fluctuation factor 0' in Eq. (17.222) is therefore 71

CI

= 24 "" 2.958.

(17A.14)

This number was calculated in Ref. [12] by solving the Schriidinger equation [20] . With the help of the WKB approximation he derived a recursion relation for the higher coefficients Ck of the expansion of 0': (17A.15) From this he calculated the next nine coefficients C2

=

315 9 84376 32 "". ,

- 65953 ~ 57 2509 1152 ~. .

C3 -

(17A.16)

Their large behavior is Ck

9(3)k "2 k! [ln(6k) + ,].

~:;;:

(17A.17)

Notes and References The path integral theory of tunneling was first discussed by A.I. Vainshtein, Decaying Systems and the Divergence of Perturbation Series, Novosibirsk Report (1964), in Russian (unpublished) , and in the context of the nucleation of first-order phase transitions by J.S. Langer, Ann. Phys. 41 , 108 (1967). The subject was studied further in the field-theoretic literature: M.B. Voloshin, I.y. Kobzarev, L.B. Okun, Yad. Fiz. 20. 1229 (1974); Sov. J. Nuc!. Phys. 20, 644 (1975) ; R. Rajaraman, Phys. Rep. 21, 227 (1975), R. Rajaraman, Phys. Rep. 21, 227 (1975), P. Frampton, Phys. Rev. Lett. 37, 1378 (1976) and Phys. Rev. D 15, 2922 (1977), S. Coleman, Phys. Rev. D 15, 2929 (1977); also in The Whys of Subnuclear Physics , Erice Lectures 1977, Plenum Press, 1979, ed. by A. Zichichi, I. Affleck, Phys. Rev. Lett. 46, 388 (1981). For a finite-temperature discussion see 1. Dolan and J. Kiskies, Phys. Rev. D 20, 505-513 (1979) . For multidimensional tunneling processes see H. Kleinert and R. Kaul, J. Low Temp. Phys. 38,539 (1979) (http://www.physik.fu-berlin. derkleinert/66) , A. Auerbach and S. Kivelson, Nuc!. Phys. B 257, 799 (1985). The fact that tunneling calculations can be used to derive the growth behavior of large-order perturbation coefficients was first noticed by A.I. Vainshtein, Novosibirsk Preprint 1964, unpublished. A different but closely related way of deriving this behavior was proposed by L.N. Lipatov, JETP Lett. 24, 157 (1976); 25, 104 (1977); 44, 216 (1977); 45, 216 (1977).

1260

17 Tunneling

A review on this subject is given in J. Zinn-Justin, Phys. Rep. 49, 205 (1979), and in Recent Advances in Field Theory and Statistical Mechanics , Les Houches Lectures 1982, Elsevier Science 1984, ed. by J.-B. Zuber and R. Stora. J. Zinn-Justin, Quantum Field Theory and Critical Phenomena, Clarendon, Oxford, 1990. The important applications to the E-expansion of critical exponents in O(N)-symmetric rp4 field theories were made by E. Brezin, J.C. Le Guillou, and J. Zinn-Justin, Phys. Rev. D 15, 1544, 1558 (1977). E. Brezin and G. Parisi, J. Stat. Phys. 19, 269 (1978). The perturbation expansion of the rp4-theory up to fifth-order was calculated in Ref. [13]. For two quartic interactions of different symmetries (one O(N)-symmetric, the other of cubic symmetry) see: H. Kleinert and V. Schulte-Frohlinde, Phys. Lett. B 342, 284 (1995) (cond-mat/9503038) A detailed discussion and a comprehensive list of the references to the original papers are contained in the textbook [15]. The calculation of the anomalous magnetic moments in quantum electrodynamics is described in T. Kinoshita and W.B. Lindquist, Phys. Rev. D 27, 853 (1983). M.J. Levine and R. Roskies, in Proceedings of the Second International Conference on Precision Measurements and Fundamental Constants, ed. by B.N. Taylor and W .D. Phillips, Nat!. Bur. Std. US, Spec. Pub!. 617 (1981). The analytic result for the semiclassical decay rate of a supercurrent in a thin wire was found by I.H. Duru, H. Kleinert, and N. Unal, J. Low Temp. Phys. 42, 137 (1981) (ibid.http174), H. Kleinert and T. Sauer, J . Low Temp. Physics 81, 123 (1990) (ibid.http/204) . The experimental situation is explained in M. Tinkham, Introduction to Superconductivity , McGraw-Hill, New York, 1975. See, in particular, Chapter 7, Sections 7.1- 7.3. For quantum corrections to the decay rate see N. Giordano, Phys. Rev. Lett. 61 , 2137 (1988); N. Giordano and E .R. Schuler, Phys. Rev. Lett. 63,2417 (1989) . For thermally driven tunneling processes see the review article P. Hiinggi, P. Talkner, and M. Borkovec, Rev. Mod. Phys. 62,251 (1990). Tunneling processes with dissipation were first discussed at T = 0 by A.O. Caldeira and A.J. Leggett, Ann. Phys. 149, 374 (1983), 153, 445 (1973) (Erratum) and for T i- 0 by H. Grabert, U. Weiss and P. Hiinggi, Phys. Rev. Lett. 52, 2193 (1984), A.I. Larkin and Y.N. Ovchinnikov, Sov. Phys. JETP 59, 420 (1984). Papers on coherent tunneling: A.J. Leggett, S. Chakravarty, A.T. Dorsey, M.P.A. Fisher, A. Garg, and W. Zwerger, Rev. Mod. Phys. 59, 1 (1987), U. Weiss, H. Grabert, P. Hiinggi, and P. Riseborough, Phys. Rev. B 35, 9535 (1987) , H. Grabert, P. Olschowski, and U. Weiss, Phys. Rev. B 36, 1931 (1987), On the use of periodic orbits see: P. Hiinggi and W . Hontscha, Ber. Bunsen-Ges. Phys. Chemie 95,379 (1991) . The individual citations refer to

Notes and References

1261

[1] See also the solutions in textbooks such as L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965, §25, PROBLEM 4. [2] L.D. Faddeev and V.N. Popov, Phys. Lett. B 25, 29 (1967) . [3] J.S. Langer, Ann. Phys. 41, 108 (1967) . [4] For a general proof see S. Coleman, Nuc!. Phys. B 298, 178 (1988) . [5] F .J. Dyson, Phys. Rev. 85, 631 (1952). [6] Compare the result in Eq. (17.345) with J.C. Collins and D.E. Soper, Ann. Phys. 112, 209 (1978). [7] The variational approach to tunneling was initiated in H. Kleinert , Phys. Lett. B 300, 261 (1993). It has led to very precise tunneling rates in subsequent work by R. Karrlein and H. Kleinert, Phys. Lett. A 187, 133 (1994) ; H. Kleinert and I. Mustapic, Int. J. Mod. Phys. A 11, 4383 (1995) , most precisely in Ref. [11]. [8] H. Kleinert, Phys. Lett. B 300, 261 (1993) (ibid.http/214) . [9] R. Karrlein and H. Kleinert, Phys. Lett. A 187, 133 (1994) (hep-th/ 9504048). [10] C.M. Bender and T.T. Wu, Phys. Rev. D 7, 1620 (1973), Eq. (5.22). [11] B. Hamprecht and H. Kleinert, Tunneling Amplitudes by Perturbation Theory, Phys. Lett. B 564, 111 (2003) (hep-th/ 0302124) . [12] J. Zinn-Justin, J. Math. Phys. 22, 511 (1981); Table III. [13] H. Kleinert, J . Neu, V. Schulte-Frohlinde, K.G. Chetyrkin, and S.A. Larin, Phys. Lett. B 272, 39 (1991) (hep-th/ 9503230) . [14] H. Kleinert , Phys. Rev. D 57 , 2264 (1998); Addendum: Phys. Rev. D 58 , 107702 (1998) (cond-mat / 9803268). [15] For details and applications to see the textbook H. Kleinert and V. Schulte-Frohlinde, Critical Phenomena in ¢4-Theory, World Scientific, Singapore, 2001 (ibid.http/b8) . [16] The path integral treatment of the decay rate of a supercurrent in a thin wire was initiated by J.S. Langer and V. Ambegaokar, Phys. Rev. 164, 498 (1967) , D.E. McCumber and B.1. Halperin, Phys. Rev. B 1, 1054 (1970) . [17] M. Tinkham, Introduction to Superconductivity , McGraw-Hill, New York, 1975. [18] H. Kleinert and T. Sauer, J . Low Temp. Physics 81 , 123 (1990) (ibid.http/204) . [19] D.E. McCumber and B.1. Halperin Phys. Rev. B 1, 1054 (1970). [20] Divide the numbers in the Table IV of Ref. [12] by 6n4 n to get our Cn .

Path, motive, guide, original, and end. (1709-1784) , The Rambler

SAMUEL JOHNSON

18 Nonequilibrium Quantum Statistics

Quantum statistics described by the theoretical tools of the previous chapters is quite limited. The physical system under consideration must be in thermodynamic equilibrium, with a constant temperature enforced by a thermal reservoir. In this situation, partition function and the density matrix can be calculated from an analytic continuation of quantum-mechanical time evolution amplitudes to an imaginary time tb - ta = -in/kBT. In this chapter we want to go beyond such equilibrium physics and extend the path integral formalism to nonequilibrium time-dependent phenomena. The tunneling processes discussed in Chapter 17 belong really to this class of phenomena, and their full understanding requires the theoretical framework of this chapter. In the earlier treatment this was circumvented by addressing only certain quasi-equilibrium questions. These were answered by applying the equilibrium formalism to the quantum system at positive coupling constant, which guaranteed perfect equilibrium, and extending the results to the quasi-equilibrium situation analytic continuation to small negative coupling constants. Before we can set up a path integral formulation capable of dealing with true nonequilibrium phenomena, some preparatory work is useful based on the traditional tools of operator quantum mechanics.

18.1

Linear Response and Time-Dependent Green Functions for T f= 0

If the deviations of a quantum system from thermal equilibrium are small, the easiest description of nonequilibrium phenomena proceeds via the theory of linear response. In operator quantum mechanics, this theory is introduced as follows. First, the system is assumed to have a time-independent Hamiltonian operator fl. The ground state is determined by the Schrodinger equation, evolving as a function of time according to the equation

(18.1) (in natural units with n = 1, kB = 1). The subscript S denotes the Schrodinger picture.

1262

18.1 Linear Response and Time-Dependent Green Functions forT

=I- 0

1263

Next, the system is slightly disturbed by adding to iI a time-dependent external interaction, (18.2)

where iIext(t) is assumed to set in at some time to, i.e., iIext(t) vanishes identically for t < to. The disturbed Schrodinger ground state has the time dependence (18.3)

where (;H (t) is the time translation operator in the Heisenberg picture. It satisfies the equation of motion (18.4)

iI'jr(t) == eiiltiIext(t)e-ifIt.

(18.5)

To lowest-order perturbation theory, the operator (;H(t) is given by

UH(t) = 1 - i A

it

dt' H'Fr(t') A

+ ....

(18.6)

to

In the sequel, we shall assume the onset of the disturbance to lie at to = -00 . Consider an arbitrary time-independent Schrodinger observable 0 whose Heisenberg representation has the time dependence

OH(t) = eiiltOe-iilt.

(18.7)

Its time-dependent expectation value in the disturbed state 11IT~ist(t)) is given by

(1IT~ist( t) 1011lT~ist( t)) = (1IT s(O) I(;k( t)eiiltOe- iilt(;H( t) 11IT s(O))

~ (1ITs(0) I (1 + i [co dt' iI'jr(t') + ...) OH(t) x =

(1- i [co dt' iI'j;t(t') + ...) 11ITs(0))

(1IT HIOH(t) 11IT H) - i(1IT HI [co dt' [OH(t), iIJ;t(t')] 11IT H)

(18.8)

+ ...

We have identified the time-independent Heisenberg state with the time-dependent Schrodinger state at zero time in the usual manner, i.e., 11IT H) == 11ITs(0) ). Thus the expectation value of 0 deviates from equilibrium by (1IT~ist (t) 10 (t) 11IT~ist (t)) - (1IT s (t) 10 (t) 11IT s (t) )

J (1IT s (t) 1011lT s (t) ) =

-i [co dt' (11TH I [OH(t) ,iI'Ft-xt(t')] 11ITH).

(18.9)

1 Note that after the replacements H -> Ho, H'j;t -> H}nt, Eq. (18.4) coincides with the equation for the time evolution operator in the interaction picture to appear in Section 18.7. In contrast to that section, however, the present interaction is a nonpermanent artifact to be set equal to zero at the end, and H is the complicated total Hamiltonian, not a simple free one. This is why we do not speak of an interaction picture here.

1264

18 Nonequilibrium Quantum Statistics

If the left-hand side is transformed into the Heisenberg picture, it becomes

so that Eq. (18.9) takes the form (18.10) It is useful to use the retarded Green function of the operators OH(t) and HH(t') in the state IWH) [compare (3.40)]:

(18.11 ) Then the deviation from equilibrium is given by the integral (18.12) Suppose now that the observable OH(t) is capable of undergoing oscillations. Then an external disturbance coupled to OH(t) will in general excite these oscillations. The simplest coupling is a linear one, with an interaction energy (18.13) where j(t) is some external source. Inserting (18.13) into (18.12) yields the linearresponse formula (18.14) where G~o is the retarded Green function of two operators

0: (18.15)

At frequencies where the Fourier transform of Goo(t, t') is singular, the slightest disturbance causes a large response. This is the well-known resonance phenomenon found in any oscillating system. Whenever the external frequency w hits an eigenfrequency, the Fourier transform of the Green function diverges. Usually, the eigenfrequencies of a complicated N-body system are determined by calculating (18.15) and by finding the singularities in w. It is easy to generalize this description to a thermal ensemble at a nonzero temperature. The principal modification consists in the replacement of the ground state expectation by the thermal average

18.2 Spectral Representations of Green Functions for T =J 0

1265

Using the free energy F

= - T log Tr (e- H/ T ),

this can also be written as (18.16)

In a grand-canonical ensemble, iI must be replaced by iI - fJN and F by its grandcanonical version Fa (see Section 1.17). At finite temperatures, the linear-response formula (18.14) becomes

J(O(t))r

= i

1:

dt' G~o(t, t')Jj(t'),

(18.17)

where G~o(t, t') is the retarded Green function at nonzero temperature defined by [recall (1.302)]

G~dt, t') == G~o(t - t') == 8(t - t') eF/TTr {e- H/ T [OH(t), OH(t')]}. (18 .18) In a realistic physical system, there are usually many observables, say Ok(t) for i = 1, 2, ... , l , which perform coupled oscillations. Then the relevant retarded Green

function is some l x l matrix (18.19) After a Fourier transformation and diagonalization, the singularities of this matrix render the important physical information on the resonance properties of the system. The retarded Green function at T =J 0 occupies an intermediate place between the real-time Green function of field theories at T = 0, and the imaginary-time Green function used before to describe thermal equilibria at T =J 0 (see Subsection 3.8.2). The Green function (18.19) depends both on the real time and on the temperature via an imaginary time.

18.2

Spectral Representations of Green Functions for T pO

The retarded Green functions are related to the imaginary-time Green functions of equilibrium physics by an analytic continuation. For two arbitrary operators 01, 01-, the latter is defined by the thermal average (18.20) where OH(r) is the imaginary-time Heisenberg operator (18.21)

1266

18 Nonequilibrium Quantum Statistics

To see the relation between G 12 (T) and the retarded Green function G~(t), we take a complete set of states In), insert them between the operators 01, 0 2 , and expand G 12 (T) for T 2: 0 into the spectral representation

G 12 (T) = eF / T L e- En /Te(En-Enf)T (nIO l ln')(n'10 2 In).

(18.22)

n,n'

Since G 12 (T) is periodic under T ---> T + liT, its Fourier representation contains only the discrete Matsubara frequencies Wm = 27fmT:

rl/T

Jo

eF / T

dTeiwmTG12(T)

L e- En /T (1 - e(En - Enf)/T) (nIO l ln')(n' 10 In) 2

n,n'

(18.23)

The retarded Green function satisfies no periodic (or antiperiodic) boundary condition. It possesses Fourier components with all real frequencies w:

i:

eF / T

dte iwt 8(t)e F/TTr {e- H/T [ok(t),01(O)]'f}

l)() dteiwt L [ e- En/Tei(En - Enf)t(nIOlln')(n'102In) o

n ,n'

=Fe- En /Te- i(En - Enf)t (nI0 2 In') (n'IOlln)] . (18.24) In the second sum we exchange nand n' and perform the integral, after having attached to W an infinitesimal positive-imaginary part iry to ensure convergence [recall the discussion after Eq. (3.84)]. The result is

Gf2(W) = eF/T L e- En /T

[1- e(En-Enf)/T] (n IO l ln')(n' 10 In) 2

n,n'

i

x W -

Enf

+ En + iry

.

(18.25)

By comparing this with (18.23) we see that the thermal Green functions are obtained from the retarded ones by replacing [1] -1

---------->------

(18.26)

A similar procedure holds for fermion operators Oi (which are not observable). There are only two changes with respect to the boson case. First, in the Fourier expansion of the imaginary-time Green functions, the bosonic Matsubara frequencies Wm in (18.23) become fermionic. Second, in the definition of the retarded Green

18.2 Spectral Representations of Green Functions for T =J 0

1267

functions (18.19), the commutator is replaced by an anticommutator, i.e., the retarded Green function of fermion operators Ok is defined by

G~(t, t') == G~(t - t') == 8(t - t')eF/TTr {e- fI / T [Ok(t), OiH(t')] + }

. (18.27)

These changes produce an opposite sign in front of the e(En - Enl)/T_term in both of the formulas (18.23) and (18.25). Apart from that, the relation between the two Green functions is again given by the replacement rule (18.26). At this point it is customary to introduce the spectral function

P12(W') =

(1 =f e-

Wl

/

T) eF/T

x Ln,n' e- En /T27r8(w - En'

+ E n )(nI0 1 In')(n'I0 2 In),

(18.28)

where the upper and the lower sign hold for bosons and fermions, respectively. Under an interchange of the two operators it behaves like (18.29) Using this spectral function, we may rewrite the Fourier-transformed retarded and thermal Green functions as the following spectral integrals:

1

00

G~(w)

- 00

dw' -2 P12(W') 7r

i W -

W

, +.

ZTJ

,

(18.30) (18.31)

These equations show how the imaginary-time Green functions arise from the retarded Green functions by a simple analytic continuation in the complex frequency plane to the discrete Matsubara frequencies, w --+ iwm . The inverse problem of reconstructing the retarded Green functions in the entire upper half-plane of w from the imaginary-time Green functions defined only at the Matsubara frequencies Wm is not solvable in general but only if other information is available [2]. For instance, the sum rules for canonical fields to be derived later in Eq. (18.66) with the ensuing asymptotic condition (18.67) are sufficient to make the continuation unique [3]. Going back to the time variables t and r, the Green functions are G~(t)

8(t)

1 dw' P12(w')e00

-2

- 00

iw't ,

(18.32)

7r

(18.33) The sum over even or odd Matsubara frequencies on the right-hand side of G12 (r) was evaluated in Section 3.3 for bosons and fermions as GP (r) = e - w(T- l/2T) 1 TLe - iwmT ..,-._-1__ w,e 2sin(w/2T) n zwm-w e- WT (1 + nw) (18.34)

1268

18 Nonequilibrium Quantum Statistics

and

TLe - iwmT -,-._-1__ n

Ga (7)

=

e- w(T- l/2T)

w,e

ZWm-W

1

2cos(w/2T)

e- WT (1 - n w ),

(18.35)

with the Bose and Fermi distribution functions [see (3.93), (7.529), (7.531)] 1

(18.36)

respectively.

18.3

Other Important Green Functions

In studying the dynamics of systems at finite temperature, several other Green functions are useful whose spectral functions we shall now derive. In complete analogy with the retarded Green functions for bosonic and fermionic operators, we may introduce their counterparts, the so-called advanced Green functions (compare page 38)

GMt, t') == Gt2(t - t') = -8(t' - t)eF/TTr {e- iI /T [a1(t) , a~(t')]'f}

. (18.37)

Their Fourier transforms have the spectral representation

Gt2(W)

1

00

=

-00

dw' -2 P12(W') 7r

i W -

,. ,

W -

zry

(18.38)

differing from the retarded case (18.30) only by the sign of the iry-term. This makes the Fourier transforms vanish for t > 0, so that the time-dependent Green function has the spectral representation [compare (18.32)] (18.39) By subtracting retarded and advanced Green functions , we obtain the thermal expectation value of commutator or anticommutator:

Note the simple relations: G~(t, t')

Gt2(t, t')

8(t - t')C12 (t, t'), -8(t' - t)Cdt, t').

(18.41) (18.42)

1269

18.3 Other Important Green Functions

When inserting the spectral representations (18.30) and (18.39) of Gf2(t) and Gf2(t) into (18.40), and using the identity (1.324),

w - w' + iTJ

- - -,- .- = w - w - ZTJ

TJ)2 W - w'

2(

+ TJ 2 = 2·m5(w -

w'),

(18.43)

we obtain the spectral integral representation for the commutator function: 2

C12 () t =

1

00 - 00

dw () -iwt 27r P12 we .

(18.44)

Thus a knowledge of the commutator function C 12 (t) determines directly the spectral function P12(W) by its Fourier components (18.45)

An important role in studying the dynamics of a system in a thermal environment is played by the time-ordered Green functions. They are defined by (18.46)

Inserting intermediate states as in (18.23) we find the spectral representation

G12(w)

+

L: L:

dt eiwt 8(t) eF/TTr

dte iwt 8( -t)eF/TTr {e-if/T6~(t)61-(0)}

1 dt eiwt L e- En /Tei(En-En,)t eF/T 1° dt eiwt L e- En / Te- i(En - En,)t(nI62In')(n'161In). e F/T

00

o

±

{e-H'jT61-(t)6~(0) }

- 00

(nI6 1 In') (n'16 2 In)

n,n'

(18.47)

n,n'

Interchanging again nand n', this can be written in terms of the spectral function (18.28) as

G 12 (w) =

I i ]. 1 -P12 W' [1e- ,/T w - Wii + iTJ + , e /T w - Wi - iTJ ( 00

- 00

dw' 27r

( ) 1 =t=

W

1 =t=

W

18.48

)

Let us also write down the spectral decomposition of a further operator expression complementary to C 12 (t) of (18.40), in which boson or fermion fields appear with the "wrong" commutator: (18.49) 2Due to the relation (18.41), the same representation is found by dropping the factor 8(t) in (18.32).

1270

18 Nonequilibrium Quantum Statistics

This function characterizes the size of fluctuations of the operators Inserting intermediate states, we find

L:

eF/ T

dteiwteF/TTr {e- H/ T

1

dt eiwt L

00

[

01

and

Ok.

[Ok(t),01-(O)t}

e- En /T ei(En - En,)t (nI01In/) (n/I02In)

n ,n'

-00

±e- En / Te- i(En - En, )t(nI02In/) (n/I01In)] . (18.50) In the second sum we exchange nand n' and perform the integral, which runs now over the entire time interval and gives therefore a J-function: A 12 (W)

= eF/T L e- En /T [1 ± e(En-En,)/T] (nI01In/)(n/I02In) n,n'

(18.51) In terms of the spectral function (18.28), this has the simple form

A 12 () W =

1 dw' 00 - 00

w' (') W () 27r tanh Of1 2TP12 w 27ru5:(W - W') = tanhOf 1 2TP12 w .

(

18.52 )

Thus the expectation value (18.49) of the "wrong" commutator has the time dependence

A 12 (t, t') -= A 12 (t - t') =

100

- 00

dw Pl2 () W e-iw(t-t') . w tanh Ofl -T 27r 2

(18.53)

There exists another way of writing the spectral representation of the various Green functions. For retarded and advanced Green functions G~, Gt2' we decompose in the spectral representations (18.30) and (18.38) according to the rule (1.325): i . 5:(W ----.= 2. [P - - =t= 27ru

w' ±

W -

27]

W -

w'

-

W

I)] ,

(18.54)

where P indicates principal value integration across the singularity, and write

G~A(W) =

i

1 dw 00

-00

/ 27rI P12(W )

[~=t= i7rJ(w W - W

WI) ] .

(18.55)

Inserting (18.54) into (18.48) we find the alternative representation of the timeordered Green function

Gl2 (W) = i

L: ~~

pdw /) [w ~ w' - i7r tanh Ofl 2';J(w - WI) ].

(18.56)

The term proportional to J(w - w') in the spectral representation is commonly referred to as the absorptive or dissipative part of the Green function. The first term proportional to the principal value is called the dispersive or fluctuation part .

1271

18.4 Hermitian Adjoint Operators

The relevance of the spectral function P12(W') in determining both the fluctuation part as well as the dissipative part of the time-ordered Green function is the content of the important fluctuation-dissipation theorem. In more detail, this may be restated as follows: The common spectral function P12(W') of the commutator function in (18.44) , the retarded Green function in (18.30) , and the fluctuation part of the time-ordered Green function in (18.56) determines, after being multiplied by a factor tanh'!'1(w'/2T) , the dissipative part of the time-ordered Green function in Eq. (18.56). The three Green functions -iG 12 (w) , -iG~(w) , and -iG~2(W) have the same real parts. By comparing Eqs. (18.30) and (18.31) we found that retarded and advanced Green functions are simply related to the imaginary-time Green function via an analytic continuation. The spectral decomposition (18.56) shows this is not true for the time-ordered Green function, due to the extra factor tanh,!,1(w/2T) in the absorptive term. Another representation of the time-ordered Green is useful. It is obtained by expressing tan,!,l in terms of the Bose and Fermi distribution functions (18.36) as tan,!,l = 1 ± 2nw. Then we can decompose

100

- 00

18.4

dw' P12(W') 271"

[i,+. W -

2T)

W

]

± 271"nw8(w - w') .

(18.57)

Hermitian Adjoint Operators

If the two operators Gk(t), G'iI(t) are Hermitian adjoint to each other,

G1(t) = [G1(t)jf ,

(18.58)

the spectral function (18.28) can be rewritten as

pdw') = (1 =r= e-w'/T)e F / T x 2: e- En / T271"8(w' - En'

+ En)l(nIG1(t)ln')112.

(18.59)

n ,n'

This shows that for bosons, (18.60)

pdw')

2: 0

for fermions.

This property permits us to derive several useful inequalities between various diagonal Green functions in Appendix 18A. Under the condition (18A.7), the expectation values of anticommutators and commutators satisfy the time-reversal relations

Gf2(t, t') A 12 (t , t') Cdt,t') Gdt, t')

=r=G~(t', t)*, ±A21 (t' , t)* , =r=C21 (t', t)*. ±G21 (t', t)* .

(18.61) (18.62) (18.63) (18.64)

1272

18 Nonequilibrium Quantum Statistics

Examples are the corresponding functions for creation and annihilation operators which will be treated in detail below. More generally, this properties hold for any interacting nonrelativistic particle fields Gk(t) = (/;p(t) , G~(t) = (/;b(t) of a specific momentum p. Such operators satisfy, in addition, the canonical equal-time commutation rules at each momentum (18.65) (see Sections 7.6, 7.9). Using (18.40), (18.44) we derive from this spectral function sum rule:

1

00

-00

dw' -P12(W /) = l. 27r

(18.66)

For a canonical free field with P12 (w') = 27rIn + 1. The operator 0 1 = a does the opposite. Hence we have 00 ,0

pdw') = 2m5(w' - 0)(1 T

L

e-fl/T)e F / T

e-(n±1/2)fl/T(n

+ 1).

(18.71)

n=O

Now we make use of the explicit partition functions of the oscillator whose paths satisfy periodic and antiperiodic boundary conditions: Z =

- F/T _

~

- ~e

fl - e

[2 sinh(0/2TW 1 2 cosh(0/2T)

- (n±l/2)fl/T _ {

-

boso~s

for

fermlOns

} . (18.72)

These allow us to calculate the sums in (18.71) as follows

E

00 e- (n+l/2)fl/T(n

1)

a +"2 + 1) = ( -T ao

e- F / T

Lo e-(n-1/2)fl/T(n + 1) = efl / 2T = (1 + e- fl / T ) -

= (1 T e- fl/Tf 1 e- F / T , 1

e- F / T .

(18.73)

n=O

The spectral function P12(w') of the a single oscillator quantum of frequency 0 is therefore given by P12(W') =

27f8(w' - 0).

(18.74)

With it, the retarded and imaginary-time Green functions become G~(t , t')

8(t - t')e-fl(t-t'),

(18.75)

G fl (7, 7')

(18.76)

c

lor

7

2, < 7,

(18.77)

with the average particle number nfl of (18.36) . The commutation function, for instance, is by (18.44) and (18.74): C 12 (t , t')

= e-ifl(t-t') ,

(18.78)

and the correlation function of the "wrong commutator" is from (18 .53) and (18.74):

Afl(t, t')

=

tanh,!,1

2~e-ifl(t-t').

(18.79)

Of course, these harmonic-oscillator expressions could have been obtained directly by starting from the defining operator equations. For example, the commutator function (18.80)

1274

18 Nonequilibrium Quantum Statistics

turn into (18.78) by using the commutation rule at different times

[aH(t), ak(t')] = e- iO(t- t') ,

(18.81)

which follows from (18.68). Since the right-hand side is a c-number, the thermodynamic average is trivial: (18.82) After this, the relations (18.41), (18.42) determine the retarded and advanced Green functions G~(t

- t') = 8(t - t')e- in(t- t') ,

G~(t

- t') = -8(t' - t)e- in(t- t').

(18 .83)

For the Green function at imaginary times

Gn(T,T') == eF/TTr [e-H/TrraH(T)ak(T')] ,

(18.84)

the expression (18.77) is found using [see (18.85)]

eHrate- Hr = ate nr , eHrae- Hr = ae- nr ,

(18.85)

and the summation formula (18.73) . The "wrong" commutator function (18.79) can, of course, be immediately derived from the definition (18.86) and (18.68), by inserting intermediate states. For the temporal behavior of the time-ordered Green function we find from (18.48) (18.87) and from this by a Fourier transformation

Gn(t, t') =

(1 =t= e- n/T)-1 8(t -

t')e-in(t-t), -

()-1 1 =t= en/T 8(t' -

t)e-in(t-t),

= [8(t - t') ± (e n/T =t= 1)- 1] e - in(t- t') = [8(t - t') ± nn] e - in(t- t').

(18.88)

The same result is easily obtained by directly evaluating the defining equation using (18.68) and inserting intermediate states:

Gn(t, t')

Gn(t - t') = eF/TTr [e- H/T raH(t)ak(t')]

8(t - t')(aat)e- iO(t- t') ± 8(t' - t)(a t a)e- in(t- t') 8(t - t')(l ± nn)e- iO(t- t') ± 8(t' - t)nne- in(t- t') ,

(18.89)

18.5 Harmonic Oscillator Green Functions for T =J 0

1275

which is the same as (18.88). For the correlation function with a and at interchanged, (18.90) we find in this way

Gn(t, t')

8(t - t')(at a,)e-in(t-tl) ± 8(t' - t) (a, a,t)e-ifl(t-t') 8(t - t')nne-in(t- tl) ± 8(t' - t)(1 ± nn)e- in(t- tl),

(18 .91)

in agreement with (18.64).

18.5.2

Real Field Operators

From the above expressions it is easy to construct the corresponding Green functions for the position operators of the harmonic oscillator x(t). It will be useful to keep the discussion more general by admitting oscillators which are not necessarily mass points in space but can be field variables. Thus we shall use, instead of x(t), the symbol t'

ti

1

G(t, t') = "2 [A(t , t')

+ C(t , t' )] = 2M0.

cosh ~ [ti,6 - i(t - t' )]

.

ti0.,6

(18.185)

smh2which coincides with the time-ordered Green function (18 .101) for t > t', and thus with the analytically continued periodic imaginary-time Green function (3.248) . The exponent in this generating functional is thus quite similar to the equilibrium source term (3.218). The generating functional (18.180) can, of course, be derived without the previous operator discussion completely in terms of path integrals for the harmonic oscillator in thermal equilibrium. Using the notation X(t) for a purely time-dependent oscillator field 0(x), we can take the generating functional directly from Eq. (3.168):

. JVX(t) {1i l dt [M2(X.2 - 0.2X2) + jX]} ta

e(i/Ii)Acl.i Fn,j (tb'

ta).

exp

=

tb

i

(XbtbIXata)b =

(18.186)

with a total classical action A:l,j

(18.187)

and the fluctuation factor (3.170), and express (18.187) as in (3.171) in terms of the two independent solutions Da(t) and Db(t) of the homogenous differential equations (3.48) introduced in Eqs. (2.226) and (2.227):

· 2· ] DM( ) [2 XbDa(tb)-XaDb(ta)-2XbXa

2 a tb

1() + -D a tb

l

tb

ta

dt [XbDa(t) +XaDb(t) ]j(t).

(18.188)

The fluctuation factor is taken as in (3.172). Then we calculate the thermal average of the forward- backward path integral of the oscillator X (t) via the Gaussian integral

Zob+,j-] =

JdXbdXa (Xbti,6IXa

O)n (XbtbI Xata)1t (XbtbIXata)k* · (18.189)

Here (Xb ti,6IXaO)n is the imaginary-time amplitude (2.409):

(Xb ti ,6I X aO) = .1 1

V 27rti/M

x exp { -

~ . h sm t

(.I

/LJJ

2~ Sin~~,60. [(X; + X~) cosh ti,60. -

2XbX a]} .

(18.190)

18.8 Path Integral Coupled to Thermal Reservoir

1289

We have preferred deriving Zo[J+, J- ] in the operator language since this illuminates better the physical meaning of the different terms in the result (18.185) .

18.8

Path Integral Coupled to Thermal Reservoir

After these preparations, we can embark on a study of a simple but typical problem of nonequilibrium thermodynamics. We would like to understand the quantummechanical behavior of a particle coupled to a thermal reservoir of temperature T and moving in an arbitrary potential V(x) [7]. Without the reservoir, the probability of going from Xa, ta to Xb , tb would be given by 7

(18.191) This may be written as a path integral over two independent orbits, to be called

x+(t) and x_(t): (Xbtblxata)(Xbtblxata)* =

JVx+(t) Vx _(t)

x exp {~l~b dt

(18.192)

[~ (x! - x=-) -

(V(x+) - V(x _))]} .

In accordance with the development in Section 18.7, the two orbits are reinterpreted as two branches of a single closed-time orbit xp(t). The time coordinate tp of the path runs from ta to tb slightly above the real time axis and returns slightly below it, just as in Fig. 18.1. The probability distribution (18.191) can then be written as a path integral over the closed-time contour encircling the interval (ta, tb):

(18.193) We now introduce a coupling to a thermal reservoir for which we use, as in the equilibrium discussion in Section 3.13, a bath of independent harmonic oscillators oo

== 1.

(18.328)

1309

lS.10 Langevin Equations

Then Tf(t) is an instantaneous random variable with zero average and a nonzero pair correlation function:

(Tf(t))f/

=

(Tf(t)Tf(t'))f/

0,

=

wJ(t - t'),

(18.329)

All higher correlation functions vanish. A random variable with these characteristics is referred to as white noise. The stochastic differential equation (18.317) with the white noise (18.329) reduces to the classical Langevin equation with inertia [16]. In the opposite limit of small temperatures, K(w') diverges like

K(w')

----+

nlw'l

T--->O

2kB T'

(18.330)

so that w K(w') has the finite limit lim wK(w') = Mlnlw'l.

(18.331)

T--->O

To find the Fourier transformation of this, we use the Fourier decomposition of the Heaviside function (1.302)

1

8( - w') -- - 1

dte -iw't -i t + iTf

00

21l'

-00

(18.332)

to form the antisymmetric combination

1 te (i - - + -i- ) + iTf iTf - 1 te -. d

00

8(w') - 8( -w') = -1

21l'

- 00

i

00

1l'

- 00

-iw't

t

d

t -

-iw't P

t

(18.333)

A multiplication by w' yields

Iw'l = w'[8(w') -

11

8( -w')]

00

- -

1l'

- -1

1l'

1

dtUte n - iw't -P -t

- 00 00

dte -iw't 2' P t

- 00

11

00

-

1l'

-00

dte -iw'tn Ut -P t

(18.334)

By comparison with (18.331) we see that the pure quantum limit of K(t - t') can be written as

,

Min

wK(t-t)=--( T=O

1l'

P t - t' )2'

(18.335)

Hence the quantum-mechanical motion in contact with a thermal reservoir looks just like a classical motion, but disturbed by a random source with temporally long-range correlations

,

(Tf(t)Tf(t ))f/ =

Min --1l'-

P

(t - t')2

(18.336)

1310

18 Nonequilibrium Quantum Statistics

The temporal range is found from the temporal average

((~t?)t

==

1: d~t (~t?K(~t)

= -

8~2K(w')lw'=o = -~ (k:T )

2.

(18.337)

Apart from the negative sign (which would be positive for Euclidean times), the random variable acquires more and more memory as the temperature decreases and the system moves deeper into the quantum regime. Note that no extra normalization factor is required to form the temporal average (18.337), due to the unit normalization of K(t - t') in (18.224). In the overdamped limit, the classical Langevin equation with inertia (18.317) reduced to the overdamped Langevin equation:

x(t) = -V'(x(t)) /M, + ry(t)/M,.

(18.338)

At high temperature, the noise variable ry(t) has the correlation functions (18.329). Then the stochastic differential equation (18.338) is said to describe a Wiener process.The first term on the right-hand side rx(x(t)) == -V'(x(t))/M, is called the drift of the process. The probability distribution of x(t) resulting from this process is calculated as in Eqs. (18.325), (18.321) from the path integral

P(xbtalxata) =

J

'Dry P[ryl6(x,.,(t) - Xb),

(18.339)

and 'Dry is normalized so that J 'Dry P[ryl = l. A path integral representation closely related to this is obtained by using the identity (18.340) which can easily be proved by time-slicing the Fourier representation of the 6functional

6[x - ryl =

J

'Dp ei Jdtp(x - ,.,)

(18.341)

and performing all the momentum integrals. This brings the path integral (18.339) to the form (18.342) For a harmonic potential V(x) = MW5x2/2, where the overdamped Langevin equation reads

x(t) =

-w5 x(t)/M, + ry(t)/M, = -lix(t)/M, + fj(t),

(18.343)

where the noise variable fj(t) == ry(t)/M, has the correlation functions

(fj(t))

=

0, (fj(t)fj(t')),., =

M~,26(t - t')

=

2D6(t - t'),

(18.344)

the calculation of the stochastic path integral yields, of course, once more the probability (18.306).

1311

lS.11 Path Integral Solution of Klein-Kramers Equation

18.11

Path Integral Solution of Klein-Kramers Equation

For a free particle at finite temperature, there exists another way of representing the solution (18.440). Consider the original path integral for the probability in Eq. (18.240) with the Hamiltonian (18.288) . Let us introduce the thermal velocity VT == VkBT/M and write the action as [compare (2.338)]

Ae=lb dt [-i(p.i+wv) + H(p,pv, v, x)], with

(18 .345)

ta

2(

. v )2

H(p,pv,V,x) = ,VT Pv - Z2Vf

,

(

. vf ) 2

+ 4Vf v + 2z~p

Vf 2

,

+ ~p + 2" .

(18.346)

In the path integral (18.240), we may integrate out x(t), which converts the path integral over p(t) into an ordinary integral, so that we obtain the integral representation (18.347) where

Pp(Vbtblvata)

=

JVv J~~v exp {l~b dt [ipvv - Hp(Pv , v)] },

(18.348)

with the p-dependent Hamiltonian governing Pv and v:

Hp(Pv, v)

== ,v~ (Pv-i2~fr + 4:f (V+2i~pr

(18.349)

;

In the associated Hamiltonian operator, the shift of Pv by - iv/2vf can be removed by a similarity transformation to

Hp(Pv, v) ==

ev2/4v~ Hpe-v2/4V~ = ,v~p~ + 4:f (V + 2i v; p)

2

, 2

(18.350)

Thus we can rewrite Pp(Vbtblvata) as (18.351) where Pp(Vbtblvata) is the probability associated with the Hamiltonian Hp(Pv,v). This describes a harmonic oscillator with frequency, around the p-dependent center at vp = -2ivfPfr . If we denote the mass of this oscillator by m = 1/2,vf, we can immediately write down the wave eigenfunctions as 'ljJn(v - vp) with 'ljJ(x) of Eq. (2.300). The energies are Thus we may express the probability distribution of x and vasa spectral representation

n,.

P(Xbvbtblxavata) =

e-(v~-V~)/4V~

J2dP f: 'ljJn(Vb - vp)'ljJn(va - vp)e-n-)'(tb-t 7r

a)

n=O

(18.352)

1312

18 Nonequilibrium Quantum Statistics

where (18.353) In the limit of strong damping, only n = 0 contributes to the sum, and we find

Integrating this over

Vb

leads to

,-2

in the exponent. A compact expression where we have neglected terms of order for the general solution will be derived in (18.440) by stochastic calculus.

18.12

Stochastic Quantization

In Eq. (18.314) we observed that the expectation value of powers of a classical variable x in a potential V (x) can be recovered as a result of a path integral associated with the Lagrangian (18.299). From Eq. (18.339) we know that the path integral (18.314) can be replaced by the stochastic path integral: (18.355) where (18.356) To simplify the equations, we have replaced the physical time by a rescaled parameter s = tiM,. Equivalently we may say that we obtain the expectation values (18.355) by solving the stochastic differential equation of the Wiener process

x'(s) = -V'(x)

+ 1](s),

(18.357)

where 1](s) is a white noise with the pair correlation functions

(1](s))r = 0,

(1](S)1](S'))r = 2kBTJ(s - s') ,

(18.358)

and going to the large-s limit of the expectation values (xn(s)) . This can easily be generalized to Euclidean quantum mechanics. Suppose we want to calculate the correlation functions (3.295)

(X(Tl)xh)··· X(Tn))

= Z-l

JVx xh)xh)··· X(Tn)

exp ( -~Ae). (18.359)

1313

18.12 Stochastic Quantization

We introduce an additional auxiliary time variable s and set up a stochastic differential equation

5A"

(18.360)

5X(T;S ) +1](T;S), where 1](T; s) has correlation functions

(1](T; s)) = 0,

(1](T; S)1](T'; s')) = 2lib(T - T')5(s - s').

(18.361)

The role of the thermal fluctuation width 2kBT in (18.358) is now played by 2ti. The correlation functions (18.359) can now be calculated from the auxiliary correlation functions of x( T, s) in the large-s limit: (18.362) Due to the extra time variable of stochastic variable x(T; s) with respect to (18.357), the probability distribution associated with the stochastic differential equation (18.379) is a functional P[Xb(T), Sb; Xa(T), sal given by the functional generalization of the path integral (18.300):

N x e-

JVX(T;S) J::

dS{ fr;

roo

dT[OsX(T;S)+

Jx(~;s)Ael - f" JX(:~S)2Ae}.

(18 .363)

This satisfies the functional generalization of the Fokker-Planck equation (18.292): (18.364) with the Hamiltonian (18.365) where p(T) == 5j5x(T) . We have dropped the subscript b of the final state, for brevity. Explicitly, the Fokker-Planck equation (18.364) reads -

lib [lib 5A,,] 1 dT 5X(T) 5X(T) + 5X(T) P[X(T) , s; Xa(T), Sal = -tiOsP[X(T) , s; Xa(T), Sal · 00 - 00

(18 .366)

For S ----t 00, the distribution becomes independent of the initial path Xa(T), and has the limit [compare (18.314)l lim P[X(T), s; Xa(T) , sal = s--> oo

JVX(T)

e-Ae [x]/Ii

'

(18.367)

and the correlation functions (18.378) are given by the usual path integral, apart from the normalization which is here such that (1) = 1.

1314

18 Nonequilibrium Quantum Statistics

As an example, consider a harmonic oscillator where Eq. (18.360) reads (18.368) This is solved by (18.369) The correlation function reads, therefore,

f"1 f"2 (2 2)(" ) (X(Tl;8 1)X(T2;8 2))=}o d8~}o d8;e M -&T+ w 81+8 2-81- 82 (1J(Tl;8~)1J(T2;8m·(18.370) Inserting (18.361), this becomes

(x( Tl; 81)X( T2; 82)) = 11,

lXJ d8

[e-M(-&;+w2)(8+181 -821) -

e- M(-&;+W 2 )(8+8 1+8 2)], (18.371)

or

For Dirichlet boundary conditions (Xb = Xa = 0) where operator (-8; + w 2 ) has the sinusoidal eigenfunctions of the form (3.63) with eigenfrequencies (3.64) , this has the spectral representation

(18.373) For large 81, 82, the second term can be omitted. If, in addition, 81 = 82 , we obtain the imaginary-time correlation function [compare (3.69), (3.301) , and (3.36)]: 8111~)x(Tl; 8)X(T2; 8))

11,

1

= (X(Tl)X(T2)) = M -8; + w 2 (Tl' T2) 11, sinhw(Tb - T» sinhw(T< - Ta)

M

(18.374)

wsinhw(Tb - Ta)

We can use these results to calculate the time evolution amplitude according to an imaginary-time version of Eq. (3.315): ( X bTib Ixa T.a )

=

C(xb, x a )e- A e (Xb,Xa .,Tb-Ta )/~"e -

fTb T

a

M.dT' (L e,n (x b , xb ))/r. 2 b

,

(18.375)

where Ae(Xb , Xa; Tb - Ta) is the Euclidean version ofthe classical action (4.87). If the Lagrangian has the standard form, then (18.376)

1315

18.13 Stochastic Calculus

and we obtain the imaginary-time evolution amplitude in an expression like (3.315). The constant of integration is determined by solving the differential equation (3.316), and a similar equation for Xa. From this we find as before that C(Xb, xa) is independent of Xb and Xa. For the harmonic oscillator with Dirichlet boundary conditions we calculate from this (18.377) Integrating this over Tb yields n(D/2) log[2sinhwh - Ta)], so that the second exponential in (18.375) reduces to the correct fluctuation factor in the D-dimensional imaginary-time amplitude [compare (2.409)]. The formalism can easily be carried over to real-time quantum mechanics. We replace t --+ -iT and A., --+ -iA, and find that the real-time correlation functions are obtained from the large-s limit (18.378) where x( t; s) satisfies the stochastic differential equation

. 8A nosx(t; s) = z~() uX tis

+ 1](t; s),

(18.379)

where the noise 1](t; s) has the same correlation functions as in (18.361), if we replace by t. This procedure of calculating quantum-mechanical correlation functions is called stochastic quantization [17].

T

18.13

Stochastic Calculus

The relation between Langevin and Fokker-Planck equations is a major subject of the so-called stochastic calculus. Given a Langevin equation, the time order of the potential V(x) with respect to j; and i: is a matter of choice. Different choices form the basis of the Ito or the Stratonovich calculus. The retarded position which appears naturally in the derivation from the forward- backward path integral favors the use of the Ito calculus. A midpoint ordering as in the gauge-invariant path integrals in Section 10.5 corresponds to the Stratonovich calculus.

18.13.1

Kubo's stochastic Liouville equation

It is worthwhile to trace how the retarded operator order of the friction term enters the framework of stochastic calculus. Thus we assume that the stochastic differential equations (18.322) and (18.323) have been solved for a specific noise function 1](t) such that we know the probability distribution Pry (x v tlxavat a ) in (18.324). Now we observe that the time dependence of this distribution is governed by a simple

1316

18 Nonequilibrium Quantum Statistics

differential equation known as Kubo's stochastic Liouville equation [18], which is derived as follows [19]. A time derivative of (18.324) yields

at P1)(xvtlx avat a) = x1)(t)J'(x1)(t) - x)J(x1)(t) - v)

+ x1)(t)J(x1)(t) - x)J'(x1)(t) - v). (18.380)

The derivatives of the J-functions are initially with respect to the arguments x1)(t) and x1)(t). These can, however, be expressed in terms of derivatives with respect to -x and -v. However, since x1)(t) depends on x1)(t) we have to be careful where to put the derivative The general formula for such an operation may be expressed as follows: Given an arbitrary dynamical variable z(t) which may be any local function (local in the temporal sense) of x(t) and x(t), and whose derivative is some function of z(t), i.e. , i(t) = F(z(t)) , then

-avo

d . a a . a dtJ(z(t)-z) = z(t) az(t) J(z(t)-z) = - a)z(t)J(z(t)-z)] = - a)F(z)J(z(t)-z)]. (18.381) To prove this formula, we multiply each expression by an arbitrary smooth test function g(z) and integrate over z. Each integral yields indeed the same result g(z(t)) = i(t)g'(z(t)) = F(z)g'(z(t)). Applying the identity (18.381) to (18.380), we obtain an equation for P1)(x v tlxavata): (18.382) We now express x1)(t) with the help of the Langevin equation (18.317) in terms of the friction force -M"(x1)(t), the force -V'(x1)(t)), and the noise 1](t). In the presence of the J-function J(x1)(t) - v), the velocity x1)(t) can everywhere be replaced by v, and Eq. (18.382) becomes (18.383) where

f(x,v) == -M"(v - V'(x)

(18.384)

is the sum of potential and friction forces. This is Kubo's stochastic Liouville equation which, together with the correlation function (18.218) of the noise variable and the prescription (18.325) of forming expectation values, determines the temporal behavior of the probability distribution P(x v tlxavata).

18.13.2

From Kubo's to Fokker-Planck Equations

Let us calculate the expectation value of P1)(x v tlxavata) with respect to noise fluctuations and show that P(x v tlxavata) of Eq. (18.325) satisfies the Fokker-Planck equation with inertia (18.243). First we observe that in a Gaussian expectation

1317

18.13 Stochastic Calculus

value (18.321), the multiplication of a functional F[1]] by 1] produces the same result as the functional differentiation with respect to 1] with a subsequent functional multiplication by the correlation function (1](t)1](t')): (18.385) This follows from the fact that 1](t) can be obtained from a functional derivative of the Gaussian distribution in (18.321) as:

J

1](t)e-i;; Jdtdt'1/(t)K- 1 (t,t')1/(t')= -w dt' K(t, t') 2Ii)

-w

~W;f [1 + 12~::~)2]

IXfi 12.

(18.573)

This time dependence is caused by spontaneous emission and induced emission and absorption. To identify the different contributions, we rewrite the spectral decompositions (18.550) and (18.551) in the x-independent approximation as

or

1342

18 Nonequilibrium Quantum Statistics

Following Einstein's intuitive interpretation, the first term in curly brackets is due to spontaneous emission, the other two terms accompanied by the Bose occupation function account for induced emission and absorption. For high and intermediate temperatures, (18.575) has the expansion

(18.576) The first term in curly brackets corresponds to the spontaneous emission. It contributes to the rate of change Ot(iljJ(t)li) a term -2M"(Lf 0, since the ,,-function allows only for decays. Second, there is an extra factor 2. Indeed, by comparing (18.574) with (18.576) we see that the spontaneous emission receives equal contributions from the 1 and the coth(Ii,wI/2k B T) in the curly brackets of (18.574), i.e., from dissipation and fluctuation terms Cb(t, tl) and Ab(t, tl) . Thus our master equation yields for the natural line width of atomic levels the equation

r

= 2M"( I>rf

IXfi1 2,

(18.577)

f 0 the time derivative

Otu(x, t)

=

il u(x, t),

at t = 0,

(18.658)

with the time evolution operator

n

2

1i == {[w x xl . V} + - (n· V) . A

2

(18.659)

The average over T) has made the operator il time-independent. For this reason, the average field u(x, t) at an arbitrary time t is obtained by the operation u(x, t) = U(t)u(x, 0) ,

(18.660)

where U(t) is a simple exponential (18.661) as follows immediately from (20A.ll) and the trivial property ilU(t) = U(t)il. Note that the operator il commutes with the Laplace operator V2 , thus ensuring that the harmonic property (18.650) of u(x) remains true for all times, i.e., (18.662)

18.24.3

Harmonic Oscillator

We now show that Eq. (20A.ll) describes the quantum mechanics of a harmonic oscillator. Let us restrict our attention to the line with arbitrary Xl == X and X2 = O. Applying the Cauchy-Riemann equations (18.651), we can rewrite Eq. (20A.ll) in the pure x-form

n 2 2 (x,t) , WXOxU 2( x,t) -'20xU

(18.663)

-w X Oxu l (X, t) + ~O; u l (X, t),

(18.664)

where we have omitted the second spatial coordinates X2 = O. Now we introduce a complex field (18.665) This satisfies the differential equation (18.666) which is the Schriidinger equation of a harmonic oscillator with the discrete energy spectrum En = (n + 1/2)nw, n = 0, 1, 2, ....

18.24.4

General Potential

The method can easily be generalized to an arbitrary potential. We simply replace (20A.l) by

±l(t) ±2(t) =

- 02 SI (X(t)) + n l T)(t), - OlSI(x(t))+n2T)(t) ,

(18.667)

where S(x) shares with u(x) the harmonic property (18.650) :

V 2 S(x) = 0,

(18.668)

1354

18 Nonequilibrium Quantum Statistics

i.e., the functions SIL(x) with fJ, = 1, 2 fulfill Cauchy-Riemann equations like uIL(x) in (18.651). Repeating the above steps we find, instead of the operator (18.659), 1i 2 == -(ihS )01 - (OlS )02 + 2(0' V) ,

A l l

7t

(18.669)

and Eqs. (18.663) and (20A.14) become:

~O; u2(x, t) ,

(18.670)

+ ~O;Ul(X, t) .

(18.671)

(OxS I )oxu2(x, t) -(OxS1)OxU1(x,t)

This time evolution preserves the harmonic nature of u(x). Indeed, using the harmonic property V 2 S(x) = 0 we can easily derive the following time dependence of the Cauchy-Riemann combinations in Eq. (18.651): Ot(OlU l - 02U2) = H(OlU l - 02U 2) - 020lSl(OlUl - 02U2) + 0~Sl(02Ul + 01U 2), Ot(02U1 + 01U2) = H(~Ul + OlU 2) - 020lSl(02Ul + OlU2) - O~SI(OIUl - 02 U2 ). Thus 01 u l - ~U2 and 02U1 + 01 u 2 which are zero at any time remain zero at all times. On account of Eqs. (18.671) , the combination 'Ij; (x , t)

== e-S'(x)/Ii [u 1 (x, t) + iu2 (x, t)l

(18.672)

satisfies the Schriidinger equation i1iot'lj;(x , t) = [-

~2 0; + V(X)]

'Ij;(x, t) ,

(18.673)

where the potential is related to Sl(x) by the Riccati differential equation (18.674) The harmonic oscillator is recovered for the pair of functions (18.675)

18.24.5

Deterministic Equation

The noise 1/(t) in the stochastic differential equation Eq. (18.667) can also be replaced by a source composed of deterministic classical oscillators qk(t), k = 1, 2, ... with the equations of motion (18.676) as

1/(t) ==

L qk(t) .

(18.677)

k

The initial positions qk(O) and momenta Pk(O) are assumed to be randomly distributed with a Boltzmann factor e-/3 H o,c/ Ii , such that (18.678) Using the equation of motion (18.679)

1355

Appendix 18A Inequalities for Diagonal Green Functions we find the correlation function

(qk(t)qk(t'))

=

w~ coswktcoswkt' (qk(O)qk(O))

=

COSWk(t - t').

+ sinwktsinwkt(pk(O)Pk(O)) (18.680)

We may now assume that the oscillators qk(t) are the Fourier components of a massless field, for instance the gravitational field whose frequencies are Wk = k, and whose random initial conditions are caused by the big bang. If the sum over k is simply a momentum integral J::O= dk, then (18.680) yields a white-noise correlation function (18.653) for ",(t). Thus it is indeed possible to simulate the quantum-mechanical wave functions 'I/! (x, t) and the energy spectrum of an arbitrary potential problem by deterministic equations with random initial conditions at the beginning of the universe. It remains to solve the open problem of finding a classical origin of t he second important ingredient of quantum theory: the theory of quantum measurement to be extracted from the wave function 'I/!(x, t). Only then shall we understand how God throws dice [37].

Appendix 18A

Inequalities for Diagonal Green Functions

Let us introduce several diagonal Green functions consisting of thermal averages of equal-time commutators and anticommutators of bosonic and fermionic field operators, elementary or composite. For brevity, we write ( .. .)r

Tr [exp ( -HIT) ... ] / Tr [exp( -HIT) ]

=

= ( .. . )T

(18A.1)

and define the averages with obvious spectral representations c

== (['I/!,, 'I/!'t l'f )r =

a

, 't == (['I/!, 'I/! l±)r =

1= _= 1=_=

dJ.,;

27rP12(W) ,

dJ.,;

(18A.2) 1 W

27r P12(W) tanh'f 2T ·

We shall also introduce a quantity obtained by integrating the imaginary-time Green function over a period r E [0, liT). This gives for boson and fermion fields the nonnegative expression [see (18.23)] 9

tiTdr ({j;(r){j;t(O))r

G(Wm = 0) = Jo

1_==

dJ.,; 1 } > -P12 W -1 { 27r () W tanh(wI2T) -

o.

(18A.3)

Note that for fermion fields, the spectral weight in this integral is accompanied by an extra factor tanh(wI2T) . This is due to the fact that Wm = 0 is no fermionic Matsubara frequency, but a "wrong" bosonic one. In fact , the sum (18.23) for G(wm ) contains a factor 1 - e(En - En ' )IT for both bosons and fermions, while P12(W) in the spectral representation (18A.3) introduces, via (18.59), a factor 1 - e- wlT for bosons and a factor 1 + ewlT for fermions, thus explaining the relative factor tanh(wI2T) in (18A.3). The integration over r leads to the factor 11w in (18A.3). This factor is also found by integrating the retarded Green function G 12 (t) and the commutator function C 12 (t) over all real times, resulting in the spectral representations i i

i: i:

dt 8(t) ([{j;(t) , (j;t(O)l'f)r dt8(t)([{j;(t),{j;t(0)l ±)T

J ~ p12(W)~, (18A.4)

1356

18 Nonequilibrium Quantum Statistics

Another set of thermal expectation values involves products of field operators with time derivatives rather than integrals. Their spectral representations contain an extra factor w. For example, the T-derivative of the expectation value in (18A.3) leads to (18A.5) The real-time derivatives of the expectation values in (18AA) have the spectral integrals d

i([~(O),~t(O)] 'f)T = J::'oo ~ P12(W)W,

e

i([~(O),~t(O)] ±)T = J::'oo ~ P12(w)wtanh'f 1 2T·

(18A.6)

The expectation values c, a, g, d, e satisfy several rigorous inequalities. To derive these, we observe that

Jl(w) =

91127r P12 (w);:;1{

1

tanh(w/2T)

}

(18A.7)

is a positive function. This follows directly from (18.59), according to which P12(W) at negative W is negative for bosons and positive for fermions. Having divided out the total integral 9 defined in (18A.3) , the integral over Jl(w) is normalized to unity, (18A.8) for both bosons and fermions. Using Jl(w), we form the following ratios: c 9

a 9

d 9

e 9

=

1: 1: 1: 1:

1 dw Jl(W)W {

dw Jl(W)W {

dw Jl (W)W 2 {

dw Jl(w)w 2 {

coth(w/2T) coth(w/2T) 1 1 coth(w/2T)

coth(w/2T) 1

}, }, }, }.

(18A.9)

(18A.1O)

(18A.ll)

(18A.12)

The inequalities to be derived are based on the Jensen-Peierls inequality for convex functions derived in Chapter [5]. Recall that a convex function f(w) satisfies (18A.13) which is generalized to (18A.14) where Jli is an arbitrary set of positive numbers with becomes

2:i Jli

= 1. In the continuum limit, this

(18A.15)

Appendix 18A Inequalities for Diagonal Green Functions

1357

It is obvious that a similar Jensen-Peierls inequality holds also for concave functions with inequality sign in the opposite direction. The Jensen-Peierls inequality (18A.15) is now applied to the function W

I(w) = wcoth 2T'

(18A.16)

which looks like a slightly distorted hyperbola coming in from infinity along the diagonal lines Iwl and crossing the I-axis at w = 0,/(0) = 2T. The second derivative of I(w) is positive everywhere, ensuring the convexity. The function (18A.16) appears in the integrand of the boson part of Eq. (18A.1O). The right-hand side of (18A.15) can therefore be written as a/g . The left-hand side is obviously equal to (c/g)coth(c/2Tg). Hence we arrive at the inequality c c coth 2Tg

:s: a.

(18A.17)

In terms of the original field operators, this amounts to (18A.18)

In the special case that {f; is a canonical interacting boson field of momentum p , the commutator is simply [{f;, (f;tl- = 1, and the inequality becomes 1 + 2({f;b{f;p)T

~

2 coth(1/2Tg) = 1 + e1 / gT _ 1 = 1 + 2ny-l ,

(18A.19)

i.e.,

(18A.20) where ny-I is the free-boson distribution function (18.36) for an energy g- l . This is quite an interesting relation. The quantity 9 is the Euclidean equilibrium Green function G(wm,p) at Wm = O. For free particles in contact with a reservoir, it is given by 2

g- 1 =

G(O,p) - 1 = :M - J.t == ~(p),

(18A.21)

i.e., it is equal to the particle energy measured with respect to the chemical potential J.t. Moreover, we know that for free particles (18A.22) so that the inequality (18A.20) becomes an equality. The content of the inequality (18A.20) may therefore be phrased as follows: For any interaction, the occupation of a state with momentum p is never smaller than for a free boson level of energy g-1 = G(O, p). Another inequality can be derived from the concave function -

I(Y)

y'y

= y'ycoth 2T '

(18A.23)

using y = w2 and the measure

1-0000

dWJ.t(w) =

100 dyr;;J.t(y'y) = 1. 0

vy

(18A.24)

1358

18 Nonequilibrium Quantum Statistics

As argued before, concave functions satisfy the inequality opposite to (18A.15), from which we derive the inequality

J

(Jor= y'y J.L( y'y)y dy

)

~

dy Jor= y'y J.L( y'y) J(y),

(18A.25)

which can be rewritten as (18A.26) Again, the right-hand side is a/g, but now it is bounded from above by

fi.) . ygfi. coth (~ 2Tyg

(18A.27)

2;9 : : ; a::::; yldgcoth (2~~)

(18A.28)

?:.::::;

9

The combined inequality c coth

may be used to derive further inequalities:

r? ::::; dg, c coth(d/2Tc)

::::;

a,

9

::::;

coth(c/2Ta),

c

::::;

dtanh(c/2Ta),

c

::::;

a tanh(d/2Tc) .

(18A.29)

For fermion fields we see that an inequality like (18A.17) holds with c and a interchanged, i.e.,

a acoth 2Tg ::::; c,

(18A.30)

which leads to

([~, ~tl_)T ::::; ([~, ~tl+)T tanh

([~, ~tAl+)rA

2T JOI/T dT (7f!(T)7f!t(O))r

(18A.3l)

For canonical fermion fields with [~ , ~tl+ = 1, this becomes

At A

1 - 2(7f! 7f!)T

::::;

2 tanh(1/2gT) = 1 - el/gT + l'

(18A.32)

i.e., the fermionic counterpart of (18A.20) :

At A 1 (7f!p7f!p)r ::::; e l / gT + 1 = ng-I,

(18A.33)

where ny-I is the free-fermion distribution function (18.36) at an energy g-l. As in the Bose case, free particles fulfill (18A.34) with g-l = ~(p), so that the inequality (18A.33) becomes an equality. The inequality implies that an interacting Fermi level is never occupied more than a free fermion level of energy g-l = G(O,p)-l. Also the second inequality in (18A.29) can be taken over to fermions which amounts to (18A.30), but with a and d replaced by c and e.

1359

Appendix 18B General Generating Functional

Appendix 18B

General Generating Functional

For a field operator a(t) of frequency n and its Hermitian conjugate at(t) , the retarded and advanced Green functions and the expectation values of commutators and anticommutators were derived in Eqs. (18.68)- (18.77) : G~(t, t'l

8(t - t')e- ifl(t- t') ,

G~(t , t'l

-8(t' - t)e- ifl(t- t') , e- in(t- t') ,

Gn(t, t'l

n )'1'1 e-in(t-t') . ( tanh 2T

An(t, t'l

(18B.l)

Introducing complex sources 1)(t) and 1)t(t) associated with these operators, the generating functional for these functions is

(18B.2) The complex sources are distinguished according to the closed-time contour branches by a subscript P. The generating functional can then be written down immediately as

ZO [1JP ,1)tl = exp { -

JJ dt

dt' 1)t(t)Gp (t , t l )1JP(tl ) }

,

(18B.3)

generalizing (18.179) , where the matrix G _ ~ ( An + G~ + G~ p 2 An + G~ - G~

An - G~ + G~ ) An - G~ - G~

(18BA)

contains the following operator expectations on the two time branches:

( (~paH(t+)ak(t~ )h (TpaH(L)ak(t~)h (

(~paH(t+ )ak(t~))T )

(TpaH(L)ak(t~)h

(TaH(t+)ak(t~)h ±Jak(L)aH(t+))T ) . (aH(L)ak(t~)) T

(TaH(L)ak(t~)) T

(18B.5)

Note that aH(t+) , ak(t+) and aH(L) ,ak(t-) obey the Heisenberg equations of motion with the Hamiltonians

~

[ak(t+ )aH(t+) ± aH(t+ )ak(t+ )] ,

-~

fL

[ak(L)aH(L) ± aH(L)ak(L)] .

(18B.6)

In the second-quantized field interpretation, they read

H+

~ak(t+)aH(t+),

H_

- ~ak(L)aH(L).

The explicit time dependence of the matrix elements of Gp(t, t'l in Eq. (18B.5) is

G (t t'l = p

,

e

-ifl(t-t') ( 8(t - t') ± nn 1 ± nn

±nn ) 8(t' - t) ± nn '

(18B.7)

1360

18 Nonequilibrium Quantum Statistics

with nn = (e n / T ± 1) - 1. This Green function can, incidentally, be decomposed as G~(t, t')

+ Gf;' (t , t'),

(18B .8)

where G~(t, t') is the Green function at zero temperature, i.e., the expression (18B.7) for nn The matrix G N (t , t') contains the expectations of the normal products :

== O.

(~aH(t+)ak(t~))T ) (NaH(L)ak(t~)h

(ak(t~)aH(t+) )T ) (ak(t~)aH(L)h

.

(18B.9)

For an arbitrary product of operators, the normal product N( ... ) is defined by reordering the operators so that all annihilation operators come to act first upon the state on the right-hand side. At the end, the product receives the phase factor (_ )F, where F is the number of fermion permutations to arrive at the normal order. A similar decomposition exists at the operator level before taking expectation values. For any pair of operators A(t), B(t') which are linear combinations of creation and annihilation operators, the time-ordered product can be decomposed as

TA(t)B(t')

=

(TA(t)B(t'))o

+ NA(t)B(t'),

(18B.1O)

where ( .. ')0 == Tr (10)(01 ... ) denotes the zero-temperature expectation. This decomposition is proved in Appendix 18C , where it is also generalized to products of more than two operators. Let us also go to the Keldysh basis here: ijp = Q1)p =

~ (~ -~) 7Jp.

(18B.ll)

Then the generating functional becomes [instead of (18.180)]:

ZO [1)p,1)p]

= exp [= exp { -

JJ ~J J

dt' (1)+,

dt

dt

_1)~)Q-l (;~ ~) Q (

dt' [(1)+

-1)~)(t)G~(t -

: ;_ )]

t')(1)+ + 1)- )(t')

+(1)+ + 1)~)(t)G~(t - t')(1)+ -1)- )(t')

+(1)~ -1)~)(t)An(t - t')(1)+ -1)- )(t')] },

(18B.12)

where we have used the notation (18B .13)

Expression (18B.12) can be simplified [as before (18.180)] using the time reversal relation (18.62) in the form

An(t, t')

=

8(t, t')An(t, t') ± 6(t' - t)An(t' , t),

leading to

ZO [1)p, 1)p]

= exp { -

~

i: / 00

x [(1)+

dt

(18B .14)

dt'

- 1)- )*(t)G~(t, t')(1)+ + 1)- )(t')

-(1)+ -1)_ )(t)G~(t, t')*(1)+ + 1)- )*(t') +(1)+ - 1)_ )*(t)An(t, t')(1)+ - 1)- )(t')

+(1)+ - 1)- )(t)An(t, t')* (1)+ - 1)- )*(t') 1}.

(18B .15)

1361

Appendix 18B General Generating Functional

For the case of a second-quantized field, this is the most useful generating functional. The expression (18B.15) can be used to derive the generating functional for correlation functions between one or more 4'(t) and the associated canonically-conjugate momenta. As an example, consider immediately a harmonic oscillator with 4'(t) = x(t) and the momentum p(t). We would like to find the generating functional Z[jp , k p] = Tr

(.0 Tp exp {i

l

dx [jp(t)xp(t)

+ kp(t)Pp(t)]})

(18B.16)

.

The position variable x(t) is decomposed as in (18.92) into a sum of creation and an annihilation operators:

J2Mnn [ae-

x(t) =

int

+ at eint ] .

(18B.17)

The inverse of this decomposition is (18B.18) and there is an analogous relation of the complex sources:

~t

{

} =

(j ± iMn k) ;V2Mnn.

(18B.19)

Inserting these sources into (18B.15) , we obtain the generating functional ZO[jp,kp] = exp {

-2~n1: dt loo dt'(j+-j_)(t) x {[ReAn(t, t')

11 1t

-"2

00

- 00

dt

+ iImG~(t, t')]j+(t')

- [Re An(t, t') - iIm G~(t, t')]j- (t')}

(18B.20)

dt' (k+ - L)(t) {[ImAn(t,t') - iReG~(t,t')]J+(t')

-00

_

[ImAn(t,t')+iReG~(t,t')]j_ (t')}+(jkMn)} .

Here it is useful to introduce the quantities

0:( t , t')

2~n

[ReAn(t, t')

+ iImG~(t, t')]

f3( t, t')

2~n

[1m An(t, t') -

,

iReG~(t, t')].

(18B.21)

Then the generating functional reads Zo[j+,j- , k+ , L] = exp

{-1:

-Mn

1:

dt loodt' (j+-j_)(t) [o:(t, t')j+(t')-o:*(t, t')j-(t')]

dt loodt' (k+-L)(t) [f3(t, t')j+(t')-f3*(t, t')j_ (t')]

+ (j ..... kMn)

+ (j ..... kMn)}. (18B.22)

If the oscillator is coupled only to the real source j, i.e., if its generating functional reads

(18B.23)

1362

18 Nonequilibrium Quantum Statistics

we can drop all but the first line in the exponent of (lSB.22) and have

Zo[j+,j- ] = exp { -

i:

dt

[~" dt' (j+ -

j _ )(t) [aCt , t')i+(t') - a'(t, t')j- (t')] }. (lSB.24)

Since (lSB.22) and (lSB.24) contain only the causal temporal order t > t' , the retarded Green function G~(t, t') in (lSB.21) can be replaced by the expectation value of the commutator [see (lS.40), (lS.41) , and (lS.42)]. Thus, for t > t', the functions aCt, t') and f3(t, t') are equal toll

aCt, t')

2~fl

[ReAn(t , t')

+ iImCn(t, t')],

t > t' ,

f3(t, t')

2~fl

[ImAn(t, t') - iReCn(t, t')] ,

t > t'.

(lSB.25)

For a single oscillator of frequency fl, we use the spectral function (lS.74) properties (lS.44) and (lS.53) of An(t, t') and Cn(t, t'), and find the simple expressions:

a(t,t')

_1_ [Re e-in(t-t') { coth 2~ 2Mfl tanh 2T

=

2~fl [COSfl(t -

}

+ iIme-in(t-t')]

n{::~ :~ }-

isinfl(t -

n] ,

(lSB.26)

.

1 [Ime - in(t- t') { coth 2~} Mfl n - zRee - in(t- t')] tanh 2T 2

f3( t, t')

= -

2~fl [sinfl(t - t') { ::~ :~

}

+ icosfl(t -

n] .

(lSB.27)

Note that real and imaginary parts of the functions a(t-t') can be combined into a single expression (13 = l/T) cosh[fl(f3/2 - i(t - t')] 1 { sinh(flf3/2) aCt - t ) - 2Mfl sinh[fl(f3/2 - i(t - t')] cosh(flf3/2) _

, _

for

bosons,

for

fermions.

(lSB.2S)

The bosonic function agrees with the time-ordered Green function (lS.lOl) for t > t' and continues it analytically to t < t' . In Fourier space, the functions (lSB .26) and (lSB.27) correspond to

a(w')

=

2:m (coth±1 ; ;

+ 1)

[8(w' - fl) - 8(w' + fl)] ,

f3(w')

= -

2~fl

+ 1)

[8(w' - fl)

(coth±l ; ;

+ 8(w' + fl)] .

Let us split these functions into a zero-temperature contribution plus a remainder

11 Note

± en /; 'f 1 [8(w' - fl) + 8(w' + fl)]j ,

a(w')

=

;;fl {8(W' - fl)

f3(w')

=

i1n {8(w' - fl) ± en /;

that aCt, t') = (x(t)x(t')h .

'f 1 [8(w' - fl) - 8(w'

+ fl)]

.

Appendix 180 Wick Decomposition of Operator Products

1363

On the basis of this formula, Einstein first explained the induced emission and absorption of light by atoms which he considered as harmonically oscillating dipoles in contact with a thermal reservoir. He imagined them to be harmonically oscillating dipole moments coupled to a thermal bath consisting of the Fourier components of the electromagnetic field in thermal equilibrium. Such a thermal bath is called a black body. The first purely dissipative and temperature-independent term in a(w') was attributed by Einstein to the spontaneous emission of photons. The second term is caused by the bath fluctuations, making energy go in and out via induced emission and absorption of photons. It is proportional to the occupation number of the oscillator state nn = (e - n/T'f 1) - 1. The equality of the prefactors in front of the two terms is the important manifestation of the fluctuation-dissipation theorem found earlier [see (18.53)] .

Appendix 18C

Wick Decomposition of Operator Products

Consider two operators A(t) and B(t) which are linear combinations of creation and annihilation operators

A(t) B(t)

ala(t) f31a(t)

+ a2at(t) , + f32a t (t) .

(18C.l)

We want to show that the time-ordered product of two operators has the decomposition quoted in Eq. (18B.I0):

TA(t)B(t) = (TA(t)B(t))o

+ N A(t)B(t) .

(18C.2)

The first term on the right-hand side is the thermal expectation of the time-ordered product at zero temperature; the second term is the normal product of the two operators. If A and B are both creation or annihilation operators, the statement is trivial with (1' AB)o = O. If one of the two, say A(t), is a creation operator and the other, B(t) , an annihilation operator, then

S(t - t')a(t)af(t') ± S(t' - t)at(t')a(t) S(t - t') [a(t)a t (t')]'f ± at (t')a(t).

(18C.3)

Due to the commutator (anticommutator) the first term is a c-number. As such it is equal to the expectation value of the time-ordered product at zero temperature. The second term is a normal product, so that we can write (18C.4) The same thing is true if a and at are interchanged (such an interchange produces merely a sign change on both sides of the equation). The general statement for A(t)B(t') follows from the bilinearity of the product. The decomposition (18C.2) of the time-ordered product of two operators can be extended to a product of n operators, where it reads

TA(tl) " .A(tn ) = tNA(tIl .. . A(ti )" . A(t n ).

(18C.5)

i= 2

A common pair of dots on top of a pair of operators denotes a Wick contraction of Section 3.10. It indicates that the pair of operators has been replaced by the expectation (TA(tIlA(ti))o, multiplied by a factor (- )F, if F = fermion permutations were necessary to bring the contracted operator to the adjacent positions. The remaining factors are contracted further in the same way. In this way, any time-ordered product (18C.6)

1364

18 Nonequilibrium Quantum Statistics

can be expanded into a sum of normal products of these operators containing successively one, two, three, etc. pairs of contracted operators. The expansion rule can be phrased most compactly by means of a generating functional (18C.7) Differentiations with respect to the source jet) on both sides produce precisely the above decompositions. By going to thermal expectation values of (18C.7) at a temperature T , we find

\T iJ= A

e

dtA(t)j(t))

=

-00

e -t

J= dtdt'j(t)G(t ,t')j(t') -00

,

(18C.8)

T

with

G(t, t') =

(T A(t)A(t'))o + eN A(t)A(t'))r.

(18C.9)

The first term on the right-hand side is calculated at zero temperature. All finite temperature effects reside in the second term.

Notes and References The fluctuation-dissipation theorem was first formulated by H.B. Callen and T.A. Welton, Phys. Rev. 83, 34 (1951). It generalizes the relation between the diffusion constant and the viscosity discovered by A. Einstein, Ann. Phys. (Leipzig) 17, 549 (1905) , and an analogous relation for induced light emission in A. Einstein, " Strahlungs-Emission und -Absorption nach der Quantentheorie", Verhandlungen der Deutschen Physikalischen Gesellschaft 18, 318 (1916) , where he derived Planck's black-body formula. See also the functioning of this theorem in the thermal noise in a resistor: H. Nyquist, Phys. Rev. 32, no (1928). K.V. Keldysh, Z. Eksp. Teor. Fiz. 47, 1515 (1964) ; Sov. Phys. JETP 20, 1018 (1965). See also V. Korenman, Ann. Phys. (N. Y.) 39, 72 (1966); D. Dubois, in Lectures in Theoretical Physics, Vol. IX C, ed. by W.E. Brittin, Gordon and Breach, New York, 1967; D. Langreth, in Linear and Nonlinear Electronic Transport in Solids, ed. by J.T. Devreese and V. Van Doren, Plenum, New York, 1976; A.M. Tremblay, B. Patton, P.C. Martin, and P. Maldague, Phys. Rev. A 19, 1721 (1979). For the derivation of the Langevin equation from the forward- backward path integral see S.A. Adelman, Chern. Phys. Lett. 40, 495 (1976); and especially A. Schmid, J. Low Temp. Phys. 49, 609 (1982) . To solve the operator ordering problem, Schmid assumes that a time-sliced derivation of the forward- backward path integral would yield a sliced version of the stochastic differential equation (18.317) 1)n == (M/€)(xn - 2Xn- l + Xn-2) + +(M'Y/2)(xn - Xn-2) + €V'(Xn-l. The matrix a1)/ax has a constant determinant (M/€)N(l + q/2)N . His argument [cited also in the textbook by U. Weiss, Quantum Dissipative Systems, World Scientific, 1993, in the discussion following Eq. (5.93)] is unacceptable for two reasons: First, his slicing is not derived. Second, the resulting determinant has the wrong continuum limit proportional to

Notes and References

1365

exp [J dt'Yj2] for E --+ 0, N = (tb - ta)jE --+ 00, corresponding to the unretarded functional determinant (18.274), whereas the correct limit should be 'Y-independent, by Eq. (18.282). The above textbook by U. Weiss contains many applications of nonequilibrium path integrals. More on Langevin and Fokker-Planck equations can be found in S. Chandrasekhar, Rev. Mod. Phys. 15, 1 (1943); N.G. van Kampen, Stochastic Processes in Physics and Chemistry , North-Holland, Amsterdam, 1981; P. Hanggi and H. Thomas, Phys. Rep. 88, 207 (1982); C.W. Gardiner, Handbook of Stochastic Methods, Springer Series in Synergetics, 1983, Vol. 13; H. Risken, The Fokker-Planck Equation, ibid., 1983, Vol. 18; R. Kubo, M. Toda, and N. Hashitsume, Statistical Physics II, Springer, Berlin, 1985; H. Grabert, P. Schramm, and G.-L. Ingold, Phys. Rep. 168, 116 (1988) . The stochastic Schrodinger equation with the Hamiltonian operator (18.521) was derived by A.O. Caldeira and A.J. Leggett, Physica A 121 , 587 (1983); A 130 374(E) (1985). See also A.O. Caldeira and A.J. Leggett, Phys. Rev. A 31, 1059 (1985). A recent discussion of the relation between time slicing and Ito versus Stratonovich calculus can be found in H. Nakazato, K. Okano, L. Schulke, and Y. Yamanaka, Nucl. Phys. B 346, 611 (1990). For the Heisenberg operator approach to stochastic calculus see N. Saito and M. Namiki, Progr. Theor. Phys. 16, 71 (1956). Recent applications of the Langevin equation to decay problems and quantum fluctuations are discussed in U. Eckern, W. Lehr, A. Menzel-Dorwarth, F. Pelzer. See also their references and those quoted at the end of Chapter 3. The quantum Langevin equation is discussed in G.W. Ford, J.T. Lewis und R.F. O'Connell, Phys. Rev. Lett. 55, 42273 (1985); Phys. Rev. A 37, 4419 (1988) ; Ann. of Phys. 185, 270 (1988). Deterministic models for Schrodinger wave functions are discussed in G. 't Hooft, Class. Quant. Grav. 16,3263 (1999) (gr-gcj9903084); hep-thj0003005; Int. J . Theor. Phys. 42, 355 (2003) (hep-thj0104080) ; hep-thj0105105; Found. Phys. Lett. 10, 105 (1997) (quant-phj9612018). See also the Lecture Ref. [37]. The representation in Section 18.24 is due to Z. Haba and H. Kleinert, Phys. Lett. A 294, 139 (2002) (quant-phj0106095) . See also F . Haas, Stochastic Quantization of the Time-Dependent Harmonic Oscillator, Int. J . Theor. Phys. 44, 1 (2005) (quantjph-0406062). M. Blasone, P. Jizba, and H. Kleinert , Phys. Rev. A 71, 2005; Braz. J . Phys. 35, 479 (2005) (quantjph-0504047); Annals Phys. 320,468 (2005) (quantjph-0504200). Another improvement is due to M. Blasone, P. Jizba, G. Vitiello, Phys. Lett. A 287, 205 (2001) (hep-thj0007138); M. Biasone, E . Celeghini, P. Jizba, G. Vitiello, Quantization, Group Contraction and Zero-Point Ene'T"!Jy , Phys. Lett. A 310,393 (2003) (quant-phj0208012). The individual citations refer to [1] Some authors define G 12 (T) as having an extra minus sign and the retarded Green function with a factor - i, so that the relation is more direct: G{l2(W) = G 12 (W m = - iw + 7]). See

1366

18 Nonequilibrium Quantum Statistics A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski, Sov. Phys. JETP 9, 636 (1959); or Methods of Quantum Field Theory in Statistical Physics , Dover, New York, 1975; also A.L. Fetter and J .D. Walecka, Quantum Theory of Many-Panicle Systems , McGraw-Hill, New York, 1971. Our definition (18.20) without a minus sign conforms with the definition of the fixed-energy amplitude in Chapter 9, Eq. (1.321), which is also a retarded Green function.

[2] E .S. Fradkin, The Green's Function Method in Quantum Statistics, Sov. Phys. JETP 9, 912 (1959) . [3] G. Baym and D. Merrnin, J. Math. Phys. 2, 232 (1961). One extrapolation uses Pade approximations: H.J. Vidberg and J.W. Serene, J . Low Temp. Phys. 29, 179 (1977); W.H. Press, S.A. Teukolsky, W .T. Vetterling, and B.P. Flannery, Numerical Recipes in Fortmn, Cambridge Univ. Press (1992) , Chapter 12.5. Since the thermal Green function are usually known only approximately, the continuation is not unique. A maximal-entropy method by RN. Silver, D.S. Sivia, and J.E. Gubernatis, Phys. Rev. B 41, 2380 (1990) selects the most reliable result. [4] J . Schwinger, J. Math. Phys. 2, 407 (1961) . [5] KV. Keldysh, Z. Eksp. Teor. Fiz. 47, 1515 (1964); Sov. Phys. JETP 20, 1018 (1965). [6] K-C. Chou, Z.-B. Su, B.-L. Hao, and L. Yu, Phys. Rep. 118, 1 (1985); also K-C. Chou et al., Phys. Rev. B 22, 3385 (1980). [7] RP. Feynman and F .L. Vernon, Ann. Phys. 24, 118 (1963); RP. Feynman and A.R Hibbs, Quantum Mechanics and Path Integmls , McGraw-Hill, New York, 1965, Sections 12.8 and 12.9. [8] The solution of path integrals with second time derivatives in the Lagrangian is given in H. Kleinert, J. Math. Phys. 27, 3003 (1986) (http://www . physik. fu-berlin. derkleinert/144). [9] H. Kleinert, Gauge Fields in Condensed Matter, op. cit., Vol. II, Section 17.3 (ibid.http/b2), and references therein. [10] See the review paper by S. Chandrasekhar, Rev. Mod. Phys. 15, 1 (1943) , or the original papers: A.A. Fokker, Ann. Phys. (Leipzig) 43, 810 (1914) ; M. Planck, Sitzber. Preuss. Akad. Wiss. p. 324 (1917); O. Klein, Arkiv Mat. Astr. Fysik 16, No. 5 (1922); H.A. Kramers, Physica 7, 284 (1940) ; M. Smoluchowski, Ann. Phys. (Leipzig) 48 , 1103 (1915). [11] H. Kleinert, A. Chervyakov, Phys. Lett. B 464, 257 (1999) (hep-thj9906156); Phys. Lett. B 477, 373 (2000) (quant-phj9912056) ; Eur. Phys. J . C 19, 743-747 (2001) (quantphj0002067); Phys. Lett. A 273, 1 (2000) (quant-phj0003095); Int. J. Mod. Phys. A 17, 2019 (2002) (quant-phj0208067) ; Phys. Lett. A 308,85 (2003) (quant-phj0204067); Int. J. Mod. Phys. A 18, 5521 (2003) (quant-phj0301081) . [12] In a first attempt to show that this functional the determinant is unity, A. Schmid, J. Low Temp. Phys. 49, 609 (1982). contrived a suitable time slicing of the differential operator in (18.326) to achieve this goal. However, since this was not derived from a time slicing of the initial forward-backwards path integral (18.230), this procedure cannot be considered as a proof. [13] The operator-ordering problem was first solved by H. Kleinert, Ann. of Phys. 291, 14 (2001) (quant-phj0008109).

1367

Notes and References [14] R. Benguria and M. Kac, Phys. Rev. Lett. 46, 1 (1981) ; Y.C. Chen, M.P.A. Fisher and A.J . Leggett, J . App!. Phys. 64, 3119 (1988) . [15] G.W. Ford, M. Kac, and P. Mazur, J . Math. Phys. 6, 504 (1965) . G.W. Ford and M. Kac, J. Stat. Phys. 46, 803 (1987) . [16] P. Langevin, Comptes Rendues 146, 530 (1908). [17] G. Parisi and Y.S. Wu, Scientia Sinica 24,483 (1981) .

[18] R. Kubo, J .Math.Phys. 4, 174 (1963) ; R. Kubo, M. Toda, and N. Hashitsume, Statistical Physics II (Nonequilibrium Statistical Mechanics), Springer-Verlag, Berlin, 1985 (Chap. 2) . [19] J . Zinn-Justin, Critical Phenomena , Clarendon, Oxford, 1989. [20] G. Parisi and N. Sourlas, Phys. Rev. Lett. 43, 744 (1979); J . de Phys. 41 , L403 (1981); Nuc!. Phys. B 206, 321 (1982); [21] A.J . McKane, Phys. Lett. A 76 , 22 (1980) . [22] H. Kleinert and S. Shabanov, Phys. Lett. A 235, 105 (1997) (quant-ph/9705042) . [23] E. Gozzi, Phys. Lett. B 201, 525 (1988). [24] A.O. Caldeira and A.J. Leggett, Physica A 121, 587 (1983); A 130 374(E) (1985). [25] More on this subject is found in the collection of articles D. Giulini, E. Joos, C. Kiefer , J. Kupsch, 1.0. Stamatescu, H.D. Zeh, Decoherence and the Appearance of a Classical World in Quantum Theory, Springer, Berlin, 1996. [26] G. Lindblad, Comm. Math. Phys. 48, 119 (1976) . This paper shows that the form (18.534) of the master equation (18.524) guarantees the positivity of the probabilities derived from the solutions. The right-hand side can be more generally 1

1

L:mn hmn ( 'iLmLnP + 'ipLmLn - LnpLm A

A

A

A

A

A

)

+ h.c..

[27] L. Diosi, Europhys. Lett. 22 , 1 (1993) . [28] C.W. Gardiner, IBM J. Res. Develop. 32, 127 (1988). [29] H. Kleinert and S. Shabanov, Phys. Lett. A 200, 224 (1995) (quant-ph/9503004) ; K. Tsusaka, Phys. Rev. E 59, 4931 (1999). [30] H.A. Bethe, Phys. Rev. 72, 339 (1947). [31] C. Cohen-Tannoudji, J. Dupont-Roc, G. Grynberg, Photons and Atoms: Introduction to Quantum Electrodynamics , Wiley, New York, 1992. [32] R. Rohrlich, Am. J . Phys. 68, 1109 (2000). [33] H. Kleinert and S. Shabanov, J . Phys. A: Math. Gen. 31, 7005 (1998) (cond-mat/9504121); R. Bausch, R. Schmitz, and L.A. Turski, Phys. Rev. Lett. 73, 2382 (1994); Z. Phys. B 97, 171 (1995) . [34] M. Roncadelli, Europhys. Lett. 16, 609 (1991) ; J . Phys. A 25, L997 (1992) ; A. Defendi and M. Roncadelli, Europhys. Lett. 21, 127 (1993) . [35] Z. Haba, Lett. Math. Phys. 37, 223 (1996) . [36] G. 't Hooft, Found. Phys. Lett. 10, 105 (1997) (quant-ph/9612018). [37] G. 't Hooft, How Does God Throw Dice? in Fluctuating Paths and Fields - Dedicated to Hagen Kleinert on the Occasion of his 60th Birthday, Eds. W. Janke, A. Pelster, H.-J. Schmidt, and M. Bachmann, World Scientific, Singapore, 2001 (http://www . physik.fu-berlin.de/-kleinert/fest.html).

Agri non omnes frugiferi sunt.

Nat all fields are fruitful. Tuse. Quaest., 2, 5, 13

CICERO,

19 Relativistic Particle Orbits Particles moving at large velocities near the speed of light are called relativistic particles. If such particles interact with each other or with an external potential, they exhibit quantum effects which cannot be described by fluctuations of a single particle orbit. Within short time intervals, additional particles or pairs of particles and antiparticles are created or annihilated, and the total number of particle orbits is no longer invariant. Ordinary quantum mechanics which always assumes a fixed number of particles cannot describe such processes. The associated path integral has the same problem since it is a sum over a given set of particle orbits. Thus, even if relativistic kinematics is properly incorporated, a path integral cannot yield an accurate description of relativistic particles. An extension becomes necessary which includes an arbitrary number of mutually linked and branching fluctuating orbits. Fortunately, there exists a more efficient way of dealing with relativistic particles. It is provided by quantum field theory. We have demonstrated in Section 7.14 that a grand-canonical ensemble of particle orbits can be described by a functional integral of a single fluctuating field. Branch points of newly created particle lines are accounted for by anharmonic terms in the field action. The calculation of their effects proceeds by perturbation theory which is systematically performed in terms of Feynman diagrams with calculation rules very similar to those in Section 3.18. There are again lines and interaction vertices, and the main difference lies in the lines which are correlation functions of fields rather than position variables x(t). The lines and vertices represent direct pictures of the topology of the worldlines of the particles and their possible collisions and creations. Quantum field theory has been so successful that it is generally advantageous to describe the statistical mechanics of many completely different types of line-like objects in terms of fluctuating fields. One important example is the polymer field theory in Section 15.12. Another important domain where field theory has been extremely successful is in the theory of line-like defects in crystals, superfluids, and superconductors. In the latter two systems, the defects occur in the form of quantized vortex lines or quantized magnetic flux lines, respectively. The entropy of their classical shape fluctuations determines the temperature where the phase transitions take place. Instead of the usual way of describing these systems as ensembles of particles with their interactions, a field theory has been developed whose Feynman 1368

1369 diagrams are the direct pictures of the line-like defects, called disorder field theory

[1]. The most important advantage of field theory is that it can describe most easily phase transitions, in which particles form a condensate. The disorder theory is therefore particularly suited to understand phase transitions in which defect-, vortex-, or flux-lines proliferate, which happens in the processes of crystal melting, superfluid to normal, or superconductor to normal transitions, respectively. In fact, the disorder theory is so far the only theory in which the critical behavior of the superconductor near the transition is properly understood [2]. A particular quantum field theory, called quantum electrodynamics describes with great success the electromagnetic interactions of electrons, muons, quarks, and photons. It has been extended successfully to include the weak interactions among these particles and, in addition, neutrinos, using only a few quantized Dirac fields and a quantized electromagnetic vector potential. The inclusion of a nonabelian gauge field, the gluon field, is a good candidate for explaining all known features of strong interactions. It is certainly unnecessary to reproduce in an orbital formulation the great amount of results obtained in the past from the existing field theory of weak, electromagnetic, and strong interactions. The orbital formulation was, in fact, proposed by Feynman back in 1950 [3], but never pursued very far due to the success of quantum field theory. Recently, however, this program was revived in a number of publications [4, 5]. The main motivation for this lies in another field of fundamental research: the string theory of fundamental particles. In this theory, all elementary particles are supposed to be excitations of a single line-like object with tension, and various difficulties in obtaining a consistent theory in the physical spacetime have led to an extension by fermionic degrees of freedom, the result being the so-called superstring. Strings moving in spacetime form worldsurfaces rather than worldlines. They do not possess a second-quantized field theoretic formulation. Elaborate rules have been developed for the functional integrals describing the splitting and merging of strings. If one cancels one degree of freedom in such a superstring, one has a theory of splitting and merging particle worldlines. As an application of the calculation rules for strings, processes which have been known from calculations within the quantum field theory have been recalculated using these reduced superstring rules. In this textbook, we shall give a small taste of such calculations by evaluating the change in the vacuum energy of electromagnetic fields caused by fluctuating relativistic spinless and spin-1/2 particles. It should be noted that since up to now, no physical result has emerged from superstring theory,l there is at present no urgency to dwell deeper into the subject. By giving a short introduction into this subject we shall be able to pay tribute to some historic developments in quantum mechanics, where the relativistic generalization of the Schrodinger equation was an important step towards the development of IThis theory really deserves a price for having the highest popularity-per-physicality ratio in the history of science, enjoying a great amount of financial support. The situation is very similar to the geocentric medieval picture of the world.

1370

19 Relativistic Particle Orbits

quantum field theory [6]. For this reason, many textbooks on quantum field theory begin with a discussion of relativistic quantum mechanics. By analogy, we shall incorporate relativistic kinematics into path integrals. It should be noted that an esthetic possibility to give a path Fermi statistics is based on the Chern-Simons theory of entanglement of Chapter 16. However, this approach is still restricted to 2 + 1 spacetime dimensions [7], and an extension to the physical 3 + 1 spacetime dimensions is not yet in sight.

19.1

Special Features of Relativistic Path Integrals

Consider a free point particle of mass M moving through the 3 + 1 spacetime dimensions of Minkowski space at relativistic velocity. Its path integral description is conveniently formulated in four-dimensional Euclidean spacetime where the fluctuating worldlines look very similar to the fluctuating polymers discussed in Chapter 15. Thus, time is taken to be imaginary, i.e., t =

. = -zx . 4/ e,

-ZT

(19.1)

and the length of a four-vector x = (x, x 4 ) is given by x 2 = x2

+ (X 4 )2

+ e2 T2.

= x2

(19.2)

If XJL(A) is an arbitrarily parametrized orbit, the well-known classical Euclidean action is proportional to the invariant length of the orbit in spacetime:

(19.3) and reads Acl,e

= MeS,

(19.4)

or, explicitly, (19.5) with (19.6) The prime denotes the derivative with respect to the parameter A. The action is independent of the choice of the parametrization. If A is replaced by a new parameter

A -+

,\ =

f(A),

X'2

-+

f'2

dA

-+

dA 1',

(19.7)

then 1

12

x ,

(19.8) (19.9)

19.1 Special Features of Relativistic Path Integrals

1371

so that ds and the action remain invariant. We now calculate the Euclidean amplitude for the worldline of the particle to run from the spacetime point Xa = (xa, CTa ) to Xb = (Xb , CTb). For the sake of generality, we treat the case of D Euclidean spacetime dimensions. Before starting we observe that the action (19.5) does not lend itself easily to a calculation of the path integral over e-Ac1.e/ li . There exists an alternative form for the classical action that is more suitable for this purpose. It involves an auxiliary field h()") and reads:

r>'b A" = JAa d)"

[MC 12() ()MC] . 2h()") x ).. + h ).. 2

(19.10)

This has the advantage of containing the particle orbit quadratically as in a free nonrelativistic action. The auxiliary field h()") has been inserted to make sure that the classical orbits of the action (19.10) coincide with those of the original action (19.5). Indeed , extremizing A in h()") gives the relation

h()") = VXI2()..). Inserting this back into

A

(19.11)

renders the classical action

Al,e

= Mc

l:

b

d)..VX I2 ()..),

(19.12)

which is the same as (19.5). At this point the reader will worry that although the new action (19.10) describes the same classical physics as the original action (19.12), it may lead to a completely different quantum physics of a relativistic partile. With a little effort, however, it can be shown that this is not so. Since the proof is quite technical, it will be given in Appendix 19A. The new action (19.10) shares with the old action (19.12) the reparametrization invariance (19.7) for arbitrary fluctuating path configurations. We only have to assign an appropriate transformation behavior to the extra field h()"). If).. is replaced by a new parameter>" = f()..) , then X'2 and d)" transform as in (19.8) and (19.9) , and the action remains invariant, if h()") is simultaneously changed as h

--+

hi1'.

(19.13)

We now set up the path integral of a relativistic particle associated with the action (19.10). First we sum over the orbital fluctuations at a fixed h()..). To find the correct measure of integration, we use the canonical formulation in which the Euclidean action reads

AlP, xl =

l:

b

d)" [-iPX l +

~~~ (p2 + M 2c2)] .

(19.14)

This must be sliced in the length parameter)... We form N + 1 slices as usual, choosing arbitrary small parameter differences En = )..n - )..n-l depending on n, and write the sliced action as (19.15)

1372

19 Relativistic Particle Orbits

This path integral has a universal phase space measure [recall (2.28)]

The momentum variables Pn are integrated out to give the configuration space integrals (setting )1N+1 == Ab, hN+1 == hb) [compare (2.74)] (19.17)

with the time-sliced action in configuration space (19.18) The Gaussian integrals over Xn in (19.17) can now be done successively using Formula (2.69), and we find [as in (2.75)] 1

exp [_MC(Xb- Xa)2 _MC L ]

J27r'hL/M/

2'h

L

(19.19)

2'h'

where L is the total sliced length of the orbit N+l

L

==

L

Enhn,

(19.20)

n=l

whose continuum limit is (19.21) The result (19.19) does not depend on the function h(A) but only on L , this being a consequence of the reparametrization invariance of the path integral. While the total A-interval changes under the transformation, the total length L of (19.21) is invariant under the joint transformations (19.7) and (19.13) . This invariance permits only the invariant length L to appear in the integrated expression (19.19), and the path integral over h(A) can be reduced to a simple integral over L. The appropriate path integral for the time evolution amplitude reads (19.22) where N is some normalization factor and [h] an appropriate gauge-fixing functional.

1373

19.1 Special Features of Relativistic Path Integrals

19.1.1

Simplest Gauge Fixing

The simplest choice for the latter is a 8-functional,

[h] = 8[h - c] ,

(19.23)

which fixes h(A) to be equal to the light velocity everywhere, and relates (19.24) This relation makes the dimension of the parameter A timelike. This Lorentzinvariant time parameter is the so-called proper time of special relativity, and should not be confused with the parameter time T contained in the Dth component x D = CT [see Eq. (19.1) for D = 4]. By analogy with the discussion of thermodynamics in Chapter 2 we shall then denote Ab - Aa by h/3 and write (19.24) as (19.25) If we further use translational invariance to set Aa path integral

= 0, we arrive at the gauge-fixed (19.26)

with

(,,(3 M .2 Ao,e = Jo dA 2 X.

(19.27)

Since A is now timelike, we use a dot to denote the derivative: X(A) == dX(A)/dA. Remarkably, the gauge-fixed action coincides with the action of a free nonrelativistic particle in D Euclidean spacetime dimensions. Having taken the trivial term Jon(3 dA M c/2h out of the action (19.14) , the expression (19.26) contains a Boltzmann weight e- (3Mc 2 /2 multiplying each particle orbit of mass M. The solution of the path integral is then given by (19.28) By Fourier-transforming the x-dependence, this amplitude can also be written as (19.29) and evaluated to (19.30)

1374

19 Relativistic Particle Orbits

Upon setting N =

AXt/2, where A~

== n/Mc,

(19.31)

is the well-known Compton wavelength of a particle of mass M [recall Eq. (4.345)] , this becomes the Green function of the Klein-Gordon field equation in Euclidean time: (19.32) In the Fourier representation (19.30), the integral over k [or the integral over in (19.28)] can be performed with the explicit result for the Green function

f3

(19.33) where Ky(z) denotes the modified Bessel function and x == Xb - Xa' In the nonrelativistic limit c -+ 00, the asymptotic behavior Ky(z) -+ V7r/2ze- z [see Eq. (1.353)] leads to (19.34) with the usual Euclidean time evolution amplitude of the free Schrodinger equation (19.35)

The exponential prefactor in (19.34) contains the effect of the rest energy M c2 which is ignored in the nonrelativistic Schrodinger theory. Note that the same limit may be calculated conveniently in the saddle point approximation to the f3-integral (19.28) . For c -+ 00 , the exponent has a sharp extremum at

and the f3-integral can be evaluated in a quadratic approximation around this value. This yields once again (19.34).

19.1.2

Partition Function of Ensemble of Closed Particle Loops

The diagonal amplitude (19.26) with Xb = Xa contains the sum over all lengths and shapes of a closed particle loop in spacetime. This sum can be made a partition function of a closed loop if we remove a degeneracy factor proportional to 1/ L from the integral over L. Then all cyclic permutations of the points of the loop are

1375

19.1 Special Features of Relativistic Path Integrals

counted only once. Apart from an arbitrary normalization factor to be fixed later, the partition function of a single closed loop reads Zl =

10

~ e-{3Mc2 /2

00

J

Inserting the right-hand integral in (19.29) for the path integral (with becomes

roo

d(3 Z = v: 1 D Jo (3 e

{3M c 2 /2

(19.37)

1JD xe -Ao.e/li.

Xb

=

J(27r)D d Dk (_(31i? k 2 ) exp 2M '

xa), this

(19.38)

where VD is the total volume of spacetime. This can be evaluated immediately. The Gaussian integral gives for each of the D dimensions a factor 1h/27rli,2(3/M, after which formula (2.496) leads to

_ looo -(3 d(3 e

Zl - VD

-{3Mc 2 /2

o

_ VD 1 DD ( )D/2 r (1- D/2), V27rli,2(3/M A~ 411' 1

(19.39)

where A~ is the Compton wavelength (19.31). With the help of the sloppy formula (2.504) of analytic regularization which implies the minimal subtraction explained in Subsection 2.15.1, the right-hand side of (19.38) can also be written as Zl = -VD

J

dDk ( 2 2 2/ 2) (27r)D log k + Me li, .

(19.40)

The right hand side can be expressed in functional form as (19.41) the two expressions being equal in the analytic regularization of Section 2.15, since a constant inside the logarithm gives no contribution by Veltman's rule (2.506). The partition function of a grand-canonical ensemble is obtained by exponentiating this: (19.42) In order to interprete this expression physically, we separate the integral into an integral over the temporal component kD and a spatial remainder, and write

J dDk/(27r)D

(19.43) with the frequencies (19.44)

1376

19 Relativistic Particle Orbits

Recalling the result (2.503) of the integral (2.489) we obtain (19.45) The exponent is the sum of two ground state energies of oscillators of energy nWk/2, which are the vacuum energies associated with two relativistic particles. In quantum field theory one learns that these are particles and antiparticles. Many neutral particles are identical to their antiparticles, for example photons, gravitons, and the pion with zero charge. For these, the factor 2 is absent. Then the integral (19.38) contains a factor 1/2 accounting for the fact that paths running along the same curve in spacetime but in the opposite sense are identified. Comparing (19.42) with (3.556) and (3.619) for j = 0 we identify -Zln with -W[O] and the Euclidean effective action f of the ensemble of loops: -Zl

19.1.3

= -W[Ol/n = fe/n.

(19.46)

Fixed-Energy Amplitude

The fixed-energy amplitude is related to (19.22) by a Laplace transformation: (19.47) where Tb, Ta are once more the time components in Xb , Xa' As explained in Chapter 9, the poles and the cut along the energy axis in this amplitude contain all information on the bound and continuous eigenstates of the system. The fixed-energy amplitude has the reparametrization-invariant path integral representation [here with the conventions of Eq. (19.10)] (19.48) with the Euclidean action (19.49) To prove this, we write the temporal XD -part of the sliced D-dimensional action (19.18) in the canonical form (19.15). In the associated path integral (19.16), we integrate out all x;;-variables, producing N 8-functions. These remove the integrals over N momentum variables p;;, leaving only a single integral over a common pD. The Laplace transform (19.47) , finally, eliminates also this integral making pD equal to -iE/c. In the continuum limit, we thus obtain the action (19.49) . The path integral (19.48) forms the basis for studying relativistic potential problems. Only the physically most relevant example will be treated here.

1377

19.2 Tunneling in Relativistic Physics

19.2

Tunneling in Relativistic Physics

Relativistic harbors several new tunneling phenomena, of which we want to discuss two especially interesting ones.

19.2.1

Decay Rate of Vacuum in Electric Field

In relativistic physics, an empty Minkowski space with a constant electric field E is unstable. There is a finite probability that a particle-antiparticle pair can be created. For particles of mass M, this requires the energy (19.50) This energy can be supplied by the external electric field. If the pair of charge ±e is separated by a distance which is roughly twice the Compton wavelength AXt- = n/Mc of Eq. (19.31), it gains an energy 2IEIAXt-e. The decay will therefore become significant when (19.51)

Euclidean Action In Chapter 17 we have shown that in the semiclassical limit where the decay-rate is small, it is proportional to a Boltzmann-like factor e- Ac1,eln, where Al,e is the action of a Euclidean classical solution mediating the decay. Such a solution is easily found. We use the classical action in the form (19.12) and choose the parameter A to measure the imaginary time A = 7 = it = X4 / c. Then the action takes the form (19.52) This is extremized by the classical equation of motion

M~ d7

X(7)

VI + x2(7)/C2

= -eE,

(19.53)

whose solutions are circles in spacetime of a fixed E-dependent radius lE:

(X

-

xo )2 + c2( 7 - 70 )2

_ (Mc2)2 = lE2 = eE '

E=IEI·

To calculate their action we parametrize the circles in the the unit vector in the direction of E, by an angle 8 as x(8)

=

lEE cos8+xo,

7(8)

E-

= lE sin8+70 . c

(19.54)

7 -plane, where

E is

(19.55)

1378

19 Relativistic Particle Orbits

A closed circle has an action

1cos() ] = MclEn: = nEECn:.

Acle, = Mc 2 -IEc127T d()cos() [--() ocos

(19.56)

The decay rate of the vacuum is therefore proportional to (19.57) The circles (19.54) are, of course, the space-time pictures of the creation and the annihilation of particle-antiparticle pairs at times TO - IE/c and TO + IE/c and positions XQ, respectively (see Fig. 19.1) . A particle can also run around the circle

(xo, CTo -IE)

Figure 19.1 Spacetime picture of pair creation at the point xo and time annihilation at the later time TO + lE/C,

TO - lE/C

and

repeatedly. This leads to the formula 00

n7TEc/ E rex:'" L..J F n e,

(19.58)

n=l

with fluctuation factors Fn which we are now going to determine. Fluctuations

As explained above, fluctuations must be calculated with the relativistic Euclidean action (19.10), in which we have to include the electric field by minimal coupling: (19.59) The vacuum will decay by the creation of an ensemble of pairs which in Euclidean spacetime corresponds to an ensemble of particle loops. For free particles, the partition function Z was given in (19.42) as an exponential of the one-loop partition function Zl in (19.37). The corresponding Zl in the presence of an electromagnetic field has the one-loop partition function

= (00 dj3 e- f3Mc2/2jV4xe - .7.e/h

Z 1

io

j3

,

(19.60)

1379

19.2 Tunneling in Relativistic Physics

with the Euclidean action (19.61) The equations of motion are now

x" = ~ Mc (Xl4 E - i

Xl X

x"=-~x/·E

B) ,

Mc

4

(19.62)

'

If both a constant electric and a constant magnetic field are present, the vector potential is A/L = -F/Lvxv /2 and the action (19.61) takes the simple quadratic form A

""e

=

lon{3 0

d)"

[M2

_X I2

- i -eF

2c

/LV

x/L X IV ]

(19.63)

,

and the equations of motion (19.62) are (19.64) In a purely electric field, the solutions are circular orbits: (19.65) with the electric version of the Landau or cyclotron frequency (2.646) E_

wL =

eE Mc '

(19.66)

The circular orbits are the same as those in the previous formulation (19.55) . If E points in the z-direction, the action (19.63) decomposes into two decoupled quadratic actions Ai12 ) + Ai34 ) for the motion in the Xl - x 2 and x 3 - X4 -planes, respectively, and the one-loop partition function (19.60) factorizes as follows: dfJ - {3Mc Joroo 7i e

1~ 00

The path integral for the trivial action

e

J

",,2 (12) _ v X e

.4(12) e

In

J

",,2 (34) V

x

_ .4.\34)

e

In

e-{3Mc2/2Z(12) (0)Z(34) (E).

Z(12)(0)

A(12)

2/2

=

collecting the fluctuations in the

rn{3 d)" M2 (Xl

Jo

1

2

(19.67) Xl - x 2

-plane have

+ Xl 2)

2'

with the trivial fluctuation determinant Det (-aD = 1, so that we obtain for the free-particle partition function in two dimensions

(19.68) Z(12) (0)

(19.69)

1380

19 Relativistic Particle Orbits

The factor ~XI~X2 is the total area of the system in the Xl - x 2 -plane. Note that upper and lower indices are the same in the present Euclidean metric. For the motion in the x 3 - X4 -plane, the quadratic fluctuations with periodic boundary conditions have a functional determinant

-8~

Det ( iwf8;.

-iwf

8;.) = Det (2) (2 E2) sinhwf1i/3 -8;. xDet -8;. +wL = 1 x wf1i/3 ,(19.70)

-8~

leading to the partition function becomes Z(34)

(E) =

~ ~ X3

X4

JM

27r1i2 /3

wf1i/3 wf1i/3 .

2 sinh

(19.71)

This result can, of course, also be obtained without calculation from the observation that the Euclidean electric path integral is completely analogous to the real-time magnetic path integral solved in Section 2.18. Indeed, with the E-field pointing in the z-direction, the action (19.63) becomes for the motion in the X3 - X4 -plane (19.72)

This coincides with the magnetic action (2.633) in real time, if we insert there the magnetic vector potential (2.634) and replace B by E. The equations of motion (19.62) reduces to (19.73)

in agreement with the magnetic equations of motion (2.670) in real time. Thus the motion in the X3 - X4 -plane as a function of the pseudotime A is the same as the real-time motion in the x - y -plane, if the magnetic field B in the z-direction is exchanged by an E-field of the same size pointing in the x3-axis. The amplitude can therefore be taken directly from (2.666) - we must merely replace the real time difference tb - ta by 1i/3. Inserting (19.69) and (19.71) into (19.67), we obtain the partition function of a single closed particle orbit in four Euclidean spacetime dimensions: (19.74)

where V == ~XI~X2~X3 is the total spatial volume. We now go over to real times by setting ~X4 = ic~t. By exponentiating the subtracted expression (19.74) as in Eq. (19.42), we go to a grand-canonical ensemble, and may identify Zl with i times the effective electromagnetic action caused by fluctuating ensemble of particle loops (19.75)

The integral over /3 in (19.74) is divergent. In order to make it convergent, we perform two subtractions. In the subtracted expression we change the integration

1381

19.2 Tunneling in Relativistic Physics

variable to the dimensionless ( = density elf _

f1.c -

ftC

13M c2 /2,

and obtain the effective Lagrangian

(MC)4 1 ['Xl d( [ E(/Ec T 4(27r)2 io (3 sin E(/ Ec -

1-

1 (E)2] _( Ec e.

(19.76)

6

The first subtraction has removed the divergence coming from the l/e-singularity in the integrand. This produces a real infinite field-independent contribution to the effective action which can be omitted since it is unobservable by electromagnetic experiments.2 After subtracting this divergence, the integral still contains a logarithmic divergence. It can be interpreted as a contribution proportional to E2 to the Lagrangian density

(Mc)4

elf_

f1.c div -

ftC

T

1 (E)2 {ood( _(_ a (ood( _( 24(27r)2 Ec io , e - 247r io , e ,

(19.77)

which changes the original Maxwell term E 2 /2 to Z A E 2 /2 with (19.78) According to the rules of renormalization theory, the prefactor is removed by renormalizing the field strength, replacing E ---+ E / Z:/.2, and identifying the replaced field with the physical, renormalized field E/Z:/.2 == E R . Due to the presence of the affective action, the vacuum is no longer time independent, put depends on time like c i (H - irTij2)6..tjTi . Thus the decay rate of the vacuum per unit volume is given by the imaginary part of the effective Lagrangian density (19.79)

In order to calculate this we replace E(/ Ec by z in (19.76), and expand in the integrand z

Z2

00

-.- = 1 + 2 2]-1)n 2 SIn

z

n=l

Z

-

2 2 n 7r

(2

00

= 2 L(-l)n~, n=l

'>

- '>n

E

(n == n7r EC

'

(19.80)

Adding to the poles the usual infinitesimal i1J-shifts in the complex plane (see p. 115), we replace with the help of the decomposition (1.325) (19.81) 2This energy would, however, be observable in the cosmological evolution to be discussed in Subsection 19.2.3.

1382

19 Relativistic Particle Orbits

The o-functions yield imaginary parts and thus directly the decay rate

r V (19.82) The principal values produce a real effective Lagrangian density

If we expand (19.84) and perform the integrals over (, we find fj.£eff

Mc)4

= nc ( h

1 [7 (E)4 4(27r)2 360 Ec

31 (E)6

+ 2520 Ec

]

+ . .. ,

(19.85)

or

(19.86) The subscript P of £ has been omitted since the imaginary part of the effective Lagrangian density (19.82) has a vanishing Taylor expansion. Each coefficient in (19.99) is exact to leading order in a. The extra expansion terms in (19.99) imply that the physical vacuum has a nontrivial E-dependent dielectric constant E(E). This is caused by the virtual creation and annihilation of particle-antiparticle pairs. Since the dielectric displacement D(E) is obtained from the first derivative of £eff, the dielectric constant is given by E(E) = D(E)j E = 8£eff j E8E. From (19.99) we find the lowest expansion terms

7a 2 (nc)3 2 3l7ra3 [ (nc)3 ] 2 4 E(E) = 1+ 90 (MC 2 )4E +-----w-5 (MC 2 )4 E + ....

(19.87)

Such a term gives rise to a small amplitude for photon-photon scattering in the vacuum, a process which has been observed in the laboratory.

Including Constant Magnetic Field parallel to Electric Field Let us see how the decay rate (19.82) and the effective Lagrangian (19.76) are influenced by the presence of an additional constant B-field. This will at first be

1383

19.2 Tunneling in Relativistic Physics

assumed to be parallel to the E-field, with both fields pointing in the z-direction. Then the action (19.68) for the motion in the Xl - x 2 -plane becomes (19.88) The partition function in the Xl - x 2 -plane will therefore have the same form as (19.71), except with wf replaced by iwf:

Z(12)(B) = Z(34)(iB).

(19.89)

Thus the B-field changes the partition function (19.74) of a single closed orbit to

Zl =

~X4 V

('Xl dj3

io

j3

J27rnM2

4

wfnJ3/2

wfnJ3/2

e-f3Mc2/2

j3 sinwfnJ3/2 sinhwfnJ3/2

(19.90)

'

and the effective Lagrangian (19.76) becomes

~£eff = nc (Mc)4 _1_

n

4(27r)2

roo d(

[ E(/ Ec B(/ Ec _ 1 _ (E2 - B2)(2] e- ( . (3 sinE(/EcsinhB(/Ec 6EJ (19.91)

io

In the subtracted free-field term (19.77), the term E2 is changed to the Lorentzinvariant combination E2 - B2. The decay rate (19.82) is modified to

r =3.Im~£eff = c (Mc)4(~)2_1_ ~ f(-l)n-l~ n7rB/E e-mrEciE (19 .92)

V

n

n

Ec

47r 3 2 n=l

n 2 sinh n7rB/ E

.

Including Constant Magnetic Field in any Direction The case of a constant magnetic field pointing in any direction can be reduced to the parallel case by a simple Lorentz transformation. It is always possible to find a Lorentz frame in which the fields become parallel. This special frame will be called center-oj-fields frame, and the transformed fields in this frame will be denoted by BCF and E CF . The transformation has the form E CF = "f (E + ~ x B) - L~ c "f +1 c BCF

= "f (B -

(~. E), C

(19.93)

~c x E) - "fL~ (~ .B) , +1c c

(19.94)

with a velocity of the transformation determined by

vic 1 + (Ivl/c)2

ExB IEI2 + IBI2 '

(19.95)

and "f == [1 - (Ivl/c)2tl/2. The fields IEcFI == c and IBcFI == j3 are, of course, Lorentz-invariant quantities which can be expressed in terms of the two quadratic

1384

19 Relativistic Particle Orbits

Lorentz invariants of the electromagnetic field: the scalar S and the pseudoscalar P defined by 1

-

P == --FJJvFJJv = EB = s/3. 4 (19.96) Solving these equation yields

{;} == /JS2 + p2 ± S

=

~J V(E2 -

B2)2 + 4(EB)2 ± (E2 - B2).

(19.97)

As a result we find that the formulas (19.91) and (19.92) are valid for arbitrary constant fields E and B if we replace E and B by the Lorentz invariants sand /3. After this we may expand the integrand in (19.91) in powers of sand /3 using Eq. (19.84) ,

4 (4

eST = -1 - -e2 ( s 2 - /32) + e - T 7c -lOs2/3 2 + 7/3 T3 sin e/3T sinh eST T3 6T 360 T3 - e6 _ _ (31s6 - 49s 4 /3 2 + 49s2/34 - 31/36) +... 1520

e/3T -1 -

4) (19.98)

and obtain the effective Lagrangian the fields generalizing (19.99):

.ceff = ~ {(E2 _ B2) 2

+

7a 2 (fic )3 (E2 _ B2)2

+ 180 (M C2 )4

3~~;3 [(~~~~4] 2 (E2 _ B2)[2(E2 _ B2)2 _ 4(EB?l + . . . },

(19.99)

For strong fields , a saddle point approximation to the generalized integral (19.91) yields the asymptotic form

.ceff == _~(E2 _ B2) log [-4e 2 (fiC)3 2 2 1927r

(MC )4

(E2 _ B2)]

+ ...

(19.100)

Spin-1/2 Particles We anticipate the small modification which is necessary to obtain the analogous result for spin-1/2 fermions: The tools for this will be developed in Subsections 19.5.119.5.3. The relevant formula has actually been derived before in Eq. (7.512). The bosonic result receives for fermions a factor -1 and a fluctuation factor which is the square root of the functional determinant (19.101) The normalization factor follows from Eqs. (7.415) and (7.418), where we found that the path integral of single complex fermion field carries a normalization factor 2. For a purely electric field , the square-root of the right-hand side of (19.101)

1385

19.2 Tunneling in Relativistic Physics

is 2det 1/2 [cos(e/Mc)Fl'v] = 2cos(eE/Mc). Multiplying the cos-factor into the expansion (19.80) , this is modified to cos z

Z2 00 (2 -2~ Z2 n21T2 ~ /'2 _ /'2 ' n=l n=l~ ~n 00

z-=1+2~ sin z ~

(19.102)

Performing now the singular integral over ( in (19.76), we obtain for the decay rate the same formula as in (19.82), except that the alternating signs are absent. The factor 4 of the fermionic determinant reduces to a factor 2 for the effective action. The resulting effective Lagrangian density for spin-l/2 fermions is therefore elf

~.csPin!

_ -

-nc

(Mc)4

T

1

2(271')2

{ood( (

Jo

(3

E(/Ec

E2(2) _(

tan E(/ Ec - 1 + 3E~

e

, (19.103)

a result first derived by Heisenberg and Euler in 1935 [8]. Its imaginary part yields the decay rate of the vacuum due to pair creation

(19.104) The reason for the reduction from 4 to 2 is that the sum over bosonic paths has to be first divided by a factor 2 to remove their orientation, before the fermionic factor 4 is applied. This procedure is not so obvious at this point but will be understood later in Subsection 19.5.2. The remaining factor 2 accounts for the two spin orientations of the charged particles. The Taylor series

Z Z2 1 4 2 6 --=I----z --z tanz

3

45

945

(19.105)

in the integrand of (19.103) leads to the expansion

2 [145 (E)4 4 (E)6 ] Ec + 315 Ec + . .. ,

Mc)4 ~.celf = nc ( T 1671'2

(19.106)

so that (19.107) The term proportional to a 2 -term implies a small amplitude for photon-photon scattering which can be observed in the laboratory [9].

1386

19 Relativistic Particle Orbits

As in the boson result (19.99), each coefficient is exact to leading order in a, and the virtual creation and annihilation of fermion-antifermion pairs gives the physical vacuum a nontrivial E-dependent dielectric constant 1 B£eff

€(E) = E BE = 1 +

8a 2 (hc)3

2

641m: 3

45 (MC 2 )4E + ~

[

(hC)3 ] 2 4 (MC 2 )4 E + .... (19.108)

If we admit also a constant B-field parallel to E , the formulas (19.104) and (19.103) for the spin-~ -particles are modified in the same way as the bosonic formulas (19.82) and (19.83), except that the determinant (19.101) in spinor space introduces a further factor cosh (eB / M c). Thus we obtain

For a general combination of constant electric and magnetic fields, we simply exchange E and B by the Lorentz invariants c and f3. From the imaginary part we obtain the decay rate

For strong fields, a saddle point approximation to the generalized integral (19.103) yields the asymptotic form (19.111)

19.2.2

Birth of Universe

A similar tunneling phenomenon could explain the birth of the expanding universe [10]. As an idealization of the observed density of matter, the universe is usually assumed to be isotropic and homogeneous. Then it is convenient to describe it by a coordinate frame in which the metric is rotationally invariant. To account for the expansion, we have to allow for an explicit time dependence of the spatial part of the metric. In the spatial part, we choose coordinates which participate in the expansion. They can be imagined as being attached to the gas particles in a homogenized universe. Then the time passing at each coordinate point is the proper time. In this context it is the so-called cosmic standard time to be denoted by t. We imagine being an observer at a coordinate point with dxi / dt = 0, and measure t by counting the number of orbits of an electron around a proton in a hydrogen atom, starting from the moment of the big bang (forgetting the fact that in the early time of the universe, the atom does not yet exist).

1387

19.2 Tunneling in Relativistic Physics

Geometry

With this time calibration, the component goo of the metric tensor is identically equal to unity goo(x)

=1,

(19.112)

such that at a fixed coordinate point, the proper time coincides with the coordinate time, dT = dt. Moreover , since all clocks in space follow the same prescription, there is no mixing between time and space coordinates, a property called time orthogonality, so that gOi(X)

=0.

(19.113)

As a consequence, the Christoffel symbol I'oolL [recall (10.7)] vanishes identically: -

roolL

1 = _gILV(oogov + oogov 2

goo)

= 0.

(19.114)

This is the mathematical way of expressing the fact that a particle sitting at a coordinate point, which has dxi/dt = 0, and thus dxIL/dt = u lL = (c, O,O,O), experiences no acceleration dUlL

dT =

-

2

-roolLc = 0.

(19.115)

The coordinates themselves are trivially comoving. Under the above condition, the invariant distance has the general form (19.116) We now impose the spatial isotropy upon the spatial metric gij' Denote the spatial length element by dl, so that (19.117) The isotropy and homogeneity of space is most easily expressed by considering the spatial curvature (3) R;jk I calculated from the spatial metric (3) gij (x). The space corresponds to a spherical surface. If its radius is a, the curvature tensor is, according to Eq. (10.161), (3)R ijkl (X)

=

:2

[(3)gil(X)(3)gjk(X) - (3)gik(X)(3)gjl(X)].

(19.118)

The derivation of this expression in Section 10.4 was based on the assumption of a spherical space whose curvature K 1/a2 is positive. If we allow also for hyperbolic and parabolic spaces with negative and vanishing curvature, and characterize these by a constant

=

k= {

~ -1

spherical } parabolic hyperbolic

universe,

(19.119)

1388

19 Relativistic Particle Orbits

then the prefactor 1/a2 in (19.118) is replaced by K == k/a 2. For k = -1 and 0, the space has an open topology and an infinite total volume. The Ricci tensor and curvature scalar are in these three cases [compare (10.163) and (10.156)] (3) Ril

= k 22 gil(X),

(3)R

a

= k~2·

(19.120)

a

By construction it is obvious that for k = 1, the three-dimensional space has a closed topology and a finite spatial volume which is equal to the surface of a sphere of radius a in four dimensions (19.121)

A circle in this space has maximal radius a and a maximal circumference 27ra. A sphere with radius ro < a has a volume (3)v: a

=

TO

[ " dr.p o

1"

dO sin 0

0

f V dr

0

1-

r2

r2 /a 2

(19.122)

3 a2r0n)02 47r ( -a arcsin -ro - - 1- . 2 a 2 a2

For small ro, the curvature is irrelevant and the volume depends on ro like the usual volume of a sphere in three dimensions: ~

(3)v:a TO

~

v: _ 47r 2 3 ro . TO -

(19.123)

For ro --+ a, however, (3)v,.~ approaches a saturation volume 27ra 3 . The analogous expressions for negative and zero curvature are obvious.

Robertson-Walker Metric In spherical coordinates, the four-dimensional invariant distance (19.116) defines the Robertson- Walker metric. c2dt2 - dl 2 dr 2 1 _ kr2 / a2

(19.124) 2

. 20d 2) + r 2(do + sm r.p.

(19.125)

It will be convenient to introduce, instead of r, the angle a on the surface of the four-sphere, such that

r = asina.

(19.126)

Then the metric has the four-dimensional angular form (19.127)

1389

19.2 Tunneling in Relativistic Physics

where for spherical, parabolic, and hyperbolic spaces, f(a) is equal to sina f(a) = { a sinha

k = 1, k=O , k =-1.

(19.128)

In order to have maximal symmetry, it is useful to absorb a(t) into the time and define a new timelike variable ry by edt = a(ry) dry ,

(19.129)

so that the invariant distance is measured by (19.130) Then the metric is simply -1

(19.131)

and the Christoffel symbols become

o a.,., i 0 . a.,., . foo = - , foo = 0, fOi = 0, fo/ = - 6/, a a

fi/ = 0, (19.132)

where the subscripts denote derivatives with respect to the corresponding variables:

da a.,.,

ada

== dry =

~ dt

a

==

(19.133)

~at.

We now calculate the 00-component of the Ricci tensor: (19.134) Inserting the Christoffel symbols (19.132) we find

a fool-' I-'

aof

01-' I-'

= -aOfiOi = -3!£ a.,., =

dry a

-3~ (a a a2 """"

. i= fl-'O v fovl-' = foo 0foo 0 + foo ifOi 0 + fiO 0foo i + fiOJfOj fool-'f vl-'V

a2 )

(19.135)

,

.,.,

(a.,.,) a 2 + 3 (a.,.,) a 2 ,(19.136)

= fooofooo + fooofiOi + fooifOiO + fOOifk/ = (;

r r, + 3 (;

(19.137)

so that 3 (aa.,.,.,., - a.,.,2) . Ro o = g oOR00 = - 4

a

(19.138)

1390

19 Relativistic Particle Orbits

The other components can be determined by relating them to the three-dimensional curvature tensor (3) Rij which has the simple form (19.118). So we calculate (19.139)

Rki/ +Roi/ (3) R;j

-

rk/r iO k + ri/r kO k + R Oi/ .

Inserting

and the above Christoffel symbols (19.132) gives (19.142)

and thus a curvature scalar

R

1 [3 2 ] 3 2 900 Roo + 9ij R;j _- -2" 2"(aaT)T) - aT)) - 4(2ka + aT)2 + aaT)T)) a a a 6 -3(aT)T) a

+ ka).

(19.143)

Action and Field Equation In the absence of matter, the Einstein-Hilbert action of the gravitational field is (19.144)

were

K,

is related to Newton's gravitational constant (19.145)

by (19.146)

A natural length scale of gravitational physics is the Planck length, which can be formed from a combination of Newton's gravitational constant (19.145), the light velocity c ~ 3 X 1010 cm/s, and Planck's constant n ~ 1.05459 x 10- 27 : lp =

(GNn3)-1/2 C

~ 1.615

x 10-

~

cm.

(19.147)

1391

19.2 Tunneling in Relativistic Physics

It is the Compton wavelength lp

== ti/mpc associated with the Planck mass (19.148)

The constant 1/", in the action (19.144) can be expressed in terms of the Planck length as 1

ti

'"

87r1~ '

(19.149)

If we add to (19.144) a matter action and vary to combined action with respect to the metric g/-,v, we obtain the Einstein equation (19.150)

where T/-,v is the energy-momentum tensor of matter. The constant.A is called the cosmological constant. It is believed to arise from the zero-point oscillations of all quantum fields in the universe. f

A single field contributes to the Lagrangian density C in (19.144) a term -A == -.A!'" which is typically of the order of ti/lt. For bosons, the sign is positive, for fermions negative, reflecting the filling of all negative-energy in the vacuum. A constant of this size is much larger than the present experimental estimate. In the literature one usually finds estimates for the dimensionless quantity

.A c2 r2 AO

== 3H6'

(19.151)

where Ho is the Hubble constant, whose inverse is roughly the lifetime of the universe

HOI ~ 14

X

109 years.

(19.152)

Present fits to distant supernovae and other cosmological data yield the estimate

[11] r2 AO

~

0.68 ± 0.10.

(19.153)

The associated cosmological constant .A has the value

_ 3H6 ~ r2 Ao ~ r2 Ao ~ r2 Ao .A - r2 Ao - c2 ~ (6.55 x 10 27)2 cm ~ (6.93 x 109 ly )2 ~ ( 2.14 Runiverse )2'

(19.154)

Note that in the presence of A, the Schwarzschild solution around a mass M has the metric (19.155)

1392

19 Relativistic Particle Orbits

with 2MG N

B(r) = 1- - 2 c r

2

2 2 M lp

r2

-Ar = 1- - - - -0'\0 2. 3 mp r 3 (2.14 Runiverse)

-

(19.156)

The A-term adds a small repulsion to Newton's force between mass points. if the distances are the order of the radius of the universe. The value of the constant A associated with (19.153) is

A= ~ =

0'\0 3HJ

l~ ~ 1O-122~.

It

c2 81f

Ii

(19.157)

Such a small prefactor can only arise from an almost perfect cancellation of the contributions of boson and fermion fields. This cancellation is the main reason for postulating a broken supersymmetry in the universe, in which every boson has a fermionic counterpart. So far, the known particle spectra show no trace of such a symmetry. Thus there is need to explain it by some other not yet understood mechanism. The simplest model of the universe governed by the action (19.144) is called Friedmann model or Friedmann universe .

19.2.3

Friedmann Model

Inserting (19.138) and (19.143) into the DO-component of the Einstein equation (19.150) , we obtain the equation for the energy

2+ ka 2) -

a34 ( a7J

A

=

0

liTo .

(19.158)

In terms of the cosmic standard time t , the general equation reads 3

at)2 + k aC22 ] [( -;;:

2

AC =

2 0 C liTo .

(19.159)

The simplest Friedmann model is based on an energy-momentum tensor Too of an ideal pressure-less gas of mass density p: (19.160) where uJ.t are the velocity four-vectors uJ.t = (,,(, "tv/c) of the particles whose components are uJ.t = (,,(, "tv/c) with "t == 1/ v 2 / c2 if the four components transform O like (dx = cdt, dx) . The gas is assumed to be at rest in our comoving coordinates. so that only the Too-component is nonzero:

VI -

(19.161) This component is invariant under the time transformation (19.129).

1393

19.2 Tunneling in Relativistic Physics

As a fortunate accident, this component of the Einstein equation has no a."."a term. Thus we may simply study the first-order differential equation

2

a34 ( a." + ka

2) -

A = c"'p.

(19.162)

Since the total volume of the universe is 27l'a3 , we can express p in terms of the total mass M as follows M

(19.163)

P = 27l'2a3 ' In this way we arrive at the differential equation

(19.164) This equation of motion can also be obtained in another way. We express the action (19.144) in terms of a(T/) using the equation (19.178) for R. We use the volume (19.121) and the relation (19.129) to rewrite the integration measure as (19.165)

The second expression arises from the first by a partial integration and ignoring the boundary terms which do not influence the equation of motion. The above matter is described by the action (19.167) Variation with respect to a yields the Euler-Lagrange equation 3

6 ( a."." + ka ) - 4Aa -

",Mc 27l'2

= O.

(19.168)

Note that in terms of the Robertson-Walker time t [recall (19.129)], the equation of motion reads .. A 1 ",Mc a = Sa - 6 27l'2a2'

(19.169)

As one should expect, the cosmological expansion is slowed down by matter, due to the gravitational attraction. A positive cosmological constant, on the other hand, accelerates the expansion. At the special value A=

",Mc

AEinstein

== -7l'2 2 3 = a

4G N M 47l'G N P = --, 7l'C a c

--2-3-

(19.170)

1394

19 Relativistic Particle Orbits

the two effects cancel each other and there exist a time-independent solution at a radius a and a density p. This is the cosmological constant which Einstein chose before Hubble's discovery of the expanding universe to agree with Hoyle's steadystate model of the universe (a choice which he later called the biggest blunder of his life ). Multiplying (19.168) by ary and integration over 'fJ yields the pseudo-energy conservation law (19.171) in agreement with (19.164) for a vanishing pseudo-energy. This equation may also be written as (19.172) where (19.173) This looks like the energy conservation law for a point particle of mass 2 in an effective potential of the universe Vuniv(a)

=

ka 2

-

a

max

a-

~a4 3'

(19.174)

at zero total energy. The potential is plotted for the spherical case k = 1 in Fig. 19.2.

Potential of closed Friedman universe as a function of the reduced radius a/amax. for Aa~ax = 0.1. Note the metastable minimum which leads to a possible solution a == ao. A tunneling process to the right leads to an expanding universe. Figure 19.2

Friedmann neglects the cosmological constant and considers the equation (19.175)

1395

19.2 Tunneling in Relativistic Physics

The solution of the differential equation for this trajectory is found by direct integration. Assuming k = 1 we obtain

TJ=

~ ~ ~ (19.176) = J =-arccos-. J J_vuniV(a) J-(a - amax /2)2 + a;;'ax/4 amax

With the initial condition a(O) = 0, this implies

a(TJ) = a~ax (1 - COSTJ).

(19.177)

Integrating Eq. (19.129) , we find the relation between TJ and the physical (=proper) time t =

~

JdTJa(TJ)

= a;:

JdTJ (1- COSTJ)

= a;: (TJ - sinTJ)·

(19.178)

The solution a(t) is the cycloid pictured in Fig. 19.3. The radius of the universe bounces periodically with period to = 7ra max /c from zero to amax . Thus it emerges from a big bang, expands with a decreasing expansion velocity due to the gravitational attraction, and recontracts to a point.

2.5

a 1.5

~---nO"n .2 ~n" 0.4'--'O"" .6~n o.'n 8 ~-l---'1."n 2~

t Ito

Figure 19.3 Radius of universe as a function of time in Friedman universe, measured in terms of the period to == 7ra max /c. (solid curve=closed, dashed curve=hyperbolic, dotted curve =parabolic). The curve for the closed universe is a cycloid. Certainly, for high densities the solution is inapplicable since the pressure-less ideal gas approximation (19.161) breaks down. Consider now the case of negative curvature with k = -1. Then the differential equation (19.175) reads

A4 =0 ary2 - a2 - amax a - -a 3 .

(19.179)

In order to compare the curves we shall again introduce a mass parameter M and rewrite the density as in (19.163), although M has no longer the meaning of the total mass of the universe (which is now infinite) . The solution for A = 0 is now

a(TJ)

amax ( cosh TJ - 1) , -2-

(19.180)

t

a;: (sinh TJ - TJ) .

(19 .181)

1396

19 Relativistic Particle Orbits

The solution is again depicted in Fig. 19.3. After a big bang, the universe expands forever , although with decreasing speed, due to the gravitational pull. The quantity a max is no longer the maximal radius nor is to the period. Consider finally the parabolic case k = 0, where the equation of motion (19.175) reads (19.182) where M is a mass parameter defined as before in the negative curvature case. Now the solution for A = is simply

°

(19.183) which is inverted to (19.184) Now we solve (19.129) with

t - amax -

3

12c TJ ,

(19.185)

so that (19.186) This solution is simply the continuation of leading term in the previous two solutions to large t.

19.2.4

Thnneling of Expanding Universe

It is now interesting to observe that the potential for the spherical universe in Fig. 19.2 allows for a time-independent solution in which the radius lies at the metastable minimum which we may call ao. The solution a == ao is a timeindependent universe. We may now imagine that the expanding universe arises from this by a tunneling process towards the abyss on the right of the potential [10]. Its rate can be calculated from the Euclidean action of the associated classical solution in imaginary time corresponding to the motion from ao towards the right in the inverted potential - v univ ( a ) . Observe that this birth can only lead to universe of positive curvature. For a negative curvature, where the a2 -term in (19.174) has the opposite sign, the metastable minimum is absent.

1397

19.3 Relativistic Coulomb System

19.3

Relativistic Coulomb System

An external time-independent potential V(x) is introduced into the path integral (19.48) by substituting the energy E by E - V(x). In the case of an attractive Coulomb potential, the second term in the action (19.49) becomes

A int = _ ('b d)" h()") (E + e2 /r? J),a 2Mc2 '

(19.187)

where r = Ixl. The associated path integral is calculated via a Duru-Kleinert transformation as follows [12]. Consider the three-dimensional Coulomb system where the spacetime dimension is D = 4. Then we increase the three-dimensional space in a trivial way by a dummy fourth component X4, just as in the nonrelativistic treatment in Section 13.4. The additional variable X4 is eliminated at the end by an integral J dx!/ra = J dla , as in (13.115) and (13.122). Then we perform a Kustaanheimo-Stiefel transformation (13.101) dxl-' = 2A(u)1-' "du". This changes X'1-'2 into 4iPU' 2 , with the vector symbol indicating the four-vector nature. The transformed action reads:

AeE = r),b d).,{ 4MiP U,2().,) + ,

J),a

2h()")

h()") [(M 2c4_ E2)U2 -2Ee2- e4 ] }. (19.188) 2Mc2U2 U2

We now choose the gauge h()") = 1, and go from)., to a new parameter s via the path-dependent time transformation d)" = fds with f = iP . Result is the DKtransformed action (19.189) It describes a particle of mass p, = 4M moving as a function of the "pseudotime" s

in a harmonic oscillator potential of frequency w

1

= --vM 2c4 - E2 .

(19.190)

2Mc

The oscillator possesses an additional attractive potential -e4 /2M c2 iP, which is conveniently parametrized in the form of a centrifugal barrier (19.191) whose squared angular momentum has the negative value (19.192) Here a denotes the fine-structure constant a == e2 /nc also a trivial constant potential v"onst = -

E

2

Mc2e .

::::; 1/137.

In addition, there is

(19.193)

1398

19 Relativistic Particle Orbits

If we ignore, for the moment , the centrifugal barrier

v.,xtra ,

the solution of the path

integral can immediately be written down [see (13.122)] : (X

11. 1 1 dL ee 2 EL/Mc?n 1411" d'V (71 LI71 0) bIx) a E = -i2Me16 0 0 la b a , 00

(19.194)

where (71bLI71aO) is the time evolution amplitude of the four-dimensional harmonic oscillator. There are no time slicing corrections for the same reason as in the threedimensional case. This is ensured by the affine connection of the KustaanheimoStiefel transformation satisfying (19.195) (see the discussion in Section 13.6). Performing the integral over "fa in (19.194), we obtain

with the variable (19.197) and the parameters v K,

2w11.Me2

=

J-tW =

211.

JM 2c4/E2 -1'

~v'M2c4 _ E2 11.e

=

E~.

11.e v

(19.198) (19.199)

As in the further treatment of (13.198) , the use of formula (13.202) 2

(X)

Io(zcos(B/2)) = - 2:)2l + 1)Pz(cosB)I21+1 (z) z 1=0

(19.200)

provides us with a partial wave decomposition

(xblxa)E

2l + 1 2:)rblra)e,I - 4- Pz(cosB) rbra 1=0 7r 1

=

(X)

-

t

r ~ f)rblra)e,l Yim(Xb)Yz;;'(Xa). b a 1=0 m=- l

(19.201)

1399

19.3 Relativistic Coulomb System

The radial amplitude is normalized slightly differently from (13.205):

n

2M

-i--~-

2Me

X

a

n

1

00

0

1 2vY dy--e sinhy

exp [-KcothY(Tb + Ta)] 121H

(19.202)

(2KVTbTa~h ) sm y

.

At this place, we incorporate the additional centrifugal barrier via the replacement 2l + 1 --+ 2l + 1 = V(2l

+ 1)2 + l~xtra,

(19.203)

as in Eqs. (8.145) and (14.239). The integration over y according to (9.29) yields

. n Mf(-/J+l+1) hlTa)El = -z2M ~ W f+l/2 (2KTb) M Vf+1/ 2 ( 2KTa). (19.204) , enK (2l+1)!'v , This expression possesses poles in the Gamma function whose positions satisfy the equations /J - l - 1 = 0,1,2,... . These determine the bound states of the Coulomb system. To simplify subsequent expressions, we introduce the small positive l-dependent parameter (19.205)

=

Then the pole positions satisfy /J = iii n - 51, with n = l + 1, l + 2, l + 3, .. . . Using the relation (19.198), we obtain the bound-state energies: Enl =

~

±Me2 [1+(n~251)2rl/2

±Me [1 _~ _a (_1 __ ~) + O(( 2n2 n 2l + 1 8n 4

2

3

6 )] .

(19.206)

Note the appearance of the plus-minus sign as a characteristic property of energies in relativistic quantum mechanics. A correct interpretation of the negative energies as positive energies of antiparticles is straightforward only within quantum field theory, and will not be discussed here. Even if we ignore the negative energies, there is poor agreement with the experimental spectrum of the hydrogen atom. The spin of the electron must be included to get more satisfactory results. To find the wave functions, we approximate near the poles /J ~ iii:

1

/J - iii

~

1 n T ! /J - iii' 2 n2K2 ( E)2 2Me2 iii 2M M e2 E2 - E;'l' Ell

(19.207)

1400

19 Relativistic Particle Orbits

with the radial quantum number nr = n -l - 1. By analogy with the nonrelativistic equation (13.207), the last equation can be rewritten as 1 1

(19.208)

K=-1/'

aH

where (19.209)

denotes a modified energy-dependent Bohr radius [compare (4.344)]. It sets the length scale of relativistic bound states in terms of the energy E. Instead of being 1/0'. >::> 137 times the Compton wavelength of the electron Ti / M c, the modified Bohr radius is equal to 1/0'. times Tic/E. Near the positive-energy poles, we now approximate (19.210)

Using this behavior and formula (9.48) for the Whittaker functions [together with (9.50)] we write the contribution of the bound states to the spectral representation of the fixed-energy amplitude as

Ti (rblra)E,1 = -M

L c n=l+l 00

(

E M 2 c

)2 E22M_c E2iTi RnI(rb)RnI(ra) + .... 2

(19.211)

nl

A comparison between the pole terms in (19.204) and (19.211) renders the radial wave functions (iii + I)! (n - l - I)!

(n - l

-= I)! e-r/fiiiH(2r/ii I aH )1+1 L~I+1 (2r/ii a ). nl - l-l I H

(ii + l)!

The properly normalized total wave functions are (19.213)

The continuous wave functions are obtained in the same way as from the nonrelativistic amplitude in formulas (13.215)- (13.223).

19.4

Relativistic Particle in Electromagnetic Field

Consider now the relativistic particle in a general spacetime-dependent electromagnetic vector field AJt(x).

1401

19.4 Relativistic Particle in Electromagnetic Field

19.4.1

Action and Partition Function

An electromagnetic field AI'(x) is included into the canonical action (19.14) in the usual way by the minimal substitution (2.642): (19.214) and the amplitude (19.22): (xblxa) =

2~c 10

00

dL

J

Vh q,[h]

J

(19.215)

VD x e-A./n,

with the minimally coupled action [compare (2.704)] (19.216) which reduces in the simplest gauge (19.23) to the obvious extension of (19.26): (19.217) with the action (19.218) The partition function of a single closed particle loop of all shapes and lengths in an external electromagnetic field is from (19.37)

Zl =

10 ~ e- f3Mc2/2 00

J

VDxe- Ae/Ii.

(19.219)

As in (19.45) and (19.46) this yields, up to a factor l/n the effective action of an ensemble of closed particle loops in an external electromagnetic field.

19.4.2

Perturbation Expansion

Since the electromagnetic coupling is rather small, we can split the exponent e-A/Ii into e-Ao/lie-Aintl li and expand the second factor in powers of Aint: e - Aintlli

=

E 00

(-ie/net

n!

!! n

[

(1if3

Jo

]

dTi

x(T;)A(x(T;)) .

(19.220)

If the noninteracting effective action implied by Eqs. (19.39), (19.37) , and (19.46) is

denoted by

fe,o= _ __ 100d(3 - f3 M2 / 2 " - Zl(3 e It

0

VD

V27rn2(3/M

__ VD

D-

A~

D(

1 _ )D/2f(1 D/2), (19.221)

47r

1402

19 Relativistic Particle Orbits

with the Compton wavelength pansion f e = fe,o _ {'>odf3 e-{3Mc2 /2

n

Jo

n

f3

AXt-

of Eq. (19.31), we obtain the perturbation ex-

f

VD D (-ie{e J27rn2f3/M n=l n.

r lIT [ r ·{3 dTi X(r;)A(X(r;))]) , \=1

Jo

0

(19.222) where ( . .. )0 denotes the free-particle expectation values [compare (3.480)- (3.483)] taken in the free path integral with periodic paths with a fixed f3 [compare (19.38) and (19.41)]: (19.223) The denominator is equal to VD/J27rn2f3/MD. The free effective action in the expansion (19.222) can be omitted by letting the sum start with n = O. The evaluation of the cumulants proceeds by Fourier decomposing the vector fields as

A(x) =

1(~:;DeikX A(k),

(19.224)

and rewriting (19.222) as fe = fe,o _

n

n x

f

n=l

roo df3 e-{3Mc /2 2

Jo

f3

(-ie/,ne n.

r fr [1 i=l

VD D J27rn2f3JM

dDk~AJLi(ki)]

(27r)

Ifr [ rn.{3 dTiXJLi(r;)eikiX(Til]) .(19.225) \ =1

Jo

0

First we evaluate the expectation values

(x( T1)eik1x (71l ... x( Tn)eiknx(Tnl) 0 .

(19.226)

Due to the periodic boundary conditions, we separate, as in Section 3.25, the path average Xo = X(T) [recall (3 .804)]' writing (19.227) and factorize (19.226) as

(ei(k 1+... +knlxQ) 0 ( 0 and (19.270) diverges. The physical resolution of this divergence problem is to assume the initial point charge eo in the electromagnetic interaction to be different from the experimentally observed e to precisely compensate the renormalization factor, i.e., a -> --

r

[1-II(V 2)] -a r

>::; -

(19.271) Thus, the result Eq. (19.270) is really obtained in terms of eo , i.e., with a replaced by aD . Then, using (19.271), we find that up to order a 2, the atomic potential is = _~ _

2~2

n J(3 l(x) . (19.272) r 40M 2c2 The second is an additional attractive contact interaction. It shifts the energies of the s-wave bound states in Eq. (19.206) slightly downwards.

Veff(x)

a

1408

19.5

19 Relativistic Particle Orbits

Path Integral for Spin-1/2 Particle

For particles of spin 1/2 the path integral formulation becomes algebraically more involved. Let us first recall a few facts from Dirac's theory of the electron.

19.5.1

Dirac Theory

In the Dirac theory, electrons are described by a four-component field 'l/Jcx (x) in spacetime parametrized by xl" = (ct, x) with /-L = 0, I, 2, 3. The field satisfies the wave equation

(inlP - Mc)'l/J(x) = 0, where IP is a short notation for anticommutation rules

"(1"01"

and

"(I"

(19.273)

are 4 x 4 Dirac matrices satisfying the (19.274)

where gl"v is now the Minkowski metric

-~o ~ )

0 -1 ( 0 0 1

gl"v =

o o

.

0

(19.275)

-1

An explicit representation of these rules is most easily written in terms of the Pauli matrices (1.445):

°

"( =

(1 0) 0 -1

.

'

"(' =

(

-

(Ji0 (Ji) 0 '

(19.276)

where 1 is a 2 x 2 unit matrix. The anticommutation rules (19.274) follow directly from the multiplication rules for the Pauli matrices: (19.277) The action of the Dirac field is A =

J

d4 xi{;(x)(inlP - Mc) 'l/J(x) ,

(19.278)

where the conjugate field i{; (x) is defined as

i{;(x) == 'l/Jt(xho.

(19.279)

It can be shown that this makes i{;(x)'l/J(x) a scalar field under Lorentz transformations, i{;(xhl"'l/J(x) a vector field, and A an invariant. If we decompose 'l/J (x) into its Fourier components (19.280)

1409

19.5 Path Integral for Spin-1j2 Particle

where V is the spatial volume, the action reads (19.281) with the 4 x 4 Hamiltonian matrix (19.282) This can be rewritten in terms of 2 x 2-submatrices as H (p)

= (

Mc

- pCT

pCT ) c.

(19.283)

-Mc

Since the matrix is Hermitian, it can be diagonalized by a unitary transformation to (19.284) where (19.285) are energies of the relativistic particles of mass M and momentum p. Each entry in (19.284) is a 2 x 2-submatrix. This is achieved by the Foldy- W outhuysen transformation (19.286) where

s=

-i'-t ·C./2,

C. = arctan (vic),

v

= plM =

(19.287)

velocity.

The vector C. points in the direction of the velocity v and has the length ( = arctan (v 1c), such that

Mc

cos(=

(19.288)

viP 2 + M22' c

A function of a vector v is defined by its Taylor series where even powers of v are scalars v 2n = v 2n and odd powers are vectors v 2n+1 = v 2n v . If v denotes as usual the direction vector v = vIlv!, the matrix I·V has the property that all even powers of it are equal to a 4 x 4 unity matrix up to an alternating sign: b· v)2n = (_l)n . Thus S = -il . v( and the Taylor series of eiS reads explictly

.

(-l)n (()n

e's :=~4, ..~

2 + b . v~=&5,...

(_l)n- l (()n n!

2

(

(

= cos 2 +1· v sin 2· (19.289)

1410

19 Relativistic Particle Orbits

,0

Now, S commutes trivially with 1 . P = 1 . ~ Ipl, while anticommuting with due to the anticommutation rules (19.274). Hence we can move the right-hand transformation in (19.286) simply to the left-hand side with a sign change of S, and obtain (19.290) It is easy to calculate e2iS : we merely have to double the rapidity in (19.289) and obtain

2'S Me e' =cos(+I·vsin(= yf 2 M22(1+ I ·pIMe) . p + e

(19.291)

Hence we obtain

H d =e2iS H=

yfp2

Me (1+I.pIMe)Me2,0(1+I.pIMe). + M2e2

(19.292)

,0

Taking the right-hand parentheses to the left of changes the sign of I . The product (1 + 1 . piM e) (1 - I . pi M e) is simply 1 + p2 I M 2e2, such that (19.293)

,0

Remembering from Eq. (19.276) shows that Hd has indeed the diagonal form (19.284). Going to the diagonal fields 'lji~(t) = eiS'ljik(t), the action becomes (19.294)

Thus a Dirac field is equivalent to a sum of infinitely many momentum states, each being associated with four harmonic oscillators of the Fermi type. The path integral is a product of independent harmonic path integrals of frequencies ±Wk. It is then easy to calculate the quantum-mechanical partition function using the result (7.419) for each oscillator, continued to real times: ZQM

=

II {2 cosh4[wk(tb -

t a )/2]} .

(19.295)

k

This can also be written as (19.296)

or as

1411

19.5 Path Integral for Spin-1j2 Particle

Since the trace is invariant under unitary transformations, we can rewrite this as ZQM

= exp {

~ Tr log [inot -

(19.298)

H(nk)] } ,

or, since the determinant of "'·l is unity, as

ZQM=exp{~Tr log[in,OOt-,."oHd(nk)] }=exp{~Tr IOg[in,OOt-nC"(k-Mc2]}. If we include the spatial coordinates into the functional trace, this can also be written as ZQM =

exp

[Tr log (in,OOt -

inC"( v -Mc2)] = exp {Tr log [c (in~ - Mc)]).

In analytic regularization of Section 2.15, the factor c in the tracelog can be dropped. Moreover, there exists a simple algebraic identity (19.299) The factors on the left-hand side have the same functional determinant since [compare (7.338) and (7.420)] Det(in~

_ Mc) =

eTrlog(ili~-Mc)

= Det(-in~

=

e V4

d4 J ~Trlog(Ii, - Mc)

=

eV4

Jd4 ~Trlog( - li, - Mc)

- Mc).

(19.300)

This allows us, as a generalization of (7.420) , to write

Det(in~ - Mc) = Det(in~ + Mc) = JDet( -n20 2 - M 2c2)14x4'

(19.301)

where 14x4 is a 4 x 4 unit matrix. In this way we arrive at the quantum-mechanical partition function (19.302) The factor 4 comes from the trace in the 4 x 4 matrix space, whose indices have disappeared in the formula. The exponent determines the effective action r of the quantum system by analogy with the Euclidean relation Eq. (19.46) . The Green function of the Dirac equation (19.273) is a 4 x 4 -matrix defined by (in~ - M c)O:{3 (xlxa){31'

= iM(D) (x - xa)Jo:1'"

(19 .303)

Suppressing the Dirac indices, it has the spectral representation ( I )-

XbXa -

J~ in (27rn)4p -Mc+ir/

-ip(x - x')/Ii

.

(19.304)

This can be written more formally as a functional matrix

in (xblxa) = (xblin~ _ Mc 1xa ), which obviously satisfies the differential equation (19.303) .

(19.305)

1412

19.5.2

19 Relativistic Particle Orbits

Path Integral

It is straightforward to write down a path integral representation for the amplitude (19.305): (19.306) with the action A[x,p]

= fos dT [-pi; + C~ - Me)],

(19.307)

where i; == dX(T)/dT. We are using the proper time T to parametrize the orbits. 3 Thus S is the total time, in contrast to the length parameter L in the Euclidean discussion in the previous section [see Eq. (19.3)]. The minus sign in front of pi; is necessary to have the positive sign for the spatial part px in the Minkowski metric (19.275). The action (19.307) can immediately be generalized to A[x,p]

= foS dT [-pi; + h(T)(~ - Me)],

(19.308)

with any function h( T) > O. This makes it invariant under the reparametrization (19.309) The path integral (19.306) contains then an extra functional integration over h(T) with some gauge-fixing functional If>[h], as in (19.22), which has been chosen in (19.307) as If>[h] = 6[h - 1]. The path integral alone yields an amplitude (19.310) and the integral over S in (19.306) produces indeed the propagator (19.305). In evaluating this we must assume, as usual, that the mass carries an infinitesimal negative imaginary part i'fJ. This is also necessary to guarantee the convergence of the path integral (19.306) . Electromagnetism is introduced as usual by the minimal substitution (2.642). In the operator version, we have to substitute

8/1-

.e

-----+

8/1- + Z ne AI'"

(19.311)

Thus we obtain the gauge-invariant action (19.312) 3It corresponds to the parameter>. in Subsection 19.2.1. From now on we prefer the letter T , since there will be no danger of confusing the proper time with the coordinate time in Subsection 19.2.1.

1413

19.5 Path Integral for Spin-1j2 Particle

Another path integral representation which is closer to the spinless case is obtained by rewriting (19.305) as

.

(xlxa) = (tnt;

in

+ Me) (xl_n282 _ M2e2Ixa),

(19.313)

where we have omitted the negative infinitesimal imaginary part for brevity, and used the fact that

-in of the mass, (19.314)

on account of the anticommutation relation (19.274). By rewriting (19.313) as a proper-time integral (19.315) we find immediately the canonical path integral (19.316) with the action (19.317) The suppressed Dirac indices of the 4 x 4 -amplitude on the left-hand side, (xlxa)",,a, are entirely due to the prefactor (int; + Me)",,a on the right-hand side. As in the generalization of (19.307) to (19.307), this action can be generalized to (19.318) with any function h(r) > 0, thus becoming invariant under the reparametrization (19.309), and the path integral (19.306) contains then an extra functional integration J'Dh(r) ~[h]. The action (19.318), is precisely the Minkowski version of the path integral of a spinless particle of the previous section [see Eq. (19.14)]. Introducing here electromagnetism by the minimal substitution (19.311) in the prefactor of (19.313) and on the left-hand side of (19.314) , the latter becomes then

(incjJ

-;4 +

Me)

(incjJ

-;4 -

Me) = n2

[(w - ;e A

r-

;e~/tVF/tv] -

M 2e2, (19.319)

where (19.320)

1414

19 Relativistic Particle Orbits

are the generators of Lorentz transformations in the space of Dirac spinors. For any fixed index /-l, they satisfy the commutation rules: (19.321) Due to the antisymmetry in the two indices, this determines all nonzero commutators of the Lorentz group. Using Eqs. (19.252), we can write the last interaction term in (19.319) as (19.322) where L: i are the generators of rotation (19.323) and O·

L: '

.

== in' == i'y

o' "

=i

(_(Ji 0

0)

(Ji

(19.324)

are the generators of rotation-free Lorentz transformations. Thus "J''''F L..J

19.5.3

= _ (

JW

cr(B + iE) 0 ) 0 cr(B - iE) .

(19.325)

Amplitude with Electromagnetic Interaction

The obvious generalization of the path integral (19.316) which includes minimal electromagnetic interactions is then _

1 [(.

(xIX a )-2M

e)

]iotx)dB JVh(T) [h] l

2nJ/J - cif.. +Mc

x =X(Tb) Xa=X(Ta)

4

V x

J(27rn)4 Te V4p

A

iA/1i

,

(19.326) with the action

The symbol T is the time-ordering operator defined in (1.241), now with respect to the proper time T, which has to be present to account for the possible noncommutativity of FlJ.vL:IJ.V /2 at different T. Integrating out the momentum variables yields the configuration-space path integral

1415

19.5 Path Integral for Spin-1j2 Particle

with the action

The coupling to the magnetic field adds to the rest energy M e2 an interaction energy fie

Hint

= --cr· B. Me

(19.330)

From this we extract the magnetic moment of the electron. We compare (19.330) with the general interaction energy (8.315), and identify the magnetic moment as (19.331) Recall that in 1926, Uhlenbeck and Goudsmit explained the observed Zeeman splitting of atomic levels by attributing to an electron a half-integer spin. However, the magnetic moment of the electron turned out to be roughly twice as large as what one would expect from a charged rotating sphere of angular momentum L, whose magnetic moment is L

~

(19.332)

= f.J,B 1i:'

where f.J,B == fie/Me is the Bohr magneton (2.647). On account of this relation, it is customary to parametrize the magnetic moment of an elementary particle of spin S as follows: (19.333) The dimensionless ratio g with respect to (19.332) is called the gyromagnetie ratio or Lande factor. For a spin-1/2 particle, S is equal to cr /2, and comparison with (19.331) yields the gyromagnetic ratio (19.334)

g=2,

the famous result found first by Dirac, predicting the intrinsic magnetic moment f.J, of an electron to be equal to the Bohr magnet on f.J,B, thus being twice as large as expected from the relation (19.332), if we insert there the spin 1/2 for the orbital angular momentum. In quantum electrodynamics one can calculate further corrections to this Dirac result as a perturbation expansion in powers of the fine-structure constant a [recall (1.502)]. The first correction to g due to one-loop Feynman diagrams was found by Schwinger:

g=2

X

(1

+ 2:)

:::; 2

X

1.001161,

(19.335)

1416

19 Relativistic Particle Orbits

where a is the fine-structure constant (1.502). Experimentally, the gyromagnetic ratio has been measured to an incredible accuracy: g

= 2 x 1.001159652193(10),

(19.336)

in excellent agreement with (19.335). If the perturbation expansion is carried to higher orders, one is able to reach agreement up to the last experimentally known digits [14] . In the literature, there exist other representations of path integrals for Dirac particles involving Grassmann variables. For this we recall the discussion in Subsection 7.11.3 that a path integral over four real Grassmann fields ()11-, fJ = 0,1,2,3 (19.337) generates a matrix space corresponding to operators el1- with the anticommutation rules (19.338) and the matrix elements (19.339) It is then possible to replace path integral (19.328) by

(19.340) with the action of a relativistic spinless particle [the action (19.329) without the spin coupling] (19.341) and an action involving the Grassmann fields

()11-:

This follows directly from Eq. (7.508). The function h(T) is the same as in the bosonic actions (19.14) and the path integral (19.22) guaranteeing the reparametrization invariance (19.13).

1417

19.5 Path Integral for Spin-1j2 Particle

After integrating out the momentum variables in the path integral (19.340), the canonical action is of course replaced by the configuration space action (19.216). In the simplest gauge (19.23), the total action reads

A[x,el'] =

foS dT [- ~:i;2 _ ~ (:i;A+i4~Fl'vel'eV) _ ~C2 + i:el'(T)(hT)].

(19.343) Note that the Grassmann variables can always be integrated out, yielding the functional determinant [compare with the real-time formula (7.512)]

J1)x

e-fr;

JdT[O"(T)iJ"(T)+(e/4Mc)F,,vO"Ov]

=

4Det 1/ 2 [bl'v OT

-

i .;c Fl'v(x( T))] . (19.344)

For a constant field tensor, and with the usual antiperiodic boundary conditions, the result has been given before in Eq. (19.101) .

19.5.4

Effective Action in Electromagnetic Field

In the absence of electromagnetism, the effective action of the fermion orbits is given by (19.302). Its Euclidean version differs from the Klein-Gordon expression in (19.38) only by a factor -2:

rf

~o = -2Tr log [_1i?02

+ M 2c2] .

(19.345)

Explicitly we have from (19.39), (19.41), and (19.46):

r~,o = 2VD roo df3 e-{3Mc2/2 Ji

Jo f3

1 D = 2 VDD 1D 2 r(1 - D /2). (19.346) J27rJi2f3/M'\~ (47r) /

The factor 2 may be thought of as 4 x 1/2 where the factor 4 comes from the free path integral over the Grassmann field,

J1) e D

e-Ae,o[Ol/1i

= 4.

(19.347)

counts the four components of the Dirac field. Recall that by (19.284), the Dirac field carries four modes, one of energy Jiwk, with two spin degrees of freedom, the other of energy -Jiwk with two spin degrees. The latter are shown in quantum field theory to correspond to an antiparticle with spin 1/2. The path integral over x (T) which counts paths in opposite directions with the ground state energy (19.39) describes particles and antiparticles [recall the remarks after Eq. 19.45]. This explains why only the spin factor 2 remains in (19.346). By including the vector potential via the minimal substitution p ~ p-(e/c)A, we obtain the Euclidean effective action from Eq. (19.219), and thus obtain immediately the path integral representation -f

re Ji

=2

roo df3 e-{3Mc2/2 J1)D X e- Ae/Ii

Jo f3

'

(19.348)

1418

19 Relativistic Particle Orbits

with the Euclidean action (19.218). This is not yet the true partition function re of the spin-1/2 particle, since the proper path integral contains the additional Grassmann terms of the action (19.343). In the Euclidean version, the full interaction is

Thus we obtain the path integral representation

~~

= 2

1~ 00

e- fJMc2/2

J J TJD x

TJDOe- {Ae,o[x,lIl+Ae,;n,[x,II]}/Ii ,

(19.350)

where the free part of the Euclidean action is Aeo[x,O] ,

19.5.5

= Aeo[x] + Aeo[O] == lo , '0

lifJ

M d72 j;2(7)

+

lolifJ 0

11, . d7 -401-'(7)8'-'(7).

(19.351)

Perturbation Expansion

The perturbation expansion is a straightforward generalization of the expansion (19.222): (19.352)

The leading free effective action coincides, of course, with the n = O-term of the sum [compare (19.346)]. The expectation values are now defined by the Grassmann extension of the Gaussian path integral (19.223): (O[x , 0])0

_ J TJD x J TJDO O[x, 0] e- Ae,o[x,IIJlIi = J TJD x e-Ae,o[xJlIi J TJDO e-Ae,o[IIJlIi' ~--;:---D

(19.353)

where the denominator is equal to (1/2)VDh/27r11,2 f3 / M x 4. There exists also an expansion analogous to (19.225) , where the vector potentials have been Fourier decomposed according to (19.224). Then we obtain an expansion just like (19.225) , except for a factor -2 and with the expectation values replaced as follows :

1419

19.5 Path Integral for Spin-1j2 Particle

The evaluation of these expectation values proceeds as in Eqs. (19.226)- (19.237), except that we also have to form Wick contractions of Grassmann variables which have the free correlation functions (19.355)

where (19.356)

is the Euclidean version of the antiperiodic Green function (3.109) solving the inhomogeneous equation (19.357)

Outside the basic interval [-11,,8, 11,,8) the function is to be continued antiperiodically. in accordance with the fermionic nature of the Grassmann variables. In operator language, the correlation function (19 .355) is the time-ordered expectation value (T{}(7){}(7'))o [recall (3.296)]. By letting 7 --+ 7' once from above and once from below, the correlation function shows agreement with the anticommutation rule (19.338). In verifying this we must use the fact that the time ordered product of fermion operators is defined by the following modification of the bosonic definition in Eq. (1.241): (19.358)

where tin' ... ,til are the times tn, . .. ,t 1 relabeled in the causal order, so that (19.359)

The difference lies in the sign factor Ep which is equal to 1 for an even and -1 for an odd number of permutations of fermion variables.

19.5.6

Vacuum Polarization

Let us see how the fluctuations of an electron loop change the electromagnetic field action. To lowest order, we must form the expectation value (19.354) for n = 0 and k1 = -k2 == k :

( [j;1'1 (71) + ;~kVltgvl (71)BI'1 (71)] eikOx (rll [j;1'2 h) -

2i~kV2BV2 h)B1'2 (72)] e-ikOXbl) ~ (19.360)

From the contraction of the velocities j;1'1 (71) and j;1'2 (72) we obtain again the spinless result (19.238) leading in (19.240) to the integrand

kl'2) ~(u - 1/2)2 . ( k12Jl'lV2 _ kl'l 11M2 1 kl'2) 1 ·C 2(71, 1') 2 = (k12Jl'lV2 - kl'l

(19.361)

1420

19 Relativistic Particle Orbits

In addition, there are the Wick contractions of the Grassmann variables:

([ 2~kVleVl(Tl)e/tl(Tl)] = -

Since E2h

-

eik8x (n)

(k~t5I"1V2 -

[2i~kV2eV2(T2)eMh)] e- ikOX (r2)) 0 ktlkt2)

~2 ~E2(TI -

T2) .

(19.362)

T2) = 1, this changes the spinless result (19.361) to

Remembering the factor -2 in the expansion (19.353) with respect to the spinless one, we find that the vacuum polarization due to fluctuating spin-1/2 orbits is obtained from the spinless result (19.249) by changing the factor 4(u - 1/2)2 = (2u - 1)2 in the integrand to -2 x 4u(u - 1) = 8u(1 - u). The resulting function II(k 2 ) has the expansion (19.364) The first term produces a renormalization of the charge which is treated as in the bosonic case [recall (19.269)- (19.272)]' which causes an additional contact interaction (19.365) There, the vacuum polarization has the effect of lowering the state 2S1 / 2 , which is the s-state of principal quantum number n = 2, against the p-state 2P1/2 by 27.3 MHz. The experimental frequency shift is positive ~ 1057 MHz [recall Eq. (18.600)]' and is mainly due to the effect of the electron moving through a bath of photons as calculated in Eq. (18.599). The effect of vacuum polarization was first calculated by Uehling [13], who assumed it to be the main cause for the Lamb shift. He was disappointed to find only 3% of the experimental result, and a wrong sign. The situation in muonic atoms is different. There the vacuum polarization does produce the dominant contribution to the Lamb shift for a simple reason: The other effects contain in a factor M / M;, where M; is the mass of the muon, whereas the vacuum polarization still involves an electron loop containing only the electron mass M, thus being enhanced by a factor (M/t/M)2 ~ 2102 over the others. The calculations for the electron in an atom have been performed to quite high orders [14] within quantum electrodynamics. We have gone through the above calculation only to show that it is possible to re-obtain quantum field-theoretic result within the path integral formalism. More details are given in the review article [5],

19.6 Supersymmetry

1421

As mentioned in the beginning, the above calculations are greatly simplified version of analogous calculations within superstring theory, which so far have not produced any physical results. If this ever happens, one should expect that also in this field a second-quantized field theory would be extremely useful to extract efficiently observable consequences. Such a theory still need development [15].

19.6

Supersymmetry

It is noteworthy that the various actions for a spin-l/2 particle is invariant under certain supersymmetry transformations.

19.6.1

Global Invariance

Consider first the fixed-gauge action (19.343). Its appearance can be made somewhat more symmetric by absorbing a factor Vti/2M into the Grassmann variables (}J.L(T), so that it reads

The correlation functions (19.355) of the (}-variables are now (19.367) with ~~(T - T') = E(T - T')/2. In this normalization, Gf(T, T') coincides, up to a sign, with the first term in the derivative 'G (T, T') of the bosonic correlation function [recall (19.230) and the first term in (19.232)] . Let us apply to the variables the infinitesimal transformations (19.368) where a is an arbitrary Grassmann parameter. For the free terms this is obvious. The interacting terms change by (19.369) Inserting FJ.LvXJ.L(T) = dAv(X(T))/dT - Ov[AJ.L(X(T))XJ.L(T)] , the first term cancels and the second is a pure surface term, such that the action is indeed invariant. Supersymmetric theories have a compact representation in an extended space called superspace. This space is formed by pairs (T, () , where ( is a Grassmann variable playing the role of a supersymmetric partner of the time parameter T. The coordinates XJ.L(T) are extended likewise by defining (19.370)

1422

19 Relativistic Particle Orbits

A supersymmetric derivative is defined by

DXI"(T)

=(~ o( + i(~) aT XI"(T)

=

iBI"(T)

+ i(i;I"(T).

(19.371)

If we now form the integral, using the Grassmann formula (7.379),

J dT ~; iXI"(T)DXI"(T) =J

dT~; i

[i;(T)+i((hT)] [iBI"(T)

+ i(i;I"(T)] ,

(19.372)

we find (19.373) which proportional to the free part of the action (19.343) . As a curious property of differentiations in superspace we note that

D2XI"(T) = ii;I"(T) - ((}I"(T),

D3XI"(T) = -(}I"(T) - (£(T),

(19.374)

such that the kinetic term (19.372) can also be written as

-JdT d ( XI"(T)D 3XI"(T).

(19.375)

21f

The interaction is found from the integral in superspace i J dT d( AI"(X(T))DX(T) 21f

= i J dT ~; [AI"(X(T)) + iOvAI"(X(T))BV(T)] [iBI"(T) + i(i;I"(T)] ,

(19.376)

which is equal to

- J dT [AI" (T) i;(T)

+ ~Fl"vBI"(T)BV(T)]

,

thus reproducing the interaction in (19.343) . The action in superspace can therefore be written in the simple form (19.377)

19.6.2

Local Invariance

A larger class of supersymmetry transformations exists for the action without gauge fixing which is the sum of the free part (19.341) and the interacting part (19.342). Absorbing again the factor vn/2M into the Grassmann variable BI"(T) , and rescaling in addition h(T) by a factor 1/c, the reparametrization-invariant action reads

A[x,p,B, h]

=

foS dT {-Pi; + ~~ [(p - ~A M· +2:iBI"(T)BI"(T) -

r-

M 2C2]

e

ih(T)~Fl"v(X(T))BI"(T)BV(T)

}.

(19.378)

1423

19.6 Supersymmetry

Let us now compose the action from invariant building blocks. For simplicity, we ignore the electromagnetic interaction. In a first step we also omit the mass term. The extra variable h(T) requires an extra Grassmann partner X(T) for symmetry, and we form the action

This action possesses a local supersymmetry. If we now perform T-dependent versions of the supersymmetry transformations (19.368)

8x l" 8h

= =

8BI" = a(T)p, 8X = 20(T).

ia(T)BI" , ia(T)x ,

8p = 0, (19.380)

If we integrate out the momenta in the path integral, the action (19.379) goes over into

where a term proportional to X2(T) has been omitted since it vanishes due to the nilpotency (7.375). This action is locally supersymmetric under the transformations

8x l" = ia(T)BI", 8h = ia(T)X,

8BI" = a( T) [ . _ i BI"] h(T) X 2X , 8X = 20(T) .

(19.382)

We now add the mass term (19.383) This needs a supersymmetric partner to compensate the variation of (19.382).

A5 =

~ foS dT

[B5(T)05(T) + MCX(T)B5(T)] .

Am

under

(19.384)

Indeed, add to (19.382) the transformation

8B5 = Mca(T) ,

(19.385)

we see that the sum AM + A5 is invariant. Adding this to (19.379) , we obtain the locally invariant canonical action

rS { h(T) h(T) M [. .] A[x,p, B, B5,h, X] = Jo dT -px+ 2MP2- -2-Mc+2i BI"(T)BI"(T) +B5(T)B5(T) +~X(T) [BIl-(T)PI"(T) + McB5(T)]}.

(19 .386)

1424

19 Relativistic Particle Orbits

Appendix 19A

Proof of Same Quantum Physics of Modified Action

Consider the sliced path integral for a relativistic point particle associated with the original action (19.12). If we set the initial and final paramters Aa and Ab equal to AO and AN+l, and slice the A-axis at the places An (n = 1,2, .. . , N), the action becomes N+1

L

Acl,e = Me

IX n

-

(19A.l)

xn-l l,

n=l

where IX n - xn-ll = V(xn - Xn _l)2 are the Euclidean distances [recall (19.2)] . The Euclidean amplitude for the particle to run from Xa = Xo to Xb = XN+l is therefore given by the product of integrals in D spacetime dimensions (xb lxa)

=Nf! [j

dDXn]

e-McL:~:,' lxn-Xn-'I/Ii.

(19A.2)

As a consequence of the reparamterization invariance of (19.12), this expression is independent of the thickness An - An-l of the slices, which we shall denote by rename as == hnE, where E is some fixed small number. We now factorize the exponential of the sum into a product of N + 1 exponentials and represent each factor as an integral [using Formulas (1.343) and (1.345)] e-McIXn-Xn - l l/1i

=

V 1 EMe

00

27[-1i

dh h-l/2e-€hnMc/21i-MC(Xn-Xn _,)2/2hn€1i

0

n

n

Absorbing constants in the normalization factor (xblxa)

=

N

II [1 n= l

00

N, we arrive at

dhnh~D-l)/2] e-Mc€E;:~1'hn/2

0

IT [J

1

X

(19A.3)

.

V27rEhN+1/McD n=l

dDXn ] V27rEhn/Me D

The second line contains only harmonic integrals over the formulas of Appendix 2B, with the result (xbLlxa O) =

Xn

,

e-McE;:~,'(Xn-Xn_l)2/2hn€

(19AA) '

which can all be done with the help of

1 D e- Mc (Xb - Xa)2 /2L\ V27rLIi/Me

(19A.5)

where L is the total parameter length

N+l L==

L

(19A.6)

Ehn ,

n=l

Replacing (19A.5) by its Fourier representation [compare with (1.329) and (1.337)], we can rewrite (19AA) as

[1

00 dh h(D- l)/2] ( X Ix ) = NNrr+l ban n

n=l

e-MCL/2IiJ~e-LP2/2MC+iP(Xb-Xa)/1i (27rIi)D .

(19A.7)

0

Thus we are left with the product of integrals over hn . Before we can perform these, we must make sure to respect the sum (19A.6). This is done by inserting an auxiliary unit integral that separates out an integral over the total length L: 1=

1

00

o

dL8(E N +1 Eh - L) = n= l

n

1 ~1 00

0

00

2AC

-ioo

du e-"(E;:~,' €hn -L)/2>'c 27ri '

(19A.8)

Appendix 19A Proof of Same Quantum Physics of Modified Action

1425

where'xc is the Compton wavelength (19.31). Then the product of integrals over h n in the brackets of (19A.7) can be rewritten as follows: (19A.9) Setting h n =

r;, we treat the curly backets as

(19A.1O) For large N , the integral over (J can be approximated by the Gaussian integral around the neighborhood of the saddle point at (J = (N + 1)(D + 1)'xc/ L:

l

i CO

- ioo

d

~

""!!""euL/2>.C (J - (N+l)(D+l)/2

27l'i

large N

[ ] (N+l)(D+l)/2 L . (19A.ll) y'27r (N + 1)(D + 1)'xc

_1_

While letting N tend to infinity, we keep L/(N + 1) side as an exponential

[(D + 1)'xc ]

1 __ y'27r

E

== l fixed . Then we may write the right-hand

- L(D+l)/2i"_ -

1 e -zL /2>.c __ y'27r ,

(19A.12)

where

z

== vlogv with v == (D + 1)'xc/l

(19A.13)

is a large number. Inserting (19A.12) back into the curly brackets of (19A.9), the constant z can be absorbed into the mass ofthe particle by replacing M by the renormalized quantity Ml = M(I+z). With this, (19A.9) becomes roo

Jo

~e- LMIC/21i 2'xc '

(19A.14)

and the path integral (19A.4) reduces to (19A.15) where all irrelevant factors are contained in the normalization factor Nil. The remaining integral over L can now be done and yields

where MR is the renomalized mass (19A.16) If we choose the normalization factor Nil = 1, this is exactly the same result as that obtained before in Eq. (19.30) from the modified action (19.10), if we use in the original action (19.12) the mass M/(1 + Z)1/2 rather than M for the calculation.

1426

19 Relativistic Particle Orbits

Notes and References Relativistic quantum mechanics is described in detail in J.D. Bjorken and S.D. Drell, Relativistic Quantum Mechanics, McGraw-Hill, New York, 1964, relativistic quantum field theory in S.S. Schweber, Introduction to Relativistic Quantum Field Theory, Harper and Row, New York, 1962; J.D. Bjorken and S.D. Drell, Relativistic Quantum Fields, McGraw-Hill, New York, 1965; C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill, New York, 1985. The individual citations refer to [1] For the development and many applications see the textbook H. Kleinert, Gauge Fields in Condensed Matter, World Scientific, Singapore, 1989; Vol. I, Superfiow and Vortex Lines (Disorder Fields, Phase Transitions); Vol. II, Stresses and Defects (Differential Geometry, Crystal Melting) (WY\IK/bl, where IJwwK is short for (http://WY\I.physik.fu-berlin.de/-kleinert). [2] See Vol. I of the textbook [1] and the original paper H. Kleinert, Lett. Nuovo Cimento 35, 405 (1982) (ibid.http/97). The theoretical prediction of this paper was confirmed only 20 years later in S. Mo, J . Hove, A. Sudbo, Phys. Rev. B 65, 104501 (2002) (cond-mat/Ol09260); Phys. Rev. B 66, 064524 (2002) (cond-mat/0202215). [3] R.P. Feynman, Phys. Rev. 80, 440 (1950). [4] There are basically two types of approach towards a worldline formulation of spinning particles: one employs auxiliary Bose variables: R.P. Feynman, Phys. Rev. 84, 108 (1989); A.O. Barut and LH. Duru, Phys. Rep. 172, 1 (1989); the other anticommuting Grassmann variables: E.S. Fradkin, Nucl. Phys. 76, 588 (1966); R. Casalbuoni, Nuov. Cim. A 33, 389 (1976); Phys. Lett. B 62, 49 (1976); F.A. Berezin and M.S. Marinov, Ann. Phys. 104, 336 (1977) ; 1. Brink, S. Deser, B. Zumino, P. DiVecchia, and P.S. Howe, Phys. Lett. B 64, 435 (1976); L. Brink, P. DiVecchia, and P.S. Howe, Nucl. Phys. B 118, 76 (1976). These worldline formulations were used to recalculate processes of electromagnetic and strong interactions by M.B. Halpern, A. Jevicki, and P. Senjanovic, Phys. Rev. D 16, 2476 (1977) ; M.B. Halpern and W. Siegel, Phys. Rev. D 16, 2486 (1977) ; Z. Bern and D.A. Kosower, Nucl. Phys. B 362, 389 (1991) ; 379, 451 (1992) ; M. Strassler, Nucl. Phys. B 385, 145 (1992); M.G. Schmidt and C. Schubert, Phys. Lett. B 331, 69 (1994) ; Nucl. Phys. Proc. Suppl. B,C 39, 306 (1995); Phys. Rev. D 53, 2150 (1996) (hep-th/9410100). For many more references see the review article in Ref. [5]. [5] C. Schubert, Phys. Rep. 355, 73 (2001) ; G.V. Dunne, Phys. Rep. 355, 73 (2002); [6] As a curiosity of history, Schriidinger invented first the relativistic Klein-Gordon equation and extracted the Schriidinger equation from this by taking its nonrelativistic limit similar to Eqs. (19.34) and (19.35) . [7] In particle physics, the Chern-Simons theory of ribbons explained in Section 16.7 was used to construct path integrals over fluctuating fermion orbits: A.M. Polyakov, Mod. Phys. Lett. A 3, 325 (1988) .

Notes and References

1427

For more details see C.H. Tze, Int. J . Mod. Phys. A 3, 1959 (1988) . [8] W. Heisenberg and H. Euler, Z. Phys. 98, 714 (1936) . English translation available at wwwK/files/heisenberg-euler . pdf. J . Schwinger, Phys. Rev. 84, 664 (1936); 93, 615; 94, 1362 (1954) . [9] E . Lundstrom, G. Brodin, J. Lundin, M. Marklund, R. Bingham, J. Collier, J .T. Mendonca, and P. Norreys, Phys. Rev. Lett. 96, 083602 (2006). [10] A. Vilenkin, Phys. Rev. D 27, 2848 (1983). [11] See the internet page http://super . colorado. edurmichaele/Lambda/links. html. [12] The path integral of the relativistic Coulomb system was solved by H. Kleinert, Phys. Lett. A 212, 15 (1996) (hep-th/9504024) . The solution method possesses an inherent supersymmetry as shown by K. Fujikawa, Nuc!. Phys. B 468, 355 (1996) . [13] E .A. Uehling, Phys. Rev. 49, 55 (1935) . [14] T. Kinoshita (ed.) , Quantum Electrodynamics, World Scientific, Singapore, 1990. [15] For a first attempt see H. Kleinert, Lettere Nuovo Cimento 4, 285 (1970) (wwwK/24) . New developments can be traced back from the recent papers I.I. Kogan and D. Polyakov, Int. J . Mod. Phys. A 18, 1827 (2003) (hep-th/0208036) ; D. Juriev, Alg. Groups Geom. 11, 145 (1994); R. Dijkgraaf, G. Moore, E. Verlinde, and H. Verlinde, Comm. Math. Phys. 185, 197 (1997).

Quandoquidem inter nos sanctissima divitiarum maiestas

Since the majesty of wealth is most sacred with us JUVENAL, Sat. 1, 113

20 Path Integrals and Financial Markets

An important field of applications for path integrals are financial markets. The prices of assets fluctuate as a function of time and, if the number of participants in the market is large, the fluctuations are pretty much random. Then the time dependence of prices can be modeled by fluctuating paths.

20.1

Fluctuation Properties of Financial Assets

Let S(t) denote the price of a stock or another financial asset . Over long time spans, i.e., if data recording frequency is low, the average over many stock prices has a time behavior that can be approximated by pieces of exponentials. This is why they are usually plotted on a logarithmic scale. This is best illustrated by a plot of the Dow-Jones industrial index over 60 years in Fig. 20.1. The fluctuations of the index have a certain average width called the volatility of the market. Over

1940

1960

1980

2000

Figure 20.1 Logarithmic plot of Dow Jones industrial index over 80 years. There are four roughly linear regimes, two of exponential growth, two of stagnation [1].

1428

1429

20.1 Fluctuation Properties of Financial Assets

100 ~==========================~ b Volatility u x 103 5.0

1.0 1984

1986

1988

1990

1992

1994

1996

Figure 20.2 (a) Index S&P 500 for I3-year period Jan. 1, 1984- Dec. 14, 1996, recorded every minute, and (b) volatility in time intervals 30 min (from Ref. [2]).

long times, the volatility is not constant but changes stochastically, as illustrated by the data of the S&P 500 index over the years 1984-1997, as shown in Fig. 20.2 [3] . In particular, there are strong increases shortly before a market crash. The theory to be developed will at first ignore these fluctuations and assume a constant volatility. Attempts to include them have been made in the literature [3]-[79] and a promising version will be described in Section 20.4. The volatilities follow approximately a Gamma dstribution, as illustrated in Fig. 20.3. w r-----~----_,r-----~----_,------~----_,----__.

• Empirical volatiJiw ._._._._._- Gauss clisb'_ fit ....... -Log-nonnal clisb'. fit Gannna clisb'_ fit

Nonnalized volatility

Figure 20.3 Comparison of best Gaussian, log-normal, and Gamma distribution fits to volatilities over 300 min (from Ref. [80]). The normalized log-normal distribution has the form Dlog- normal(z) = (21T0'2z2) - lj2e- (logz - I')2j2a2. The Gamma distribution will be discussed further in Subsection 20.1.5.

1430

20 Path Integrals and Financial Markets

An individual stock will in general be more volatile than an averge market index, especially when the associated company is small and only few shares are traded per day.

20.1.1

Harmonic Approximation to Fluctuations

To lowest approximation, the stock price S(t) satisfies a stochastic differential equation for exponential growth

S(t) S(t) = rs

+ 1](t),

(20.1)

where rs is the growth rate, and 1](t) is a white noise variable defined by the correlation functions

(1](t)) = 0,

(1](t)1](t'))

=

(j20(t - t').

(20.2)

The standard deviation 17 is a precise measure for the volatility of the stock price. The squared volatility v == 17 2 is called the variance. The quantity dS(t)/ S(t) is called the return of the asset . From financial data, the return is usually extracted for finite time intervals ~t rather than the infinitesimal dt since prices S(t) are listed for certain discrete times tn = to + n~t. There are, for instance, abundant tables of daily closing prices of the market S(t n ), from which one obtains the daily returns ~S(tn)/S(tn) = [S(tn+l) - S(tn)]/S(tn). The set of available S(tn) is called the time series of prices. For a suitable choice of the time scales to be studied, the assumption of a white noise is fulfilled quite well by actual fluctuations of asset prices, as illustrated in Fig. 20.4.

S(w)

Figure 20.4 Fluctuation spectrum of exchange rate DM/US$ as function of frequency in units 1lsec, showing that the noise driving the stochastic differential equation (20.1) is approximately white (from [13]) .

1431

20.1 Fluctuation Properties of Financial Assets

For the logarithm of the stock or asset pricel

x(t) == log S(t)

(20.3)

this implies a stochastic differential equation for linear growth [14, 15, 16, 17]

.

x(t) =

1 2 ) SS - 20' =rx+1](t,

(20.4)

where

rx == rs -

1 2 -0' 2

(20.5)

is the drift of the process [compare (18.405)] . A typical set of solutions of (20.4) is shown in Fig. 20.5. 6

Figure 20.5 Behavior of logarithmic stock price following the stochastic differential equation (20.3).

The finite differences ~x(tn) = x(tn+l)-x(tn) and the corresponding differentials

dx are called log-returns. The extra term 0'2/2 in (20.5) is due to Ito's Lemma (18.413) for functions of a stochastic variable x(t). Recall that the formal expansion in powers of dt:

dx dx(t) = dSdS(t) .

1 d2x

+ 2dS2dS .

2

(t)

+ ...

2

S(t) 1 [S(t)] 2 S(t) dt - 2 S(t) dt + ...

(20.6)

may be treated in the same way as the expansion (18.426) using the mnemonic rule (18.429), according to which we may substitute x 2 dt -+ (x 2 )dt = 0'2 , and thus (20.7) The higher powers in dt do not contribute for Gaussian fluctuations since they carry higher powers of dt . For the same reason the constant rates rs and rx in S(t)/S(t) and x(t) do not show up in [S(t)/ S(t)J2 dt= x 2 (t)dt. ITo form the logarithm, the stock or asset price S(t) is assumed to be dimensionless, i.e. the numeric value of the price in the relevant currency.

1432

20 Path Integrals and Financial Markets

In charts of stock prices, relation (20.5) implies that if we fit a straight line through a plot of the logarithms of the prices with slope r x, the stock price itself grows on the average like (20.8) This result is, of course, a direct consequence of Eq. (18.425). The description of the logarithms of the stock prices by Gaussian fluctuations around a linear trend is only a rough approximation to the real stock prices. The volatilities depend on time. If observed at small time intervals, for instance every minute or hour, they have distributions in which frequent events have an exponential distribution [see Subsection 20.1.6]. Rare events, on the other hand, have a much higher probability than in Gaussian distributions. The observed probability distributions possess heavy tails in comparison with the extremely light tails of Gaussian distributions. This was first noted by Pareto in the 19th century [18], reemphasized by Mandelbrot in the 1960s [19], and investigated recently by several authors [20, 22]. The theory needs therefore considerable refinement. As an intermediate generalization we shall introduce, beside the heavy power-like tails, also the so-called semi-heavy tails, which drop off faster than any power, such as e- xa x b with arbitrarily small a > 0 and any b > O. We shall see later in Section 20.4 that semi-heavy tails of financial distribution may be viewed as a consequence of Gaussian fluctuations with fluctuating volatilities. Before we come to these we may fit the data phenomenologically with various non-Gaussian distributions and explore the consequences.

20.1.2

Levy Distributions

Following Pareto and Mandelbrot we may attempt to fit the distributions of the price changes l:iSn = S(t n+1) - S(t n), the returns l:iSn/S(tn), and the log-returns l:ix n = x(tn+1) - x(tn) for a certain time difference l:it = tn+1 - tn approximately with the help of Levy distributions [19, 22, 24, 23]. For brevity we shall, from now on, use the generic variable z to denote any of the above differences. The Levy distributions are defined by the Fourier transform (20.9) with (20.10) For an arbitrary distribution D(z) , we shall write the Fourier decomposition as

- J-

D(z) =

dp etpZ . D(p)

271"

(20.11 )

1433

20.1 Fluctuation Properties of Financial Assets

and the Fourier components D(p) as an exponential (20.12) where H(p) plays a similar role as the Hamiltonian in quantum statistical path integrals. By analogy with this we shall also define H(z) so that (20.13) An equivalent definition of the Hamiltonian is

10 ' P(z)

10-< 10~5~0~-----:_2;C;:5;----;!;0-~~2';;-5~-----:'50 Z/(7

Z/(7

Figure 20.6 Left: Levy tails of the S&P 500 index (1 minute log-returns) plotted against z/8. Right: Double-logarithmic plot exhibiting power-like tail regions of the S&P 500 index (1 minute log-returns) (after Ref. [23]) e- H(P)

==

(e - ipz ).

(20.14)

For the Levy distributions (20.9), the Hamiltonian is (20.15) The Gaussian distribution is recovered in the limit A --+ 2 where the Hamiltonian simply becomes (J'2p2/2. For large z, the Levy distribution (20.9) falls off with the characteristic power-law -,x

L(72(z)

,x

--+

A

AU 2IzI 1+,x'

(20.16)

This power falloff is the heavy tail of the distribution discussed above. It is also called power tail, Paretian tail, or Levy tail). The size of the tails is found by approximating the integral (20.9) for large z, where only small momenta contribute, as follows: (20.17)

1434

20 Path Integrals and Financial Markets

with

A;2 =

(JA100 dp' (JA _p'A cos p' = ~ sin(7rA/2) r(1 2/\ 0 7r 27r /\

-,

+ A).

(20.18)

The stock market data are fitted best with A between 1.2 and 1.5 [13], and we shall use A = 3/2 most of the time, for simplicity, where one has

A 3/ 2 = ~ (J3/2 ,,2

(20.19)

4 V27f '

The full Taylor expansion of the Fourier transform (20.10) yields the asymptotic series _A

_

L,,2(Z) -

00

(-1)n

1 L -,n=O n. 0

00 dp(JAnpAn

_ --2 cospz n 7r

00

L n=O

(_1)n+1 (JAn

, n.

2n r(l 7r

sin "2A

+ nA) I II+>.· (20.20) Z

This series is not useful for practical calculations since it diverges. In particular, it is unable to reproduce the pure Gaussian distribution in the limit A ---> 2. There also exists an asymmetric Levy distribution whose Hamiltonian is (20.21) where E(p) is the step function (1.312), and F

_{

A,,,(P) -

tan(7rA/2) -(1/7r) logp2

forA-j.1, for A = 1.

(20.22)

The large-izi behavior of this distribution is again given by (20.17), except that the prefactor (20.18) is multiplied by a factor (1 + (3).

20.1.3

Truncated Levy Distributions

Mathematically, an undesirable property of the Levy distributions is that their fluctuation width diverges for A < 2, since the second moment (20.23) is infinite. If one wants to describe data which show heavy tails for large log-returns but have finite widths one must make them fall off at least with semi-heavy tails at very large returns. Examples are the so-called truncated Levy distributions [22]. They are defined by (20.24)

1435

20.1 Fluctuation Properties of Financial Assets

with a Hamiltonian which generalizes the Levy Hamiltonian (20.15) to

H(p)

a?- A = 20'2 oX(loX) [(a + ip)A + (a - ip)A - 2a A] a

2 (a 2 + p2)A/2 cos [oX arctan(p/a)] - a A aA- 2oX(1 _ oX) .

(20.25)

The asymptotic behavior of the truncated Levy distributions differs from the power behavior of the Levy distribution in Eq. (20.17) by an exponential factor e- az which guarantees the finiteness of the width a and of all higher moments. A rough estimate of the leading term is again obtained from the Fourier transform of the lowest expansion term of the exponential function e -H(P):

---> z->oo

2saAf(1 e

. (\)

-alzl

\) sm 7r /\ _e __ +/\ 7r slzlHA'

(20.26)

where 0'2

s =2

a 2- A oX(l- oX)'

(20.27)

The integral follows directly from the formulas [25]

1

00

-00

dp . 8(z) c az 27r e'PZ (a + ip)A = f(-oX) ZHA'

1

00

-00

dp . 8( -z) e- a1zl 27r e'PZ (a - ip)A = f(-oX) Iz IHA' (20.28)

and the identity for Gamma functions 2 1 f( -z) = -f(l

.

+ z) sm(7rz)/7r.

(20.29)

The full expansion is integrated with the help of the formula [26]

1

00

- 00

dp 27r

eipz

(a + ip)A(a - iPt

1 { f( \ ) W(v-A)/2,(HA+v)/2(2az) z > 0, - (2a)A/2+v/2 for IzIHA/2+v/2 1 f( -v) W(A-v)/2,(HA+V)/2(2az) z < 0,

(20.30)

where the Whittaker functions W(v-A)/2 ,(HA+v)/2(2az) can be expressed in terms of Kummer's confluent hypergeometric function IF1(a; b; x) of Eq. (9.45) as

W", .. (x) (20.31) 2M. Abramowitz and 1. Stegun, op. cit., Formula 6.1.17.

1436

20 Path Integrals and Financial Markets

as can be seen from (9.39), (9.46) and Ref. [27]. For 1/ = 0, only z > 0 gives a nonzero integral (20.30) , which reduces, with W-A/2 ,1/2+.X/2(Z) = Z-A/2 e -z/2, to the left equation in (20.28). Setting.x = 1/ we find

100 -dpe - 00

ipz (2

a +p

21f

2)" _ (

-

2a

),,/2

1 1 W. (I I) -11 () 01/2+" 2az . 1+" - r Z -1/ '

(20.32)

Inserting (20.33) we may write

dp ipz 2 2 ,,_ ( 2a ) i0021f e (a +p) - ~

1/2+"

00

For

1/

=

1

(20.34)

y'7iT(_I/)K1/ 2+,,(a zl). 1

-1 where K - 1/ 2(X) = K 1/ 2(X) = V1f/2xe- x, this reduces to

100 dp - 00

eipz _ _ 1_ =

a 2 + p2

21f

~e-alz l . 2a

Summing up all terms in the expansion of the exponential function

(20.35) e-H(p):

yields the true asymptotic behavior L-(A2 ,a)(z) "

____ z ..... oo

(2- 2A)saAr(1

e

alzl . (.x) \)sm 1f _e-__ + /\ 1f S IzI 1+ A'

(20.37)

which differs from the estimate (20.26) by a constant factor (see Appendix 20A for details) [28] . Hence the tails are semi-heavy. In contrast to Gaussian distributions which are characterized completely by their width (7 , the truncated Levy distributions contain three parameters (7, .x, and a. Best fits to two sets of fluctuating market prices are shown in Fig. 20.7. For the S&P 500 index we plot the cumulative distributions

for the price differences z = b.S over b.t = 15 minutes. For the ratios of the changes of the currency rates DM/$ we plot the returns z = b.S/ S with the same b.t. The plot shows the negative and positive branches Pd-z), and P>(z) both plotted on the positive z axis. By definition:

Pd-oo) P>(-oo)

0, PdO) = 1/2, Pdoo) = 1, = 1, P>(O) = 1/2, P>(oo) = O. =

(20.39)

1437

20.1 Fluctuation Properties of Financial Assets

P> (±z)

P>(±z)
"al(z) = ~e-H(Pll

=

00

-00

dp4

,,2

C2

C 4

= a 2,

+ 3c22 ,

(20.42)

(20.43)

p=O

and so on:

(Z6) = c6 + 15c4c2 + 15c~,

(zs) = Cs + 28c6C2 + 35c~ + 21Oc4c~ + 105c~, .. .. (20.44)

In a first analysis of the data, one usually determines the so-called kurtosis , which is the normalized fourth-order cumulant _ _ _ C4 (Z4)c (Z4)c r;, = C4 = ~ = (Z2)~ = 7 · (20.45) It depends on the parameters a , A, a as follows r;,

(2 - A)(3 - A) a2a 2 .

=

Given the volatility a and the kurtosis equation

(20.46)

we extract the Levy parameter a from the

r;"

a =.!. / (2 - A)(3 - A) .

aV

In terms of

r;,

(20.47)

r;,

and A, the normalized expansion coefficients are I /

'

..

\ ~:- .' "\~

%1

t

\~

L~;"(z)

0.3 {kl //;1

/1 ;/1;

..~ \~\

\z\

0.1

d/'/

-2

\~

0.2

\\

~\:;

~

z

-1

1 2 0 3 Figure 20.8 Change in shape of truncated Levy distributions of width a = 1 with increasing kurtoses r;, = 0 (Gaussian, solid curve), 1, 2,5, 10. -3

_

Cs =

n/2- 1 r;,

r(n - A)/r(4 - A) (3 _ A)n/2-2(2 _ A)n/2-2·

2 r;,

(7 - A)(6 - A)(5 - A)(4 - A) (3 - A)2(2 - A)2 ,

(20.48)

1439

20.1 Fluctuation Properties of Financial Assets

For ,\

= 3/2, the second equation in (20.47) becomes simply

a ~Vu~~,

(20.49)

=

and the coefficients (20.50):

_ 5·7 2 Ct>=-3-~' r(n - 3/2)/r(5/2) 3n/2- 2/2n- 4

----'---;::-'--::-';'----'.,-'--'-~

n/2-1

(20.50)

At zero kurtosis, the truncated Levy distribution reduces to a Gaussian distribution of width u. The change in shape for a fixed width and increasing kurtosis is shown in Fig. 20.8. From the S&P and DM/US$ data with time intervals tl.t = 15 min one extracts u 2 = 0.280 and 0.0163, and the kurtoses ~ = 12.7 and 20.5, respectively. This implies a ;::,j 0.46 and a ;::,j 1.50, respectively. The other normalized cumulants (C6, C8," ' ) are then all determined to be (1881.72, 788627.46, ... ) and (-4902.92,3.3168 x 106 , ... ), respectively. The cumulants increase rapidly showing that the expansion needs resummation. The higher normalized cumulants are given by the following ratios of expectation values

(20.51)

In praxis, the high-order cumulants cannot be extracted from the data since they are sensitive to the extremely rare events for which the statistics is too low to fit a distribution function .

20.1.4

Asymmetric Truncated Levy Distributions

We have seen in the data of Fig. 20.7 that the price fluctuations have a slight asymmetry: Price drops are slightly larger than rises. This is accounted for by an asymmetric truncated Levy distribution. It has the general form [24]

(20.52) with a Hamiltonian function

H(p) ==

u2

a 2-

A

2 '\(1- ,\) 2 (a 2

u

[(a + ip)A(1 + {3) + (a -

ip)A(1- {3) - 2a A]

+ p2)A/2 {cos['\ arctan(p/a)] + i,Bsin['\ arctan(p/a)]} a A- 2 ,\(1 _,\)

a A (2053) .

.

1440

20 Path Integrals and Financial Markets

This has a power series expansion .

1

.1

2

1

3

.1

4

5

H () P = tC1P + "2C2P - t3iC3P - 4fC4P + t5fC5P + ... There are now even and odd cumulants

_ Cn

-

en =

(20.54)

-inH(n)(o) with the values

2f(n-A) 2- n{1 t n=even, (J f(2 _ A) a fJ or n = odd.

(20.55)

The even cumulants are the same as before in (20.41). Similarly, the even expectation values (20.42)- (20.44) are extended by the odd expectation values:

(z)==

['>0 dzzL~~,a,(3)(z) = 00

i:

= Cl,

e-H(p) I

P

p=o

_~e-H(p)1 d2

=

C 2

+ c2l'

= -i.!!!..-eH(p) I = (z3)==1°O dz z3 DA,a,(3)(z) ,,2 d3 -00 P p=O

C 3

+ 3c2 C1 + c3l'

dz Z2 L(>\,a,(3)(z) = (z2)=1°O ,,2

P

-00

p=O

dz Z4 DA,a,(3)(z) = ~e-H(P)I =c4 + 4c3 C1 + (z4)=1°O ,,2 d4 -00 P p=O

3r~ + 6c2 c21 + c41 , - '" (20.56)

The inverse relations are C1

= (Z)c = (Z),

= (Z2)c = (Z2) C3 = (Z3)c = (Z3) C4 = (Z4)c = (Z4) = ((z - (Z)c))4 C2

(Z)2 = ((Z - (Z)c))2, 3(Z)(Z2) + 2(Z)3 = ((Z - (Z)c))3, 3(Z2? - 4(Z)(Z3) + 12(z)2(z2) - 6(Z)4 3(Z2 - (Z)~)2 = ((z - (Z)c))4 - 3c~.

(20.57)

These are, of course, just simple versions of the cumulant expansions (3.582) and (3.584). The distribution is now centered around a nonzero average value: JL

== (z) = C1·

(20.58)

The fluctuation width is given by

(J2 == (Z2) - (z? = ((z - (Z))2) = C2.

(20.59)

For large z, the asymmetric truncated Levy distributions exhibit semi-heavy tails, obtained by a straightforward modification of (20.28):

-

L~~,a)(z) ~

1 dp e'PZ. {

---> z-->oo

00

-00

27r

(J2 a 2- A 1- 2 A(I-A) [(a

+ ip)A(I+fJ) + (a -

>. sin(7r A) e-aizi (J2e 2sa f(I+A)--s--[I+fJsgn(z)]. 7r Izl1+A

ip)A(I-fJ) - 2a A]

}

(20.60)

1441

20.1 Fluctuation Properties of Financial Assets

In analyzing the data, one uses the skewness (20.61 ) It depends on the parameters a, A, (3, and

D!

or

Ii

as follows

"'-

~

~-4'---'""-:";'~":"'/-'::"2---+---~2'---':::::''''''-4''''' z - (z)

Figure 20.9 Change in shape of truncated Levy distributions of width a = 1 and kurtosis K = 1 with increasing skewness s = a (solid curve), 0.4, 0.8. The curves are centered around (z) .

s=

(2 - A)(3 aD!

.

(20.62)

The kurtosis can also be defined by [compare (20.45)]3

(20.63) From the data one extracts the three parameters volatility a, skewness s, and kurtosis Ii, which completely determine the asymmetric truncated Levy distribution. The data are then plotted against z - {z} = z - /1, so that they are centered at the average position. This centered distribution will be denoted by L~~,o:,(3)(z), i.e. (20.64) The Hamiltonian associated with this zero-average distribution is

fJ(p) == H(p) - H'(O)p,

(20.65)

and its expansion in power of the momenta starts out with p2, i.e. the first term in (20.54) is subtracted. 3Some authors call the ratio ((z - (z) ))4/0.4 in (20.63) kurtosis and the quantity K the excess kurtosis. Their kurtosis is equal to 3 for a Gaussian distribution, ours vanishes.

1442

20 Path Integrals and Financial Markets

In terms of (J, s, and _ C -

n/2- 1 /1,

n-

/1"

the normalized expansion coefficients are

f(n - '\)/f(4 _,\) { 1 (3 - ,\)n/2-2(2 - ,\)n/2-2 J(3 _ ,\)/(2 _

n = even,

---;-c:------'--:-:---;-;o-~::c-'-____:_:----';-;;--;;-

,\)/1,S

for

(20.66) n = odd.

The change in shape of the distributions of a fixed width and kurtosis with increasing skewness is shown in Fig. 20.9. We have plotted the distributions centered around the average position z = (Cl) which means that we have removed the linear term iCIP from H(p) in (20.52), (20.53) , and (20.54) . This subtracted Hamiltonian whose power series expansion begins with the term C2p2/2 will be denoted by - ( ) -Hp _ () -H'( Op=-C2P ) 1 2 . 1 3 1 4 . 1 5 Hp - Z+Z-C5P + . .. 2 3! C3P - -C4P 4! 5!

20.1.5

(20.67)

Gamma Distribution

For a Hamiltonian H+(p)

= v log (1 -

(20.68)

ip/ M)

one obtains the normalized Gamma distribution of mathematical statistics: D-G amma(z) jt,v

1 lIvzv - 1e- jtz = __

r(v)'"

,

ioroo dz fyGamma(z) /-t,V

which is restricted to positive variables z. Expanding H+(p) in a power series - L:~=l invM-npn /n = identify the cumulants

= 1

(20.69)

,

L:~=l inCnpn

In! , we (20.70)

so that the lowest moments are (20.71) The maximum of the distribution lies slightly below the average value

zmax = (v -l) /M =

z at

z -l/M·

(20.72)

The Gamma distribution was seen in Fig. 20.3 to yield an optimal fit to the fluctuating volatilities when is it is plotted as a distribution of the volatility (J = VZ: (20.73) As a normalized distribution of the volatility (J rather than the variance v = calls this function a Chi distribution.

(J2,

one

1443

20.1 Fluctuation Properties of Financial Assets

In the limit v 8-function:

----+ 00

at fixed

z = v / fl. , the Gamma distribution becomes a Dirac

iJGamma(z) JJ,V

----+

8(Z - ~) fl.

(20.74)

If we add to H+(p) the Hamiltonian

H_(p) = v log(l + ip/ fl.)

(20.75)

we obtain the distribution of negative z-values: D-Gamma(z) JJ ,V

= _ 1_ II- vlzIV- 1e- JJlzl f(v)""

,

0

(20.76)

z:-S:.

By combining the two Hamiltonians with different parameters fl. one can set up a skewed two-sided distribution.

20.1.6

Boltzmann Distribution

The highest-frequency returns b.S of NASDAQ 100 and S&P 500 indices have a special property: they display a purely exponential behavior for positive as well as negative z, as long as the probability is rather large [32]. The data are fitted by the Boltzmann distribution S&P500 2004-05 (1min)

NASDAQ100 2001-02 (1min)

-1

-1

-2

-2

-3

-3

-4

-4

_5L-__ -1

~

____l -_ _

-0.5

0.5 z in %

~

-5

-4

logP(z)

.. -2

\'; 2 0 z in %

4

Figure 20.10 Boltzmann distribution of S&P 500 and NASDAQ 100 high-frequency log-returns recorded by the minute.

(20.77) We can see in Fig. 20.10, that only a very small set of rare events of large Izl does not follow the Boltzmann law, but displays heavy tails. This allows us to assign a temperature to the stock markets [33]. The temperature depends on the volatility of the selected stocks and changes only very slowly with the general economic and political environment. Near a crash it reaches maximal values, as shown in Fig. 20.11.

1444

20 Path Integrals and Financial Markets

It is interesting to observe the historic development of Dow Jones temperature over the last 78 years (1929-2006) in Fig. 20.12. Although the world went through a lot of turmoil and economic development in the 20th century, the temperature remained almost a constant except for short heat bursts. The hottest temperatures occurred in the 1930's, the time of the great depression. These temperatures were never reached again. An especially hot burst occurred during the crash year 1987. Thus, in the long run, the market temperature is not related to the value of the index but indicates the riskiness of the market. Only in bubbles, high temperatures go along with high stock prices. An extraordinary increase in market temperature before a crash may be useful to investors as a signal to go short. The Fourier transform of the Boltzmann distribution is B(p)

=

1

00

dz eiPZ J..- e-l zl/T

2T

-00

=

1 1 + (Tp)2

= e- H(P)

,

(20.78)

so that we identify the Hamiltonian as

H(p) = log[1 + (Tp)2].

(20.79)

This has only even cumulants:

C2n = -( _1)n H(2n) (0) = 2(2n - 1)! T 2n ,

n = 1,2, . ...

(20.80)

The Boltzmann distribution is a special case of a two-sided Gamma distribution discussed in the previous subsection. It has semi-heavy tails. We now observe that the Fourier transform can be rewritten as an integral [recall (2.497)] (20.81) S&P500 index 1990-2006

NASDAQ100 index 1990-2006

1500~Ind ex 1000

2000

500 O L-----~------~------~~

1990

6000~Ind ex

4000

1995

2000

2005

O L-----~------~----~~~

1990

1995

2000

2005

4~

:[1T;1;rI111111tTT 1990 1995 2000 2005 year

: Tn';;nIIllrITT! 1990

1995

year

2000

2005

Figure 20.11 Market temperatures of S&P 500 and NASDAQ 100 indices from 1990 to 2006. The crash in the year 2000 occurred at the maximal temperatures TS&P500 ~ 2 x 10- 4 and TNASDAQ ~ 4 x 10- 4 •

1445

20.1 Fluctuation Properties of Financial Assets

10000

Dow Jones Index 1929- 2006

1000

100 41 1950

1960

1970

1980

1990

2000

4 3

Market Temperature 104 T

1987

2 1

0 1930 year

Figure 20.12 Dow Jones index over 78 years (1929-2006) and the annual market temperature, which is remarkably uniform, except in the 1930's, in the beginning of the great depression. An other high extreme temperature occurred in the crash year 1987.

This implies that Boltzmann distribution can be obtained from a superposition of Gaussian distributions. Indeed, the prefactor of p2 is equal to twice the variance v = (j2 of the Gaussian distribution. Hence we change the variable of integration from T to (j2, substituting T = (j2 /2T 2 , and obtain (20.82) The Fourier transform of this is the desired representation of the Boltzmann distribution as a volatility integral over Gaussian distributions of different widths: (20.83) The volatility distribution function is recognized to be a special case of the Gamma distribution (20 .69) for I-l = 1/2T2 , V = 1.

1446

20.1.7

20 Path Integrals and Financial Markets

Student or Tsallis Distribution

Recently, another non-Gaussian distribution has become quite popular through work of Tsallis [58, 59, 60]. It has been proposed as a good candidate for describing heavy tails a long time ago under the name Student distribution by Praetz [61] and by Blattberg and Gonedes [62]. This distribution can be written as (20.84) where N8 is a normalization factor JJf(1/8) N8 = f(1/8 - 1/2) ' and (JK

(20.85)

== (JJ1 - 38/2, where (J is the volatility of the distribution. The function

e8(z) is an approximation of the exponential function called 8-exponential: e8Z

= [1 -

J:

uZ

]-1/8

(20.86)

.

In the limit 8 ---+ 0, this reduces to the ordinary exponential function eZ • Remarkably, the distribution of a fixed total amount of money W between N persons of equal earning talents follows such a distribution. The partition functions is given by (20.87) After rewriting this as

where E is an infinitesimal positive number, and a binomial expansion of (e - iAW l)N eiAW , we can perform the integral over A, using the formula 4

1 /\ (-iA + 00

d\

-00

27r

1

- ipx -

E)Ve

v- I

P

- f(v) e

- €P8() p ,

-

(20.89)

where 8(z) is the Heaviside function (1.300), we obtain ZN(W)

E (N)

W N- 1 N- l

= f(N)

(_l)k

N _ k

(N - k _l)N- l

48ee 1.8. Gradshteyn and I.M. Ryzhik, op. cit., Formula 3.382.7.

(20.90)

1447

20.1 Fluctuation Properties of Financial Assets

The sum over binomial coefficients adds up to unity, due to a well-known identity for binomial coefficients5 , so that (20.91) If we replace W by W - W n , this partition function gives us the unnormalized probability that an individual owns the part Wn of total wealth. The normalized probability is then

P (

) = Z-l (W -

N Wn

W n )N-2 =

r(N _ 1)

N

N - 1 ( _ Wn)N W 1 E

2

(20.92)

Defining W == W/(N -2) , which for large N is the average wealth per person, PN(wn ) can be expressed in terms of the .(()]2} (K>.(()

(2

K>.(()

K>.(()

.

(20.122)

Using the identity [50] 211

(20.123)

Kv+1(z) - Kv-1 (z) = -Kv(z),

z

the latter equation can be expressed entirely in terms of p

=

(,) = K 1+>.(() p ~ K>.(()

(20.124)

as (20.125) Usually, the asymmetry of the distribution is small implying that

Cl

is small, implying a small (3.

It is then useful to introduce the symmetric variance

(20.126) and write Cl

The cumulants

C3

= (3u.,

and

C4

C2

( ) 1)2 = U 2 = Us2 + (32 [1)4 (2 + 2 1 + .>. (2 Us -

2]

Us'

(20.127)

are most compactly written as

(20.128) and

(20.129)

1453

20.1 Fluctuation Properties of Financial Assets The first term in

C4

is equal to

0'; times the kurtosis of the symmetric distribution (20.130)

Inserting here

0'; from (20.126), we find 3 1'(.8 == r2()

6

+ (1 + >.) (r()

(20.131)

- 3.

Since all Bessel functions Kv(z) have the same large-z behavior Kv(z) --t V7r/2ze - z and the small-z behavior Kv(z) --t r(lI)/2(z/2)V , the kurtosis starts out at 3/>. for ( = 0 and decreases monotonously to 0 for ( --t 00. Thus a high kurtosis can be reached only with a small parameter >.. The first term in C3 is (31'(.80';, and the first two terms in C4 are 1'(.80'; + 6 (C3/ (3 - 1'(.80';) . For a symmetric distribution with certain variance and kurtosis 1'(.8 we select some parameter>' < 3/1'(.., and solve the Eq. (20.131) to find (. This is inserted into Eq. (20.126) to determine

0';

0'; ( p() .

(20.132)

If the kurtosis is larger, it is not an optimal parameter to determine generalized hyperbolic distributions. A better fit to the data is reached by reproducing correctly the size and shape of the distribution near the peak and allow for some deviations in the tails of the distribution, on which the kurtosis depends quite sensitively. For distributions which are only slightly asymmetric, which is usually the case, it is sufficient to solve the above symmetric equations and determine the small parameter (3 approximately by the skew s = C3/0'3 from the first line in (20.128) as

(20.133) This approximation can be improved iteratively by reinserting (3 into the second equation in (20.127) to determine, from the variance 0'2 of the data, an improved value of 0';. Then Eq. (20.129) is used to determine from the kurtosis I'(. of the data an improved value of 1'(.., and so on. For t he best fit near the origin where probabilities are large we use the derivatives

G(O)

8)>'-1/2 (C>'L,

G'(O)

(3G(O),

G"(O)

_

(20.134)

'Y

(~) >.-3/2 C>'k+ + (~) >.- 1/2 8- 2C>' (1 -

(20.135) 2>' _ (328 2) L ,

-~ [3 (~r-3/2 C>'k+ + (~r-1/2 8- 2C>' (3 -

6>' - (328 2)] k_,

(20.136)

where we have abbreviated (20.137) The generalized hyperbolic distributions with>' = 1 are called hyperbolic distributions. The option prices following from these can be calculated by inserting the appropriate parameters into an interactive internet page (see Ref. [51]) . Another special case used frequently in the literature is>. = - 1/2 in which case one speaks of a normal inverse Gaussian distributions also abbreviated as NIGs.

1454

20 Path Integrals and Financial Markets

20.1.12

Debye-Waller Factor for Non-Gaussian Fluctuations

At the end of Section 3.10 we calculated the expectation value of an exponential function (e PZ ) of a Gaussian variable which led to the Debye-Waller factor of Bragg scattering (3.308). This factor was introduced in solid state physics to describe the reduction of intensity of Bragg peaks due to the thermal fluctuations of the atomic positions. It is derivable from the Fourier representation (e PZ )

== jdz __1_e- z2/2u2ePz = jdz j dp v'27ra 2

e- u2p2/2eipz+Pz

=

eu2P2/2 . (20.138)

27r

There exists a simple generalization of this relation to non-Gaussian distributions, which is (20.139)

20.1.13

Path Integral for Non-Gaussian Distribution

Let us calculate the properties of the simplest process whose fluctuations are distributed according to any of the general non-Gaussian distributions. We consider the stochastic differential equation for the logarithms of the asset prices

x(t) = rx

+ 1](t) ,

(20.140)

where the noise variable 1](t) is distributed according to an arbitrary distribution. The constant drift rx in (20.140) is uniquely defined only if the average of the noise variable vanishes: (1](t)) = O. The general distributions discussed above can have a nonzero average (x) = Cl which has to be subtracted from 1](t) to identify r x. The subsequent discussion will be simplest if we imagine r x to have replaced Cl in the above distributions, i.e. if the power series expansion of the Hamiltonian (20.54) is replaced as follows:

H(p)

---+

HrJp)

H(p) - H'(O)p + irxp == fJ(p)

+ irxp

. 1 2 .1 3 1 4 .1 5 zrX p +2"C2P -z3fC3P -4fC4P +zSTC5P + .... (20.141) Thus we may simply work with the original expansion (20.54) and replace at the end (20.142) The stochastic differential equation (20.140) can be assumed to read simply

x(t)

=

1](t).

(20.143)

With the ultimate replacement (20.142) in mind, the probability distribution of the endpoints Xb = X(tb) for the paths starting at a certain initial point Xa = x(ta) is given by a path integral of the form (18.342):

P(Xbtblxata) = j'D1] {x(tb)=X b'Dxexp [- (tb dtiI(1](t))] 8[x -1]] . ix(ta)=Xa ita

(20.144)

1455

20.1 Fluctuation Properties of Financial Assets

The function H('fJ) is the negative logarithm of the chosen distribution of log-returns [recall (20.13)] (20.145)

H('fJ) = -log D('fJ).

For example, H('fJ) is given by -log LS~'

t large

V27ra 2

(20.176)

The Levy measure has been calculated explicitly for many non-Gaussian distributions. As an example, the generalized hyperbolic case has for A :::: 0 a Levy measure [76] (20.177) where h,(z) and Y.\(z) are standard Bessel functions. The decomposition of the Hamiltonian according the Levy-Khintchine formula and the additivity of the associated noises form the basis of the LVy-Ito theorem which states that an arbitrary stochastic differential equation with Hamiltonian (20.175) can be decomposed in the form i: = rxt + 1)G +1)9 +1»1,

(20.178)

where 1)G is a Gaussian noise 1)9 =

r

dz (e ipz -

J1xl9

1) F(z)

(20.179)

a superposition of discrete noises called Poisson point process with jumps smaller than or equal to unity, and 1»1 =

r

dz (e iPZ -1) F(z)

J1xl9

(20.180)

a noise with jumps larger than unity. Consider the simplest noise of the type 1). + k) TZ ~ P(n T ) = e- TZ(T {) )keTZ = _-'::::-oz':--'-,----'(nk) = ~nkP(n L.. ' TZ ) = e- TZ(TZ {)T % )ke L.. ,Z Z T% r(>.z) .

n=O

n=O

(20.184) Thus (n) = >'z, (n 2 ) = >'z(>'z+l) , (n 3 ) = >'z(>'z+1)(>'z+2), (n 4 ) = >'z(>'z+1)(>'z+2)(>'z+3) , so that (/2 = >'z, s = 2/"J>:Z, and f(. = 6/>.z. As typical noise curve is displayed in Fig. 20.16. An arbitrary Levy weight F(z) in 1]9 may

20

15

06

0.8

1.0

Figure 20.16 Typical Noise of Poisson Process. always be viewed as a superposition F(z) = J~I dZ F(Z)8(z - Z) so that the distribution function becomes a superposition of Poisson noises:

1 1

D(z) =

20.1.18

- I

00

n

dZ F(Z) ~ e- TZ ~ 8(z - nZ) .

(20.185)

Semigroup Property of Asset Distributions

An important property of the probability (20.144) is that it satisfies the semigroup equation (20.186)

In the stochastic context this property is referred to as Chapman-Kolmogorov equation or Smoluchowski equation. It is a general property of processes which a without

1462

20 Path Integrals and Financial Markets

memory. Such processes are referred to a Markovian [12]. Note that the semigroup property (20.221) implies the initial condition (20.187) In Fig. 20.17 we show that the property (20.221) is satisfied reasonably well by experimental asset distributions, except for small deviations in the low-probability tails. 10" .------~-------,

IO~

10 '

u S&P 500 " S&P 500 D S&P 500 1 day " S&P 500 5 days

IO~

z

Figure 20.17 Cumulative distributions obtained from repeated convolution integrals of the 15-min distribution (from Ref. [13]). Apart from the far ends of the tails, the semigroup property (20.221) is reasonably well satisfied. One may verify the semigroup property also using the high-frequency distributions of S&P 500 and NASDAQ 100 indices in Fig. 20.10 recorded by the minute, which display the Boltzmann distribution (20.77) . If we convolute the minute distribution an integer number t times, the result aggress very well with the distribution of t-minute data. Only at the heavy tails of rare events do not follow this pattern Note that due to the semigroup property (20.221) of the probabilities one has the trivial identity

JdXb f(Xb, tb)P(Xbtblxata) Jdx [j dXb f(Xb ,tb)P(Xbtblxt)] P(xtlxata).

(20 .188)

In the mathematical literature, the expectation value on the left-hand side of Eq. (20.188) is written as (20.189)

20.1 Fluctuation Properties of Financial Assets

1463

With this notation, the second line can be re-expressed in the form (20.190) so that we obtain the property of expectation values (20.191) In mathematical finance, this complicated-looking but simple property is called the towering property of expectation values. Note that since P(Xbtblxata) is equal to 8(Xb - xa) for ta = tb, the expectation value (20.189) has the obvious property (20.192)

20.1.19

Time Evolution of Moments of Distribution

From the time-dependent distribution (20.165) it is easy to calculate the time dependence of the moments: (20.193) Inserting (20.165), we obtain

(xn)(t) =

1 dp e- tH(p) 1 dx xneipx = 1 dpe- tH(p) (-iOp)n8(p). 00

00

00

27l'

00

-00

(20 .194)

00

After n partial integrations, this becomes

(Xn)(t) = (i8 te- tH(p) I p

p=O

.

(20.195)

All expansion coefficients en of H(p) in Eq. (20.141) receive the same factor t, so that the cumulants of the moments all grow linearly in time: (20.196)

-1

-1

-2

10

-3 L-_-'---~_---'---_---' -50 50

Figure 20.18 Gaussian distributions of S&P 500 and NASDAQ 100 weekly log-returns.

1464

20.1.20

20 Path Integrals and Financial Markets

Boltzmann Distribution

As a an example, consider the Boltzmann distribution (20.77) of the minute data. The subsequent time-dependent variance and kurtosis are found by inserting the cumulants Eq. (20.80) into (20.196), yielding

K(t)

C4

= -

tc§

3

(20.197)

= - .

t

The first increases linearly in time, the second decreases like one over time, which makes the distribution more and more Gaussian, as required by the central limit theorem (20.167) . The two quantities are plotted in Figs. 20.19 and 20.20. The agreement with the data is seen to be excellent. S&P 500 2004-2005

S&P 500 2004-2005

1.2

10

()"2(t)

deviation in %

0.8

..... 0.4 -5 0

0

5

10

15

20

-10

0

Time lag t (hours)

5

10

15

20

Time lag t (hours)

NASDAQIOO 2001-02 18

NASDAQIOO 2001-02 10

()"2(t)

deviation in %

12

-5 0

0

5

10

15

Time lag t (hours)

20

-10

0

10

15

20

Time lag t (hours)

Figure 20.19 Variance of S&P 500 and NASDAQ 100 indices as a function of time. The slopes (x 2 )c(t)/t are roughly 1.1/20 x 1/60 min and 18.1/20 x 1/60 min, respectively, so that Eq. (20.198) yields the temperatures Ts P500 :::; 0.075 and TNASDAQlOO :::; 0.15. The right-hand side shows the relative deviation from the linear shape in percent.

Note that if the minute data are not available, but only data taken with a frequency lito (in units 1/min) and an expectation (x 2)c(t0), then we can find the temperature of the minute distribution from the formula (20.198) Since K goes to zero like lit, the distribution becomes increasingly Gaussian as the time grows, this being a manifestation of the central limit theorem (20.167) of statistical mechanics according to which the convolution of infinitely many arbitrary distribution functions of finite width always approaches a Gaussian distribution.

1465

20.1 Fluctuation Properties of Financial Assets S&P500 2004-2005

S&P 500 2004-2005 deviation in % . 10

I«t)

o • -10

oL~~==~':'::l

o

10

15

20

Time lag t (hours)

-20 ' - - - - - - - - - - - - - - " o 10 15 20 Time lag t (hours)

NASDAQ100 2001-02 600

cr-~~-------;

NASDAQlOO 2001-02 10 ,---~--------;

deviation in % 400

.. . ..

200 -5

oO'--------:------:IC:O -----:1C:5 -----:'20

-10 O'--------:------:IC:O -----:1~5-----:'20

Time lag t (hours)

Time lag t (hours)

Figure 20.20 Kurtosis of S&P 500 and NASDAQ 100 indices as a function of time. The right-hand side shows the relative deviation from the lit behavior in percent.

This result is in contrast to the pure Levy distribution in Subsection 20.1.2 with A < 2, which has no finite width and therefore maintains its power falloff at large distances. If we omit the time argument in the cumulants (xn)c(l), for brevity, and insert these into the decompositions (20.56) we obtain the time dependence of the moments:

(x) (t) = t(x)c, (x 2)(t) =t(x2)c + t2(x)~, (x 3)(t) = t(x 3)c + 3t2 (x)c(x 2)c + t 3 (x)~, (x 4)(t) =t(x4)c + 3t2(X2)~ - 4t2(x)c(x3)c + 6t3(X)~(X2)c + t4(X)~,

(20.199)

Let us now calculate the time evolution of the Boltzmann distribution (20.77) of the minute data. The Hamiltonian H(p) was identified in Eq. (20.79) , so that 1

e-tH(p) =

(20.200)

[1 + (TpF]t' The time-dependent distribution is therefore given by the Fourier integral P(x, t)

=

1

00

-00

dp eipx-tH(p)

27r

=

1

00

-00

dp

27r (1

1

+ T2p2)t

eipx.

(20.201)

1466

20 Path Integrals and Financial Markets

The calculation proceeds most easily rewriting (20.200), by analogy with (20.81) and using again formula (2.497), as an integral (20.202)

The prefactor of p2 is equal to t2 = tv a Gaussian distribution in x-space. We therefore change the variable of integration from T to v, substituting T = tv/2T 2, and obtain [compare (20.82)] e

roo dvv v-tv/2T2 -tvp2/2 ee .

-tH(p) _ (_t_)t _1_ - 2T2 r(t) io

(20 .203)

t

The Fourier transform of this yields the time evolution of the Boltzmann distribution as a time-dependent superposition of Gaussian distributions of different widths [compare (20.83)]: P( X t ) -_ ( - t ) , 2T2

t

- 1

r(t)

loo -dv v e-tv/2T2-1-e-x2/2tv t

0

v

V27rtv

.

(20.204)

The integral can be performed using the integral formula (2.557), and we find the time-dependent distribution _

P(x, t) -

Loo 00

dp ipx-tH(p) _ 1 27r e - T y'1fr(t)

Ix I) t - l/2 ( 2T K t- 1/ 2(l x l/T) ,

(20.205)

where t is measured in minutes. For t = 1, we reobtain the fundamental distribution (20.77) , recalling the explicit form of K 1/ 2(Z) in Eq. (1.345). In the limit of large t , the integral over v in (20.204) can be evaluated in the saddle-point approximation of Section 4.2. We expand the weight function in the integrand around the maximum at Vm = 2T2(1- l/t) as vt-le-tv/2T2 = [2T2(1 - l/t)

r-

1

e- (t-l) e-t8v2/2[2T2(1 -1/t)]2 [1

+ O( ov3 )] , (20.206)

where Ov = v - vm , and O( ov 3 ) is equal to -tov 3 /3 [2T 2(1- l/t)j3. Then we observe that for large t, the Gaussian can be expanded into derivatives of o-functions as follows:

e- tz2 / 2 [1 +

~z3 + . . .J = V27r / t

[O(z) + ;/'(z) +

~O'(z) ...J .

This can easily be verified by multiplying both sides with J(z) = J(O) Z2 f"(0)/2 + .. . and integrating over z. Thus we find to leading order e-t8v2/2[2T2(1-1/t)j2 ---.

~ 2T2(1 -

l/t) o(ov) .

(20.207)

+ z!,(O) + (20.208)

1467

20.1 Fluctuation Properties of Financial Assets

Using the large-t limit (1 - l/t)t (20.207), we obtain vt- le- tv/2T2

----+

----+

e- 1 , and including the first correction from

j¥(2T2)t e- t {b(bV) +

[2T2(12~ l/tW b//(bv) + .. .} .

(20.209)

If we now employ Stirling's formula (17.286) to approximate tt/f(t) ----+ vt/27ret, we see that the integral (20.204) converges for large t against the Gaussian distribution e- x2 /2tvm / .J27rtvm with the saddle point variance Vm ----+ 2T2. The behavior near the peak of the distribution can be calculated from the small-z expansion7

( Z)v

2"

[f(l-

7r v) Z2 4 2V] K Az)=2sin7rvf(1-v) 1+1!f(2-v)4 +O(z ,z) .

(20.210)

For large v and small z, this can be approximated by [recall (20.29)]

Gr

f~) e- z2 /4(t- 3/2).

Kv(z) ;::.:,

(20.211)

We now use Stirling's formula (17.286) to find the large-t limit r(t - 1/2)/f(t) leading to the Gaussian behavior near the peak: 1 e- x2 /22T 2(t- 3/2). .J27r 2T2 t

P(x, t);::.:, small lx l

----+

1,

(20.212)

This corresponds to the approximation of the Fourier transform (20.200) by e- tT2p2. The Gaussian shape reaches out to x ;::.:, T0. For very large t, it holds everywhere, as it should by virtue of the central limit theorem (20.167).

20.1.21

Fourier-Transformed Tsallis Distribution

For the Hamiltonian defined in Eq. (20.97), the time evolution is given by the Fourier transform

which can be rewritten, by analogy with (20.99), as

e-tHo.(3(P)=~ f(t/b)

roo dv

10

vt/oe-I'Ve-vp2/2

v

J-l

= 1/(Jb.

'

(20.214)

Taking the Fourier transform of this yields the time-dependent distribution function P 0,(3

t o (x t) = -J-l-/ , f (t / b)

1

00

0

-dv v t/ oe- I'V _1_ e -

V

.J27rv

X

2

/2v

'

7M. Abramowitz and 1. Stegun, op. cit., Formulas 9.6.2 and 9.6.10.

J-l

= 1/ (Jb,

(20.215)

1468

20 Path Integrals and Financial Markets

which becomes with the help of Formula (2.497): p,t/Ii

PIi,{3(X, t)

1

(X2 ) t/21i-l/4

= r (t/ b)..;2ir 2p,

2Kt / Ii- 1/ 2 ( f'iP,x),

1

p, = (3b·

(20.216)

For b = 1 and p, = 1/2T2 , this reduces to the time-dependent Boltzmann distribution (20.205). At the origin, its value is [compare (20.104)]

PIi,{3(O, t) = V!ir(t/b - 1/2)/r(t/b). 20.1.22

(20.217)

Superposition of Gaussian Distributions

All time-dependent distributions whose Fourier transforms have the form e-tH(p) possess a path integral representation (20.161) and which obeys the semigroup equation (20.221) . In several examples we have seen that the Fourier transforms e-H(p) can be obtained from a superposition of Gaussian distributions of different variances e- vp2 / 2 with some weight function w(v): e- H(p)

=

looo dvw(v)e- vp2 / 2.

(20.218)

This was true for the Boltzmann distribution in Eq. (20.81), for the Fouriertransformed Tsallis distribution in Eq. (20.100), and for the Boltzmann Distribution of a relativistic particle in Eq. (20.105). In all these cases it was possible to write a similar superposition for the time-dependent distributions: e- tH(p)

=

looo dV'wt(v')e- V'p2/ 2 = looo dvwt(v)e- tvp2/2,

(20.219)

where we have introduced the time-dependent weight function (20.220) See Eqs. (20.202), (20.214), and (20.105) for the above three cases. The question arises as to general property of the time dependence of the weight function Wt(v) which ensures that that the Fourier transform of the superposition (20.219) obtained as in Eq. (20.164) satisfies the semigroup condition (20.221) [52]. For this, the superposition has to depend on time as

or e-tH(p) = [e-H(p)]t. This implies that the weight factor Wt(v) has the property

J

dV12Wt,+t2(v12)e- V12P2/2

=

J

dV2Wt2(v2)e- V2p2/2

J

dVIWtl (vl)e - V1P2/2.

(20.221)

For the Laplace transforms Wt(v) of Wt(v) (20.222)

1469

20.1 Fluctuation Properties of Financial Assets

this amounts to the factorization property

Wh +t2 (pv)

=

(20.223)

Wt2 (Pv)WtJ (pv) ,

which is fulfilled by the exponential

Wt(Pv) = e-tHv(pv),

(20.224)

with some Hamiltonian Hv(Pv). Since the integral over v in (20.222) runs only over the positive v-axis, the Hamiltonian Hv(Pv) is analytic in the upper half-plane of PV' It is easy to relate Hv(Pv) to H(p). For this we form the inverse Laplace transformation, the so-called Bromwich integral [53] 'Y + ioo dp Wt(v) = ~ePvv-tHv(pv), (20.225) 'Y- ioo 27r2 in which 'Y is some real number larger than the real parts of all singularities in cPvv-tHv(Pv) . Inserting this into (20.219) and performing the v-integral yields dp 1 - tH(p) _ 'Y+ ioo _v - tHv(Pv) (20.226) e . 2/ e . 'Y-ioo 27r2 P 2 - Pv Closing the integral over Pv in the complex pv-plane and deforming the contour to an anticlockwise circle around Pv = p2/2 we find the desired relation:

l

l

(20.227) The time-dependent distribution associated with any Hamiltonian H (p) via the Fourier integral (20.170) can be represented as a superposition of Gaussian distributions with the help of Eq. (20.219) if we merely choose Hv(Pv) = H( V'2Pv) . Recalling the derivation of the Fourier representation (20.165) from the path integral (20.153) and the descendence of that path integral from the stochastic differential equation (20.140) with the noise Hamiltonian H(p) we conclude that the Hamiltonian Hv(ipv) governs the noise in a stochastic differential equation for the volatility fluctuations. Take, for example, the superposition of Gaussians (20.105). The Laplace transform of the weight function w/3( v) is

W/3(Pv) =

J

dv e- vPv w/3( v) =

J

dv e- vpv

J !3

27rv 3 The Laplace transform of w/3(v) is related to this by

J

dv'e- v'Pv w/3(v') =

e-/3/2v = e- V2 /3Pv .

J

dv e-/3vPv w/3(v) = W/3(!3pv).

(20.228)

(20.229)

Hence it is given by W/3(Pv) = e- /3..fti);; = e-/3Hv(Pv), which satisfies the relation (20.223). Moreover, Hv(p2/2) is equal to H(p) = vir, as required by Eq. (20.227) and in accordance with Eq. (20.105) for M = O. Note that instead of the Bromwich integral (20.226) it is sometimes more convenient to use Post's Laplace inversion formula [54] -1' (_1)k k+1 &kWt(X) I Wt (v ) - 1m k' x !:l k k-+oo.

uX

x=k/v

(20.230)

1470

20.1.23

20 Path Integrals and Financial Markets

Fokker-Planck-Type Equation

From the Fourier representation (20.170) it is easy to prove that the probability satisfies a Fokker-Planck-type equation (20.231)

Indeed, the general solution 'I/; (x, t) of this differential equation with the initial condition 'I/; (x, 0) is given by the path integral generalizing (20.144) (20.232)

This satisfies the Fokker-Planck-type equation (20.231) . To show this, we take

'I/;(x, t) at a slightly later time t + E and expand 'I/;(X,t+E)

=

JD1Jexp[-l~bdtH(1J(t))] 'I/; (x-l~dt'1J(t')-lt+ 0 had to be carried only up to the second order in !::"x(t) == ftt+< dt' x(t'), we must now keep all orders. Evaluating the noise averages of the multiple integrals on the righthand side with the help of the correlation functions (20.155)- (20.262), we find the time dependence of the expectation value of an arbitrary function of the fluctuating variable x(t)

After the replacement equation:

Cl -> T

x, the function f(x(t)) obeys therefore the following

(j(x(t)))

=

-HrJiOx)(f(x(t))) .

(20.255)

Separating the lowest-derivative term, this takes a form generalizing Eq. (18.412) :

(j(x(t))) = (f'(x(t)))(x(t)) - HrJiox)(f(x(t))).

(20.256)

This might be viewed as the expectation value of the stochastic differential equation

j(x(t)) = f'(x(t))x(t) - Hrx(iox)f(x(t)) ,

(20.257)

which, if valid, would be a simple direct generalization of Ito's Lemma (18.413). However, this conclusion is not allowed. The reason lies in the increased size of the fluctuations of the higher expansion terms [!::"x(t)]n for n ~ 2. In the harmonic case of Subsection 18.13.3, these were negligible in comparison with the leading term Zl(t) in Eq. (18.413). Let us see what happens in the present case, where all higher

1474

20 Path Integrals and Financial Markets

cumulants en in the Hamiltonian (20.141) may be nonzero. For the argument we proceed as in Eq. (18.405) by working with the stochastic differential equation

i:(t) = (i:(t))

+ 1J(t) = Cl + 1J(t).

(20.258)

rather that (20.169) . Then the correlation functions of 1J(t) consist only of the connected parts of (20.155)-(20.262):

(1J(td) (1J(t 1)1J(t2)) (1J(t 1 )1J(t2)1J(t3)) (1J(t 1 )1J(t2)1J(t3)1J(t4 ))

=

(20.259)

CI,

= c28(tl -

t2)

(20.260)

= C38(tl -t2)8(tl -t3), = c4 8(t1 -t2)8(tl -t3)8(tl -t4 )

(20.261) (20.262)

+ C~ [8(tl-t2)8(h-t4)+8(tl-t3)8(t2-t4)+8(tl-t4)8(h-t3) ].

We now estimate the size of the fluctuations Zn of [~x(t)]n. The first contribution Z2,1 to Z2 defined in Eq. (18.415) is still negligible since it is smaller than the leading one Zl(t) == ftHfdt'1J(t') by a factor E. The second contribution Z2,3 to Z2 defined in Eq. (18.415), however, has now a larger variance, due to the c4-term in (20.262) , which contains one more 8-function than the c~-term, so that

This is larger than the harmonic estimate (18.420) by a factor liE, which makes Z2,2(t), in general, as large as the leading fluctuation Zl(t) of Xl(t) . It can therefore not be ignored. The subtraction of the second term (Z2,2(t))2 = E2C~ in the variance does not help since that is ignorable. A similar estimate holds for the higher powers. Take for instance [~x(t)P: (20.264)

and calculate the variance of the strongest fluctuating last term in this: ({[Zl(t)P}2) - ([Zl(t)P)2). The first contribution is equal to

Itt+fdt It+fdt2 Itt+fdt3 I t+fdt4 Itt+fdt5 I t+fdt6 (1J(tl)1J(t2)1J(t3)1J(t4)1J(t5)1J(t6)) , 1

t

t

t

(20.265) and becomes, after an obvious extension of the expectation values (20 .155)- (20.262) to the six-point function: (20.266)

The second contribution ([Zl (t)P)2 in (20.264) is equal to E2C~ and can be ignored for E --+ O. Thus, due to (20.266) , the size of the fluctuations [Zl (t)P is of the same order as of the leading one Xl(t) [compare again (18.421)].

1475

20.2 Iti5-like Formula for Non-Gaussian Distributions

This is an important result which is the reason why we cannot derive a direct generalization (20.257) of the Ito formula (18.413) to non-Gaussian fluctuations, but only the weaker formula (20.256) for the expectation value. For an exponential function f(x) = ePx this implies the relation (20.267) and it is not allowed to drop the expectation values. A consequence of the weaker Eq. (20.267) for P = 1 is that the rate rs with which the average of the stock price S(t) = ex(t) grows is given by formula (20.1), where the rate r x is related to rs by

rs

rx - fI(i)

=

rx - [H(i) - iH'(O)]

=

=

-Hrx(i),

(20.268)

which replaces the simple Ito relation rs = rx + (72/2 in Eq. (20.5). Recall the definition of fI(p) = H(p) - H'(O)p in Eq. (20.141). The corresponding version of the left-hand part of Eq. (20.4) reads

(~) =

(:i;(t)) - fI(i)

=

(:i;(t)) - [H(i) - iH'(O)]

=

(:i;(t)) - rx - Hrx(i).

(20.269)

The forward price of a stock must therefore be calculated with the generalization of formula (20.8), in which we assume again TJ(t) to fluctuate around zero rather than rx:

(S(t))

=

S(O)erst

=

S(O) (erxt+J; dt'1)(t'))

=

S(O)e- Hrx(i)t

If TJ(t) fluctuates around rx this takes the simpler form

(S(t))

=

S(O)e rst

=

S(O) (e J; dt'1)(t'))

=

=

S(O)e{rxt- [H(i) - iH'(O)]t}. (20.270)

S(O)e-Hrx(i)t.

(20.271)

This result may be viewed as a consequence of the following generalization of the Gaussian Debye-Waller factor (18.425):

(e P J; dt'1)(t')) = e-Hrx(iP)t.

(20.272)

Note that we may derive the differential equation (20.255) of an arbitrary function f(x(t)) from a simple mnemonic rule, expanding sloppily [similar to Eq. (18.426) but restricted to the expectation values]

(f(x(t + dt)))

=

(f(x(t)))

+ (f'(x(t))) (:i;(t)) dt + ~ (f" (x(t))) (:i;2(t)) dt 2

+ ~(f(3)(x(t)))(:i;3(t))dt3 + . . . ,

(20.273)

(:i;2(t))dt 2 --+ C2dt,

(20.274)

3.

and replacing

(:i;(t))dt

--+

c1dt,

(:i;3(t))dt 3 --+ C3dt, ....

In contrast to Eq. (18.429), this replacement holds now only on the average. For the same reason, portfolios containing assets with non-Gaussian fluctuations cannot be made risk-free in the continuum limit E --+ 0, as will be seen in Section 20.7.

1476

20.2.2

20 Path Integrals and Financial Markets

Discrete Times

The prices of financial assets are recorded in the form of a discrete time series x(tn ) in intervals 6.t = tn+l - tn, rather than the continuous function x(t). For stocks with a large turnover, or for market indices, the smallest time interval is typically 6.t = 1 minute. We have seen at the end of Section 18.13.3 that this makes Ito's Lemma (18.413) an approximate statement. Without the limit 6.t -+ 0, the fluctuations of the higher expansion terms in (20.253) no longer disappear but are merely suppressed by a factor (Tv'6.t n . In quiet economic periods this is usually quite small for 6.t = 1 minute, so that the higher-order fluctuations can usually be neglected after all. While this approximate validity is a disadvantage of the discrete time series over the continuous one, it is an advantage with respect to processes with nonGaussian noise, at least as long as the financial markets are not in turmoil. Then the non-Gaussian higher-order fluctuations of the expansion terms in (20.253) are suppressed by the same order (T.,;;s:i as the corrections to Ito's rule in the Gaussian case [recall (18.432)]. As a typical example, take the Boltzmann distribution (20.77). Its cumulants en carry a factor Tn [see Eq. (20.80)] where the market temperature T is a small number of the order of a few percent in the natural time units of one minute used in this context [see Fig. (20.11)]. The smallness of T is, of course, a consequence that the minute data do not usually possess large volatility, except near a crash. In the natural units of minutes, the time interval E in the calculations after Eq. (20.257) is unity, so that powers of E can no longer be used for size estimates. This role is now taken over by T. In fact, since en is of order Tn, the fluctuations of zn(t) are of the order Tn. Thus, in non-Gaussian fluctuations of a discrete time series, the smallness of T leads to a suppression of the higher fluctuations after all. Moreover, the suppression is just as good as for time series with Gaussian fluctuations, where the corrections are of the order ((T.,;;s:i)n- l for zn(t) with n 2: 2. These have the size of Tn- l for the Boltzmann distribution [recall (20.198)]. A similar estimate holds for all non-Gaussian noise distributions with semi-heavy tails as defined on at the end of Subsection 20.1.1. For all semi-heavy tails for which the cumulants Cn decrease with power Tn of a small parameter such as the temperature T, we may drop the expectations values of the Ito-like expansion (20.256), and remain with an approximate Ito formula for the fluctuating difference 6.f(x(tn )) = f(x(tn+l)): (20.275) which is a direct generalization of the discrete version (18.432) of Ito's Lemma (18.413). A characteristic property of the Boltzmann distribution (20.77) and others with semi-heavy tails is that their Hamiltonian possesses a power series in p2 of the type (20.141) with finite cumulants Cn of the order of (Tn, where (T is the volatility of the minute data. This property is violated only by power tails of price distributions which do exist in the data (see Fig. 20.10). These do not go away upon multiple

1477

20.3 Martingales

convolution which turns the Boltzmann distribution more and more into a Gaussian as required by the central limit theorem (20.167). These heavy tails are caused by drastic price changes in short times observed in nervous markets near a crash. If these are taken into account, the discrete version (18.432) of Ito's Lemma (18.413) is no longer valid. This will be an obstacle to setting up a risk-free portfolio in Section 20.7.

20.3

Martingales

In financial mathematics, an often-encountered concept is that of a martingale. The name stems from a casino strategy in which a gambler doubles his stake each time a bet is lost. A stochastic variable m(t) is called a martingale, if its expectation value is time-independent. A trivial martingale is provided by any noise variable m(t) = rJ(t) .

20.3.1

Gaussian Martingales

For a harmonic noise variable, the exponential m(t) = eJ; dt'ry(t /)-rr 2 t/2 is a nontrivial martingale, due to Eq. (20.8). For the same reason, a stock price S(t) = ex(t) with x(t) obeying the stochastic differential equation (20.140) can be made a martingale by a time-dependent multiplicative factor associated with the average growth rate rs, i.e., (20.276) is a martingale. The prefactor e- rst is referred to as a discount factor with the rate rs. If we calculate the probability distribution (20.170) associated with the stochastic differential equation (20.140), which for harmonic fluctuations of standard deviation (J corresponds to a Hamiltonian

(J2

H(p) = irxp+

2 p2 ,

(20.277)

The integral representation (20.170) for the probability distribution,

prx(Xbtblxata) =

L: ~~

exp [ip(xb - xa) -

~t (irxp + (J2p2 /2) 1

(20.278)

with ~t = tb -ta has obviously a time-independent expectation value of e- rstS(t) = e- rstex(t). Indeed, the expectation value at the time tb is given by the integral over Xb

which yields

1478

20 Path Integrals and Financial Markets

where we have used the Ito relation rs = r x that

+ (>2/2 of Eq.

(20.5). The result implies (20.280)

Since this holds for all tb we may drop the subscripts b, thus proving the time independence and thus the martingale nature of e-rstS(t). In mathematical finance, where the expectation value in Eq. (20.280) is written according to the definition (20.189) as lE[crs(tb-ta)eXblxataJ, the martingale property of e-rs(tb-ta)S(t) is expressed as

lE[e- rs(tb- ta)S(tb)lxataJ = S(ta) (20.281) or as lE[e- rs(tb- ta)eXb IXataJ = eXa . For the martingale f(Xb, tb) = e- rs(tb- ta)eXb, the expectation values on the righthand side of the towering formula (20.191) reduces with the help of (20.281) and (20.192) to 1E[IE[f(Xb, tb) Ix tJlxataJ

=

1E[IE[f(x, t) Ix tJlxataJ

=

lE[f(x , t) IXataJ.

(20 .282)

Using once more the relations (20.281) and (20.192) to leads to (20.283) Performing the momentum integral in (20.278) gives the explicit distribution

P rx( Xb tbiXat ) a =-

1

J21m2(tb - t a)

exp { - [Xb - Xa 2(- rx(tb -) taW} . 2(> tb - ta

(20.284)

We may incorporate the discount factor e- rs(tb - ta) into the probability distribution (20.284) and define a martingale distribution for the stock price (20.285) whose normalization falls off like e-rs(tb-ta) . If we define expectation values with respect to pM,rx(Xbtblxata) by the integral (without normalization)

(f(Xb))M,rx ==

JdXb f(Xb)pM,rx(Xbtblxata) ,

(20.286)

then the stock price itself is a martingale: (20.287) Note that there exists an entire family of equivalent martingale distributions 00 dp p(M,r)(Xbtblxata) = e- rllt - exp [ip(Xb - Xa) - 6.t Hr(p)J , (20.288) - 00 211" with an arbitrary rate rand rx == r + i"I(i). Indeed, multiplying this with eXb and integrating over Xb gives rise to a .t

3 - ---, 180°i.-- - - - T - - - - - + - - - --+ 160

- dlogP(xt lxata}}/dx

140 120 100

80 60

40 20

......... -----------------------------------10

20

30

40

50

60

80

70

!:>.t

Figure 20.24 Solid curve showing slope -dlogP(xtlxata}/dx of exponential tail of distribution as a function of time. The dots indicate the analytic short-time approximation (20.369) to the curve (from [79]). This shows that the probability distribution P(xt IXata) has exponential tails (20.361) for large l~xl. Note that in the present limit ,~t » 1 the slopes of the exponential tails are independent of ~t . The presence of Po causes the slopes for positive and negative ~x to be different, so that the distribution P(xt IXata) is not up-down symmetric with respect to price changes. From the definition of Po in Eq. (20.353) we see that this asymmetry increases for a negative correlation p < 0 between stock price and variance. In the second case l~xl « ,v~t/E, by Taylor-expanding y in (20.358) near its minimum and substituting the result into (20.360) , we obtain (20.362)

where N'(~t) = N(~t) exp( -Wolv~t/E2) . Thus, for small l~xl, the probability distribution P(xt IXata) is a Gaussian, whose width 1J2

= (1 - p2 hv ~t

(20.363)

Wo grows linearly with

~t .

The maximum of P(x t IXata) lies at

~xm(t) = ~rs~t,

with

~rs == -

,V

2wo

[1 + 2p(wo -I)] .

(20.364)

E

The position moves with a constant rate ~rs which adds to the average growth rate rs of S(t) removed at the beginning of the discussion. The true final growth rate of S(t) is fs = rs + ~rs·

1492

20 Path Integrals and Financial Markets

The above discussion explains the property of the data in Fig. 20.22 that the logarithmic plots of P(xt Ixat a) are linear in the tails and quadratic near the peaks with the parameters specified in Eqs. (20.361) and (20.362). As time progresses, the distribution broadens in accordance with the scaling form (20 .357) and (20.358) . In the limit f:j.t ---> 00, the asymptotic expression (20.362) is valid for all f:j.x and the distribution becomes a pure Gaussian, as required by the central limit theorem [22] . It is interesting to quantify the fraction f(f:j.t) ofthe total probability contained in the Gaussian portion of the curve. This fraction is plotted in Fig. 20.25. The precise way of defining and calculating the fraction f(f:j.t) is explained in Appendix 20B. The inset in Fig. 20.25 illustrates that the time dependence of the probability density at the maximum Xm approaches f:j.r 1/ 2 for large time, a characteristic property of evolution of Gaussian distributions.

10

2

3

4

5

6

7

8

9

10

11

0.6 0.5 0.4

0.3 0.2

0.1

ooL----4~0----~80L----1~ 20----~ 16~ 0 ----2~00----~ 24~ 0

t>.t Figure 20.25 Fraction f(t:.t) of total probability contained in Gaussian part of P( x t IXata) as function of time interval t:.t. The inset shows the time dependence of the probability density at maximum P(xmt IXata) (points), compared with the falloff ex: t:.r 1/ 2

of a Gaussian distribution (solid curve).

20.4.6

Tail Behavior for all Times

For large If:j.x I, the integrand in (20.347) oscillates rapidly as a function of p, so that the integral can be evaluated in the saddle point approximation of Section 4.2. As in the evaluation of the integral (17.9) we shift the contour of integration in the complex p-plane until it passes through the leading saddle point of the exponent ipf:j.x - f:j.tH(p, f:j.t). To determine its position we note that the function H(p, f:j.t) in Eq. (20.348) has singularities in the complex p-plane, where the argument of the

1493

20.4 Origin of Semi-Heavy Tails

Imp

p+ 1

--------.--------P,

--------~--------

r-------.-------~------~~----~ Rep 0 -iq~

p; P2

Figure 20.26 Singularities of H(p, ~t) in complex p-plane (dots). Circled crosses indicate the limiting positions ±iq"!' of the singularities for 'Y~t » 1. The cross shows the saddle point Ps located in the upper half-plane for ~x > O. The dashed line is the shifted contour of integration to pass through the saddle point Ps. logarithm vanish. These points are located on the imaginary p-axis and are shown by dots in Fig. 20.26. The relevant singularities are those lying closest to the real axis. They are located at the points pt and PI, where the argument of the second logarithm in (20.348) vanishes. Near these zeros, we can approximate H(p, t:.t) by the singular term:

-

2,v

±

H(p, t:.t) ~ C;2t:.t log(p - PI ). With this approximation, the position of the saddle point Ps by the equation 1

• A

tuX

t:.t) I = ut dH(p, d A

P

p=p,

~

2,v - 2 C;

X

{ Ps - PI+, 1

Ps - PI

(20.365)

= Ps(t:.x) is determined t:.x> 0, (20.366)

t:.x < O.

The solutions are indicated in Fig. 20.26 by the cross. The approximation (20.366) is obviously applicable since for a large It:.xl satisfying the condition It:.xprl » C;2 , the saddle point Ps is very close to one of the singularities. Inserting the approximation (20.366) into the Fourier integral (20.347) , we obtain the asymptotic expression for the probability distribution

,v/

t:.x> 0, t:.x < 0,

(20.367)

1494

20 Path Integrals and Financial Markets

qr

where == ~ipf(t) are real and positive. Thus the tails of the probability distribution P(x t IXata) for large l.6.xl are exponential for all times t. The slopes of the logarithmic plots in the tails q± == ~ d log( x t IXata) / dx are determined by the positions Pf of the singularities closest to the real axis. These positions Pf depend on the time interval .6.t. For times much shorter than the relaxation time (,t « 1), the singularities lie far away from the real axis. As time increases, the singularities approach the real axis. For times much longer than the relaxation time (,.6.t » 1) , the singularities approach the limiting points Pf -+ ±iq;, marked in Fig. 20.26 by circled crosses. The limiting values are

q.

±

=

±po +

e

Wo vr=tJ2 -P

for

,.6.t» 1.

(20.368)

The slopes q±(.6.t) approach these limiting slopes monotonously from above. The behavior is shown in Fig. 20.24. The slopes (20.368) are of course in agreement with Eq. (20.361) in the limit ,.6.t » 1. In the opposite limit of short time b.6.t « 1), we find the analytic time behavior

±( ) 2 4, q .6.t :=::;j ±po + Po + e 2 (1 _ p2).6.t

,.6.t« 1.

for

(20.369)

This approximation is shown in Fig. 20.24 as dots.

20.4.7

Path Integral Calculation

Instead of solving the Fokker-Planck equation (20.330) we may also study directly the path integral for the probability distribution using the Hamiltonian (20.321). Note the extra two terms at the end in comparison with the operator expression (20.319) . They account for the symmetric operator order implied by the path integral. The distribution .Pp(Vb, tblvata) at fixed momentum introduced in Eq. (20.329) has the path integral representation

.Pp(Vb, tblvata)

=

JVV

~~v eAp[Pv,vl ,

(20.370)

with the action (20.371) The path integral (20.370) sums over all paths Pv(t) and v(t) with the boundary conditions v(t a) = Va and V(tb) = Vb . It is convenient to integrate the first term in the (20.371) by parts, and to separate H (p, Pv , v) into a v-independent part hvpv - h /2 - ipEp/2 and a linear term [8H (p,Pv , v) /8v ]v. Thus we write

A,[pv, v] = i[pv(tb)Vb - pv(ta)va] - hv

l

tb

ta

- l~b dt [ipv(t) + 8~~)] v(t).

dt Pv(t)(tb - ta) (20.372)

1495

20.4 Origin of Semi-Heavy Tails

Since the path v(t) appears linearly in this expression, we can integrate it out to obtain a delta-functional 8 [Pv(t) - Pv(t)], where Pv(t) is the solution ofthe Hamilton equationpv(t) = i8H/8v(t). This, however, coincides exactly with the characteristic differential equation (20.336) which was solved by Eq. (20.337) with a boundary condition Pv(tb) = Pv' Taking the path integral over Vpv removes the delta-functional, and we find (20.373) where J denotes the Jacobian J

= Det- 1

(i8 + 82~~::~, V)). t

(20.374)

From (20.321) we see that

z, + pEp.

8 2 H(p,pv , v) _- -' 8Pv 8V

(20.375)

According to Formula (18.254), this is equal to

J=

e-(7- i p ta. For non-Gaussian fluctuations with semi-heavy tails, there exists an approximate solution whose Fourier representation is P(x t b b

Ix t ) = e-rw(tb-taljoo a a

- 00

dp eip(Xb-Xale- [R(pl+irxwp]Ub-tal 27r

'

(20.420)

with fI(p) of Eq. (20.141). Recalling the discussion in Section 20.3, this distribution function is recognized as a member of the equivalent family of martingale distributions (20.293) for the stock price S(t) = ex(tl . It is the particular distribution in which the discount factor r coincides with the riskfree interest rate rw.

20.7.3

Option Pricing for Gaussian Fluctuations

For Gaussian fluctuations where H(p) = U 2p2/2, the integral in (20.419) can easily be performed and yields

This probability distribution is obviously the solution of the path integral P(Xbtblxata) = 8(tb - ta)e-rw(tb- tal

J

V x exp {- 2~2lb [x - rxw]2}. (20.422)

The distribution function (20.421) is recognized as the riskfree member of the family of Gaussian martingale distributions (20.285) for the stock price S(t) = ex(tl. It is the particular distribution in which the discount factor r equals the riskfree interest rate rw , i.e., the expression (20.421) coincides with the martingale distribution p(M,rwl(Xbtblxata) of Eq. (20.288), whose explicit form is the expression (20.289) fro r = rw. This distribution is referred to as the risk-neutral martingale distribution. An option is written for a certain strike price E of the stock. The value of the option at its expiration date tb is given by the difference between the stock price on expiration date and the strike price: (20.423)

1506

20 Path Integrals and Financial Markets

where

XE == logE.

(20.424)

The Heaviside function in (20.423) accounts for the fact that only for Sb > E it is worthwhile to execute the option. From (20.423) we calculate the option price at an arbitrary earlier time using the time evolution probability (20.421)

O(Xa, t a) =

i:

dXb O(Xb, tb) p(M,rw)(Xbtb lxata).

(20.425)

Inserting (20.423) we obtain the sum of two terms

O(xa, t a) = OS(xa, t a) - OE(X a, ta),

(20.426)

where

and

OE(X a, t a) = Ee-rw(tb-ta)

1 V27ra 2 (tb- t a)

1 dXb 00

[Xb - Xa - rxw(tb - taW} 2a 2 (tb - t a) .

exp {

XE

(20.428) In the second integral we set _ Xa + rxw(tb - ta) = Xa + ( rW - 2"a 1 X_ =

2) (tb - ta) ,

(20.429)

and obtain (20.430)

OE(X a, t a) = e-rw(tb-ta) E N(y_),

(20.431)

where N(y) is the cumulative Gaussian distribution function

N(y) ==

jY ~e-e/2 , - 00

(20.432)

.J2ir

evaluated at

y-

X_ - XE

va

2

log[S(ta)j E]

va

(tb - ta)

log[S(ta)j E] + (rw -

va

2

(t a

-

!a t b)

2)

2

+ rxw(tb - t a)

(tb - ta) (tb - ta)

(20.433)

20.7 Option Pricing

1507

The integral in the first contribution (20.427) to the option price is found after completing the exponent in the integrand quadratically as follows :

Introducing now (20.435) and rescaling Xb as before, we obtain (20.436) with y+

X+ - XE

10g[S(ta)j E] + (rxw + 0'2) (tb - t a)

V0'2(ta - tb)

V0'2(ta - tb)

10g[S(ta) j E]

+ (rw + ~0'2) (tb -

t a)

(20.437)

V0'2(ta - tb) The combined result (20.438) is the celebrated Black-Scholes formula of option pricing. In Fig. 20.28 we illustrate how the dependence of the call price on the stock price varies with different times to expiration tb - ta and with different volatilities 0'. Floor dealers of stock markets use the Black-Scholes formula to judge how expensive options are, so they can decide whether to buy or to sell them. For a given riskfree interest rate rw and time to expiration tb - t a, and a set of option, stock, and strike prices 0, S, E, they calculate the volatility from (20.438). The result is called the implied volatility, and denoted by I:(x - XE)' As typical plot as a function of x - XE is shown in Fig. 20.29. If the Black-Scholes formula were exactly valid, the data should lie on a horizontal line I:(x - XE) = 0'. Instead, they scatter around a parabola which is called the smile of the option. The smile indicates the presence of a nonzero kurtosis in the distribution of the returns, as we shall see in Subsection 20.7.7. It goes without saying that if the integral (20.427) is carried out over the entire Xb axis, it becomes independent of time, due to the martingale character of the risk-neutral distribution p(M,rw) (Xbtblxata).

1508

40 35 30 25 20 15 10 5

20 Path Integrals and Financial Markets

35

~~ '

1;// // ,

0(8)

//~~" ' / /

/~// /

////,,'

'

~

"

/// / ' -----:

20

---- ---

40

--

80 8

60

40 35 30 25 20 15 10 5

", ,---~~?

5 20

40

~~

~

-

60

80

-

100

E

0(8)

/ /'

--

-

20

40

60

80

8

F igure 20.28 Left: Dependence of call price 0 on stock price S for different times before expiration date (increasing dash length: 1, 2, 3,4, 5 months). T he parameters are E = 50 US$, a = 40%, rw = 6% per mont h . Right : Dependence on t he strike price E for fixed stock price 35 US$ and t he same t imes to expiration (increasing with dash length). Bottom: Dependence on t he volatilities (from left to right : 80%,60%,20%, 10%, 1%) at a fixed time t b - ta = 3 months before expiration.

7.0

.----r~--.___-~--.___-~--.___-~._____,

6.0

5.0

4.0

3.0

L--_~_-------'-

-4.0

_ _~_---'-_ _ __

0.0

-2.0 Xb

- xa

Figure 20.29 Smile deduced from options (see [13]).

--'---_~.L------'

2.0

4.0

1509

20.7 Option Pricing

20.7.4

Option Pricing for Boltzmann Distribution

The above result can easily be extended to price the options for assets whose returns obey the Boltzmann distribution (20.204). Since this is a superposition of Gaussian distributions, we merely have to perform the same superposition over the BlackScholes formula (20.438). Thus we insert (20.438) into the integral (20.204) and obtain the option price for the Boltzmann-distributed assets of variance 0'2:

roo dvv ve -tvja20v( Xa , ta ) ,

_1_ O B( X a , ta ) -_ (~)t 0'2 r(t) Jo

t

(20.439)

where OV(xa , t a ) is the Black-Scholes option price (20.438) of variance v. The superscript indicates that the variance 0'2 in the variables of y+ and y_ in (20.433) and (20.437) is now exchanged by the integration variable v. Since the Boltzmann distribution for the minute data turns rapidly into a Gaussian (recall Fig. 20.15), the price changes with respect to the Black-Scholes formula are relevant only for short-term options running less than a week. The changes can most easily be estimated by using the expansion (20.209) to write (20.440)

20.7.5

Option Pricing for General Non-Gaussian Fluctuations

For general non-Gaussian fluctuations with semi-heavy tails, the option price must be calculated numerically from Eqs. (20.425) and (20.423). Inserting the Fourier representation (20.419) and using the Hamiltonian Hrxw(p)

==

H(p) +irxwp

(20.441)

defined as in (20.141), this becomes O(xa , t a )

=

1

00

dXb (e Xb - eXE)P(Xbtblxata)

XE

e-rw(tb-ta)

1

00

dXb (eXb _ eXE)

XE

1

00

-00

dp eip(Xb-Xa)-Hrxw (P)(tb - ta). (20.442)

21l'

The integrand can be rearranged as follows :

1 1

O( X a , t a ) = e-rw(tb- ta )

00

XE

dXb

00

-00

dp [exaei(P-i)(Xb- xa ) _ eXE eiP(Xb-xa)] e -Hrxw (p)(tb -ta) .

21l'

(20.443) Two integrations are required. This would make a numerical calculation quite time consuming. Fortunately, one integration can be done analytically. For this purpose we change the momentum variable in the first part of the integral from p to p + i and rewrite the integral in the form O(xa , t a )

=

e-rw(tb-ta)

1 1 00

XE

dXb

00

-00

dP eip(Xb-Xa) f(p), 2

1l'

(20.444)

1510

20 Path Integrals and Financial Markets

with (20.445) We have suppressed the arguments X a , XE, tb - ta in f(p), for brevity. The integral over Xb in (20.444) runs over the Fourier transform (20.446)

J:::

of the function f(p) . It is then convenient to express the integral dXb in terms of the Heaviside function e(Xb - XE) as J::"oo dXb e(Xb - XE) and use the Fourier representation (1.308) of the Heaviside function to write (20.447) Inserting here the Fourier representation (20.446), we can perform the integral over Xb and obtain the momentum space representation of the option price (20.448) For numerical integrations, the singularity at p employ the decomposition (18.54) 1 p+zTJ

= 0 is inconvenient. We therefore

P

-. = - - i7rt5(p) , p

(20.449)

to write

We have used the fact that the principal value of the integral over lip vanishes to subtract the constant f(O) from eip(XE - Xa) f(p). After this, the integrand is regular and does not need any more the principal-value specification, and allows for a numerical integration. For Xa very much different from XE, we may approximate

1 dp 00

-

- 00

where E(X)

27r

eip(XE - Xa ) f(p)

p

- f(O)

1

::::>

-E(Xa - xE)f(O), 2

(20.451)

== 2e(x) - 1 is the step function (1.311), and obtain (20.452)

20.7 Option Pricing

1511

-0.4

Figure 20.30 Difference between call price 0(8, t) obtained from truncated Levy distribution with kurtosis fi, = 4 and Black-Scholes price OBs(8, t) with (}'2 = V as function of stock price 8 for different times before expiration date (increasing dash length: 1,2,3,4,5 months). The parameters are E = 50US$, (}' = 40%, rw = 6% per month.

Using (20.418) we have e-Hrxw(i) = e rw , and since e-Hrxw(O) = 1 we see that O(xa , t a) goes to zero for Xa ---> -00 and has the large-xa behavior (20.453) This is the same behavior as in the Black-Scholes formula (20.438) . In Fig. 20.30 we display the difference between the option prices emerging from our formula (20.450) with a truncated Levy distribution of kurtosis K = 4, and the Black-Scholes formula (20.438) for the same data as in the upper left of Fig. 20.28. For truncated Levy distributions, the Fourier integral in Eq. (20.442) can be expressed directly in terms of the original distribution function which is the Fourier transform of (20.52): (20.454) By inspecting Eq. (20.53) we see that the factor tb - ta multiplying Hrxw (p) in (20.442) can be absorbed into the parameters CT, A, 0:, f3 of the truncated Levy distributions by replacing (20.455) Let us denote the truncated Levy distribution with zero average by L~~,a,(3))(x). It is the Fourier transform of e-H(P):

L~~,a,(3)(x)

=

1

00

-00

2dP eipxn:

H(p).

(20.456)

The Fourier transform of e -H(P)(tb - ta ) is then simply given by L~~'~~L)(x) . The additional term rxw in the exponent of the integral (20.442) via (20.441) leads to a drift r xw in the distribution, and we obtain

1

00

- 00

dp eipx-[H(p)+irxwp](tb-ta) = LCA,a,(3)) (x - r (t - t )) 2n: 172(tb - t a) Xw b a '

(20.457)

1512

20 Path Integrals and Financial Markets

Inserting this into (20.419), we find the riskfree martingale distribution to be inserted into (20.442) : (20.458) The result is therefore a truncated Levy distribution of increasing width and uniformly moving average position. Since all expansion coefficients Cn of H (p) in Eq. (20.40) receive the same factor tb - t a, the kurtosis /'i, = C4/ c~ decreases inversely proportional to tb - tao As time proceeds, the distribution becomes increasingly Gaussian, this being a manifestation of the central limit theorem of statistical mechanics. This is in contrast to the pure Levy distribution which has no finite width and therefore maintains its power falloff at large distances. Explicitly, the formula (20.442) for the option price becomes

O(xa, t a) =

e - rw(tb - ta)

1 dXb 00

XE

(e Xb -

eXE)L~~'~~ta) (Xb - Xa - r Xw (tb - ta)) . (20.459)

This and similar equations derived from any of the other non-Gaussian models lead to fairer formulas for option prices.

20.7.6

Option Pricing for Fluctuating Variance

If the fluctuations of the variance are taken into account, the dependence of the price of an option on v(t) needs to be considered in the derivation of a time evolution equation for the option price. Instead of Eq. (20.407) we write the time evolution as

dO = dt

1 [ dt O(x(t)

80

. +. x(t) dt, v(t) + v(t) dt, t + dt) -

80.

80.

1820,2

O(x(t) , v(t), t) ]

8 20. .

1820

.2

(

= 7ft + 8x x + 8v v + "2 8X2 x dt + 8x8v xv dt + "2 8v 2 v dt + .... 20.460

)

The expansion can be truncated after the second derivative due to the Gaussian nature of the fluctuations. We use Ito's rule to replace

x 2 -------+ v(t),

v2 -------+ f 2V(t),

xv

-------+

pw(t) .

(20.461)

These replacements follow directly from Eqs. (20.306) and (20.308) and the correlation functions (20.310) . Thus we obtain

~~ = ~t [O(x(t) + x(t) dt, v(t) + v(t) dt, t + dt) -

O(x(t), v(t), 80 80. 80. 1 8 20 8 20 1 8 20 2 = 7ft + 8x x + 8v v + "2 8x2 V + 8x8v pw + "2 8v2 f V.

t)] (20.462)

This is inserted into Eq. (20.403). If we adjust the portfolio according to the rule (20.404), and use the Ito relation SIS = x + v/2, we obtain the equation [compare (20.409)]

8 00) = - No -80 (.x+ -V2 ) +NoO. Norw ( - + 8x 8x 2

20.7 Option Pricing

1513

As before in Eq. (20.409), the noise in :i; has disappeared. In contrast to the singlevariable treatment, however, the noise variable 'T/v remains in the equation. It can only be removed if we trade a financial asset V whose price is equal to the variance directly on the markets. Then we can build a riskfree portfolio containing four assets

W(t) = Ns(t)S(t) + No(t)O(S, t) + Nv(t)V(t) + NB(t)B(t),

(20.464)

instead of (20.396). Indeed, by adjusting

Nv(t) No(t)

80(S(t) , v(t), t) 8v(t)

(20.465)

we could cancel the term v in Eq. (20.463). There is definitely need to establish trading in such an asset. Without this, we can only reach an approximate freedom ofrisk by ignoring the noise 'T/v(t) in the term v and replacing v by the deterministic first term in the stochastic differential equation (20.308):

v(t)

--->

- , [v(t) - v].

(20.466)

In addition, we can account for the fact that the option price rises with the variance as in the Black-Scholes formula by adding on the right hand side of (20.466) a phenomenological correction term -,\v called price of volatility risk [78, 11]. Such a term has simply the effect of renormalizing the parameters, and v to

" = , +'\,

and

v' = ,v;" .

(20.467)

Thus we find the Fokker-Planck-like differential equation [compare (20.417)]

80

at = rw O -

+,

(V) 80, _, 80 V 8 20 8 20 E2 V 8 20 rw -"2 8x [v(t) - v ]8v - "2 8X2 - pEV 8x8v - 2 8v2 .

(20.468) On the right-hand side we recognize the Hamiltonian operator (20.319), with, and v replaced by,' and v', in terms of which we can write

~~ =

rw (0 - 8x O)

+ (H' + " + pE8x + E2 8v )

O.

(20.469)

The solution of this equation can easily be expressed as a slight modification of the solution P(Xb, Vb, tblxavata) in Eq. (20.340) of the differential equation (20.318). Since it contains only additional first-order derivatives with respect to (20.318), we simply find, with x == Xa and t == t a, the solution PV(Xb, Vb, tblxavata) satisfying the initial condition (20.470)

1514

20 Path Integrals and Financial Markets

as follows : (20.471) where the subscript sh indicates that the arguments

Xb -

Xa

and

Vb - Va

are shifted: (20.472)

The distribution (20.470) may be inserted into an equation of the type (20.425) to find the option price at the time ta from the price (20.423) at the expiration date tb. If we assume the variance Va to be equal to v, and the remaining parameters to be "(*

= 2, v = 0.01,

E

=

O.I ,rw

= 0,

(20.473)

the price of an option with strike price E = 100 one half year before expiration with the stock price S = E (this is called an option at-the-money) is 2.83 US$ for p = -0.5 and 2.81 US$ for p = 0.5. The difference with respect to the Black-Scholes price is shown in Fig. 20.31.

20.7.7

Perturbation Expansion and Smile

A perturbative treatment of any non-Gaussian distributions D(x), which we assume to be symmetric, for simplicity, starts from the expansion (20.474)

0(8, v, t) - OBs(8, t) p = 0.5 0.1

0.05

" 8

- 0.05 -0.1

Figure 20.31 Difference between option price 0(8, V, t) with fluctuating volatility and Black-Scholes price OB8(8, t) with 0"2 = V for option of strike price 100 US$. The parameters are given in Eq. (20.473) . The noise correlation parameter is once p = -0.5 and once p = 0.5. For an at-the-money option the absolute value is 2.83 US$ for p = -0.5 and 2.81 US$ for p = 0.5 (after Ref. [78]).

1515

20.7 Option Pricing where

(20.475) which can also be expressed as a series

(20.476)

(20.477) The quantities cn contain the kurtosis K, in the case of a truncated Levy distribution with the powers r:;nj2 - 1. If the distribution is close to a Gaussian, we may re-expand all expressions in powers of the higher cumulants. In the case of a truncated Levy distribution, we may keep systematically all terms up to a certain maximal power of r:; and find D(x)

{1_;:2(~_~+2~~+~~+5~~~_~~~)

=

X4 ( C4

+ 0'4

17c~

Ct;

24 - 48 +

96 +

Cs

C4 Ct;

239c1 )

192 + 288 - 9216

x 6 ( C6 7c~ Cs C4Ct; 7c1 ) 720 - 288 - 1440 - 5760 + 2304

+ 0'6

xs ( ~

+ O's

Cs c~ ) 1152 + 40320 - 9216 +. . .

} e -x2j2u2

(20.478)

J27l'uf '

where we have introduced the modified width -2 = 0'1 -

2 (1 _ 0'

~ C6 4 + 24

_

13c~ _ Cs 96 192

_ C4 C6 _ 31c1

64

512 +...

)

.

(20.479)

(20.480)

(20.481)

1516

20 Path Integrals and Financial Markets

(20.482) The exponential can be re-expressed compactly in the quasi-Gaussian form _ D(x) =

1

V27r.,a) (x) = a e-2a ,,2

dYeiaYXea [(l - iY)'+(l+iy)' 1 271" •

(20A.l)

Expanding the last exponential in a Taylor series, we obtain

J 00

ae- 2a

- 00

ae- 2a JOO dYeiaXy 271" -00

00

dy . an _e,yax ' " -[(1 - iyl 271" ~ n!

+ (1 + iy)>.t =

n=O

f n= O

an

n!

t (n) m= O

(1 _ iy)>.(n- ml(1 + iy)>.m

m

(20A.2)

containing binominal coefficients. Changing the order of summations yields

(20A.3)

This can be written with the help of the Whittaker functions (20.30) as (20A.4)

1518

20 Path Integrals and Financial Markets

After converting the Gamma functions of negative arguments into those of positive arguments, this becomes -(A,a) L(72

_ a -2a~ n..[!--. 2An / 2f(1+Am) sin(7rAm) (x) - --e L.Ja L.J ( )l+An/ 2 I( _ )1 WAn/2-Am ,(An+l)/2(2ax) , 7r n= l m=O ax m. n m.

(20A.5) The Whittaker functions W A,'1'(x ) have the following asymptotic expansion (20A.6) For

r == (An +

1)/2 and A == A(n - 2m)/2, the product takes the form

k

II[r 2 - (A -

+ 1/2)2]

j

j=l k

II (Am + j)(An+1 -

(20A.7)

Am - j).

j=l

Inserting this into (20A.6) and the result into (20A.5) , we obtain the asymptotic expansion for large x : 1 _2ae- ax ~ n2 An ..[!--. f(1+Am)Sin(7rAm)(2 )- Am - - e - - L.J a L.J ax 7r X n= l m= l m!(n - m)!

j}A ,a)( X ) (72

x

[

~ m =1(Am+ j )(An+1-Am- j )]

1 + L.J

k!(2ax)k

.

(20A.8)

k=l

We have raised the initial value of the index of summation m by one unit since sin( 7r Am) vanishes for m = O. If we define a product of the form n~=l to be equal to unity, we can write the term in the last bracket as

...

1 k!(2ax)k

00

L

k=O

k

II (j + Am)(l -

+ An) .

Am - j

(20A.9)

j=l

Rearranging the double sum in (20A.8) , we write L~~,a)(x) as (' )

L_~,a (x)

1 _2a e- ax 00 f(l+Am)sin(7rAm) 00 an2An '" '" -x- ~ m!(2ax)Am n~ (n - m)!

=

- ;e

v

00

X {;

1 k!(2ax)k

D k

(j

+ Am)(l- Am - j + An) ,

and further as 1 _2a Cax ~ f(l + Am)sin(7rAm) (2 )- Am - ;e ~ L.J m! ax m=l 00

x

00

~(2Aa)n+m~ {; k!(2~X)k

D k

(j

+ Am)(l- j + An) =

(20A.1O)

1519

Appendix 20A Large-x Behavior of Truncated Levy Distribution

f:

f:

_~ e _2a e-ax 1 f(I+Am)sin(7rAm)(2'\a)m 7r x k=Ok!(2ax)k m=l (2ax),\mm! k

x

II (j' + Am) L 00

j'=1

(2'\a)n

----;J

n= O

k

II (An + 1 -

j) .

(20A.Il)

j= l

The last sum over n in this expression can be re-expressed more efficiently with the help of a generating function (20A.12) whose Taylor series is (20A.13) leading to

x (20A.14) This can be rearranged to

(20A.15) Thus we obtain the asymptotic expansion 0.1 II 0.08

,

I;

0.06 "

\

\

\\

L;;"a) (K)(X)

',\1

0.04

',"~

"" ~

" ,, ::,:~::

0.02

~--~--~~~~~~~2 x 1.2

1.4

1.8

Figure 20.32 Comparison of large-x expansions containing different numbers of terms (with K = 0, 1, 2, 3, 4, 5, with increasing dash length) with the tails of the truncated Levy distribution for A = V2, (]' = 0.5, Q = 1.

1520

20 Path Integrals and Financial Markets

(20A.16) where B

_ r(I+Am+k)sin(7rAm)

m!

km-

.

(20A.17)

We shall denote by L~~,a)(K)(x) the approximants in which the sums over K are truncated after the Kth term. To have a definite smallest power in lx i, the sum over m is truncated after the smallest integer larger than (K - k + A)/A. The leading term is

_ e- 2a e-ax !(O)( -8(2af') r(1 + A) sin(7rA)( -8) 7r xl+A ea's(2- 2') r(1 + A)sin(7rA) -ax 8 7r Xl+A e

(20A.18)

The large-x approximations L~~,a)(o)(x) are compared with the numerically calculated truncated Levy distribution in Fig. 20.32.

Appendix 20B

Gaussian Weight

For simplicity, let us study the Gaussian content in the final distribution (20.347) only for the simpler case p = O. Then we shift the contour of integration in (20.347) to run along p + i / 2, and we study the Fourier integral

P(x t IXata) = e-6.X/2j+00 2dP eip6.x-H(p,6.t) , -00 7r where

_

H(P,t:..t) =

(20B.l)

,2vt + 2, v In [cosh-Of + 0 2+,2.smh -Ot] ,

- -2-

'"

-2

'"

--n-

2

2,,,

2

(20B.2)

and (20B.3) The function H(p, t:..t) is real and symmetric in p. The integral (208.1) is therefore a symmetric function of t:..x. The only source of asymmetry of P(x t Ixat a) in t:..x is the exponential prefactor in (20B.l) . Let us expand the integral in (208.1) for small t:..x:

P(x t IXata)

>::J

e-6.x/2 [ItO -

~lt2(t:..X)2]

>::J

Ito e-1l.x/2e-1'2(1l.x)2 /21'0 ,

(20B.4)

where the coefficients are the first and the second moments of exp[H(p, t:..t)]

J ~~

+00

Ito(t:..t) =

e- H(p,1l.t) ,

(20B.5)

If we ignore the existence of semi-heavy tails and extrapolate the Gaussian expression on the righthand side to t:..x E (-00,00), the total probability contained in such a Gaussian extrapolation will be the fraction

(20B.6)

Appendix 20C Comparison with Dow-Jones Data

1521

This is always less than 1 since the integral (20B.6) ignores the probability contained in the semiheavy tails. The difference 1 - f(llt) measures the relative contribution of the semi-heavy tails. The parameters J.1.o (Ilt) and J.1.2 (Ilt) are calculated numerically and the resulting fraction f (Ilt) is plotted in Fig. 20.25 as a function of Ilt . For Ilt --t 00, the distribution becomes Gaussian, whereas for small Ilt , it becomes a broad function of p.

Appendix 20C

Comparison with Dow-Jones Data

For the comparison of the theory in Section 20.4 with actual financial data shown in Fig. 20.22, the authors of Ref. [79] downloaded the daily closing values of the Dow-Jones industrial index for the period of 20 years from 1 January 1982 to 31 December 2001 from the Web site of Yahoo [108]. The data set contained 5049 points S(t n ), where the discrete time variable tn parametrizes the days. Short days before holidays were ignored. For each tn , they compiled the log-returns Il x(tn ) = InS(tn+l)/S(tn). Then they partitioned the x-axis into equally spaced intervals of width Ilx and counted the number of log-returns Ilx(t n ) falling into each interval. They omitted all intervals with occupation numbers less than five, which they considered as too few to rely on. Only less than 1% of the entire data set was omitted in this way. Dividing the occupation number of each bin by Ilr and by the total occupation number of all bins, they obtained the probability density for a given time interval Ilt = 1 day. From this they found p(DJ)(xt Ixat a) by replacing

Ilx

--t

Ilx - rsllt.

Assuming that the system is ergodic, so that ensemble averaging is equivalent to time averaging, they compared p(DJ) (x t IXata) with the calculated P(x t IXata) in Eq. (20.347). The parameters of the model were determined by minimizing the mean-square deviation LL>x L>t 1I0g p(DJ) (x t IXata) - log P(x t IXataW , with the sum taken over all available Ilx and over ~t = 1, 5, 20, 40, and 250 days. These values of Ilt were selected because they represent different regimes: 'Yllt« 1 for t = 1 and 5 days, 'Yilt"'" 1 for t = 20 days, and 'Yilt» 1 for t = 40 and 250 days. As Figs. 20.22 and 20.23 illustrate, the probability density P(x t IXata) calculate from the Fourier integral (20.347) with components (20.348) agrees with the data very well, not only for the selected five values of time t , but for the whole t ime interval from 1 to 250 trading days. The comparison cannot be extended to Ilt longer than 250 days, which is approximately 1/20 of the entire range of the data set, because it is impossible to reliably extract p(DJ) (x t IXata) from the data when Ilt is too long. The best fits for the four parameters 'Y, V, c:, J.1. are given in Table 20.1. Within the scattering of the data, there are no discernible differences between the fits with t he correlation coefficient p being zero or slightly different from zero. Thus the correlation parameter p between the noise terms for stock price and variance in Eq. (20.310) is practically zero. This conclusion is in contrast with the value p = - 0.58 found in [109] by fitting the leverage correlation function introduced in [110]. Further study is necessary to understand this discrepancy. All theoretical curves shown in the above figures are calculated for p = 0, and fit the data very well. The parameters 'Y, V, c:, J.1. have the dimensionality of l/time. One row in Table 20.1 gives their values in units of l/day, as originally determined in our fit . The other row shows the annualized values of the parameters in units of l/year, where one year is here equal to the average number of 252.5 trading days per calendar year. The relaxation time of variance is equal to Ih = 22.2 trading days = 4.4 weeks"", 1 month, where 1 week = 5 trading days. Thus one finds that the variance has a rather long relaxation time, of the order of one month, which is in agreement with an earlier conclusion in Ref. [109]. Using the numbers given in Table 20.1, the value of the parameter c: 2/2'Yv is "'" 0.772 t hus satisfying the smallness condition Eq. (20.322) ensuring that v never reaches negative values. The stock prices have an apparent growth rate determined by the position Xm(t) where the probability density is maximal. Adding this to the initially subtracted growth rate rs we find that the apparent growth rate is f's = rs - 'Yv/2wo = 13% per year. This number coincides with the apparent average growth rate of the Dow-Jones index obtained by a simple fit of the data points St n with an exponential function of tn. The apparent growth rate f's is comparable to the

1522

20 Path Integrals and Financial Markets

Table 20.1 Parameters of equations with fluctuating variance obtained from fits to DowJones d ata. The fit yields p ~ 0 for the correlation coefficient and 1; ' = 22.2 trading days for the relaxation time of variance. Units l/day l/year

'Y

4.50 x 10 -· 11.35

v 8.62 x 10 0.022

e: -0

2.45 x 10 0.618

J..l -0

5.67 x 10 - 4 0.143

vrv

= 14.7%. Moreover, the parameter (20.328) which average stock volatility after one year (T = characterizes the width of the stationary distribution of variance is equal to Vrnax/W = 0.54. This means that the distribution of variance is broad, and variance can easily fluctuate to a value twice greater than the average value v. As a consequence, even though the average growth rate of the stock index is positive, there is a substantial probability of J~oo df).xP(xt IXata) ~ 17.7% to have negative growth for f).t = 1 year. According to (20.368), the asymmetry between the slopes of exponential tails for positive and negative f).x is given by the parameter Po , which is equal to 1/2 when p = 0 [see also the discussion of Eq. (20B.l) in Appendix 20B]. The origin of this asymmetry can be traced back to the transformation from S(t)/S(t) to x(t) using Ito's formula. This produces a term v(t)/2 in Eq. (20.306), which leads to the first term in t he Hamiltonian operator (20.319). For p = 0 this is the only source of asymmetry in f).x of P(xt Ixata ). In practice, the asymmetry of the slopes Po = 1/2 is quite small (about 2.7%) compared to the average slope ~ wo/e: = 18.4.

q;

Notes and References Option pricing beyond Black and Scholes via path integrals was discussed in Ref. [111], where strategies are devised to minimize risks in the presence of extreme fluctuations, as occur on real markets. See also Ref. [11 2]. Recently, a generalization of path integrals to functional integrals over surfaces has been proposed in Ref. [113] as an alternative to the Heath-Jarrow-Morton approach of modeling yield curves (see http : //risk.ifci.ch/OOOl1661.htm). The individual citations refer to [1] The development of the Dow Jones industrial index up to date can be found on internet sites such as http://stockcharts.com/ charts/historical/ dj ia1900. html. Alternatively they can be plotted directly using Stephen Wolfram's program Mathematica (v.6) using the small program t=FinancialData["-DJI" ,All];l = Length[t] tt = Table[t[[kJ] [[1]] [[l]]+(t[[k]] [[1]] [[2]]-1)/12,t[[kJ] [[2]], {k,l,l}] ListLogPlot [tt] Note that the last stagnation period was predicted in the 3rd edition of this book in 2004. [2] P. Fizeau, Y . Liu, M. Meyer, C.-K. Peng, and H.E . Stanley, Volatility Distribution in the Sf3P500 Stock Index, Physica A 245, 441 (1997) (cond-mat/9708143). [3] B.E. Baaquie, A Path Integral Approach to Option Pricing with Stochastic Volatility: Some Exact Results , J . de Physique I 7, 1733 (1997) (cond-mat/9708178) . [4] J.P. Fouque, G. Papanicolaou, and K.R. Sircar, Derivatives in Financial Markets with Stochastic Volatility (Cambridge University Press, Cambridge, 2000); International Journal of Theoretical and Applied Finance, 3, 101 (2000). [5] J. Hull and A. White, Journal of Finance 42,281 (1987); C.A. Ball and A. Roma, Journal of Financial and Quantitative Analysis 29, 589 (1994); R. Schobel and J . Zhu, European Finance Review 3, 23 (1999).

Notes and References

1523

[6] E.M. Stein and J.C. Stein, Review of Financial Studies 4, 727 (1991). [7] J .C. Cox, J .E. Ingersoll, and S.A. Ross, Econometrica 53, 385 (1985) . In the mathematical literature such an equation runs under the name Feller process. See Ref. [4] and W. Feller, Probability Theory and its Applications, sec. ed., Vol. 11., John Wiley & Sons, 1971. See also related work in turbulence by B. Holdom, Physica A 254,569 (1998) (cond-mat/9709141) . [8] P. Wilmott, Derivatives, John Wiley & Sons, New York, 1998. [9] R. Courant and D. Hilbert, Methods of Mathematical Physics, vol. 2 (John Wiley & Sons, New York, 1962). [10] C.M. Bender and S.A. Orszag, Advanced Mathematical Methods for Scientists and Engineers, Springer, New York, 1999. [11] C.G. Lamoureux and W.D. Lastrapes, Forecasting Stock-Return Variance: Toward an Understanding of Stochastic Implied Volatilities, Rev. of Financial Studies 6, 293 (1993). See also D.T. Breeden, An Intertemporal Asset Pricing Model with Stochastic Consumption and Invertment Opportunities, Jour. of Financial Economics 7, 265 (1979) . [12] This is named after the Russian mathematician Andrei Andrejewitsch Markov who published in 1912 the German book on Probalility Theory cited on p. 1081 with the German spelling Markoff. [13] J.-P. Bouchaud and M. Potters, Theory of Financial Risks , Prom Statistical Physics to Risk Management, Cambridge University Press, 2000. [14] J. Hull, Options, FUtures and Other Derivatives , Prentice-Hall Int., 1997. [15] R. Rebonato, Interest-Rate Option Models, John Wiley & Sons, Chichester, 1996. [16] M.W. Baxter and A.J.O. Rennie, Financial Calculus, Cambridge Univ. Press, Cambridge, 1996. [17] J. Voigt, The Statistical Mechanics of Financial Markets, Springer, Berlin, 2001. [18] V. Pareto, Giornale degli Economisti, Roma, January 1895; and Cours d'economie politique, F . Rouge Editeur, Lausanne and Paris, 1896; reprinted in an edition of his complete works (Vol III) under the title Ecrits sur la courbe de la repartition de la richesse , Librairie Droz, Geneva, 1965 (http://213 .39 .120 .146: 8200/droz/FMPro). [19] B.B. Mandelbrot, Jilraetals and Scaling in Finance, Springer, Berlin, 1997; J. of Business 36, 393 (1963) . [20] T. Lux, Appl. Financial Economics 6, 463 (1996); M. Loretan and P.C.B. Phillips, J. Empirical Finance 1, 211 (1994). [21] J .P. Nolan, Stable Distributions, American University (Wahington D.C.) lecture notes 2004 http://academic2.american.edu/-jpnolan/stable/chap1.pdf. [22] R.N. Mantegna and H.E. Stanley, Stochastic Process with Ultraslow Convergence to a Gaussian: The Truncated Levy Flight , Phys. Rev. Lett. 73, 2946 (1994). [23] P. Gopikrishnan, M. Meyer, L.A.N. Amaral, and H.E. Stanley, Europ. Phys. Journ. B 3, 139 (1998) ; P. Gopikrishnan, V. Plerou, L.A.N. Amaral, M. Meyer, and H.E. Stanley, Scaling of the distribution of fluctuations of financial market indices , Phys. Rev. E 60, 5305 (1999) ; V. Plerou, P. Gopikrishnan, L.A.N. Amaral, M. Meyer, and H.E. Stanley, Scaling of the

1524

20 Path Integrals and Financial Markets distribution of price fluctuations of individual companies , Phys. Rev. E 60, 6519 (1999); V. Plerou, P. Gopikrishnan, X. Gabaix, H.E. Stanley, On the Origin of Power-Law Fluctuations in Stock Prices, Quantitative Finance 4 , C11 (2004).

[24] 1. Koponen, Analytic Approach to the Problem of Convergence of Truncated Levy Flights Towards the Gaussian Stochastic Process, Phys. Rev. E 52, 1197-1199 (1995). [25] 1.S. Gradshteyn and 1.M. Ryzhik, op. cit., Formulas 3.382.6 and 3.382.7. [26] ibid., op. cit., Formula 3.384.9. [27] ibid., op. cit., Formula 9.220.3 and 9.220.4. [28] This result was calculated by A. Lyashin when he visited my group in Berlin. Note that other authors have missed the prefactor in the asymptotic behavior (20.37), for example A. Matacz, Financial Modeling and Option Theory with the Truncated Levy Process, condmat/9710197. The prefactor can be dropped only for a = O. [29] ibid., op. cit., Formulas 3.462.3. [30] ibid., op. cit., Formula 9.246. [31] ibid., op. cit., Formula 9.247.2. [32] Exponential short-time behavior has been observed in Ref. [13] and numerous other authors: L.C. Miranda and R Riera, Physica A 297, 509 (2001) ; J.L. McCauley and G.H. Gunaratne, Physica A 329, 178 (2003); T . Kaizoji, Physica A 343, 662 (2004) ; R Remer and R Mahnke, Physica A 344,236 (2004); D. Makowiec, Physica A 344,36 (2004) ; K. Matia, M. Pal, H. Salunkay, and H.E. Stanley, Europhys. Lett. 66, 909 (2004); A.C. Silva, RE. Prange, and V.M. Yakovenko, Physica A 344,227 (2004) ; R Vicente, C.M. de Toledo, V.B.P. Leite, and N. Caticha, Physica A 361,272 (2006) ; A.C. Silva and V.M. Yakovenko, (physics/0608299) . [33] H. Kleinert and T. Xiao-Jiang to be published. [34] B. Grigelionis, Processes of Meixner Type , Lith. Math. J. 39, 33 (1999) . [35] W. Schoutens, Meixner Processes in Finance, Report 2001-002, EURANDOM, Eindhoven (wvw.eurandom.tue.nl/reports/2001/002wsreport.ps). [36] O. Barndorff-Nielsen, T . Mikosch, S. Resnick, eds., Levy Processes tions Birkhiiuser, 2001.

Theory and Applica-

[37] O. Barndorff-Nielsen, Infinite Divisibility of the Hyperbolic and Generalized Inverse Gaussian Distributions, Zeitschrift fur Wahrscheinlichkeitstheorie und verwandte Gebiete 38, 309-312 (1977). [38] O. Barndorff-Nielsen, Processes of Normal Inverse Gaussian Type, Finance & Stochastics, 2, No.1, 41-68 (1998). [39] O. Barndorff-Nielsen, Normal Inverse Gaussian Distributions and Stochastic Volatility Modeling, Scandinavian Journal of Statistics 24, 1-13 (1977) . [40] O. Barndorff-Nielsen, N. Shephard, Modeling with Levy Processes for Financial Econometrics, MaPhySto Research Report No. 16, University of Aarhus, (2000). [41] O. Barndorff-Nielsen, N. Shephard, Incorporation of a Leverage Effect in a Stochastic Volatility Model, MaPhySto Research Report No. 18, University of Aarhus, (1998). [42] O. Barndorff-Nielsen, N. Shephard, Integrated Ornstein Uhlenbeck Processes , Research Report, Oxford University, (2000).

Notes and References

1525

[43] J. Bertoin, (1996) Levy Processes, Cambridge University Press. [44] J. Bretagnolle, Processus a increments independants , Ecole d'Ete de Probabilites, Lecture Notes in Mathematics, Vol. 237, pp 1-26, Berlin, Springer, (1973). [45] P.P. Carr, D. Madan, Option Valuation using the Fast Fourier Transform, Journal of Computational Finance 2, 61-73 (1998). [46] P.P. Carr, H. Geman, D. Madan, M. Yor, The Fine Structure of Asset Returns: an Empirical Investigation, Working Paper, (2000). [47] T. Chan, Pricing Contingent Claims on Stocks Driven by Levy Processes, Annals of Applied Probability 9, 504-528, (1999). [48] R. Cont, Empirical Properties of Asset Returns: Stylized Facts and Statistical Issues , Quantitative Finance 1, No.2, (2001) . [49] R. Cont, J .-P. Bouchaud, M. Potters, Scaling in Financial Data: Stable Laws and Beyond, in B. Dubrulle, F. Graner & D. Sornette (eds.): Scale invariance and beyond, Berlin, Springer, (1997) . [50] ibid., op. cit., Formula 8.486.10. [51] See http://www.fdm.uni-freiburg.de/UK. [52] P. Jizba, H. Kleinert, Superposition of Probability Distributions, Phys. Rev. E 78, 031122 (arXiv:0802.069) . [53] G. Arfken, Mathematical Methods for Physicists, 3rd ed., Academic Press, Orlando, FL. See t15.12 Inverse Laplace Transformation", pp. 853-861 , 1985. [54] E. Post, Trans. Amer. Math. Soc. 32 (1930) 723. [55] H. Kleinert, Stochastic Calculus for Assets with Non-Gaussian Price Fluctuations, Physica A 311 , 538 (2002) (cond-mat/0203157). [56] R.F. Pawula, Phys. Rev. 162, 186 (1967). [57] See http://www . physik. fu-berlin. derkleinert/b5/f iles [58] L. Borland, A Theory of Non-Gaussian Option Pricing, Quantitative Finance 2, 415 (2002) (cond-mat/0205078). [59] C. Tsallis, J. Stat. Phys. 52,479 (1988); E.M.F. Curado and C. Tsallis, J. Phys. A 24, L69 (1991) ; 3187 (1991); A 25, 1019 (1992). [60] C. Tsallis, C. Anteneodo, L. Borland, R. Osorio, Nonextensive Statistical Mechanics and Economics, Physica A 324,89 (2003) (cond-mat/030130). [61] P. Praetz The Distribution of Share Price Changes , Journal of Business 45, 49 (1972). [62] R. Blattberg and N. Gonedes, A Comparison of the Stable and Student Distributions as Statistical Models of Stock Prices, Journal of Business 47, 244 (1972). [63] R.S. Liptser and A.N. Shiryaev, Theory of Martingales, Kluwer, 1989. [64] D. Duffie, Dynamic asset pricing theory , Princeton University Press, 2001, p. 22. [65] J .M. Steele, Stochastic Calculus and Financial Applications , Springer, New York, 2001 , p. 50. [66] H. Kleinert, Option Pricing from Path Integral for Non-Gaussian Fluctuations. Natural Martingale and Application to Truncated Levy Distributions, Physica A 312, 217 (2002) (cond-mat/0202311).

1526

20 Path Integrals and Financial Markets

[67] F. Esscher, On the Probability FUnction in the Collective Theory of Risk , Skandinavisk Aktuarietidskrift 15, 175 (1932) . [68] H.U. Gerber and E .S.W. Shiu, Option Pricing by Esscher Transforms , Trans. Soc. Acturaries 46, 99 (1994) . [69] J .M. Harrison and S.R Pliska, Martingales and Stochastic Integrals in the Theory of Continuous Trading, Stoch. Proc. Appl. 11, 215 (1981) ; A Stochastic Calculus Model of Continuous Trading Complete Markets, 13, 313 (1981); [70] E. Eberlein, J. Jacod, On the Range of Options Prices, Finance and Stochastics 1 , 131, (1997). [71] E. Eberlein, U. Keller, Hyperbolic Distributions in Finance , Bernoulli 1, 281-299, (1995). [72] K Prause, The Generalized Hyperbolic Model: Estimation, Financial Derivatives, and Risk Measures , Universitat Freiburg Dissertation, 1999 (http://t,IT,lw . freidok. unifreiburg . de/volltexte/15/pdf/15_1.pdf). [73] E. Eberlein, U. Keller, K Prause, New Insights into Smile, Mispricing and Value at Risk: the Hyperbolic Model, Journal of Business 71 , No.3, 371-405, (1998) . [74] E . Eberlein, S. Raible, Term Structure Models Driven by General Levy Processes , Mathematical Finance 9 , 31-53, (1999) . [75] See the Wikipedia page en.wikipedia.org/wiki/Levy_skew_alpha-stable_distribution. [76] S. Raible, Levy Processes in Finance, Ph.D. Thesis, Univ. Freiburg, (http://Wt.IT.l.freidok.uni-freiburg.de/volltexte/15/pdf/51_1.pdf).

2000

[77] M.H.A. Davis, A General Option Pricing Formula , Preprint, Imperial College, London (1994). See also Ref. [70] and T. Chan, Pricing Contingent Claims on Stocks Driven by Levy Processes, Ann. Appl. Probab. 9, 504 (1999). [78] S.L. Heston, A Closed-Form Solution for Options with Stochastic Volatility with Applications to Bond and Currency Options, Review of Financial Studies 6, 327 (1993). [79] A.A. Dragulescu and V.M. Yakovenko, Quantitative Finance 2, 443 (2002) (condmat / 0203046) . [80] P. Jizba, H. Kleinert, and P. Haener, Perturbation Expansion for Option Pricing with Stochastic Volatility, Berlin preprint 2007 (arXiv:0708.3012). [81] S. Micciche, G. Bonanno, F . Lillo, RN. Mantegna, Physica A 314, 756 (2002) . [82] D. Valenti, B. Spagnolo, and G. Bonanno, Physica A 314, 756 (2002) . [83] H. Follmer and M. Schweizer, Hedging and Contingent Claims under Incomplete Information, in Applied Stochastic Analysis, edited by M.H.A. Davis and RJ. Elliot, 389 Gordon and Breach 1991. [84] RF. Engle, Autoregressive conditional heteroskedasticity with estimates of the variance of UK inflation , Econometrica 50, 987 (1982); Dynamic conditional correlation-A simple class of multivariate GARCH models , J . of Business and Econ. Stat., 20, 339 (3002) (http://Wt.IT.l.physik.fu-berlin.de/-kleinert/finance/englel.pdf); RF. Engle and KF. Kroner, Multivariate simultaneous generalized ARCH, Econometric Theory 11, 122 (1995); RF. Engle and K Sheppard, Theoretical and empirical properties of dynamic conditional correlation multivariate GARCH, Nat. Bur. of Standards working paper No. 8554 (2001)

Notes and References

1527

(http://wvw.nber.org/papers/w8554.pdf). See also the GARCH toolbox for calculations at http://wvw.kevinsheppard.com/ research/ucsd_garch/ucsd_garch.aspx. [85] T. Bollerslev Generalized autoregressive conditional heteroskedasticity, J. of Econometrics 31,307 (1986) ; Modelling the coherence in short-run nominal exchange rates: A multivariate generalized ARCH model, Rev. Economics and Statistics 72, 498 (1990). [86] L. Bachelier, Theorie de la Speculation, L. Gabay, Sceaux, 1995 (reprinted in P. Cootner (ed.), The Random Character of Stock Market Prices, MIT Press, Cambridge, Ma, 1964, pp.17- 78. [87] A. Einstein, Uber die von der molakularkinetischen Theorie der Wiirme geforderte Bewegung von in ruhenden Fliissigkeiten suspendierten Teilchen, Annalen der Physik 17, 549 (1905). [88] N. Wiener, Differential-Space, J . of Math. and Phys. 2, 131 (1923) . [89] F. Black and M. Scholes, J. Pol. Economy 81, 637 (1973) . [90] R.C. Merton, Theory of Rational Option Pricing, Bell J. Econ. Management Sci. 4, 141 (1973) . [91] These papers are available as CNRS preprints CPT88/PE2206 (1988) and CPT89/PE2333 (1989) . Since it takes some effort to obtain them I have placed them on the internet where they can be downloaded as files dashl.pdf and dash2.pdf from http://wvw . physik.fu-berlin.de/-kleinert/b3/papers. [92] S. Fedotov and S. Mikhailov, preprint (cond-mat/980610l). [93] M. Otto, preprints (cond-mat/9812318) and (cond-mat/9906196). B.E. Baaquie, L.C. Kwek, and M. Srikant, Simulation of Stochastic Volatility using Path Integration: Smiles and Frowns, cond-mat/0008327 [94] R. Cond, Scaling and Correlation in Financial Data, (cond-mat/9705075) . [95] A.J. McKane, H.C. Luckock, and A.J. Bray, Path Integrals and Non-Markov Processes. l. General Formalism, Phys. Rev. A 41, 644 (1990) ; A.N. Drozdov and J.J. Brey, Accurate Path Integral Representations of the Fokker-Planck Equation with a Linear Reference System: Comparative Study of Current Theories , Phys. Rev. E 57, 146 (1998) ; V. Linetsky, The Path Integral Approach to Financial Modeling and Options Pricing, Computational Economics 11 129 (1997); E.F. Fama, Efficient Capital Markets, Journal of Finance 25, 383 (1970) . A. Pagan, The Econometrics of Financial Markets , Journal of Empirical Finance 3 , 15 (1996) . C.W.J. Granger, Z.X. Ding, Stylized Facts on the Temporal Distribution Properties of Daily Data from Speculative Markets, University of San Diego Preprint, 1994. [96] H. Geman, D. Madan, M. Yor, Time Changes in Subordinated Brownian Motion , Preprint, (2000) . [97] H. Geman, D. Madan, M. Yor, Time Changes for Levy Processes , Preprint (1999). [98] J. Jacod, A.N. Shiryaev, Limit Theorems for Stochastic Processes , Berlin, Springer, (1987). [99] P. Levy, Theorie de l'addition des variables aliatoires , Paris, Gauthier Villars, (1937) . [100] D. Madan, F. Milne, Option Pricing with Variance Gamma Martingale Components, Mathematical finance, 1, 39-55, (1991). [101] D. Madan, P.P. Carr, E.C. Chang, The Variance Gamma Process and Option Pricing, European Finance Review 2, 79-105, (1998).

1528

20 Path Integrals and Financial Markets

[102] D. Madan, E. Seneta, The Variance Gamma Model for Share Market Returns , Journal of Business 63, 511-524, (1990) . [103] B.B. Mandelbrot, Fractals and Scaling in Finance , Berlin, Springer, (1997) . [104] P. Protter, Stochastic Integration and Differential Equations: a new approach , Berlin, Springer, (1990) . [105] T .H. Rydberg, The Normal Inverse Gaussian Levy Process: Simulation and Approximation, Cornmun. Stat., Stochastic Models 13(4), 887-910 (1997). [106] G. Samorodnitsky, M. Taqqu, Stable Non-Gaussian Random Processes, New York, Chapman and Hall (1994). [107] K. Sato, Levy Processes and Infinitely Divisible Distributions, Cambridge University Press, (1999). [108] Yahoo Finance http://finance.yahoo.com. To download data, enter in the symbol box: -OJ!, and then click on the link: Download Spreadsheet. [109] J . Masoliver and J . Pere1l6, Physica A 308, 420 (2002) (cond-mat/0111334) ; Phys. Rev. E 67, 037102 (2003) (cond-mat/0202203); (physics/0609136). [110] J .-P. Bouchaud, A. Matacz, and M. Potters, Phys. Rev. Letters 87, 228701 (2001). [111] J .-P. Bouchaud and D. Sornette, The Black-Scholes Option Pricing Problem in mathematical finance: Generalization and extensions for a large class of stochastic processes , J. de Phys. 4, 863 (1994); Reply to Mikheev's Comment on the Black-Scholes Pricing Problem, J . de Phys. 5, 219 (1995) ; [112] J .-P. Bouchaud, G. lori, and D. Sornette, Real- World Options , Risk 9, 61 (1996) (http://xxx.lanl.gov/abs/cond-mat/9509095); J .-P. Bouchaud, D. Sornette, and M. Potters, Option Pricing in the Presence of Extreme Fluctuations, in Mathematics of Derivative Securities, ed. by M.A.H. Dempster and S.R. Pliska, Cambridge University Press, 1997, pp. 112-125; The Black-Scholes Option Pricing Problem in Mathematical Finance: Generalization and Extensions for a Large Class of Stochastic Processes, J . de Phys. 4, 863 (1994). [113] P. Santa-Clara and D. Sornette, The Dynamics of the Forwam Interest Rate Curve with Stochastic String Shocks , Rev. of Financial Studies 14, 149 (2001) (cond-mat / 9801321) .

Index hyperspherical harmonics, 718 Legendre polynomials, 712 spherical harmonics, 712 trigonometric, 726 ADELMAN, S.A. , 1364 adjoint Hermitian operator, 16 adjoint representation, 746 advanced Green function , 1268 affine connection, 777 in Coulomb system, 934 in dionium atom, 1015 Riemann, 86 Riemann-Cartan, 778 AFFLECK, 1. , 1259 AHARONOV, Y., 1158 Aharonov-Bohm effect, 640, 641, 1088, 1096, 1138, 1158 Airy function, 178 AITCHISON, 1.J.R., 456, 893, 1161 Alexander -Conway knot polynomial, 1112 knot polynomial, 1103, 1106-1108, 1160 generalized to links, 1120 ALEXANDROV, A.S., 569 algebra Dirac, 671 Lie, 56 of dynamical group of dionium atom, 1016 of dynamical group of hydrogen atom, xiii, 969 rotation group, 58 Pauli,667 algebraic topology, 1103 ALLILUEV, S.P., 457 ALONSO, D., 456 ALvAREZ-GAUME, L., 893 AMARAL, L.A.N ., 1523 AMBEGAOKAR, V., 1261 ambient isotopy of knots, 1104, 1153, 1156 AMELINO-CAMELIA, G. , 1158 American option, 1500 AMIT, D.J., 1083 Ampere law, 875, 1406 amplitude, see also time evolution, 43, 94 evolution, 974

ABARBANEL, H .D.1., 208 ABO-SHAEER, J.R., 693 ABRAHAM, R., 87 ABRAMOWITZ, M ., 50, 70, 174, 178, 244,246, 410, 505, 715, 758, 761, 826, 1174, 1435, 1467 ABRIKOSOV, A .A ., 1366 absence of extra R-term in curved-space Schriidinger equation, 801, 897, 906, 909, 938 absorption, 1341, 1342, 1363 absorptive part influence functional, 1290 of Green function, 1270, 1283 action, 1 canonical, 3 Chern-Simons, 1124, 1127, 1138, 1140, 1147,1160 classical, 1 DeWitt, 890 effective, 300- 303, 305, 308, 871 effective classical, 679 Einstein-Hilbert, 1390 Euclidean, 137, 238, 242, 1204 Faddeev-Popov, 868, 871 , 1050, 1055 Jacobian, 795, 797, 799, 901, 902, 904, 905, 909, 939, 940, 949, 950 kink, 1178, 1188 Maxwell, 1405 midpoint, 793 nonlocal, 263 particle in magnetic field , 179, 181 postpoint, 792, 802 prepoint, 793 pseudotime-sliced, 924-927 quantum-statistical, 137 super, 1332 time-sliced, 91 curvilinear coordinates, 773 Wess-Zumino action, 747 activation energy, 176 ADAMS, B.G ., 489, 567 ADAMS, D.H. , 1161 addition theorem Bessel functions , 753, 762 Gegenbauer polynomials, 716

1529

1530 fixed-energy, 45, 50, 391, 921, 929, 974, 985 Duru-Kleinert transformation, 983 spectral representation, 752 fixed-pseudoenergy, 977 free particle, 101 from w - t 0 -oscillator, 761 imaginary-time evolution, 141 spectral decomposition, 759 with external source, 238 integral equation, 895 near group space, 739- 742, 751 near spinning top, 742 near surface of sphere, 722, 723, 729, 734, 736, 739, 740 of spinning particle, 743 of spinning top, 741, 742 on group space, 739, 741, 751 on surface of sphere, 722, 734- 736, 738, 741, 800, 999 time-dependent frequency, 127 particle in magnetic field, 179 spectral representation, 766 pseudotime evolution, 929 radial, 699, 704, 705 Coulomb system, 994 oscillator, 993 scattering, 190 eikonal approximation, 71 first correction to eikonal, 341 time evolution, 43, 46, 89, 94, 100, 235, 752, 929, 1262 fixed path average, 237 of free particle, 109, 110 of freely falling particle, 177 of particle in magnetic field , 179, 181, 183 perturbative in curved space, 846 with external source, 232 time-sliced, 89, 101 configuration space, 97 in curvilinear coordinates, 773 momentum space, 94 phase space, 91 Ampere law, 71 analysis, spectral, 131 analytic regularization, 159 ANDERSON, M.H., 693 ANDERSON, P.W. , 1083 ANDREWS, M .R. , 693 angle Euler, 60, 62, 64 tilt, 955, 957

Index angular barrier, 724, 726, 732, 994 four-dimensional, 997 momentum, 56 conservation law, 436 decomposition, 697, 704, 705, 713, 719, 722 anharmonic oscillator, xxv, 467, 1213 effective classical potential, 470 anholonomy, objects of, 887 annihilation operator, 644, 955, 956, 969 anomalous dimension, 517 magnetic moment, 1145, 1210 square-root trick, 518 anomaly, eccentric, 438 ANSHELEVIN, V.V. , 1159 ANTENEODO, C ., 1525 anti-instanton, 1168 anticausal, 38 time evolution operator, 38 anticommutation rules fields, 654 Grassmann variables, 654, 664 anticommuting variables, 654, 673 antikink, 1167, 1168 antiperiodic boundary conditions, 224, 225, 231, 346 functional determinant, 349 Green function , 224, 225, 248, 1266 anyons, xi, 640, 1100 anypoint time slicing, 793 approximation Born, 70, 74, 191, 344 eikonal, 194, 340, 369, 417, 443 Feynman-Kleinert, 464 Ginzburg-Landau, 305 isotropic for effective classical potential, 481 mean-field, 306, 312, 338 Pade, 1076 saddle point, 377, 1206, 1241 semiclassical, 369, 1166, 1167 Thomas-Fermi, 440 tree, 305 Wentzel-Kramers-Brillouin (WKB) , 369, 372, 374, 396, 1225, 1259 arbitrage of financial asset, 1500 statistical, 1500 ARCH model, 1496 ARFKEN, G., 1525 ARNOLD, P., 695

1531 AROVAS, D.P. , 1161, 1163 ARRIGHINI, G.P., 207 ARTHURS, A.M., 206 ARVANITIS, C . , 566

asset, see also financial asset, asymmetric spinning top, 85 truncated Levy distribution, 1439 asymptotic series, 701 of perturbation theory, 273, 378, 505, 633, 1209 atom hydrogen, 910, see also Coulomb systern, 931 one-dimensional, 376, 451 hydrogen-like, 72 Thomas-Fermi, 432 atomic units, 482, 954, 1238 attempt frequency, 1207 AUERBACH , A . , 1259 AUST, N . , 1163 autoparallel, 777 coordinates, 790 auxiliary nonholonomic variation, 785 average energy, 78 functional, 209, 249 particle number, 78 AVRON , J.E., 489, 567, 568 axial gauge, 878 BAAQUIE, B.E. , 1522, 1527 BABAEV, E ., x BABCENCO , A . , 589, 973 BACHELlER, L., 1499, 1527 BACHMANN, M . , 255, 367,

548, 568, 590, 1367 background field, 321 field method for effective action, 321, 871 BAGNATO, V . S., 625 BAKER, H .F ., 207 Baker-Campbell-Hausdorff formula, 90, 201, 207, 350, 651 BALL, C . A . , 1522 BALLOW, D .D., 751 BALSA , R., 567 BANERJEE, K., 567 BANK , P., x BANKS , T. , 566, 1161 BARNDORFF-NIELSEN , 0 ., 1524 BARNES , T., 567 barrier

angular, 724, 726, 732, 994 four-dimensional, 997 centrifugal, 700, 702, 706, 714, 720, 721, 918 time-sliced, 703, 706 height, 1165 high, semiclassical tunneling, 1166 low, sliding regime, 1219 BARUT, A . O., xiii, 973, 1017, 1018, 1426 basis complete in Hilbert space, 21 functions, 20 local, 19 multivalued tetrads, 780 triads, 778, 780 tetrads, 779 multivalued, 780 reciprocal, 779 triads, 776 multi valued, 778, 780 reciprocal, 776 BASTIANELLI, F., 892 BATEMAN, H . , 716 bath Ohmic for oscillator, 268 oscillators, 262 photons, 266 for oscillator, 271 master equation, 1340 thermal photons, 268 thermal for quantum particles, 262 BATICH, C., 1159 BAUR, H . , x BAUSCH, R., 1367 BAXTER, M.W. , 1523 BAYM, G. , 694, 695, 1366 BELOKUROV, V .V., 893 BEN-EFRAIM, D.A . , 1159 BENDER, C.M., 353, 456, 566, 1261, 1523 Bender-Wu recursion relations, 353 BENGURIA, R. , 1367 BEREZIN , F.A., 693, 1426 BERGMAN , 0. , 1158 BERN, Z., 1426 Bernoulli numbers, 170, 633 polynomials, 245 BERRY, M .V . , 405, 455, 1158, 1160 BERTOIN, J., 1525 Bessel function , 50, 156, 169, 410, 1021, 1033

1532 addition theorem, 753, 762 as regulator, 977, 983, 985, 993, 997 modified, 50, 698 representation of distributions (generalized functions), 825 BEssIs, D. , 566 Beta function , 430, 690, 1135 BETHE, H.A., 1367 Bianchi identity, 876 BJJLSMA , M., 694 bilocal density of states, 409 BINGHAM, R. , 1427 Biot-Savart energy, 886 bipolaron, 544 BIRELL , N .D. , 973 BJORKEN, J.D. , 1426 black body, 1363 holes, 774 BLACK , F., 1499, 1527 Black-Scholes formula, 1499, 1507, 1511, 1513, 1514, 1516, 1528 blackboard framing, 1151 BLAIZOT, J.-P ., 694, 695 BLASONE, M., 1365 BLATTBERG, R. , 1525 BLINDER, S.M. , 973 BLM/Ro knot polynomials, 1109, 1153, 1156 Bloch theorem, 635 BLOORE, F .J ., 692 BOHM, M., 751, 1017 Bogoliubov transformation, 675 BOGOLJUBov , N.N ., 568, 694, 695, 893 BOGOMONLY, E .B . , 456 BOHM, D ., 1158 Bohr magneton, 181, 1415 radius, 421, 467, 486, 629, 954, 1344, 1400 Bohr-Sommerfeld quantization rule, 373, 375, 398, 450, 451 BOITEUX , M. , 973 BOLLERSLEV, T ., 1527 Boltzmann constant, 76 distribution, 94, 136, 138, 154 in financial markets, 1443 factor, 76 local, 330, 458, 459 quantum, 1207, 1254 BONANNO, G ., 1526 bond length, 1019 effective, 1032

Index BONESS, A.J., 1499 Borel resummability, 1210 transform, 1210 BORKOVEC , M., 1260 BORLAND, L. , 1525 BORMANN , P ., 693 Born approximation, 70, 74, 191, 344 BORN , M., 591 BOROWITZ, S. , 420 Bose -Einstein condensate, 88, 591, 599, 603, 618 distribution, 222, 248, 1268 normal part, 618 fields fluctuating, 644 quantized, 641 occupation number, 222, 248 particles ensemble of orbits, 592 partition function , 647 bosons, 222, 248, 591 , 592, 636, 638 field quantization, 641 free energy, 676 free particle amplitude, 637 integration, 646 many orbits, 592 Nambu-Goldstone, 311, 324, 325 nonequilibrium Green functions , 1266 quantization of particle number, 641 second quantization, 641 BOUCHAUD, J.-P., 1523, 1525, 1528 bounce solution, 1201 bound states Coulomb system, 935 poles, 1017 boundary condition antiperiodic, 224, 225, 231, 346 Dirichlet, 103, 126, 153, 213, 216, 229, 260, 262, 340, 346, 837 in momentum space, 154 functional determinant, 349 Neumann, 153, 230, 1049 periodic, 126, 167, 219, 222, 242, 247, 250,256 box, particle in, 579, 581, 582 Boz , M., 1158 bra-ket formalism of Dirac, 18, 21, 663 for probability evolution, 1327 BRAATEN, E . , 893 bracket Kauffman knot polynomial, 1109

1533 Lagrange, 7, 8 Poisson, 4, 8, 9, 39, 56, 663 BRADLEY, C.C ., 693 Bragg peaks, 1322 reflection, 12 scattering, 252, 1454 BRANDT, S.F., 368 BRAY, A .J ., 1527 BREEDEN , D.T ., 1523 BRERETON, M.G. , 1159, 1163 BRETAGNOLLE, J., 1525 BRETIN, V., 695 BREY, J.J ., 1527 BREZIN , E., 1260 BRILLOUIN, L. , 279, 455 Brillouin-Wigner perturbation theory, 279 BRINK , L., 1426 BRITTIN , W .E. , xiii, 973, 1364 BRODIMAS, G.N ., 207 BRODIN , G ., 1427 Bromwich integral, 1469 BROSENS, F ., x, 569, 693, 695 Brownian bridge, 1350 motion, 1349 BRUDNER, H .J ., 420 BRUSH, S.G. , 206 bubble critical, xxvi, 1201, 1207, 1243, 1244, 1247- 1252, 1254 in Minkowski space, 1255 instability, 1202 radius, 1247, 1250-1252 wall, 1250, 1253 decay&equency, 1207 solution, 1201 BUCKLEY, I.R.C ., 566 BUDNYJ , B. , x BUND , S., 693 Burgers vector, 782 BURGHARDT, I. , 456 CABRERA, B., 1163 CAGE, M.E., 71 CAl, J .M ., 1017 CAl, P .Y ., 1017, 1018 CALAGAREAU, G. , 1159 Calagareau-White relation, 1122, 1123, 1159 calculus Ito, 1315 stochastic, 189, 1319 Stratonovich, 1315

CALDEIRA, A .O ., 368, 1260, 1365, 1367 call option, 1498, 1500 CALLEN, H.B ., 87, 1364 CALOGERO, F ., 1017 CAMETTI, F., 893 CAMPBELL, J.E. , 207 CAMPBELL, W.B ., 208, 457 canonical action, 3 anti commutation relations, 664 commutation relations, 15, 39, 40, 92 ensemble, 77 Laplacian, 55, 56 path integral correlation functions, 255 quantization, 39, 55- 57, 66 transformation, 6, 8, 9 generating function, 10 CARR, P.P ., 1525, 1527 Cartan curvature tensor, 934 CASALBUONI, R., 1426 CASATI, G. , 893 CASTELLI, C. , 1498 CASWELL, W., 567 catenane, self-entangled polymer ring, 1158 CATICHA , N. , 1524 causal, 38 ordering, 35 time evolution amplitude, 43 operator, 43 causality, 221 , 591 , 921, 1292, 1339 caustics, 112, 129, 129 CELEGHINI, E., 1365 central limit theorem, 1457, 1464, 1467, 1482, 1492, 1512 generalized, 1459 centrifugal barrier, 700, 702, 706, 714, 720, 721, 918 time-sliced, 703, 706 CEPERLEY, D ., 694 chain diagram, 284, 812, 816, 823, 831, 841, 850 random, 1019 stiff, 1024 CHAKRABARTY, D ., 567 CHAKRAVARTY, S. , 368, 1260 champaign bottle potential, 310 CHAN, T ., 1525, 1526 CHANDLER, D ., 566 CHANDRASEKHAR, S., 1081, 1365, 1366 CHANG, B .K., 565

1534 CHANG, E.C., x, 1527 CHANG, L.-D. , 368 chaos hard,405 smooth, 405 Chapman-Kolmogorov equation, 1461, 1471 character expansion, 740 charge quantization Dirac, 747 charged particle in magnetic field fixed-energy amplitude, 765 wave functions, 763, 766 radial, 766 CHAUDHURI, D. , 1082 chemical potential, 77, 597, 1071, 1357 CHEN, Y.-H., 1160 CHEN, Y.C ., 1367 CHENG , B .K ., 207 CHENG , K.S. , xiv, 88, 890, 916, 917 CHERN , S., 1160 Chern-Simons action, 1124, 1127, 1138, 1140, 1147 theory, 1137, 1147, 1160 nonabelian, 1149, 1155 of entangled polymers, xi, 1124, 1127 CHERVYAKOV, A. , 247, 590, 892, 893, 1366 CHETYRKIN, K.G ., 1083, 1261 CHEVY, F ., 695 Chi distribution, 1442 CHOU, K.-C., 1366 Christoffel symbol, 11,86, 776, 777, 780, 792, 799 circle, particle on, 571 , 574, 581 classes of knot topology, 1101 classical action, effective, 679 Boltzmann factor, 155, 329, 333 effective, 329, 333 differential cross section, 446 effective action, 309, 679 effective potential, 328, 333, 679 eikonal, 371 limit, 154 local density of states, 409 mechanics, 1 momentum, local, 369 motion in gravitational field, 775 orbit, 1 particle distribution, 138 partition function, 76 path,2

Index path integral, 384, 1331 potential effective, xxii, 328, 333, 336, 463, 464, 467, 471, 473, 474, 476, 679 solution, 1174, 1212, 1213, 1250 almost, 1190 tunneling, 1167 statistics, 76, 1248 CLAY, M. , 567 closed-path variations in action principle, 784 closed-time path integral, 1294, 1340 closure failure, 782, 787 cluster decomposition, 294 coefficients strong-coupling expansion, 512 virial, 1100, 1100 COHEN-TANNOUDJI , C ., 1367 coherence length, 1235 coherent states, 350, 651 COLEMAN , S., 455, 1259, 1261 collapse of path fluctuations, 701, 726, 918, 919, 931, 1241 collective excitations, 614, 681 field , 682 fields , 681 phenomena, 681 variables, 681 COLLIER, J ., 1427 COLLINS , J .C. , 1261 COLLINS, P .D.B. , 1017 commissions in financial markets, 1500 commutation rules canonical, 15, 39, 40, 92 equal-time, 40 field , 642 commuting observables, 4 complete basis, 21 completeness relation, 19, 21- 23, 27, 28, 31 , 46, 47, 571, 764, 768 basis Dads, 51, 776 Dirac, 21 lattice, 107 composite field, 306, 1238 knot , 1107 composition law for time evolution amplitude, 90, 700 composition law for time evolution operator, 37,38, 72 compound knots, 1102, 1105, 1108 inequivalent, 1103, 1104

1535 Compton wavelength, 421 , 1374, 1375, 1377, l400, 1402 COND, R ., 1527 condensate Bose-Einstein, 88, 591, 599, 603, 618 critical temperature, 596, 602 superconductor, 1238 energy, 1240, 1242 superfluid helium, 606 critical temperature, 606 condition Schwarz integrability, 7, 180, 639, 777, 778, 780, 848, 879 Wentzel-Kramers-Brillouin (WKB), 370,373 configuration space, 98 confluent hypergeometric functions, 757, 961 conformal group, 970 invariance in field theory, 963 transformation Weyl, 962, 963 conformal transformation, 456 conformally flat , 962, 962 conjugate points, 129 connected correlation functions, 288 generating functional, 288 diagram, 284, 294 n-point function, 292, 303 two-point function, 302 connectedness structure of correlation functions, 289 connection affine, 777 in Coulomb system, 934 Riemann-Cartan, 778 Riemann, 86, 776 rules, Wentzel-Kramers-Brillouin (WKB),372 spin, 888 conservation law angular momentum, 436 current, 17 energy, 14, 75, 1169 momentum, 301 , 1086 probability, 16, 1292, 1297, 1300, 1334, 1335, 1340 constant Boltzmann, 76 cosmological, 1391 coupling dimensionally transmuted, 1237

in Ginzburg-Landau expansion, 1235 dielectric, 534 diffusion, 1293 Euler-Mascheroni, 156, 542, 1198 fine-structure, 71, 421, 629, 1397, 1406, 1415, 1416 Hubble, 1391 Lamb, 1345 Planck, 13 constraint geometric, 579, 801 topological, 571, 1085 CONT, R., 1525 continuity law, 17 continuous spectrum, 47 Coulomb system, 961 continuum limit, 92, 93 contortion tensor, 779 contraction tensors appearing in Wick expansion, 708, 942, 1022, 1041 Wick pair, 251 convention, Einstein summation, 2, 4 functional, 290 convergence proof for variational perturbation expansion, 1233 radius of strong-coupling expansion, 1235 vanishing radius in perturbation series, 1209 convex effective potential, 338 function, 338, 460 Conway knot polynomial, 1109 CONWAY, J .H ., 1111, 1160 Conway-Seifert knot , 1108 Cooper pair, 1236 field, 1238 coordinate -dependent mass, 871 autoparallel, 790 curvilinear, 697, 774 time-sliced amplitude in, 773 cyclic, 572 generalized, 1 geodesic, 790 independence, 804, 809, 811 , 831, 834, 841, 851, 854 of path integral in time-sliced formulation, 800 normal, 790

1536 Riemann, 790 parabolic, Coulomb wave functions , 957 radial,793 transformation, 776, 975, 977, 978, 982, 984 nonholonomic, 778 COOTNER, P. , 1527 core, repulsive in 3He potential, 1238 CORINALDESI, E., 1158 CORNELL, E .A ., 693 CORNISH, F.H.J ., 973 CORNWALL, J.M., 367 CORRADINI, 0., 892 correction fluctuation, 412 semiclassical expansion, 407, 451 tracelog, 416 fluctuations in tunneling process, 1167 Langer, 983 time slicing, 979, 981, 992, 994 correlation functions, 209, 249, 250 connected, 288 connectedness structure, 289 from vacuum diagrams, 298 in canonical path integral, 255 in magnetic field , 254 one-particle irreducible, 300 subtracted, 222, 225, 244, 264, 327, 333, 334, 336, 860 correspondence principle, 15, 17, 31 , 55, 56, 62, 66,67 group, 56 Heisenberg, 40, 41 CORWIN, A.D., 695 cosmic standard time, 1386, 1392 cosmological constant, 1391 Einstein, 1393 evolution, 1381, 1392 cotorsion of polymer, 1109 COTTA-RAMUSINO, P ., 1161 Coulomb amplitude D = 2, 938 D = 3,948 polar decomposition, 957 energies, 954 Hamiltonian, 16 potential, 918 scattering, 71 system, 467, 475, 481, 894, 910, 931 affine connection, 934 and oscillator, 1399

Index bound states, 935 continuous spectrum, 961 curvature and torsion after transformation, 931 D = 1, energies, 451 D = 2,933 D = 2, amplitude, 937 D = 2, time-slicing corrections, 938 D = 3, amplitude, 948 D = 3, energies, 954 D = 3, time-slicing corrections, 943, 949 dynamical group 0(4, 2) , 969, 972 eccentricity of orbit, 438 effective classical potential, 483, 567 energy eigenvalues, 954 in magnetic field, 484 one-dimensional, 376, 451 particle distribution, xxii, 484 path integral, xiii, 931 pseudotime-sliced action, 932 amplitude, 932 radial, 985, 987 relativistic path integral, xi, 1397 solution in momentum space, 964 time-slicing corrections, 951 torsion, 934 transformation to oscillator, xiii, 935, 936, 946, 947, 952, 953, 956, 957, 959, 969 wave functions, 467, 937, 954, 957 algebraic aspects, xiii, 969 parabolic coordinates, 957 coupling constant dimensionally transmuted, 1237 in Ginzburg-Landau expansion, 1235 magnetic, 905 minimal, 905 strong, 518 weak, 272, 518 COURANT, R., 1523 COURTEILLE, P .W., 625 covariant derivative, 780 functional, 873 fluctuations, 873 Laplacian, 896, 900, 907 -Weyl,963 perturbation expansion, 865

1537 Taylor expansion, 791 variations, 873 COWLEY, E.R. , 566 COWLEY, R.A ., 614 Cox, J.C. , 1523 CRAIGIE, N.S. , 1162 creation operator, 644, 955, 956, 969 CRICK, F .H .C ., 1159 critical bubble, xxvi, 1201, 1207, 1243, 1244, 1247- 1252, 1254 in Minkowski space, 1255 instability, 1202 radius, 1247, 1250-1252 wall, 1250, 1253 current, 1149, 1241, 1242 exponent of field theory, 1227 of polymers, 1024 exponent, polymers, 1062, 1069, 1075, 1077, 1083 phenomena, 1261 temperature Bose-Einstein, 596, 602 of superconductor, 1236 superfluid helium, 606 CROOKER, B.C., 695 cross section classical, 446 semiclassical, 447 crossings in knot graph, 1085, 1103, 1104, 1105, 1107, 1113, 1114, 1119, 1121 crystals, quantum, 566 CUCCOLI, A ., 565, 566 cumulants, 274 expansion in perturbation theory, 274, 278, 294, 493 polymer distribution, 1023 truncated Levy distribution, 1437 cumulative distribution, 1436 Gauss, 1506 CURADO, E .M .F ., 1525 current, 209, 240, 250 conservation law, 17 critical, 1149, 1241, 1242 density, 17 Hall, 1144 periodic, 247 super, 1240 CURTRIGHT, T .L. , 893 curvature, 783 effective potential, 909 in transformed H-atom, 931

scalar, 66, 87 of spinning top, 87 Riemann-Cartan, 781 sphere, 800 tensor of disclination, 783 Riemann, 780 Riemann-Cartan, 779, 934 curvature and torsion space with, 773 Schriidinger equation, 894 curved spacetime, 10 curvilinear coordinates, 697, 774 time-sliced amplitude in, 773 cutoff infrared (IR), 814 ultraviolet (UV), 805 cycles in permutations, 595 cyclic coordinate, 10, 572 permutations, 595 variable, 571 , 574 cyclotron frequency, 181, 1379 cylinder function , parabolic, 1052 d'Alembert formula, 124 DA SILVA, A.J ., 1158 DALIBARD, J ., 695 DANIELL, P.J. , 206 Daniels distribution for polymers, 1042 DANIELS, H.E., 1082 DASH, J.W. , 1501 DASHEN, R ., 455 DAVID, F., 590 DAVIES, P.C.W., 973 DAVIS, K.B ., 693 DAVIS, M .H .A ., 1526 DE BOER, J ., 892 DE DOMINICIS, C., 367 DE RAEDT, B. , 208 DE RAEDT , H ., 208 DE SOUZA CRUZ, F.F., 695 DE TOLEDO, C.M. , 1524 de Broglie thermal wavelength, 139, 595 wavelength, 370 Debye -Waller factor , 252, 1322, 1475 non-Gaussian fluctuations, 1454 function, 1031, 1049 temperature, 1236 decay bubble, frequency, 1207

1538 of supercurrent by tunneling, 1235 rate, 1200, 1245 thermally driven, 1255 via tunneling, 1199, 1200, 1213, 1242, 1245- 1247, 1254 decoherence, 1334 decomposition, angular momentum, 697, 704, 705, 713, 719, 722 in D dimensions, 714 in four dimensions, 730 defect crystal, 781 , 783 field, 1369 DEFENDI, A. , 208, 1367 definition of path integral perturbative, 287 time-sliced , 89 degeneracy of spherical harmonics, 716 degenerate limit, 631 DEGENNEs, P .G ., 1082, 1083 DEKKER, H ., xiv, 88, 890, 917 DELOS, J.B. , 455 Delta, 1502 hedging, 1504 Delta of option, 1501, 1503, 1504 8-function and Heaviside function, 44 Dirac, 24, 44 path integral, 770 would-be, 710 DEMPSTER, M .A.H., 1528 density current, 17 matrix, 33, 140 of states, 270, 601 bilocal, 409 local classical, 399, 401, 409 local quantum-mechanical, 406 local semiclassical, 409 Thomas-Fermi, 419 of supercoiling in DNA, 1118 operator, 33, 1277 particle, 137 partition function, 136, 463, 548 probability, 17 spin current, 907 states, 82 derivative assets, 1498 covariant, 780 expansion, 174, 407, 412, 875 functional, 209 covariant, 873

Index lattice, 104 Radon-Nikodym, 1455 DES CLOIZEAUX, J., 1083 DESER, S., 1160, 1426 desired velocity, 189 DESITTER, J. , 569 determinant Faddeev-Popov, 863, 866- 869 fluctuation, 110 easy way, 1185 functional free particle, 110 from Green function, 344 oscillator, 116 time-dependent frequency, 120 Van Vleck-Pauli-Morette, 388, 390, 916,917 Wronski, 122, 124, 214, 345 DEVORET, M .H. , 368 DEVREESE, J.T., x, xiv, 276, 569, 589, 695, 973, 1364 DeWitt -Seeley expansion, 851 action, 890 extra R-term, 909 DEWITT, B .S., xiv, 88, 367, 890, 893, 916, 917, 973 DEWITT-MoRETTE, C., 207, 388, 589, 917 DHAR, A., 1082 diagram chain, 284, 812, 816, 823, 831, 841, connected, 284 disconnected, 284 Feynman, 282, 1368 local, 811 loop, 283 nonlocal, 812 one-particle irreducible (lPI), 300, 304, 318, 501,871 reducible, 318 tadpole, 499 tree, 304, 307, 310, 315 watermelon, 284, 812, 816, 823, 841, 850 dielectric constant, 534 difference equation, stochastic, 1496 differential cross section classical, 446 semiclassical, 447, 447 Mott scattering, 449 differential equation

909,

693,

911, 692,

850

319,

831 ,

1539 first-order, 219 Green function, 219 for time-dependent frequency, 226 Hamilton-Jacobi, 10, 370, 384 Riccati, 167, 370 stochastic, 1308 Sturm-Liouville, 123 Thomas-Fermi,426 diffraction pattern, 29 diffusion constant, 1293 matrix, 1297 DIJKGRAAF, R., 1427 dilatations, 970 local, 970 dilation operator, 955, 957 dilute-gas limit, 1190 dimension, anomalous, 517 dimensionally transmuted coupling constant, 1237 DINEYKHAN, M., 569 DING, Z.X., 1527 dionium atom, 974 affine connection, 1015 dynamical group 0(4, 2), 1016 path integral, 1004 time slicing corrections absense, 1009 DIOSI, L ., 1367 Dirac -Fermi distribution, 225 algebra, 671 bra-ket formalism, 21 for probability evolution, 1327 brackets, 18, 663 charge quantization, 747 8-function, 24, 44 and Heaviside function , 44 path integral, 770 interaction picture generating functional, 1286 time evolution operator, 1279 string, 641 , 881, 1090, 1093 theory of magnetic monopoles, 881 DIRAC, P.A.M ., 87, 206, 751, 1017, 1162 Dirichlet boundary conditions, 103, 126, 153, 213, 216, 229, 260, 262, 340, 346, 837 in momentum space, 154 disclinations and curvature, 783 disconnected diagrams, 284 discontinuity fixed-energy amplitude, 47

discount factor in financial distributions, 1477 dislocations and torsion, 781 disorder field, 1369 dispersion relation, 1211 dispersive part of Green function, 1270 displacement field , electric, 534 dissipation, 262 -fluctuation theorem, 1271, 1275, 1276, 1336, 1363, 1364 Drude, 265, 268, 1292, 1299, 1300, 1302, 1307 Ohmic, 1293 dissipative part in influence functional, 1290 of Green function , 1270, 1271, 1283 distribution Boltzmann, 94, 138, 154 in financial markets, 1443 Bose-Einstein, 222, 1268 Chi, 1442 classical of particles, 138 Dirac 8, 25 Fermi-Dirac, 225, 1268 financial, 1432 heavy tails, 1432 Gamma, 1442 Gauss cumulative, 1506 in financial markets, 1432 Heaviside, 25 Levy, 1432, 1432, 1458 asymmetric, 1434 truncated, 1434, 1437 asymmetric, 1439 cumulants, 1437 martingale, 1478 Maxwell, 1325 Meixner in financial markets, 1450 Pareto-Levy-stable, 1458 particle, 155, 182 Student in financial markets, 1447 Tsallis in financial markets, 1447 Weibull, 1449 distributions (generalized functions), 25, 45 as limits of Bessel functions, 825 extension to semigroup, 821 products of, 830 DI VECCHIA, P., 1426 divergence

1540 infrared (IR), 814 of perturbation series, 1209, 1233 ultraviolet (UV), 159, 805 dividends of financial asset, 1500 DMjUS$ exchange rate, 1439 DNA molecules, 1117, 1117, 1118, 1120, 1159 circular, 1117 DODONOV, V .V., 207 DOLAN, L. , 1259 DOLL, H., 1153, 1160 DOMB, C., 1082 DORDA, G ., 1161 DORSEY, A.T., 1260 double -well potential, 472, 473, 522, 1164, 1165, 1168, 1169, 1190 convex effective potential, 338 particle density, 479 helix, 1117, 1117, 1118, 1120, 1159 circular, 111 7 double-slit experiment, 12 Dow-Jones industrial index, 1428 DOWKER, J .S., 589, 890 DRAGULESCU, A.A., x, 1481, 1526 DRELL, S.D. , 1426 drift, 1431 Wiener process, 1310 DROZDOV, A.N., 1527 Drude dissipation, 265, 268, 1292, 1299, 1300, 1302, 1307 relaxation time, 265 duality transformation, 168, 169 DUBOIS, D ., 1364 DUBRULLE, B., 1525 DUFFIE, D., 1525 Dulong-Petit law, 176, 327, 605- 607, 625, 634 DUNCAN, A. , 566 DUNHAM, J.L., 373 DUNNE, G.V., 1161, 1426 DUPONT-Roc, J., 1367 DURANTE, N.L., 207 DURFEE, D.S., 693 DURU, I.H ., xiv, 920, 973, 1260, 1426 Duru-Kleinert equivalence, 976 angular barrier and Rosen-Morse potential,994 D-dimensional systems, 1003 extended Hulthen potential general Rosen-Morse potential, 1002 four-dimensional angular barrier and general Rosen-Morse potential, 997

Index Hulthen potential and general RosenMorse potential, 1000 radial Coulomb and Morse system, 985 radial Coulomb and radial oscillator, 987 radial oscillator and Morse system, 983 Duru-Kleinert transformation, 925, 931, 974, 978, 982- 985, 993, 995, 998, 1000, 1001, 1012 D = 1, 974 effective potential, 976 fixed-energy amplitude, 983 of Schr6dinger equation, 982 radial Coulomb action, 986 oscillator, 987 time-slicing corrections, 976 dynamical group, 969 group 0(4, 2) of Coulomb system, 969, 972 of dionium atom, 1016 metric, 381 dynamical hedging, 1504 Dyson series, 35, 203 DYSON, F .J ., 1261 DZYALOSHINSKI, I.E., 1366 EBERLEIN, E. , x, 1526 eccentric anomaly, 438 eccentricity of Coulomb orbit, 438 ECKER, G., 893 ECKERN, U. , xvi, 1365 ECKHARDT, B ., 456 EDMONDS, A.R., 88, 973 EDWARDS, S.F., 750, 1081, 1083, 1159 effect Aharonov-Bohm, 640, 641, 1088, 1096, 1138, 1158 excluded-volume in polymers, 1062, 1063, 1069, 1070 Meissner, 312, 1148 quantum Hall, 1146, 1161 fractional, xi, 1143, 1146, 1160 effective action, 300-303, 305, 308, 871 background field method, 321, 871 classical approximation, 309, 679 two loops, 315 bond length, 1032 classical action, 679 Boltzmann factor, 329, 333 free energy, 467

1541 potential, xxii, 328, 333, 336, 459, 463, 464, 467, 471, 473, 474, 476, 679 energy, 302, 305 potential, 308, 336, 337, 900, 916 convex in double well, 338 convexity, 338 due to curvature, 909 Duru-Kleinert transformation, D = 1, 976 from effective classical potential, 336 in space with curvature and torsion, 799 mean-field, 338 on sphere, 801 range, 610 efficient markets, 1500 EFIMOV, G.V., 569 eikonal, 370, 371 approximation, 194, 340, 369, 417, 443 Einstein -Bose distribution, 222, 1268 equation, 1391, 1392 equation for gravity, 781 equivalence principle, 774, 775 invariance, 887 summation convention, 2, 4, 290, 309 tensor, 781 EINSTEIN , A., 1364, 1527 Einstein-Hilbert action, 1390 electric displacement field, 534 electrodynamics, quantum (QED) , 1369, 1420 electromagnetic field, 883, 1406 forces, 883 self-energy, 1406 units, 71 ELIEZER, D. , 1162 elliptic eigenvalue of stability matrix, 404 elliptic theta function, 607, 687 ELWORTHY, K.D., 917 eInission, spontaneous, 1341, 1342, 1363 end-to-end distribution, polymer, 1019, 1020 cumulants, 1023 exact, 1024 Gaussian approximation, 1029 moments, 1021 saddle point approximation, 1028 short-distance expansion, 1026 ENDRIAS, S., xii, xvi energy -entropy argument for path collapse, 919

-momentum tensor symmetric, 781 activation, 176 average, 78 Biot-Savart, 886 conservation, 1169 conservation law, 14, 75 density, Thomas-Fermi, 420 effective, 302, 305 excitation, 613, 614 Fermi, 419, 631 free, 78 functional Ginzburg-Landau, 306 ground state anharmonic oscillator, 466 hydrogen atom, 467 internal, 78 of condensate in superconductor, 1240, 1242 Rydberg, 72 self-, 305 shift, 274, 276, 278- 280, 352 Thomas-Fermi, 429, 431, 432, 435 variational, xxvii zero-point, 146, 332, 675, 1219 energy-momentum tensor, 1391 ENGLE, R.F ., 1526 ensemble Bose particle orbits, 592 canonical, 77 Fermi particle orbits, 592 grand-canonical, 78, 80 ENSHER, J.R. , 693 entangled polymer, xi Chern-Simons theory, xi entanglement paths, 1084, 1088, 1101 Chern-Simons theory, 1124, 1127 polymers, 1084, 1088, 1101 Chern-Simons theory, 1124, 1127 entropy -energy argument for path collapse, 919 equal-time commutation rules, 40 equation Chapman-Kolmogorov, 1461, 1471 Einstein, 1391, 1392 Einstein for gravity, 781 Euler-Lagrange, 2, 3, 5, 6, 11, 235, 1295 first and second London, 1157 Fokker-Planck, 1295, 1304, 1365 for financial assets, 1494 with inertia, 1297, 1318

1542 overdamped, 1318 Hamilton-Jacobi, 10, 370, 384 Klein-Kramers, 1295, 1297 overdamped, 1304 Landau-Lifshitz, 749 Langevin, 1289, 1365 operator form, 1308 quantum, 1308 semiclassical, 1308 with inertia, 1309 Lindblad, 1336, 1340 Liouville, 33 Lippmann-Schwinger, 73, 74, 344, 610, 1091, ll58 master, 1335, 1336 photon bath, 1340 Maxwell, 1406 of motion, 41 Hamilton, 3, 4, 41 Heisenberg, 41, 749 Poisson, 421, 422, 1406 Riccati differential, 370 Schr6dinger, 15, 16, 18, 25, 34, 35, 3840, 43, 44, 51, 53, 897, 909, 952, 1262 Duru-Kleinert transformation, 982 in space with curvature and torsion, 894 time-independent, 16, 929 Smoluchowsi, 1295 Smoluchowski, 1304, 1461, 1471 Thomas-Fermi differential, 426 Wentzel-Kramers-Brillouin (WKB), 371 equilibrium, thermal, 249 equipartition theorem, 327, 462 equivalence Duru-Kleinert, 976 angular barrier and Rosen-Morse potential, 994 D-dimensional systems, 1003 extended Hulthen potential and general Rosen-Morse potential, 1002 four-dimensional angular barrier and general Rosen-Morse potential, 997 Hulthen potential and general RosenMorse potential, 1000 radial Coulomb and Morse system, 985 radial Coulomb and radial oscillator, 987 radial oscillator and Morse system, 983 principle Einstein, 774, 775

Index new, 784, 1347 quantum, 798, 910 equivalent knots, llOl martingales, 1481 path integral representations, 901 ERIs, T., xvi Esscher martingales, 1481 transform, 1480, 1480 ESSCHER, F. , 1480, 1526 ESTEVE, D ., 368 ESTEVE, J.G., 567 Euclidean action, 137, 238, 242, 256, 1204 Green function, 251 group, 56 Legendre transform, 137 periodic Green function, 240 source term, 238 space, metric, 1370 time evolution amplitude, 141 Euler -Heisenberg formula, 1385 -Lagrange equations, 2, 3, 5, 6, ll, 235, 1295 -Maclaurin formula, 174 -Mascheroni constant, 156, 542, ll98 angles, 60, 62, 64 relation, thermodynamic, 81 EULER, H., 1427 European option, 1500 evolution, see also time, 34 cosmological, 1381, 1392 exceptional knots, ll08 excess kurtosis, 1441 exchange interaction, 428 excitation energy, 613, 614, 681 excluded-volume effects in polymers, 1062, 1063, 1069, 1070 expanding universe, 1386 expansion asymptotic, 273, 378, 505, 633 character, 740 cumulant in perturbation theory, 274, 278, 294, 493 derivative or gradient, 167, 407, 412, 875 DeWitt-Seeley, 851 fluctuations, 103, ll2 Ginzburg-Landau, 1235 gradient, 168, 875

1543 large-stiffness, 1041, lO42, 1047, lO57, lO59 Lie, 61, 1143 loop, 307 Magnus, 35, 203 midpoint, 790 Neumann-Liouville, 35, 203 normal modes, 1172 perturbation, 272 covariant, 865 large-order, 1208 path integral with 8-function potential, 770 postpoint, 790 prepoint, 790 Robinson, 172, 173, 600 saddle point, 377 semiclassical, 391 around eikonal, 371 small-stiffness, lO41 , 1042 strong-coupling, xxii, xxvii, 512- 515, 543, 566, 1233- 1235 coefficients, 512 Taylor, 2 covariant, 791 virial, lO99 weak-coupling, 543 Wick, 209, 249, 251, 251 , 1322, 1363 expectation filter , 1455 expectation value, 32, 209, 249 local, 460 experiment double-slit, 12 exponent critical of field theory, 1227 of polymers, 1024, 1062, lO69, 1075, 1077, lO83 Wegner, 518 exponential integral, 156, 1209 extended zone scheme, 575, 593, 1012 extension of theory of distributions (generalized functions) , 821, 830 external force , 209 potential, 802 source second quantization, 672, 673 EZRA, G.S ., 456 factor Boltzmann, 76

Debye-Waller,1475 non-Gaussian fluctuations, 1454 fluctuation, lO3 tunneling, 1167 Lande, 1415 structure of polymer, 1030, lO33 FADDEEV, L .D ., 192, 893, 1163, 1261 Faddeev-Popov action, 868, 871 , lO50, lO55 determinant, 863, 866-869 gauge-fixing functional, 192, 858, 1150, 1178 failure of closure, 782, 787 FAINBERG , V ., 1158 false vacuum, 1254 FAMA, E.F., 1527 FEDORIUK, M.V., 115, 455 FEDOTOV, S., 1527 Feller process, 1523 FELLER, W. , 1523 F ENG NEE, LIM, xii FERANSHUK, I.D ., 568 Fermi -Dirac distribution, 225, 248, 1268 energy, 419, 631 fields fluctuating, 654 quantized, 654 liquid, 1238 momentum, 419, 631 occupation number, 225, 248 particle orbits, 592 sphere, 419, 599, 1238 temperature, 632 fermions, 225, 248, 591, 592, 636- 638, 640, 654 field quantization, 654 free energy, 676 free particle amplitude, 637 integration, 657 many orbits, 637 nonequilibrium Green functions , 1266 partition function , 660 quantization of particle number, 654 second quantization, 654 statistics interaction, 637 F ERRARI, F., 1163 ferromagnetism, classical Heisenberg model, lO35 FESHBACH, H., 132, 205, lO17 FETTER, A .L ., 693, 1160, 1366 Feynman diagrams, 282, 1368

1544 integrals, 364 path integral formula, 91 rule, 809, 832, 835 FEYNMAN, R.P ., xiii, xv, 206, 207, 565, 567569, 692, 693, 893, 1366, 1426 Feynman-Kleinert approximation, 464, 475, 476,478 field anticommutation relations, 654 background, 321 background method for effective action, 321 , 871 collective, 681, 682 commutation relations, 642 composite, 306, 1238 Cooper pair, 1238 defect, 1369 disorder, 1369 electric displacement, 534 electromagnetic, 883 energy, 1124 gauge, 875 minimal coupling, 883 Klein-Gordon, 1374 Green function, 1374 magnetic, 179, 484 operator, 641 order, 1235, 1247 quantization bosons,641 external source, 672, 673 fermions, 654 relativistic, 1278 statisto-magnetic, 1140, 1142, 1143, 1146 super, 1332 theory conformal invariance, 963 critical exponents, 1227 effective classical, 679 polymer, 1070 quantum, 676 relativistic quantum, 591, 1368 vierbein, 783, 886 vortex, 1369 weak magnetic, 484, 488 filter of expectation value, 1455 financial asset arbitrage, 1500 statistical, 1500 dividends, 1500 Fokker-Planck equation, 1494

Index Hamiltonian, 1488 hedging of investment, 1498 kurtosis of data, 1438, 1441, 1453, 1511, 1515 log-return, 1431 option, 1428 price, 1498 return, 1430 skewness of data, 1441 smile of data, 1507, 1516 strategy, 1501 time series of data, 1430 utility function, 1481 variance of data, 1430, 1450 volatility of data, 1428, 1430, 1432, 1481, 1482, 1507, 1508 risk, 1513 fine-structure constant, 71, 421, 629, 1210, 1397, 1406, 1415, 1416 FINKLER, P ., 208, 457 first quantization, 677 first-order differential equation, 219 Green function, 219 antiperiodic, 225 periodic, 222 time-dependent frequency, 226 FISHER, M.P.A. , 1260, 1367 fixed-energy amplitude, 45, 50, 391, 921, 929, 974, 985 charged particle in magnetic field , 765 discontinuity, 47 Duru-Kleinert transformation, 983 free particle, 752- 754 discontinuity, 754 spectral representation, 752 oscillator radial, 755 spectral representation, 756 Piischl-Teller potential, 996 Rosen-Morse potential, 996 fixed-pseudoenergy amplitude, 977 FIZEAU, P. , 1522 FIZIEV, P ., 891, 917 FLACHSMEYER, J., 693 FLANNERY, B .P., 1366 flat conformally, 962, 962 space, 775 flexibility of polymer, 1051, 1054 FLIEGNER, D ., 457 Flory theory of polymers, 1069 FLORY, P.J., 1082 fluctuation

1545 -dissipation theorem, 1271, 1275, 1276, 1336, 1363, 1364 Bose fields, 644 correction, 412 semiclassical expansion, 407, 451 tracelog, 416 tunneling, 1167 correction to tunneling, xxv, 1171- 1173, 1181, 1201, 1212, 1244 covariant, 873 Debye-Waller factor , 1454 non-Gaussian, 1454 determinant, 110, 385, 1185 easy way, 1185 ratio, 117 expansion, 103, 112 factor, 103 free particle, 110 oscillator, 112- 115, 117- 119 tunneling, 1167 Fermi fields, 654 fields, 592 formula, 116 kinks would-be zero modes, 1187 zero modes, 1174, 1177, 1180, 1183, 1187, 1201, 1202, 1245 part of Green function , 1270, 1271 part of influence functional, 1290 quantum, xiii, 100, 103, 328, 369, 377, 463,491 thermal, 100, 249, 328, 463, 491, 1194 translational, 386 width local, 462 longitudinal, 525 transversal, 525 FLUGGE, S., 372, 751 , 1017 flux magnetic, 1090 quantization, 1088, 1090 in superconductor, 1090 tube, 1090 FOLLMER, H., 1526 FOKKER, A.A. , 1366 Fokker-Planck equation, 1295, 1296, 1304, 1347, 1365 for financial assets, 1494 with inertia, 1297, 1318 overdamped, 1318 Foldy-Wouthuysen transformation, 1409 FOMIN, V.M., 569 forces

electromagnetic, 883 external, 209 gravitational, 774 magnetic, 179 statisto-magnetic, 1140 FORD, G.W., 368, 1365, 1367 FORD, K .W ., 456 formalism Hamilton, 3 Lagrange, 2, 1295 formula Baker-Campbell-Hausdorff, 90, 201, 207, 350, 651 Black-Scholes, 1507, 1511, 1513, 1514, 1516, 1528 d ' Alembert, 124 Euler-Maclaurin, 174 fluctuation, 116 Fresnel integral, 48, 97, 109, 113, 114, 145 Gelfand-Yaglom, 119, 121, 121, 122, 124, 126, 1185 Gelfand-Yaglom-like, 150 Gutzwiller trace, 404 Heron, 744 Laplace inversion, Post, 1469 level shift, 520 splitting, 1194 Levy-Khintchine, 1460 Lie expansion, 61 Magnus, 203 Mehler, 132, 205, 554 Poisson, 29, 156 Post, Laplace inversion, 1469 Rutherford, 442 smearing, 463 Sochocki,47 Stirling, 505, 589, 1209, 1467 Trotter, 93, 93, 208 Veltman, 161, 228 Wigner-Weisskopf for natural line width, 1337, 1342 Zassenhaus, 202 forward- backward path integral, 1294, 1333 path order, 1283 time order, 1283 FouQuE, J .P., 1522 four-point function, 295 Fourier space, measure of functional integral, 151 transform, 752 fractional

1546 quantum Hall effect, xi, 641, 1143, ll46, ll60, ll61 statistics, 640, 1096, 1097, llOO FRADKIN, E., 751, ll61 FRADKIN, E.S., 893, 1366, 1426 frame linking number, ll26 framing, 1126, ll50, ll51 blackboard, ll51 FRAMPTON, P., 1259 FRANK-KAMENETSKII, M .D., ll59, ll60 FRANKE, G., 693 FRASER, C.M., 456, 893 free energy, 78, 471, 473, 475 bosons,676 effective classical, 467 fermions, 676 free particle amplitude for bosons, 637 for fermions, 637 from w --t 0 -oscillator, 761 fixed-energy amplitude, 752- 754 discontinuity, 754 spectral representation, 752 fluctuation factor, llO from harmonic oscillator, 140, 187 functional determinant, llO path integral, 101, 103, 135 quantum-statistical, 135 radial propagator, 762 wave function, 755 time evolution amplitude, 101, 109, llO momentum space, llO wave functions, 131 from w --t 0 -oscillator, 761 FREED , K .F ., 1081 FREEDMAN, D.Z. , 893 freely falling particle time evolution amplitude, 177 FREIDKIN, E. , 368 frequency cyclotron, 181, 1379 insertion, 306 Landau, 181, 181, 485, 1379 magnetic, 181, 485 Matsubara, 143, 144, 151, 154, 155 of wave, 12 optimal in variational perturbation theory,497 Rydberg, 955 shift, 264 Fresnel integral, 48, 97, 109, ll3, ll4, 145

Index FREY, E., 1082 FREYD, P., ll60 friction coefficient, 270 Drude, 1292, 1299, 1300, 1302, 1307 force , 265 FRIEDEN, B .R., 893 Friedmann model, 1392 universe, 1392 FRIEDRICH, H ., 456 FRISCH, H .L. , ll59 FROMAN, N., 372 FROMAN, P.O ., 372 fugacity, 597 local, 617, 677 FUJII, M., 1082 FUJIKAWA, K., 973, 1427 FUJITA, H ., 1082 FULLER, F.B., ll59 function Airy, 178 basis, 20 Bessel, 50, 156, 169, 410, 1021, 1033 modified, 50, 698 regulating, 977, 983, 985, 993, 997 Beta, 430, 690, ll35 confluent hypergeometric, 961 convex, 338, 460 correlation, 209, 249 connectedness structure, 289 in canonical path integral, 255 in magnetic field, 254 subtracted, 222, 225, 244, 264, 327, 333, 334, 336, 860 Debye, 1031, 1049 elliptic theta, 607, 687 Gamma, 160 Gelfand-Yaglom, 124-126, 128, 150 generalized zeta, 83 generating for canonical transformations, 10 Green, 43, 123, 211, 213, 214, 218 harmonic oscillator, 212 on lattice, 249 spectral representation, 217 summing spectral representation, 229 Hankel, 50 Heaviside, 43, 100, 166 Hurwitz zeta, 599 hypergeometric, 63, 715 confluent, 757 Kummer, 757, 758, 760, 961 , 1435

1547 Langevin, 1028 Legendre, 728, 729 Lerch,599 operator zeta, 83 parabolic cylinder, 1052 Polygamma, 1450 Poly logarithmic, 599, 687 proper vertex, 301 regulating, 923, 925, 951, 977, 984 Riemann zeta, 83, 163, 170 test, 25, 45, 710 vertex, 301 wave, 12, 46, 47, 133, 752 Whittaker, 755, 757, 766, 961, 1435 Wigner, 33, 1333 functional average, 209, 249 derivative, 209 covariant, 873 determinant antiperiodic boundary conditions, 349 free particle, 110 from Green function, 344 oscillator, 116 time-dependent frequency, 120 periodic boundary conditions, 349 gauge-fixing, 192, 858, 1128, 1150, 1178,1372 generating, 209, 243, 249, 250, 275, 340 canonical path integral, 259 Dirichlet boundary conditions, 258, 259 for connected correlation functions, 288 for vacuum diagrams, 294 momentum correlation functions , 255 influence, 1290, 1292, 1338, 1339 integral measure in Fourier space, 151 time-sliced, 101 integral, extension of path integral, 809 matrix, 39, 211 , 242, 254 fundamental composition law, 721 identity, 848 FURRY, W.H., 372,1017 GABAIX, X., 1524 GABAY, L., 1527 Gamma, 1502 distribution, 1442 function, 160 of option, 1503, 1504

GANBOLD, G., 569 GARCH model, 1496 GARDINER, C .W ., 1365, 1367 GARG, A ., 1260 GARROD, C., 207, 750 gas phase, 1247 GASPARD, P ., 456 gauge -fixing functional, 192, 858, 1125, 1128, 1150, 1178, 1372 -invariant coupling, 905 axial,878 field, 875 minimal coupling, 883 statistics interaction, 639 invariance, 1125, 1412 monopole, 882 London, 1157 nonholonomic transformation, 881 transformation, 185, 1124 nonholonomic, 778 transverse, 1125 Gauss distribution cumulative, 1506 in financial markets, 1432 integral, 48, 101, 112, 139, 145, 160, 186 invariant integral topological, 1116, 1117, 1120-1123, 1126, 1137, 1152, 1159 limit of stiff polymer structure factor, 1031 link invariant, 1115 martingale, 1505 polymer, end-to-end distribution, lO29 Gauss law, 1406 GAUSS, G.F. , 1159 Gaussian tails, 1432 GAVAZZI, G .M., xiv, 88, 891, 917 Gegenbauer polynomials, 715, 718, 728, lO39 addition theorem, 716 GELFAND, I.M., 88, 120, 206 Gelfand-Yaglom -like formula, 150 formula, 119, 121, 121, 122, 124, 126, 148, 345, 385, 1185 function, 124- 126, 128, 150 GEMAN, H., 1525, 1527 generalized coordinates, 1 functions (distributions), 25, 45 as limits of Bessel functions, 825

1548 hyperbolic distribution in financial markets, 1451 Piischl-Teller potential, 734 Rosen-Morse potential, 998 zeta function, 83 generating function for canonical transformations, 10 generating functional, 209, 243, 249, 250, 275,340 canonical path integral, 259 Dirichlet boundary conditions, 258, 259 for connected correlation functions, 288 for vacuum diagrams, 294 for vertex functions, 300 momentum correlation functions, 255 nonequilibrium Green functions, 1286 geodesic, 11, 776 coordinates, 790 geometric constraint, 579, 801 quantization, 88 GERBER, H.U ., 1526 GERMAN , G. , xvi GERRY, C.C ., 207, 750 GERVAIS , J.L., 892 GEYER, F., 568 GHANDOUR, G.1. , 567 ghost fields, 1330 states, 666 GIACCONI , P., 1158 GIACHETTI, R ., 565 GIACOMELLI , G., 1162 GILLAN , M.J. , 566 GILLES, H .P. , 1081 Ginzburg-Landau approximation, 305 energy functional, 306 expansion, 1235 GIORDANO, N ., 1260 GIULINI , D., 1367 glass, Vycor, 612 GLASSER, M .L., 567 GLAUM , K ., x GOBUSH , W. , 1082 GODDARD , P., 1162 GOEKE, K ., 893 GOHBERG , 1., 208 GOLDBERGER, M.L., 279, 373 GOLDSTEIN , H., 87 Goldstone-N ambu boson, 311, 324, 325 theorem, 311, 324, 325

Index GOMES, M ., 1158 GOMPPER, G ., 590 GONEDES, N., 1525 GOOVAERTS, M.J ., 207, 276, 569, GOPIKRISHNAN, P. , 1523, 1524 GORDUS, A., 1081 GORKOV, L.P., 1366 GOSSARD , A.C ., 1161 GOZZI, E ., 1367 GRABERT , H. , 368, 1260, 1365 GRACEY , J.A., 892

589, 973

gradient expansion of tracelog, 167 representation of magnetic field, 885, 886 torsion, 963 gradient expansion, 168, 174, 407, 412, 875 GRADSHTEYN, 1.S . , 49, 109, 114, 116, 133, 146, 156, 163, 169, 171, 178, 206, 245, 246, 269, 375, 408, 411, 633, 638, 660, 715, 729, 733, 754, 755, 757, 759, 762, 827, 829, 1021, 1033, 1040, 1045, 1052, 1211, 1258, 1446, 1447, 1489, 1524 grand-canonical ensemble, 78, 80 Hamiltonian, 78 quantum-statistical partition function, 77 GRANER, F. , 1525 GRANGER, C .W .J ., 1527 granny knot , 1103, 1114 Grassmann variables, 654, 693 anticommutation rules, 654 complex, 656 integration over, 654, 655, 656 nilpotency, 654, 661 gravitational field, classical motion in, 775 forces , 774 universality, 774 mass, 774 Greeks of options, 1503, 1504 Green function, 43, 123, 211-214, 218-220, 224 Schwinger-Keldysh theory, 1277 advanced, 1268 and functional determinant, 344 antiperiodic, 224, 225, 1266 first-order differential equation, 219 antiperiodic, 225 periodic, 222 time-dependent frequency, 226 harmonic oscillator, 212

1549 imaginary-time, 252, 1265 Klein-Gordon field , 1374 on lattice, 249 periodic, 220, 223 real-time for T f= 0, 1262, 1265 retarded, 214, 216, 226, 267, 1264, 1355 spectral representation, 217 summation, 229 time-ordered, 1269 Wronski construction Dirichlet case, 213 periodic, 231 GRIGELIONIS , B. , 1524 GRIGORENKO, 1., x GROSCHE, C ., 589, 772, 891 GROSJEAN, C.C. , 207 ground state lifetime, 1213 energy anharmonic oscillator, 466 hydrogen atom, 467 group conformal, 970 correspondence principle, 56 dynamical, 969 Euclidean, 56 knots, 1102, 1103 Lorentz, 970 permutations, 591 Poincare, 970 quantization, 56, 59 renormalization, 1237 space, amplitude on, 741 growth parameters perturbation expansion, 1214 precocious of perturbation expansion, 511 rate of stock, 1430 retarded of perturbation expansion, 511 G RUTER, P ., 694 GRYNBERG, G. , 1367 GREMAUD , B., 456 GUADAGNINI , E ., 1161 GUARNERI, 1. , 893 GUBERNATIS, J.E., 1366 GUIDA, R ., 566, 568 GUIDOTTI, C ., 207 GULYAEV, Y.V., 750 GUNARATNE, G .H ., 1524 GUTH, E ., 1081 Gutzwiller trace formula, 404 GUTZWILLER, M.C. , 129, 401,456, 736, 1017

gyromagnetic ratio, 749, 1415 HAAKE, F. , 368 HAAS, F., 1365 HABA, Z. , 1365, 1367 HABERL, P ., 457 HANGGI, P ., 368, 1260, 1365 HAENER, P. , 1526 HAGEN , C.R. , 891, 1158 HALDANE, F .D.M. , 1161 half-space, particle in, 575, 576, 578 Hall current, 1144 effect fractional quantum, 641 quantum, 71 resistance, 1145, 1158 HALPERIN, B .1., 1160, 1161, 1261 HALPERN, M .B ., 1426 HAMEL, G ., 87 Hamilton -Jacobi differential equation, 10, 370, 384 equation of motion, 3, 4, 41 formalism, 3 Hamiltonian, 2 Coulomb, 16 grand-canonical, 78 modified, 921 of financial fluctuations, 1488 pseudotime, 975 standard form , 90 HAMPRECHT, B. , 590, 1082, 1261 HANKE, A., 368 Hankel function, 50 HANNA, C ., 1160 HANNAY, J .H. , 1159 HAO, B.-L ., 1366 hard chaos, 405 HARDING, A.K., 567 harmonic hyperspherical, 716, 753 addition theorem, 718 oscillator, see also oscillator, 111 spherical, 59, 988 addition theorem, 712 in one dimension, 579 in three dimensions, 712 HARRISON, J.M., 1526 HARTLE, J., 890 HASHITSUME, N., 1365, 1367 HASSLACHER, B., 455 HATAMIAN, T.S. , x

1550 HATZINIKITAS, A., 892 HAUGERUD, H ., 567 HAUSDORFF, F ., 207 HAWKING , S., 890 HAYASHI, M., 567 HE, J ., 695 heat bath, 262 general particle in, 262 Ohmic, 268 photons, 268 master equation, 1340 oscillator in, 271 particle in, 266 Heaviside function , 43, 44, 100, 166, 221 heavy tails in financial distributions, 1432 HEBRAL, B. , 695 hedging Delta, 1504 dynamical Delta, 1504 of investment, 1498 Heisenberg -Euler formula, 1385 correspondence principle, 40, 41 equation of motion, 41, 749 Euler formula, 1385 matrices, 39, 41 model of ferromagnetism, 1035 operator, 40 picture, 39, 40, 41, 1263 for probability evolution, 1326 in nonequilibrium theory, 1263, 1272 spin precession, 750 uncertainty principle, 14 HEISENBERG, W ., 1427 HELFRICH , W. , 590 helium, superfluid, 591 , 606, 607, 612 helix double, DNA, 1117, 1117, 1118, 1120, 1159 circular, 1117 H ELLER, E.J. , 456 HENNEAUX, M. , 693 HERBST, I.W., 568 HERMANS, J.J ., 1082 Hermite polynomials, 132, 205, 353, 760 Hermitian -adjoint operator, 16 operator, 16 HEROLD, H., 568 Heron formula, 744 Hessian metric, 3, 54, 65, 86, 872 Heston model, 1481

Index HESTON, S.L., 1526 HIBBS, A.R, xiii, 207, 1366 high-temperature superconductor, xi, 545, 1147,1160 Hilbert space, 18 HILBERT, D., 1523 HILLARY, M ., 567 HIOE, F.T., 567 Ho, R ., xiv, 973 HOHLER, S., 568 HOLDOM , B ., 1523 HOLLISTER, P., x HOLM, C., xvi HOLSTEIN, B.R, 207 HOLZMANN, M. , 694, 695 HOMFLY knot polynomials, 1103, 1107, 1109, 1113, 1114, 1120, 1121, 1151, 1156, 1160 homogeneous universe, 1386 HONERKAMP, J ., 893 HONTSCHA, W. , 1260 Hopf link, 1110, 1110, 1112 HORNIG, D.F. , 567 HORTON , G .K. , 566 HORVATHY, P.A ., xvi, 207, 589, 692 HOSTE, J., 1153, 1160 HOSTLER, L.C., 973, 1017 HOVE, J. , 1426 HOWE, P.S., 893,1426 HSIUNG, A. , 1081 HSIUNG, C. , 1081 HSUE, C ., 890 HUANG , K. , 694 HUBBARD, J. , 696 Hubbard-Stratonovich transformation, 682, 1063, 1072, 1078 Hubble constant, 1391 HULET, RG. , 693 HULL, J ., 1522, 1523 HULL, T.E., 973 Hulthen potential, 1000 and Rosen-Morse system, 1000 extended, 1002 HURLEY, K. , 567 Hurwitz zeta function, 599 hydrogen -like atom, 72 atom, see also Coulomb system, 931 D-dimensional, 987 energy eigenvalues, 954 one-dimensional, 376, 451, 933 three-dimensional, 943, 949 two-dimensional, 933, 938

1551 hyperbolic distribution in financial markets, 1451 eigenvalue of stability matrix, 404 hypergeometric functions, 63, 715 confluent, 757 hyperspherical harmonics, 716, 753 addition theorem, 718 identical particle orbits, 592 identity Bianchi, 876 fundamental, 848 Jacobi, 4 resolution of, 651, 652 Ward-Takakashi, 325 i7]-prescription, 46, 47, 115 ILIOPOULOS , J., 893 ILLUMINAT I, F., 695 imaginary-time evolution amplitude, 141 spectral decomposition, 759 with external source, 238 Green function, 252, 1265 impact parameter, 71, 194 implied volatility, 1507 independence of coordinates, 809 index critical of field theory, 1227 Dow-Jones industrial, 1428 Maslov-Morse, 115, 129, 388, 396, 401, 403 Morse, 129 Nikkei-225, 1450 S&CP 500, 1429, 1436, 1487 indistinguishable particles, 591 induced emission, 1341, 1342, 1363 and absorption, 1363 metric, 775, 777 inequality for nonequilibrium Green functions, 1271, 1356 Jensen-Peierls, 460, 493, 544, 569, 1356 inequivalent compound knots, 1103, 1104 knots, 1108 simple knots, 1104 inertial mass, 774 INFELD, L., 973 infinite wall potential, 575, 576, 578, 579, 581

influence functional, 1290, 1292, 1338, 1339 dissipative part, 1290 fluctuation part, 1290 infrared (IR) cutoff, 814 divergence, 814 INGERSOLL , J .E., 1523 INGERSON, J., 1501 INGOLD, G.-L. , 368, 1365 INOMATA, A ., xiv, 207, 736, 750, 973, 1017, 1018, 1158 insertion of frequency, 306 of mass, 306, 366 instability of critical bubble, 1202 of vacuum, 1255 instanton, 1168, 1259 integrability condition, Schwarz, 7, 180, 639, 777, 778, 780, 848, 879, 1141 integral -equation for amplitude, 895 kernel for Schriidinger equation, 895 Bromwich, 1469 exponential, 156 Feynman, 364 Fresnel, 48, 97, 109, 113, 114, 145 functional, extension of path integral, 809 Gaussian, 48, 101, 112, 139, 145, 160, 186 principal-value, 47, 1270, 1381 Wilson loop, 1150 integration by parts, 105, 111 on lattice, 111 over boson variable, 646 over complex Grassmann variable, 656 over fermion variable, 657 over Grassmann variable, 654, 655 interaction, 272 exchange, 428 local, 1286 magnetic, 179 picture (Dirac), 42, 72, 1279 generating functional, 1286 time evolution operator, 1279 statistic, 592, 635, 639 for fermions, 637 gauge potential, 639 topological, 637, 639, 1089, 1137 interatomic potential in 3He, 1238

1552 interest rate, riskfree, 1500, 1505 internal energy, 78 interpolation, variational, 518 intersections of polymers, 1084 intrinsic geometric quantities, 792 invariance Einstein, 887 gauge, 1125, 1412 Lorentz, 887 monopole, magnetic gauge, 882 Poincare, 887 scale, 1075 under coordinate transformations, 804, 809, 811, 831, 834, 841, 851, 854 under path-dependent time transformations, 925 invariant for knots, 1153 for ribbons, 1153, 1155, 1156 Gauss integral for links, 1115 topological, 1085, 1089, 1116-1118, 1120, 1122 inverse hyperbolic eigenvalue of stability matrix, 404 Langevin function, 1029 parabolic eigenvalue of stability matrix, 404 lORI, G., 1528 ISERLES, A. , 203 isotopy of knots ambient, 1104, 1153, 1156 regular, 1104, 1153, 1156 isotropic approximation in variational approach,481 isotropic universe, 1386

Ito calculus, 1315 integral, 1351 lemma, 1319, 1321, 1431, 1473 relation of growth rates, 1478 ITO, K, 1082 It6-like lemma, non-Gaussian, 1473 ITZYKSON, C., 208, 287, 456, 893, 1426 J .ILIOPOULOS, 456 JACKIW, R., 367, 891 , 1158, 1160, 1161, 1163 JACKSON, J.D., 123, 755 Jacobi action, 795, 797, 799, 901 , 902, 904, 905, 909, 939, 940, 949, 950

Index identity, 4 polynomials, 63, 715, 728, 733 JACOD, J., 1526, 1527 JAENICKE, J ., 566 JAIN, J.K, 1161 JANKE, W. , 208, 367, 476, 565, 567, 568, 579, 590,1367 JANNER, A ., 207 JANUSSIS, A .D ., 207 Jensen-Peierls inequality, 460, 472, 493, 544, 569, 1356 JEVICKI, A. , 751, 892, 1426 JIZBA, P., x, 1365, 1525, 1526 JOHNSTON, D., xvi JONA-LASINJO, G., 893 Jones knot polynomial, 1109, 1111, 1114 JONES, C.E., 208, 457 JONES, R .F., 566 JONES, V ., 1160 JONES, W .F .R., 1160 Joos, E., 1367 Jordan rule, 15 JORDAN, P. , 591 JUNKER, G., 207, 695, 736, 751, 1017 JURIEV, D ., 1427 KAC, M. , 206-208, 1367 KAIZOJI, T ., 1524 KALLIN , C., 1161 KALLSEN, J., x KAMO, R ., xiv, 88, 890, 917 KARRLEIN , R. , 1261 KASHURNIKOV, V.A., 695 KASPI, V.M., 567 KASTENING, B., x, 367 KATO, T ., 208 Kauffman bracket polynomial, 1109, 1112 polynomial, 1109, 1109, 1110, 1153 relation to Wilson loop integral, 1153 KAUFFMAN, L.R., 1160 KAUL , R., 1259 KAWAI , T ., xiv, 88, 890, 917 KAZAKOV, D.L, 893 KAZANSKII, A.K., 367 KEHREIN, S., 368 KELDYSH , KV., 1364, 1366 KELLER, V., 1526 KENNEDY, J., 973 KENZIE, D.S., 1082 Kepler law, 437 KETTERLE, W., 693 KHANDEKAR, D.C., 207

1553 KHARE , A. , 1161 KHOLODENKO, A ., 1082, 1115, 1159 KHVESHCHENKO, D.V., 1160 KIEFER, C ., 1367 KIKKAWA, K., 456, 893 hlnk, 1167, 1168, 1170, 1171 action, 1178, 1188 KINOSHITA, T., 1260, 1427 Kinoshita-Terasaka knot , 1108 KIVELSON, S. , 1161, 1259 KLAUDER, J.R. , 695, 736, 750, 751, 1017 KLEIN , 0 ., 1366 Klein-Gordon equation, 1278 field, 1374 Green function , 1374 Klein-Kramers equation, 1295, 1297 overdamped, 1304 KLEINERT, A ., x, xii, xvi KLEINERT, H. , xiii-xv, 11, 67, 101, 161, 175, 207, 247, 255, 287, 301 , 367, 368, 390, 456, 476, 548, 565-569, 579, 590, 641 , 693- 695, 750, 751, 891893, 907, 913, 917, 920, 930, 973, 1017, 1018, 1082, 1083, 1160, 1162, 1163, 1238, 1259-1261, 1365-1367, 1426, 1427, 1524- 1526 KLIMIN , S.N. , 569 KNEUR, J.L., 695 knot composite, 1107 compound, 1102, 1105, 1108 inequivalent, 1103, 1104 Conway-Seifert, 1108 crossings in graph, 1085, 1103, 1104, 1105, 1107, 1113, 1114, 1119, 1121 equivalent, 1101 exceptional, 1108 granny, 1103, 1114 graph crossing, 1085, 1103, 1104, 1105 overpass, 1106 underpass, 1104, 1106, 1120 group, 1102, 1103 inequivalent, 1108 invariant, 1153 Kinoshita-Terasaka, 1108 multiplication law, 1102 polynomial, xi, 1103 Alexander, 1103, 1106-1108, 1160 Alexander-Conway, 1112 BLM/Ho, 1109 Conway, 1109

HOMFLY, 1109 Jones, 1109, 1111 Kauffman, 1109, 1109, 1110, 1153 Kauffman bracket, 1109, 1112 and Wilson loop integral, 1153 X, 1109, 1109 prime, 1102, 1108 simple, 1102, 1107, 1108 inequivalent, 1104 square, 1103, 1114 stereoisomer, 1108 trefoil, 1101, 1101 KOBZAREV, LY., 1259 KOGAN , 1.1. , 1160, 1427 KOMAROV , L.L , 568 KONISHI, K., 566, 568 KOPONEN , 1. , 1524 KORENMAN , V ., 1364 KORNILOVITCH , P. , xvi KOSOWER, D.A., 1426 Kosterlitz-Thouless phase transition, 602 KOUVELIOTOU , C., 567 KOYAMA , R., 1082 KRAMER, M ., x, 695 KRAMERS , H.A. , 455, 1366 Kramers-Moyal, 1472 Kramers-Moyal equation, 1472 KRATKY, 0. , 1082 KRAUTH, W. , 694 KRIEGER, J.R., 373 KROLL, D.M., 590 KRONER, K.F ., 1526 KRUIZENGA , R., 1499 Kubo stochastic Liouville equation, 1316, 1336, 1347 KUBO , R ., 1365, 1367 KUCHAR, K ., 891 , 917 KURZINGER, W., 568, 569 KUHNO, T., 1160 Kummer functions, 757, 758, 760, 961, 1435 KUPSCH , J ., 1367 KURN , D .M ., 693 kurtosis excess, 1441 kurtosis of financial data, 1438, 1441, 1453, 1511, 1515 KUSTAANHEIMO, P. , 972 Kustaanheimo-Stiefel transformation, 943, 953, 1397, 1398 KWEK, L.C ., 1527 Lagrange brackets, 7, 8

1554 formalism, 2, 1295 multiplier, 738 LAGRANGE, J.L., 87 Lagrangian, 1 Laguerre polynomials, 759, 960 LAIDLAW, M.G.G., 589, 692 LALOE, F ., 694, 695 Lamb constant, 1345 shift, 1337, 1342, 1345, 1420 LAMBERT, J .H., 439 LAMOUREUX, C .G., 1523 Landau -Ginzburg expansion, 1235 approximation, 305 frequenc~ 181, 181, 485, 1379 level, 766 orbit, 181 , 1144 radius, 767 LANDAU, L.D ., 87, 88, 179, 306, 373, 568, 751,759,766,960,961, 1017, 1261 Landau-Lifshitz equation, 749 LANDWEHR, G ., 569 Lande factor , 1415 Langer correction, 376, 983 LANGER, J ., 696 LANGER, J.S., 1259, 1261 LANGER, R.E. , 372,1017 Langevin equation, 1289, 1365 operator form, 1308 quantum, 1308 semiclassical, 1308 with inertia, 1309 function , 1028 LANGEVIN, P., 1367 LANGGUTH, W., 750 LANGHAMMER, F., xvi LANGRETH , D., 1364 Laplace -Beltrami operator, 53, 55, 56, 59, 66, 896, 907, 909, 1348 transform, 752 Laplacian, 52, 56 canonical, 55, 56 covariant, 896, 900, 907 covariant, Weyl, 963 lattice, 106 large-order perturbation theory, 1208, 1209, 1211, 1212, 1214, 1218 large-stiffness expansion, 1041, 1042, 1047, 1057, 1059 LARIN, S.A., 1083, 1261

Index LARKIN, A.I., 368, 1260 LARSEN, D ., 569 LASTRAPES, W.D., 1523 lattice completeness relation, 107 derivative, 104 Green function, 249 Laplacian, 106 models quantum field theories, 155 statistical mechanics, 518 orthogonality relation, 107 LAUGHLIN, R.B., 1160, 1161 law Ampere, 875, 1406 Ampere, 71 angular momentum conservation, 436 composition for time evolution amplitude, 90, 700 continuity, 17 current conservation, 17 Dulong-Petit, 176, 327, 605-607, 625, 634 energy conservation, 14, 75 energy conservation law, 1169 for multiplication of knots, 1102 Gauss, 1406 Kepler, 437 momentum conservation, 301, 1086 Newton's first, 774 probability conservation, 16, 1292, 1297, 1300, 1334, 1335, 1340 scaling for polymers, 1024, 1062, 1069, 1075, 1082 LAWANDE, S.V ., 207 LAX, M., 751 LAZZIZZERA, I., 1163 LE GUILLOU, J.C. , 1260 LEDERER, K. , 1163 LEE, T.D ., 694 Legendre functions, 728, 729 polynomials, 712, 727 associated, 724, 725 Legendre transform Euclidean, 137 LEGGETT, A .J ., 368, 1260, 1365, 1367 LEGUILLOU , J.C., 568 LEHR, W ., 1365 LEIBBRANDT, G. , 159 LEINAAS, J .J., 589, 1161 LEITE, V.B .P ., 1524 lemma

1555 Ito, 1319, 1321, 1431, 1473 Ita-like, non-Gaussian, 1473 Riemann-Lebesgue, 73 LEMMENS, L .F ., 569, 693, 695 length bond, 1019 effective, 1032 classical of oscillator, 140 coherence, 1235 oscillator, 536 persistence, 1036, 1040 Planck, 1390, 1391 quantum of oscillator, 133 scattering, 610 thermal, 139, 595, 602 Lenz-Pauli vector, 963 Lerch function, 599 LERDA, A ., 1161 level -splitting formula, 1194 Landau, 766 shift due to tunneling, 1166 formula, 279, 520 operator, 279 level-splitting, 1190 quadratic fluctuations, 1194 Levi-Civita tensor, 783 transformation, 933, 933, 935 LEVINE, M.J., 1260 Levinson theorem, 1182 Levy -Ito theorem, 1460 -Khintchine formula, 1460 -stable distributions, 1458 distribution, 1432, 1432, 1458 asymmetric, 1434 truncated, 1434, 1437 tail, 1433 weight, 1460 LEWIS, J .T., 1365 LI, X .L ., 368 LI-MING, CHEN, x LIANG, W .Y., 569 LICKORISH, W.B .R., 1153 Lie algebra, 56 rotation group, 58 expansion formula, 61, 1143 lifetime metastable state, 1213

universe, 1391 LIFSHITZ, E.M., 87, 88, 179, 373, 568, 751, 759, 766, 960, 961, 1017, 1261 light scattering, 1030 velocity, 13 light tails, 1432 LILLO, F., 1526 limit classical, 154 degenerate, 631 strong-coupling, xi, 494 thermodynamic, 288 limit theorem, central, 1457, 1464, 1467, 1482, 1492, 1512 generalized, 1459 Lindblad equation, 1336, 1340 LINDBLAD, G., 1367 LINDQUIST, W .B., 1260 line width, 1337, 1341 linear oscillator, see also oscillator, 111 response theory, 141, 1262, 1264, 1277 space, 25 LINETSKY, V., 1527 link, 1120, 1159, 1160 Hopf, 1110, 1110, 1112 polynomial Alexander, 1120 simple, xxviii, 1119, 1121 linked curves, 1116 linking number, 1116, 1117, 1120 frame, 1126 Liouville equation stochastic Kubo, 1316, 1336, 1347 Wigner equation, 33 Liouville equation, 33 LIPATOV, L.N., 1259 LIPOWSKI, R., 590 Lippmann-Schwinger equation, 73, 74, 344, 610, 1091, 1158 LIPTSER, R.S. , 1525 liquid Fermi, 1238 phase, 1247 Lm, S., 566 Lm, Y ., 1522 local, 98 basis functions, 19 Boltzmann factor, 330, 458, 459 classical momentum, 369 conservation law, 17

1556 density of states, 399, 401 classical, 409 quantum-mechanical, 406 diagram, 811 dilatations, 970 expectation value, 460 field energy, 1124 fluctuation square width, 462 fugacity, 617, 677 interaction, 1286 pair correlation function, 524, 525 partition function, 459, 490 supersyrnrnetry, 1423 transformation, 5 of coordinates, 884 trial action, 459 U(l) transformations, 884 locality, 5, 591 LOEFFEL, J.J., 566 log-return of financial asset, 1431 London equations, 1157 gauge, 1157 LONDON, F ., 1162 LONDON, H., 1162 longitudinal fluctuation width, 525 projection matrix, 310, 479 trial frequency, 479 loop diagram, 283 expansion, 307 integral Gauss, for links, 1115 Wilson, 1150 Lorentz frame, 887 group, 970 invariance, 887 transformations, 887 LORETAN, M., 1523 loxodromic eigenvalue of stability matrix, 404 LOZANO, G., 1158 Lu, W.-F., x LUCKOCK, H .C ., 1527 LUKASHIN, A .V ., 1159, 1160 LUNDIN, J., 1427 LUNDSTROM, E ., 1427 LUTTlNGER, J .M ., 694 Lux, T., 1523 LYASHIN, A., 1524

Index LYKKEN, J.D., 1161 LEVY, P., 1527 MACCHI, A., 566 MACKENZIE, R., 456, 893 MACMILLAN, D ., 567 MADAN, D., 1525, 1527, 1528 MAGALINSKY, V.B., 567 MAGEE, W.S., 1082 magnetic field, 179, 484 correlation function, 254 polaron in, 544 time evolution amplitude of particle, 179, 181, 183 flux quantization, 1088, 1090 forces, 179 frequency, 181, 485 interaction, 179, 905 moment anomalous, 1145, 1210 electron, 1415 monopole, 748, 875, 1005 Dirac theory, 881 susceptibility, 1036, 1041 trap for Bose-Einstein condensation, 615 anisotropic, 620 magnetization, 336, 337, 338 magneton, Bohr, 181, 1415 Magnus expansion, 35, 203 MAGNUS, W., 207, 208 MAHESHWARI, A., 207 MAHNKE, R. , 1524 MAKI, K., 301 MAKOWIEc, D. , 1524 MALBOUISSON, J.M.C., 1158 MALDAGUE, P ., 1364 MALTONI, F ., 1158 MANDELBROT, B.B., 1523, 1528 MAN'KO, V .I., 207 MANNING, R .S., 456 MANTEGNA, R.N., 1523, 1526 MANUEL, C ., 1158 many -boson orbits, 592 -fermion orbits, 637 mapping from flat to space with curvature and torsion, 789 nonholonomic, 781, 789, 894 MARADUDIN, A., 565 MARINOV, M.S., xiv, 891, 909, 917, 1426 market, efficient, 1500

1557 MARKLUND, M., 1427 MARKOFF , A.A., 1081

Markov process, 1462 MARSDEN, J .E., 87 MARSHALL, J.T. , 569 MARTELLINI, M . , 1161 MARTHINSEN, A ., 203 MARTIN, A. , 456, 566, 893 MARTIN , P.C. , 367, 1364 MARTINEZ PENA , G.M., 567 martingale, 1477, 1481 distribution, 1478 Esscher, 1481 Gaussian, 1505 natural, 1481, 1495, 1496 MARTINIS, J .M . , 368 Maslov -Morse index, 115, 129, 388, 396, 401, 403 MASLOV , V .P., 115,455 MAS OLIVER, J., 1528 mass coordinate-dependent, 871 gravitational, 774 inertial, 774 insertion, 306, 366 polaron, 542 term, 1236 time-dependent, 871 master equation, 1335, 1336 photon bath, 1340 MATACZ , A . , 1524, 1528 material waves, 11 MATIA, K., 1524 matrix density, 33, 140 diffusion, 1297 functional, 39, 211, 242, 254 Heisenberg, 39, 41 Hessian, 3, 54, 65, 86, 872 normal, 647 Pauli spin, 63, 748 projection longitudinal, 310, 479 transversal, 310, 479 representation of spin, 743 scattering, 67, 190 scatteringT, 191, 192, 610 scatterinT, 344 stability, 403 symplectic unit, 7 matrix T , 74

Matsubara frequencies, 143, 144, 151, 154, 155, 219, 224, 243, 248 even, 143 odd, 224 MATTHEWS, M .R. , 693 Maupertius principle, 380, 968 Maxwell action, 1405 equations, 1406 theory, 1124 Maxwell distribution, 1325 MAZUR, P. , 1367 MCCAULEY , J .L., 1524 MCCUMBER, D .E . , 1261 MCGURN, A .R. , 565 McKANE, A . J., 1083, 1367, 1527 McKEA N, H .P ., 893, 911, 1082 McLAUGHLIN , D .W., 750 mean motion, 437 mean-field approximation, 306, 312, 338 effective potential, 338 measure functional integral in Fourier space, 151 time-sliced, 101 of path integral in space with curvature and torsion, 794,799 transformation of, 979, 991 , 993, 994 of path integration, 773, 890 of perturbatively defined path integral in space with curvature, 847 process, 1455 mechanics classical, 1 quantum, 1, 11 level shift due to tunneling, 1166 quantum-statistical, 76 statistical, 76 Mehler formula, 132, 205, 554 MEINHARDT, H ., 590 Meissner effect, 312, 1148 MELLER, B ., xvi

melting process, 1369 MENDONQA , J.T . , 1427 MENDOZA , H.V ., 567 MENOSSI, E ., 1158 MENSKII, M .B. , 890 MENZEL-DoRWARTH, A., 1365 MERMIN, D., 1366 MERTON , R . C., 1499, 1501, 1527 MERZBACHER, E. , 87, 373

1558 MESSIAH , A., 87 metastable phase, 1247 state, 1248 metric, 52 -affine space, 773, 784, 793, 799, 901 dynamical, 381 Euclidean space, 1370 Hessian, 3, 54, 65, 86, 872 induced, 775 Minkowski space, 669, 1370 Robertson-Walker, 1388 tensor, 52 MEWES, M .-O . , 693 Mexican hat potential, 310 MEYER, H., 566 MEYER, M., 1522, 1523 MICCICHE, S., 1526 MICHELS, J.P.J., 1114, 1159 midpoint action, 793 expansion, 790 prescription, 794 MIELKE, A., 368 MIKHAILOV, S., 1527 MIKOSCH , T ., 1524 MILLER, W.H., 566 MILLET, K.C., 1153, 1160 MILLS, L.R., 569 MILNE , F ., 1527 MILTON , K.A., 1017 minimal coupling, 905 gauge field, 883 substitution, 180, 905, 1141 subtraction, 160 Minkowski space, 779, 970 critical bubble, 1255 metric, 669, 1370 MINTCHEV, M . , 1161 MIRANDA, L .C ., 1524 MISHELOFF , M.N. , 208, 457 MITTER, H., 567 MIURA, T ., 891 MIYAKE, S.J ., 568 MIZRAHI, M.M., 890 mnemonic rule, 1180 beyond Ito, 1475 for free-particle partition function, 140, 187 Ito, 1431 Mo, S. , 1426 MOATS, R.K., 567

Index mode negative-eigenvalue for decay, 1203, 1204 zero, 217 model Black-Scholes, 1499 Drude for dissipation, 1292 for tunneling processes, 1164 Friedmann, 1392 Ginzburg-Landau, 1235 Heisenberg, of ferromagnetism, 1035 Heston, 1481 lattice quantum field theories, 155 statistical mechanic, 518 nonlinear IJ, 738, 748, 805, 1038 random chain for polymer, 1019 Thomas-Fermi, 419 modified Bessel function, 50 Hamiltonian, 921 Poschl-Teller potential, 996 time evolution operator, 921 Moebius strip, 1117 molecules DNA, 1117, 1117, 1118, 1120, 1159 circular, 1117 moment magnetic electron, 1415 moments in polymer end-to-end distribution, 1021 topological, 1127 momentum angular, 56 conservation law, 301, 1086 Fermi, 419, 631 local classical, 369 operator, 92 space path integral of Coulomb system, 964 wave functions in, 28 transfer, 70 monopole, magnetic, 748, 875, 882, 1005 Dirac theory, 881 gauge invariance, 882 gauge invariance, 882 gauge field, 882 gauge invariance, 1005 spherical harmonics, 743, 1008 MONTROLL, E.W ., 567 MOORE, G., 695, 1427 MORANDI, G., 589, 692, 1158

1559 Morse -Maslov index, 115, 129, 388, 396 index, 129 potential, 983, 985 MORSE, P.M., 132, 205, 1017 MOSER, J.K., 87 motion Brownian, 1349 equation of, 41 mean, 437 Mott scattering, 448 MOUNT, K .E., 455 move, Reidemeister in knot theory, 1103, 1104, 1151 MUHLSCHLEGEL, B., 696 MULLENSIEFEN, A., 568 MUELLER, E .J ., 695 MUKHI, S. , 893 MUKHIN, S., x multiplication law for knots, 1102 multiplicity, 283 multiply connected spaces, 1084, 1088 multivalued basis tetrads, 780 triads, 778, 780 MURAKAMA , H ., 1082 MUSTAPIC, I., xvi, 750, 1017, 1261 MYRHEIM, J. , 589 N0RSETT, S.P ., 203 NAGAI, K. , 1081 NAHM, W., 1162 NAKAzATo, H. , 1365 N ambu-Goldstone boson, 311, 324, 325 theorem, 311, 324, 325 NAMGUNG, W., 567 NAMIKI, M ., 1365 natural martingale, 1481, 1495, 1496 units, 470 atomic, 954 NEDELKO, S.N. , 569 negative-eigenvalue solution, 1203, 1204, 1213, 1214, 1251 NELSON, B .L ., 207, 917 NELSON, E ., 206, 208 NETZ, R.R., 590 NEU, J., 1083, 1261 Neumann boundary conditions, 153, 230, 1049 NEUMANN, M ., 566 Neumann-Liouville expansion, 35, 203

neutron scattering, 1030 stars, 484, 774 NEVEU, A. , 455 Newton's first law, 774 NEWTON, I. , 87 Nikkei-225 index, 1450 nilpotency of Grassmann variables, 654, 661 node, in wave function, 1166 noise, 1307, 1346 quantum, 1349 white, 1309, 1319, 1349, 1352, 1355, 1430 NOLAN, J.P ., 1523 non-Gaussian fluctuation Debye-Waller factor , 1454 nonequilibrium Green function bosons, 1266 fermions, 1266 generating functional, 1286 inequalities, 1271, 1356 perturbation theory, 1286 spectral representation, 1265 Heisenberg picture, 1263, 1272 quantum statistics, 1262, 1277, 1283, 1286 Schrodinger picture, 1263 nonholonomic coordinate transformation, 778 gauge transformation, 778, 881 mapping, 781, 789, 894 objects, 887 variation, 784 auxiliary, 785 nonintegrable mapping, 781, 789, 894 nonlinear a-model, 738, 748, 805, 1038 nonlocal action, 263 diagram, 812 NORISUYE, T., 1082 normal -mode expansion, 1172 coordinates, 790 matrix,647 part of Bose gas, 618 product, 1360, 1363 NORREYS, P. , 1427 NORTHCLIFFE, A., 456 n-point function , 250, 251, 292, 294 connected, 292, 303 vertex function, 303 number

1560

Index

Bernoulli, 170, 633 Euler-Ma8cheroni, 156, 542, ll98 frame linking, ll26 linking, 1117, 1120 Tait, ll09 twist , ll51 winding, 594, 1085 writhing, ll22, 1122, 1163 NYQUIST , H ., 1364 objects of anholonomy, 887 observables commuting, 4 operators, 31 OCNE ANU, A., 1160 O ' CONNELL, R.F ., 368, 567, 1365 O ' GORMAN, E.V., 566 Ohmic dissipation, 265, 1293 OKANO , K. , 1365 OKOPINSKA, A., 567 OKUN, L.B., 1259 OLAUSSEN ,

K ., 456

old-fa8hioned perturbation expansion, 276 OLSCHOWSKI , P., 368, 1260 OMOTE, M., 890 one-dimensional oscillator, 759, 760 radial wave functions, 759, 760 one-particle irreducible (lPI) correlation fllllctions , 300 diagrams, 300, 304, 318, 319 vacUUIll, 322, 501, 871 vertex functions, 300 one-particle reducible diagrams, 318, 498 one-point function, 292, 301 operation, skein, llll operator annihilation, 644, 955, 956, 969 creation, 644, 955, 956, 969 density, 33, 1277 dilation, 955, 957 field , 641 Heisenberg, 40 Hermitian, 16 Laplace-Beltrami, 53, 55, 56, 59, 66, 896, 907, 909, 1348 level shift, 279 momentum, 92 observable, 31 ordering problem, xiv, 17,55, 794,1297, 1364 solved, 794, 804 position, 92 pseudotime evolution, 923, 924

resolvent , 45 tilt, 955 tilting, 567 time evolution, 34, 35, 37- 40, 43, 72, 77, 89, 90, 94, 250 interaction picture, 42 retarded, 38 time-ordering, 35, 37, 229 zeta function , 83 optimization in variational perturbation theory, 464, 486, 488, 491- 493, 500, 516,544 option American, 1500 call, 1498, 1500 Delta, 1501, 1503, 1504 European, 1500 Gamma, 1503, 1504 Greeks, 1503, 1504 of financial a8set, 1428 price, 1498 Black-Scholes formula, 1507 strike, 1505 put, 1498, 1500 smile, 1507 Theta, 1503, 1504 Vega, 1503, 1504 orbits cla8sical, 1 identical particles, 592 Landau, 1144 many-boson, 592 many-fermion, 637 order field, 1235, 1247 of operators, causal, 35 parameter, 1235, 1243 superconductor, 1243 problem for operators, xiv, 17, 55, 794, 1297, 1364 solved, 794, 804 ORSZAG , S.A. , 566, 1523 orthogonality of time and space, 1387 relation, 19 ba8is Dads, 51, 776 lattice, 107 orthonormality relation, 19 oscillator anharmonic, 467 D=l

spectral representation, 133

1561 fixed-energy amplitude radial, 755 spectral representation, 756 fluctuation factor , 112-115, 117-119 free particle amplitude from w --> 0 limit, 761 from Coulomb system, xiii, 935, 936, 946, 947, 952, 953, 956, 957, 959, 969, 1399 functional determinant, 116 in heat bath of photons, 271 in Ohmic heat bath, 268 length scale, 536 classical, 140 quantum, 133 path integral, 111, 143 radial, 983, 987 principal quantum number, 757 wave function , 757, 758 wave functions for D = 1, 759, 760 radial amplitude, 985, 993 time evolution amplitude, 111 time-dependent frequency functional determinant, 120 path integral, 127, 147 wave function , 131 wavelength classical, 140 quantum, 133 OSORIO, R. , 1525 OTEO, J .A., 203, 207 OTTO, M., 1527 OTTO , P ., 567 O UVRY, S. , 1158 OVCHINNIKOV, Y.N. , 368, 1260 overcompleteness relation, 651 overdamped Fokker-Planck equation with inertia, 1318 Langevin equation with inertia, 1310 overdamping, 1303 overheated phase, 1247 overpass in knot graph, 1106 PACHECO, A.F. , 567 packet, wave, 14 Pade approximation, 1076 PAGAN , A. , 1527 pair Cooper, 1236 correlation function , 524, 525 field, 1238

terms in second quantization, 674 in superconductivity, 675 Wick contraction, 251 PAK, N.K. , 1016, 1158 PAL, M., 1524 PALDUS , J ., 567 PANIGRAHI, P .K. , 1161 PAPADOPO ULOS, G., 893 PAPANICOLAOU, G., 1522 PAPANICOLAOU, N ., 751 parabolic coordinates, Coulomb wave functions, 957 cylinder function , 1052 eigenvalue of stability matrix, 404 parameter impact, 71 order, 1235, 1243 skewness, 1459 stability, 1459 Pareto -Levy-stable distributions, 1458 distributions in financial markets, 1432 tail, 1433 PARETO, V ., 1523 PARISI, G., 1083, 1260, 1367 PARKER, C .S., 566 partial integration, 105, 111 lattice, 111 summation, 105, 111 particle density, 137 distribution, 137, 138, 155, 182 classical, 138 Coulomb system, xxii, 484 free radial propagator, 762 in a box, 579, 581, 582 in half-space, 575, 576, 578 in heat bath, 262 in heat bath of photons, 266 in magnetic field action, 179, 181 fixed-energy amplitude, 765 radial wave function, 766, 769 spectral representation of amplitude, 764, 766 time evolution amplitude, 179, 181, 183 wave function , 763, 766 indistinguishability, 591 number, average, 78

1562 on a circle, 571, 574, 581 on sphere, effective potential, 801 on surface of sphere, 57 orbits ensemble of bosons, 592 ensemble of fermions, 592 identical, 592 relativistic, 1368 and stiff polymer, 1370 path integral, 1370, 1371 particle, spinning amplitude, 743 particles, many at a point, 654 partition function, 660 Bose particles, 647 classical, 76 density, 136, 463, 548 fermions, 660 grand-canonical quantum-statistical, 77 local, 459, 490 quantum-mechanical, 77 relativistic, 1410 quantum-statistical, 77 path classical, 2 closed, in action principle, 784 collapse, xv, 701 , 706, 726, 732, 751, 918, 919, 931, 1241 energy-entropy argument, 919 fixed average time evolution amplitude, 237 in phase space, 97, 137 order in forward- backward path integral, 1283 path integral, 89, 93- 100 classical amplitude, 384 probability, 1331 coordinate invariance in time-sliced formulation, 800 perturbative definition, 809 Coulomb system, xiii, 931 relativistic, xi, 1397 equivalent representations, 901 Feynman's time-sliced definition, 89 divergence, 918 for probability, 1289 for zero Hamiltonian, 91 forward-backward, 1333 path order, 1283 free particle, 101, 103, 109, 110 momentum space, 110

Index freely falling particle, 177 in dionium atom, 1004 measure, 773, 890 in space with curvature and torsion, 794,799 oscillator, III time-dependent frequency, 127, 147 particle in magnetic field, 179, 181, 183 perturbative definition, 287 calculations in, 804 measure of path integration, 847 quantum-statistical, 135 oscillator, 143 radial, 700 relativistic particle, 1370 and stiff polymer, 1370 reparametrization invariance, 1371 solvable, 101, 111, 974 spinning particle, 743 spinning top, 741, 742 stable for singular potentials, 921 time-sliced Feynman,91 in space with curvature and torsion, 798 velocity, 189, 192 path-dependent time transformation (DK), 982, 984, 985, 993, 995, 998, 1000, 1001, 1012 reparametrization invariance of, 925 pattern, diffraction, 29 PATTON, B ., 1364 Pauli algebra, 667 exclusion principle, 591 spin matrices, 63, 718, 748 Pauli vector, 963 PAULI, W. , 388 Pawula theorem, 1472 PAWULA, R.F., 1525 PEAK, D ., 750 PEARSON, K., 1081 PECHUKAS, P. , 456 PEETERS, B., 892 PEETERS, F.M. , x, 569 PELSTER, A., x, 255, 367, 368, 548, 568, 569, 590,695,891, 1017, 1367 PELZER, F. , 1365 PEPPER, M ., 1161 PERCIVAL, I.C., 456 periodic boundary conditions, 126, 167, 219, 222, 242, 247, 250, 256

1563 functional determinant, 349 current, 247 Green function, 220, 221, 223, 248 Euclidean, 240 permutation group, 591 persistence length, 1036, 1040 perturbation coefficients precocious growth, 511 retarded growth, 511 expansion Bender-Wu, 353 covariant, 865 large-order, 1208 path integral with ,,-function potential, 770 theory, 272, 1277 Brillouin-Wigner, 279 cumulant expansion, 274, 278, 294, 493 large-order, 1209, 1211, 1212, 1214, 1218 nonequilibrium Green functions , 1286 Rayleigh-Schriidinger, xi, 276, 276, 280 scattering amplitude, 340 variational, xi, 458, 496, 496 via Feynman diagrams, 276 perturbative definition of path integral, 287, 804 coordinate invariance, 809 measure of path integration, 847 phase gas, 1247 liquid, 1247 metastable, 1247 overheated, 1247 shifts, 1175, 1177, 1181, 1184 Shubnikov, 1090 slips in thin superconductor, 1242 space, 3, 98 paths in, 97, 137 transition, 1246 Kosterlitz-Thouless, 602 phenomena collective, 681 entanglement, 1084, 1088, 1101, 1127 phenomena, critical, 1261 PHILLIPS, P.C.B., 1523 PHILLIPS, W.D., 1260 photoelectric-effect, 13 photon bath master equation, 268, 1340

physics of defects, 781, 783 PI, S.-Y., 891, 1158 picture Heisenberg, 39, 40, 41, 1263 for probability evolution, 1326 in nonequilibrium theory, 1263, 1272 interaction (Dirac), 42, 72 generating functional, 1286 time evolution operator, 1279 Schriidinger, 40, 41 in nonequilibrium theory, 1263 PINTO, M.B. , x, 695 PIPPARD, A .B. , 1162 PITAEVSKI, L .P., 87 PITMAN, J., 750 Planck constant, 13 length, 1390, 1391 PLANCK, M., 1366 plane wave, 13 PLASTINO, A., 893 PLEROU, V ., 1523, 1524 PLISKA, S.R ., 1526, 1528 PLO, M., 567 PODOLSKY, B ., 88 POSCHL, G., 751 , 1017 Piischl-Teller potential, 729 general, 734 Poincare group, 970 POINCARE, H., 751 point conjugate, 129 transformation, 5 turning, 129 Poisson brackets, 4 , 8, 9, 39, 56, 663 equation, 421 , 422, 1406 summation formula, 29, 156, 264, 573, 575, 580, 582 polar coordinates, 697, 705, 946 decomposition of Coulomb amplitude, 957 polaron, 533, 536 in magnetic field, 544 mass, 542 polaronic exciton, 544 poles from bound states, 1017 POLLOCK, E.L., 695 POLYAKOV, A .M ., xv, 1159, 1426 POLYAKOV, D. , 1427 POLYCHRONAKOS, A., 1161

1564 Polygamma functions , 1450 Poly logarithmic functions , 599, 687 polymer Chern-Simons theory, 1124, 1127 critical exponent, 1024, 1062, 1069, 1075, 1077, 1083 end-to-end distribution, 1019, 1020 cumulants, 1023 Daniels, 1042 exact, 1024 Gaussian approximation, 1029 moments, 1021 rod-limit, 1032 saddle point approximation, 1028 short-distance expansion, 1026 entangled, 1084, 1088, 1101 excluded-volume effects, 1062, 1063, 1069, 1070 field theory, 1070 flexibility, 1051, 1054 Flory theory, 1069 Gaussian random paths structure factor , 1031 linked, 1116 moments arbitrary stiffness, 1047 Gaussian limit, 1032 rod-limit, 1032 physics, 1019 rod limit, 1032 structure factor , 1033 scaling law, 1024, 1062, 1069, 1075, 1082 self-entangled ring, 1158 semiclassical approximation, 1065 stiff, 1032 polynomial Alexander, 1103, 1106- 1108, 1160 generalized to links, 1120 Bernoulli, 245 BLM/ Ho, 1153, 1156 Gegenbauer, 715, 728, 1039 addition theorem, 716 Hermite, 132, 205, 760 HOMFLY, 1103, 1107, 1113, 1114, 1120, 1121, 1151, 1156, 1160 Jacobi, 63, 715, 728 Jones, 1114 knot , xi, 1103 Alexander, 1108 Alexander-Conway, 1112 BLM/Ho, 1109 Conway, 1109 HOMFLY, 1109

Index Jones, 1109, 1111 Kauffman, 1109, 1109, 1110 Kauffman bracket, 1109, 1112 and Wilson loop integral, 1153 X , 1109, 1109 Laguerre, 759, 960 Legendre, 712, 727 associated , 724, 725 PoPov, V .N., 192, 893, 1163, 1261 POROD, G., 1082 portfolio riskfree, 1504 position operator, 92 Post Laplace inversion formula, 1469 POST, E ., 1525 postpoint action, 792, 802 expansion, 790 prescription, 794 postulate, Feynman, 809, 832, 835 potential champaign bottle, 310 chemical, 77, 597, 1071, 1357 double-well, 472, 473, 522, 1164, 1165, 1168, 1169, 1190 convex effective potential, 338 particle density, 479 effective, 308, 336, 337, 900, 916 derivation, 978 in space with curvature and torsion, 799 on sphere, 801 effective classical, 328, 333, 463 Coulomb, 483, 567 external, 802 general Rosen-Morse, 997, 998, 1000, 1002 Hulthen, 1000 extended, 1002 infinite wall, 575, 576, 578, 579, 581 interatomic in 3He, 1238 Mexican hat, 310 Rosen-Morse, 994, 996, 1183, 1195, 1213 singular, 918 statisto-electric, 1140 vector, 802 in Fokker-Planck equation, 1304 statisto-electromagnetic, 1140 statisto-magnetic, 1124 time-sliced action, 802 POTTERS, M., 1523, 1525, 1528 PRAETZ, P., 1525 PRANGE, R .E ., 1524

1565 K ., 1526 precession, Thomas, 1145 precocious growth of perturbation expansion, 511 premium, 1498 prepoint action, 793 expansion, 790 prescription, 794 prescription i'f/, 46, 47, 115 midpoint, 794 postpoint, 794 prepoint, 794 PRESILLA , C., 893 PRESS , W.H., 1366 pressure, 81 price of option, 1498 strike, 1505 prime knot, 1102, 1107, 1108 principal quantum number radial oscillator, 757 principal-value integral, 47, 1270, 1381 principle correspondence, 15, 17, 31, 55, 56, 62, 66, 67 equivalence Einstein, 774, 775 new, 784 Maupertius, 380, 968 Pauli exclusion, 591 probability conservation law, 16, 1292, 1297, 1300, 1334, 1335, 1340 end-to-end distribution in polymers, 1019, 1020 exact , 1024 Gaussian approximation, 1029 moments, 1021 saddle point approximation, 1028 short-distance expansion, 1026 evolution bra-ket formalism, 1327 Heisenberg picture, 1326 path integral for , 1289 problem entanglement, xi, 1084, 1088 operator-ordering, xiv, 17, 55, 794, 1297, 1364 solved, 794, 804 topological, 1084, 1088 unitarity, 906 process PRAUSE,

Feller, 1523 Markov, 1462 measure, 1455 melting, 1369 non-Gaussian, 1454 Wiener, 1310, 1319 product normal of operators, 1363 scalar, 19 in space with torsion, 906 time-ordered of operators, 250, 1363 PROKOF' EV , N.V., 695 propagator, see also Green, 43 proper time Schwinger formula, 160 vertex functions, 301 proper time, 11, 1373, 1412 PROTTER, P ., 1528 pseudo-Hamiltonian, 928 pseudoenergy spectrum, 954 pseudotime, 439 action, 924- 927 Coulomb system, 932 amplitude, 924 Coulomb system, 932 evolution amplitude, 929, 984, 987 operator, 923, 926 Hamiltonian, 975 Schrodinger equation, 928 put option, 1498, 1500 quadratic completion, 211, 242, 256 fluctuations level-splitting, 1194 tunneling, xxv, 1171- 1173, 1181, 1194, 1201, 1212, 1244 quantization Bohr-Sommerfeld, 373, 375, 398, 450, 451 canonical, 39, 55- 57, 66 field , 677 first, 677 geometric, 88 group, 56, 59 of charge, 747 of magnetic flux, 1088, 1090 in superconductor, 1090 particle number bosons, 641 fermions, 654 second, 642, 643, 677

1566 semiclassical, 373, 398 stochastic, 1306, 1312 quantum -statistical action, 137 partition function, 77 path integral, 135, 143 with source, 237 Boltzmann factor , 1207, 1254 crystals, 566 electrodynamics (QED), 1210, 1369, 1420 equivalence principle, 798, 910 field theory, 591 lattice models, 155 relativistic, 591, 1368 fluctuation, xiii, 100, 103, 328, 369, 377, 463,491 Hall effect, 71, 641, 1146, 1161 fractional, xi, 1143, 1146, 1160 Langevin equation, 1308, 1336 mechanics, 1, 11 level shift due to tunneling, 1166 partition function, 77 with source, 209 noise, 1349 number principal, 757 radial, 756 in relativistic atom, 1400 statistics, 76 nonequilibrium, 1262, 1277, 1283, 1286 wire, 933 quantum field theory, 676, 679 QUESNE, C., 208 radial amplitude, 699, 704, 705, 713, 719, 720 oscillator, 985, 993 coordinates, 793 Coulomb, 985, 987 oscillator, 983, 987 principal quantum number, 757 path integral, 700 propagator free particle, 762 quantum number, 756 relativistic atom, 1400 wave functions free particle, 755 oscillator, 757, 758 particle in magnetic field, 769

Index radius Bohr, 421,467,486, 629, 954, 1344, 1400 critical bubble, 1247, 1250-1252 of convergence strong-coupling expansion, 1235 vanishing in perturbation series, 1209 Radon-Nikodym derivative, 1455 RAFELI, F., 1158 RAIBLE, S., 1526 RAJARAMAN, R., 455, 1259 RAMAN , C ., 693 RAMOS , R.O ., 695 RANDJBAR-DAEMI , S. , 1161 random chain, 1019 range, effective, 610 rapidity, 1409 RASHBA, E ., 569 rate decay, 1200, 1245 DMjUS$ exchange, 1439 growth of stock, 1430 riskfree interest, 1500, 1505 ratio gyromagnetic, 749, 1415 of fluctuation determinants, 117 RAUNDA , F ., 567 RAY, R . , 1161 RAYLEIGH, L. , 1081 Rayleigh-Ritz variational method, 466 Rayleigh-Schrodinger perturbation theory, xi, 276, 276, 280 scattering amplitude, 342 real-time Green function for T i- 0, 1262, 1265 REBONATO, R., 1523 reciprocal basis tetrads, 779 basis triads, 776 recursion relations Bender-Wu, 353 REED, J.F ., 567 reflection, Bragg, 12 REGGE, T., 1017 regular isotopy of knots, 1104, 1153, 1156 regularization, analytic, 159 regulating Bessel function, 977, 983, 985, 993, 997 function in path integral, 923, 925, 951, 977, 984 REIBOLD, R., 368 Reidemeister moves in knot theory, 1103, 1104, 1151 REINHART , P.-G., 893

1567 relation Calagareau-White, 1122, 1123, 1159 canonical anticommutation, 664 commutation, 40 completeness, 19, 21- 23, 27, 28, 31 , 46, 47,571, 764 basis Dads, 51, 776 Dirac, 21 Euler, 81 Ito, 1478 orthogonality, 19 basis Dads, 51, 776 orthonormality, 19 overcompleteness, 651 skein, 1111, 1150, 1153, 1155, 1156, 1156, 1160 uncertainty, 32 unitarity, 68 relativistic fields , 1278 particle, 1368 and stiff polymer, 1370 path integral, 1370 path integral Coulomb system, xi, 1397 reparametrization invariance, 1371 quantum field theories, 591 REMER, R. , 1524 RENNIE, A .J .O ., 1523 renormalization group, 1237 renormalized potential, 265 reparametrization invariance of configuration space, 804, 809, 811, 831 , 834, 841, 851 , 854 of relativistic path integral, 1371 under path-dependent time transformations,925 replica trick, 1074 REPPY, J.D., 695 representation adjoint, 746 matrices, 743 spectral, 46, 132, 759 nonequilibrium Green functions , 1265 spin matrices, 743 repulsive core in 3He potential, 1238 resistance, Hall, 1145, 1158 RESNICK, S., 1524 resolution of identity, 651 , 652 resolvent, 921, 923, 925, 974 operator, 45 retarded, 38

Green function , 214, 216, 226, 267, 1264, 1355 growth of perturbation expansion, 511 time evolution amplitude, 43 operator, 38, 43 return of financial asset, 1430 log, 1431 REYES SANCHEZ, R. , 567 REZENDE, J., 349 ribbon, 1117, 1118, 1120, 1159 circular, 1117 invariant, 1153, 1155, 1156 Riccati differential equation, 167, 370 Ricci tensor, 87 Riemann-Cartan, 781 RICE, T.M., 696 RICHTER, K ., 456 Riemann -Cartan connection, 778 curvature tensor, 779, 934 space, 773 -Lebesgue lemma, 73 connection, 86, 776 spinning top, 87 coordinates, 790 curvature tensor, 780 space, 722, 848 zeta function, 83, 163, 170 RIERA, R. , 1524 RINGWOOD, G.A. , xiv RISEBOROUGH , P., 368,1260 risk-neutral, 1507 martingale distribution, 1505 RISKEN, H ., 1365 riskfree interest rate, 1500, 1505 portfolio, 1504, 1513 RITSCHEL, D. , 567 ROBERTS, M.J., 456 Robertson-Walker metric, 1388 Robinson expansion, 172, 173, 600 ROBINSON , J.E., 172 rod limit of polymer, 1032 structure factor, 1033 ROSSLER, J ., 569 ROEPSTORFF, G ., 207 ROHRLICH , R., 1367 ROMA , A., 1522 RONCADELLI, M ., 208, 1367 ROSEN , N ., 1017

1568 Rosen-Morse potential, 994, 996, 996, 1183, 1195, 1213 general, 997, 998, 1000, 1002 ROSENFELDER, R., x, 193, 208, 342, 569 ROSENZWEIG, C. , 373 ROSKIEs , R., 1260 Ross , S.A., 1523 ROST, J .M ., 456 rotation, 56 symmetry, 697, 734 R-term in curved-space Schrodinger equation absence, 801 , 897, 906, 909, 938 Cheng, 916 DeWitt, 909 RUBIN, R.J. , 1082 RUDER, H ., 568 RUIJSENAARS, S., 1158 rule Feynman, 809, 832, 835 Ito, 1321, 1431, 1473 Ito-like, non-Gaussian, 1473 Jordan, 15 semiclassical quantization, 398 smearing, 463 Veltman, 16 1, 228, 814, 816, 818, 820, 831 Wick, 209, 249, 251, 251, 1220, 1322, 1363 RUNGE , K.J. , 695 Runge-Lenz-Pauli vector, 963 Rutherford formula, 442 scattering, 441, 442 Rydberg energy, 72 frequency, 955 RYDBERG, T.H ., 1528 RYZHIK , LM., 49, 109, 114, 116, 133, 146, 156, 163, 169, 171, 178, 206, 245, 246, 269, 375, 408, 411 , 633, 638, 660, 715, 729, 733, 754, 755, 757, 759, 762, 827,829,1021, 1033, 1040, 1045, 1052, 1211, 1258, 1446, 1447, 1489, 1524 RZEWUSKI, J ., 367, 693 S&P 500 index, 1429, 1436, 1487 SACKETT, C.A., 693 saddle point approximation, 377, 1206, 1241 for integrals, 376 expansion, 377, 390 SAITO, N. , 1082, 1365

Index SAITOH, M ., 569 SAKITA, B., 1161 SAKODA, S., 973 SALAM , A ., 1161 SALJE, E.K.H., 569 SALOMONSON , P ., 892 SALUNKAY, H ., 1524 SAMMELMAN, G .S. , 917 SAMORODNITSKY, G ., 1528 SAMUEL, J., 1082 SAMUELSON, P ., 1499 SANTA-CLARA , P ., 1528 SARKAR, S., 342 SATO, K ., 1528 SAUER, T ., xvi, 208, 1260, 1261 scalar curvature, 66, 87 Riemann-Cartan, 781 sphere, 800 product, 19 in space with torsion, 906 scale invariance, 1075 SCALETTAR, R., x scaling law for polymers, 1024, 1062, 1069, 1075, 1082 scattering amplitude, 190 eikonal approximation, 71 first correction to eikonal, 341 perturbation expansion, 340 Bragg, 1454 Coulomb, 71 length, 610 light, 1030 matrix, 190 Mott, 448 neutron, 1030 Rutherford, 442 SCHAKEL, A., 693- 695, 1160 SCHALM , K ., 892 SCHEIFELE, G ., 972 SCHERER, P ., 456 SCHIFF, L .L , 87, 372 SCHMID, A ., 1364, 1366 SCHMIDT, H.-J. , 1367 SCHMIDT, M .G ., 457, 1426 SCHMIDT, S., x, 695 SCHMITZ, R. , 1367 SCHNEIDER, C.K .E., 973, 1018 SCHOBEL, R., 1522 SCHOLES, M., 1499, 1527 SCHOUTEN, J .A ., 11, 778 SCHOUTENS, W. , 1524

1569 SCHRAMM, P., 1365 SCHREIBER, A .W., 569 SCHRIEFFER, J .R., 1161 SCHRODINGER, E ., 972 Schrodinger equation, 15, 16, 18, 25, 34, 35, 38-40, 43, 44, 51 , 53, 897, 909, 928, 952, 1262 Duru-Kleinert transformation, 982 in space with curvature and torsion, 894 integral kernel, 895 pseudotime, 928 time-independent, 16, 929 time-slicing corrections, 982 picture, 40, 41 in nonequilibrium theory, 1263 wave function , 16 SCHROER, B. , 591 SCHUBERT, C., x, 457, 893, 1426 SCHULKE, L ., 1365 SCHUTZ , M ., 1082 SCHULER, E.R., 1260 SCHULMAN, L.S., 207, 589, 692, 750, 1158 SCHULTE-FROHLINDE, V ., 161, 287, 568, 694, 750, 892, 1083, 1260, 1261 SCHULTZ, T.D. , 569 SCHWARTZ, E .L ., 456 SCHWARTZ, L., 88 Schwarz integrability condition, 7, 180, 639, 777, 778, 780, 848, 879, 1141 SCHWARZ, H .A., 7, 88 SCHWEBER, S.S., 367, 1158, 1426 SCHWEIZER, M., x, 1526 Schwinger -Keldysh formalism, 1277 proper-time formula, 160 SCHWINGER, J. , 456,1017,1162,1366,1415, 1427 SCULLY, M .O., 567 second quantization, 592, 642, 643, 677 bosons, 641 external source, 672, 673 fermions, 654 pair terms, 674 SEELEY, R.T., 893, 911 Seeley-DeWitt expansion, 851 Seifert surfaces, 1151 self -energy, 305 of electromagnetic field, 1406 -entangled polymer ring, 1158 -financing strategy, 1502

-interaction in field theory, 1074, 1077 in polymers, 1126 -intersections of polymers, 1084 SELYUGIN, O.V. , 568 SEMENOFF, G.W., 1162 semi-heavy tail, 1432, 1476 semiclassical approximation, 369, 1166, 1167 polymers, 1065 density of states, 409 differential cross section, 447 Mott scattering, 449 expansion, 376, 391 around eikonal, 371 Langevin equation, 1308 quantization rule, 373, 398 time evolution amplitude, 388 SEMIG, L. , xvi SENA, P. , 695 SENETA, E ., 1528 SENJANOVIC, P ., 1426 SERENE, J.W., 1366 series asymptotic, 273, 378, 505, 633, 701 Dyson, 35, 203 perturbation, 272 large-order, 1208 path integral with 8-function potential, 770 strong-coupling, 543, 1233- 1235 Taylor, 2 weak-coupling, 543 SERVUSS, R.M., 590 SEURIN, Y., 695 SEZNEC, R., 566 SHABANOV, S. , 1367 SHAH, S., 1163 SHAPERE, A., 693, 751, 1161, 1163 SHAVERDYAN, B.S ., 567 SHAW, S. , 1159 SHEPHARD, N., 1524 SHEPPARD, K. , 1526 SHERRINGTON, D. , 696 SHEVCHENKO, O.Y., 567 shift Lamb, 1337, 1342, 1345, 1420 operator for energy, 279 phase, 1175, 1177, 1181, 1184 SHILOV, G.E., 88 SHIRKOV, D .V ., 893 SHIRYAEV, A.N., 1525, 1527 SHlU, E.S .W. , 1526

1570 Shubnikov phase, 1090 Shubnikov phase, 1090 SIEGEL, C .L ., 87 SIEGEL, W., 1426 O'-model, nonlinear, 738, 748, 805, 1038 SILVA , A.C ., 1524 SILVER, R.N. , 1366 SILVERSTONE, H.J ., 567 SIMON, B. , 566, 568, 1160 simple knots, 1102, 1107, 1108 inequivalent, 1104 links, xxviii, 1119, 1121 SINGER, 1.M., 893, 911 SINGH , L.P ., 693 SINGH , V.A. , 750, 1158 singular potentials, 918 stable path integral, 921 SINHA, S., 1082 SIRCAR, K.R., 1522 SISSAKIAN, A.N. , 567 SIVIA, D.S., 1366 skein operations, 1111 relation, 1111, 1150, 1153, 1155, 1156, 1156,1160 SKENDERIS, K., 892 skewness of financial data, 1441 parameter, 1459 SKYRME, T.H.R., 751 sliding decay, 566, 1219 slip of phase in thin superconductor, 1242 small bipolaron, 544 small-stiffness expansion, 1041, 1042 smearing formula, 463 SMILANSKY, D ., 893 smile in financial data, 1507, 1516 Smoluchowsi equation, 1295 Smoluchowski equation, 1304, 1461, 1471 SMOLUCHOWSKI, M ., 1366 SMONDYREV, M.A., 568 smooth chaos, 405 Sochocki formula, 47 SOKMEN , 1., 1016 SOLDATI, R ., 1158 SOLOVTSOV, 1.L., 567 solution bounce, 1201 classical, 1174 almost, 1190 tunneling, 1167 critical bubble, 1201

Index negative-eigenvalue for decay, 1203, 1204 solvable path integral, 101, 111, 974 SOMMERFELD, A., 87, 1017 SOMORJAI, R .L ., 567 SOPER, D .E., 1261 SORNETTE, D. , 1528 source, 209, 210, 212, 232-237, 242, 247, 249 in imaginary-time evolution amplitude, 238 in quantum mechanics, 209 in quantum-statistical path integral, 237 in time evolution amplitude, 232 SOURIAU, J.M., 207 SOURLAS, N ., 1367 space -time curved, 10 Minkowski, 779 configuration, 98 extended time, 809 fiat, 775 Hilbert, 18 linear, 25 metric-affine, 773, 793, 901 Minkowski, 970 multiply connected, 1084, 1088 phase, 3, 98 reparametrization invariance, 804, 809, 811, 831, 834, 841, 851, 854 Riemann, 722, 848 Riemann-Cartan, 773 super, 1332, 1421 space with curvature and torsion, 773 mapping to, 789 path integral, 789 measure, 794, 799 time-sliced , 798 scalar product, 906 Schrodinger equation, 894 SPAGNOLO, B., 1526 spectral analysis, 131 density, 265 of bath, 263 function sum rule, 1272 representation, 46, 132, 759 amplitude of particle in magnetic field , 764, 766 dissipative part, 1271 fixed-energy amplitude free particle, 752 oscillator, 756 nonequilibrium Green functions, 1265

1571 of Green function , 217, 229 spectrum continuous, 47 Coulomb, 935, 961 bound-state, 935 continuum, 961 pseudoenergy, 954 sphere amplitude near surface, 722, 723, 729, 734, 736, 739,740 on surface, 734- 736, 738, 741, 999 curvature scalar, 800 Fermi, 419, 599, 1238 particle on surface, 57 surface in D dimensions, 79, 715 spherical -hyper harmonics, 716, 753 addition theorem, 718 components of vector, 1250 harmonic in one dimension, 579 in t hree dimensions, 712 harmonics, 59, 713, 718, 723, 988 addition theorem, 712 degeneracy in D dimensions, 716 monopole, 743, 1008 spin and torsion, 774 connection, 888 current density, 907 matrix representation, 743 Pauli matrices, 63, 748 precession, Heisenberg, 750 spinning particle amplitude, 743 path integral, 743 spinning top, 57, 59, 64- 66, 77, 85, 717, 734, 739, 741, 742 amplitude, 741, 742 curvature scalar, 87 path-integral, 741 Ricci tensor, 87 Riemann connection, 87 spontaneous emission, 1341, 1342, 1363 square knot , 1103, 1114 root trick, 498 anomalous, 518 width of local fluctuations, 462 SQUIRES, E .J ., 1017 SRIKANT, M. , 1527 SRIVASTAVA, S. , 565

stability matrix, 403 eigenvalue direct hyperbolic, 404 direct parabolic, 404 elliptic, 404 inverse hyperbolic, 404 inverse parabolic, 404 loxodromic, 404 stability parameter, 1459 stable path integral for singular potentials, 921 STAMAT ESCU , 1.0 ., 1367 STANCU, 1. , 567 standard cosmic time, 1386, 1392 form of Hamiltonian, 90 tetrads, 886 STANLEY , H.E. , 1082, 1523, 1524 stars, neutron, 774 states coherent, 350, 651 density, 82, 601 classical, 409 local, 409 local classical, 399, 401 local quantum-mechanical, 406 metastable, 1248 Schriidinger, 16 stationary, 16, 33 statistical mechanics, 76 lattice models, 518 statistics, 591 classical, 76, 1248 fractional, 640, 1096, 1097, 1100 interaction, 592, 635, 637, 639 for anyons, 640 for bosons, 635 for fermions, 637 gauge potential, 639 quantum, 76 statisto -electric field, 1140 potential, 1140 -electromagnetic vector potential, 1140 -magnetic field, 1140, 1142, 1143, 1146 forces, 1140 vector potential, 1124 steady-state universe, 1394 STEELE, J.M., 1525 S T EEN , F.H., 751

1572 STEGUN, I. , 50, 70, 174, 178, 244, 246, 410, 505, 715, 758, 761, 826, 1174, 1435, 1467 STEIN, E.M., 1523 STEIN, J .C., 1523 STEINBERGER, J. , 567 STEINER, F., 693, 750, 891 , 1017 STELLE, K.S., 893 STEPANOW, S., 1082 stereoisomer knots, 1103, 1108 STEVENSON, P.M., 567 STEWART, I., 1160 STIEFEL, E ., 972 stiff chain, 1024 polymer, 1032 Stirling formula, 505, 589, 1209, 1467 stochastic calculus, 189, 1315, 1319 difference equation, 1496 differential equation, 1307, 1308, 1346 Liouville equation Kubo, 1316, 1336, 1347 quantization, 1306, 1312 Schr6dinger equation, 1333 STOCK, V.S. , 695 STOCK MAYER, W.H., 1082 Stokes theorem, 782, 783, 877, 879 STONE, M., 751 , 1161 STOOF, H.T.C., 694 STORA, R. , 1260 STORCHAK, S.N., 1017 STORER, R .G ., 484 STORMER, H.L., 1161 straightest lines, 777 STRASSLER, M., 1426 strategy of portfolio manager, 1501 Stratonovich calculus, 1315 integral, 1351 STRATONOVICH, R ., 696 STRECLAS, A., 207 STREIT, L., 736, 1017 strike price of option, 1505 string Dirac, 641 , 881, 1090, 1093 super, 1369 theory, 1369, 1421 strip, Moebius, 1117 strong-coupling behavior, 512 expansion, xxii, xxvii, 512- 515, 518, 543, 566, 1233- 1235

Index coefficients, 512 limit, xi, 472, 473, 494 structure factor of polymer, 1030, 1033 Gaussian limit, 1031 rod limit, 1033 Student distribution in financial markets, 1447 Sturm-Liouville differential equation, 123 Su , Z.-B. , 1366 substitution, minimal, 180, 905, 1141 subtraction correlation function , 222, 225, 244, 264, 327, 333, 334, 336, 860 subtraction, minimal, 160 SUDARSHAN, E .C.G ., 589, 692 SUDB0, A. , 1426 summation by parts, 105, 111 convention, Einstein, 2, 4, 290, 309 formula, Poisson, 29, 156, 264, 573, 575, 580,582 super atom, 606 field, 1332 geometry, 1421 selection rule, 591 space, 1332, 1421 string, 1369, 1421 symmetry, 650, 1421 local, 1423 superaction, 1332 supercoil, 1117, 1118 density, 1118 superconductor, 1090, 1163, 1237, 1243 condensate, 1238 critical temperature, 1236 high-temperature, xi, 545, 1147, 1160 order parameter in, 1243 pair terms, 675 thin wire, 1235 type II, 1090 supercurrent, 1240, 1245 superfluid, 1238 helium, 591, 606, 607, 612 superheated water, 1248 supersymmetry, 1330 surface of sphere amplitude near, 722, 723, 729, 734, 736, 739, 740 amplitude on, 722, 734-736, 738, 741, 800,999 in D dimensions, 79, 715

1573 particle on, 57 Seifert, 1151 terms in partial integration, 2 susceptibility, magnetic, 1036, 1041 SUZUKI, H., 566, 568 SUZUKI, M., 208 SVIDZINSKIJ, A.V., 696 SVISTUNOV, B .V ., 695 symbol Christoffel, 11, 86 Levi-Civita, 783 symmetry energy-momentum tensor, 781 rotations, 697 translations, 1174, 1197, 1201, 1206, 1218 symplectic coordinate transformations, 7 unit matrix, 7 T matrix, 74, 191, 192, 344, 610 TABOR, M., 405 tadpole diagrams, 498, 499 tail heavy, 1433 Levy, 1433 Pareto, 1433 power, 1433 semi-heavy, 1432, 1476 tails Gaussian, 1432 Tait number, 1109 TAIT, P.G., 1109, 1163 TAKAHASHI, K ., 1082 TALKNER, P., 1260 TANGUI, C., x TANNER, G. , 456 TAQQU, M. , 1528 TARRACH, R. , 567, 1158 TATARU, L., 893 Taylor expansion, 2, 99 covariant, 791 TAYLOR, B .N. , 1260 TEITELBOIM , C. , 693 TELLER, E., 751, 1017 temperature critical Bose-Einstein, 596, 602 superconductor, 1236 superfluid helium, 606 Debye,1236 Fermi, 632 TEMPERE, J. , 695

TEMPLETON, S., 1160 TENNEY, M., x tensor contortion, 779 curvature of disclination, 783 Riemann-Cartan, 934 Einstein, 781 energy-momentum, 1391 Levi-Civita, 783 metric, 52 of contractions in Wick expansion, 708, 942, 1022, 1041 Ricci, 87 Riemann-Cartan, 781 Riemann curvature, 780 Riemann-Cartan curvature, 779 symmetric energy-momentum, 781 torsion, 778 of dislocation, 781 test function, 25, 45, 710 tetrads basis, 779 multivalued, 780 reciprocal, 779 standard, 886 TEUKOLSKY, S.A. , 1366 THEIS, W., xvi theorem Bloch, 635 central limit, 1457, 1464, 1467, 1482, 1492, 1512 generalized, 1459 equipartition, 327, 462 Levinson, 1182 Levy-Ito, 1460 Nambu-Goldstone, 311, 324, 325 Pawula, 1472 Stokes, 782, 783, 877, 879 virial, 427, 440 theory Chern-Simons, 1137 nonabelian, 1149, 1155 Flory, of polymers, 1069 growth parameters of large-order perturbation coefficients, 1214 linear response, 141, 1262, 1262, 1264, 1277 Maxwell, 1124 mean-field, 306, 312 perturbation, 272, 1277 large-order, 1209, 1211, 1212, 1214, 1218

1574 quantum field, 676, 679 Schwinger-Keldysh, 1277 string, 1369, 1421 thermal de Broglie wavelength, 139, 595 driven decay, 1255 equilibrium, 249 fluctuations, 100, 249, 328, 463, 491, 1194 length scale, 139, 595, 602 wavelength, 595, 602 thermodynamic limit, 288 relation, Euler, 81 theta function elliptic, 607, 687 Theta of option, 1503, 1504 THISTLETHWAITE, M .B., 1160 THOMA, M.H. , 567 Thomas -Fermi approximation, 419, 440 atom, 432 density of states, 419 differential equation, 426 energy, 429, 431, 432, 435 energy density, 420 model of neutral atoms, 419 precession, 1145 THOMAS, H ., 1365 THOMCHICK, J. , 569 THOMSON, J.J ., 1017 'T HOOFT, G ., 159, 809, 892, 1365, 1367 three-point function , 302 tilt angle, 955, 957 operator, 567, 955 transformation, 957 time -dependent density matrix, 1333 mass, 871 -independent Schriidinger equation, 929 -ordered Green function, 1269 operator product, 1363 product , 250 -ordering in forward- backward path integral, 1283 operator, 35, 37, 229 -slicing corrections, 976

Index from Schriidinger equation, 982 general, 977 cosmic standard, 1386, 1392 extended space, 809 orthogonality, 1387 proper, 1373, 1412 series, 1496 series of financial data, 1430 slicing any point, 793 correction, 979, 981, 992, 994 transformation path-dependent, 925 path-dependent (DK), 982, 984, 985, 993, 995, 998, 1000, 1001, 1012 time evolution amplitude, 43, 46, 89, 94, 100, 235, 752, 928, 929, 974, 1262 causal, 43 composition law, 90, 700 fixed path average, 237 free particle, 101, 109, 110 freely falling particle, 177 oscillator, 111 particle in magnetic field, 179, 181, 183 perturbative in curved space, 846 retarded, 43 semiclassical, 388 with external source, 232 Euclidean amplitude spectral decomposition, 759 operator, 34, 35, 37- 40, 43, 77, 89, 90, 94, 250 anticausal, 38 causal,43 composition law, 37, 38, 72 interaction picture, 42, 1279 modified, 921 retarded, 38, 43 time-sliced action, 91 curvilinear coordinates, 773 amplitude, 89 configuration space, 97 in curvilinear coordinates, 773 momentum space, 94 phase space, 91 Feynman path integral, 89 divergence, 918 measure of functional integral, 101 path integral coordinate invariance, 800

1575 in space with curvature and torsion, 798 TINKHAM, M ., 1260, 1261 TODA, M. , 1365, 1367 TOGNETTI, V ., 565, 566 TOLLET, J.J. , 693 TOMASIK, B. , 695 TOMBOULIS, E., 367 TONINELLI, F., 893 top, spinning, 57, 59, 64-66, 77, 85, 717, 734, 739, 741, 742 amplitude, 741 , 742 asymmetric, 85 curvature scalar, 87 Ricci tensor, 87 Riemann connection, 87 topoisomerase, 1118 topological constraint, 571, 1085 interaction, 637, 639, 1089, 1137 invariant, 1085, 1089, 1116-1118, 11201123, 1126, 1137, 1152, 1159 moment, 1127 problems, 1084, 1088 topology algebraic, 1103 classes of knots, 1101 torsion and curvature, space with, 773 and spin density, 774 gradient, 963 in Coulomb system, 934 in transformed H-atom, 931 of curve, 1122 tensor, 778 of dislocation, 781 towering property, 1463 TOYODA, T., 694 TRACAS, N.D., 892 trace formula Gutzwiller, 404 tracelog, 83, 411 gradient expansion, 167 transfer of momentum, 70 transformation Bogoliubov, 675 canonical, 6, 8, 9 conformal, 456 Weyl, 962, 963 coordinate, 975, 977, 978, 982, 984 local, 884 duality, 168, 169

Duru-Kleinert, 925, 931, 974, 978, 982985, 993, 995, 998, 1000, 1001, 1012 D = 1, 974 fixed-energy amplitude, 983 of radial Coulomb action, 986 of radial oscillator, 987 of Schriidinger equation, 982 Esscher, 1480, 1480 Foldy-Wouthuysen, 1409 Fourier, 752 gauge, 185, 1124 nonholonomic, 778 Hubbard-Stratonovich, 682, 1063, 1072, 1078 Kustaanheimo-Stiefel, 953 Laplace, 752 Levi-Civita, 933, 935 local U(l) , 884 Lorentz local, 887 of coordinates, 776 of measure in path integral, 979, 991, 993,994 path-dependent time (DK) , 982, 984, 985, 993, 995, 998, 1000, 1001, 1012 Poincare, 887 point, 5 tilt, 957 translation, 56 fluctuation, 386 symmetry, 1174, 1197, 1201, 1206, 1218 transversal fluctuation width, 525 gauge, 1125 projection matrix, 310,479 trial frequency, 479 trap, magnetic for Bose-Einstein condensation,615 anisotropic, 620 tree approximation, 305 diagrams, 304, 307, 310, 315 trefoil knot , 1101, 1101 TREIMAN , S.B., 1163 TRELOAR, L.R.G. , 1081 TREMBLAY, A .M ., 1364 triads basis, 776 multi valued, 778, 780 reciprocal, 776 trial frequency longitudinal, 479

1576 transversal, 479 partition function, 459 trick anomalous square-root, 518 Faddeev-Popov, 192, 858, 1150, 1178 replica, 1074 square-root, 498 trigonometric addition theorem, 726 'Trotter formula, 93, 93, 208 TROTTER, E ., 208 TRUGENBURGER, C. , 1161 truncated Levy distribution, 1434, 1437 asymmetric, 1439 cumulants, 1437 truncated Levy distribution, 1439, 1450, 1451, 1455, 1457, 1511, 1512, 1515, 1520 Tsallis distribution in financial markets, 1447 TSALLIS, C ., 1525 TSEYTLIN , A .A ., 893 TSUI, D.C ., 1161 TSUSAKA, K ., 1367 tube, flux , 1090 tunneling, 1164, 1166, 1212 and decay, 1199, 1200, 1213, 1242, 12451247, 1254 of supercurrent, 1235 quadratic fluctuations, xxv, 1171-1173, 1181, 1201, 1212, 1244 rate formula, 1207 variational approach, xi, 1219 turning points, 129 TURSKI, L.A. , 1367 twist, 1122 number, 1109, 1151 two-point function , 251 , 292 connected, 302 type II superconductor, 1090 TZE, C.H. , 1160, 1427 U(l) local transformations, 884 UNAL, N., 1260 UEHLING , E.A. , 1427 ULLMAN, R. , 1082 ultra -local functional, 98 -spherical harmonics, 716 ultraviolet (UV) cutoff, 805 divergence, 159, 805 uncertainty principle, 14

Index relation, 32 underpass in knot graph, 1104, 1106, 1120 unit matrix, symplectic, 7 unitarity, 38 problem, 906 relation, 68 units atomic, 482, 954, 1238 electromagnetic, 71 natural, 470 universality of gravitational forces, 774 universe expanding, 1386, 1394 Friedmann, 1392 homogeneous, 1386 isotropic, 1386 lifetime, 1391 steady-state, 1394 USHERVERIDZE, A .G. , 567 utility function of financial asset, 1481 vacuum, 596 diagrams, 284 correlation functions , 298 generating functional , 294 one-particle irreducible, 322 false, 1254, 1254 instability, 1210, 1255 VAIA , R. , 565, 566 VAINSHTEIN , A.I. , 1259 VALATIN, J.G., 695 VALENTI, C.F. , 751 VALENTI, D ., 1526 VAN DEN BOSSCHE, B ., 368, 390 VAN DOREN , V., 1364 VAN KAMPEN , N .G. , 1365 VAN NIEUWENHUIZEN , P ., 892 VAN ROYEN, J ., 569 VAN VLECK, J.H., 388 VAN VUGT, M. , X VAN WINTER, C. , 88 VAN DRUTEN, N .J ., 693 Van Vleck-Pauli-Morette determinant, 388,390,909,916,917 variable anticommuting, 654, 673 collective, 681 complex Grassmann integration over, 656 cyclic, 571, 574 Grassmann, 654, 693 integration over, 654, 655 variance of financial data, 1430, 1450

1577 variation auxiliary nonholonomic, 785 covariant, 873 in action principle, 2, 784 nonholonomic, 784 variational approach, 458, 475 to tunneling, xi, 1219 energy, xxvii Rayleigh-Ritz method, 466 interpolation, 518 perturbation theory, xi, 458, 496, 496 convergence proof, 1233 optimization, 464, 486, 488, 491-493, 500, 516, 544 VASSILIEV, A .N., 367 VAUTHERIN, D ., 694, 695 vector Burgers, 782 potential, 802 in Fokker-Planck equation, 1304 statisto-electromagnetic, 1140 statisto-magnetic, 1124 time-sliced action, 802 spherical components, 1250 Vega, 1503 Vega of option, 1503, 1504 velocity desired, 189 light, 13 path integral, 189, 192 Veltman rule, 161, 228, 814, 816, 818, 820, 831 VELTMAN , M. , 159,809,892 VERLINDE, E ., 1427 VERLINDE, H ., 1427 VERNON, F.L., 1366 vertex functions, 301 generating functional for, 300 one-particle irreducible (lPI) , 300 proper, 301 vertices, 283 VETTERLING , W.T ., 1366 VICENTE, R., 1524 VIDBERG , H .J ., 1366 vierbein fields , 783 , 886 VILENKIN, A ., 1427 VILENKIN, N.H., 716 VILKOVISKI, G.A., 893 VINETTE, F ., 514, 567 virial coefficient, 1100, 1100

expansion, 1099 theorem, 427, 440 VITIELLO, G ., 1365 VLACHOS, N.D., 892 VOGELS, J .M ., 693 VOIGT, J., 1523 volatility implied, 1507 of financial data, 1428, 1430, 1432, 1481, 1482, 1507, 1508, 1513 risk, 1513 VOLOGODSKII , A .V., 1159, 1160 VOLOSHIN, M.B., 1259 VON KLITZING, K., 1161 vortex lines, 606 vortex field, 1369 VOTH, G .A., 566 VRSCAY, E ., 567 Vycor glass, 612 WALECKA, J .D ., 693, 1366 wall of critical bubble, 1253 WALLACE, S.J., 342 WANG , J .C. , 1159 WANG, M.C ., 1081 WANG, P ., 1018 WANG , P .S., 456 Ward-Takakashi identity, 325 WASSERMAN, E ., 1159 watermelon diagram, 284, 812, 816, 823, 831 , 841,850 WATSON, G.N., 728, 732, 958 WATSON, K.M., 279, 373 wave frequency, 12 function, 12, 46, 47, 133, 752, 767, 768 charged particle in magnetic field, 763, 766 Coulomb, 467, 937, 954 free particle, 131 free particle from w --t 0 -oscillator, 761 momentum space, 28 node, 1166 oscillator, 131 particle in magnetic field , 767 radial, free particle, 755 oscillator, 757, 758 particle in magnetic field, 769 Schr6dinger, 16 Wentzel-Kramers-Brillouin (WKB), 373

1578 material, 11 packet, 14 plane, 13 wavelength classical of oscillator, 140 Compton, 421, 1374, 1375, 1377, 1400, 1402 de Broglie, 370 oscillator, 536 quantum, 133 thermal, 139, 595, 602 WAXMAN , D. , 368 weak -coupling expansion, 272, 518, 543 -field expansion, 484, 488 Wegner exponent, 518 WEGNER, F .J ., 569 Weibull distribution, 1449 W EIERSTRASS , K ., 88 weight, Levy, 1460 WEINBERG , S., 11 WEISS, D., 368, 1260, 1364 WEISSTEIN, E.W. , 207, 693 WEIZEL, W ., 87 WELTON , T.A. , 1364 WENIGER, E.J. , 567 WENTZEL, G ., 455 Wentzel-Kramers-Brillouin (WKB) approximation, 369, 372, 374, 396, 1225, 1259 condition, 370, 373 connection rules, 372 equations, 371 wave function, 372, 373 W ESS, J ., 1160 Wess-Zumino action, 747 Weyl covariance, 963 order of operators, 794 WEYL, H ., 794 WHEELER, J.A ., 456 white dwarfs, 484 noise, 1309, 1319, 1349, 1352, 1355, 1430 WHITE, A ., 1522 WHITE, J.H., 1159 WHITENTON, J., xvi Whittaker functions , 755, 757, 766, 961, 1435 WHITTAKER, E.T., 206 Wick expansion, 209, 249, 251, 251, 1220, 1322, 1363 width

Index fluctuation local, 462 longitudinal, 525 transversal, 525 line, natural, 1337, 1341 WIEGEL, F.W ., 207, 1114, 1159 WIEMAN , C.E. , 693 Wiener process, 1310, 1319 drift , 1310 WIENER, N. , 206, 1527 Wigner function, 33, 1333 Liouville equation, 33 Weisskopf natural line width, 1337, 1342 WIGNER, E .P ., 279, 567 WILCZEK , F ., 456, 589, 693, 751, 893, 1160, 1161, 1163 WILHELM , J ., 1082 WILKENS, M ., 695 WILLET, R ., 1161 WILLIAMS, D. , 750 WILMOTT, P., 1523 Wilson loop integral, 1150 WILSON, R. , 973, 1018 winding number, 594, 1085 WINDWER, S., 1115, 1159 WINTGEN, D. , 456 wire, superconducting, 1235 WITTEN, E ., 751, 893, 1160, 1161, 1163 WOLOVSKY, R ., 1159 WOODHO USE, N .M .J. , 88 WOODS, A .D ., 614 would-be 8-function, 710 zero eigenvalue, 1185 writhe, 1109, 1151, 1153 writhing number, 1122, 1122, 1163 Wronski construction of Green function Dirichlet case, 213 periodic and antiperiodic, 231 determinant, 122, 124, 214, 345 Wu , T.T., 353, 566, 1261 Wu, Y.S. , 693, 1367 WUlXIAOGUANG, 569 WUNDERLIN, A., 1017 WUNNER, G ., 568 X-polynomial of knots, 1109, 1109 XIAO-JIANG , T ., 1524 YAGLOM, A.M., 120, 206 YAKOVENKO, V.M., 1481, 1524, 1526

1579 YAMAKAWA ,

H ., 1082

YAMANAKA, Y., 1365 YAMAZAKI , K ., 567 YANG, C.N. , YETTER, D. ,

694 1160 YOR, M ., 750, 1525, 1527 Yu , L. , 1366 YUKALOV , V.I., 566, 625 Yukawa potential, 481 YUNOKI , Y ., 1082 ZAANEN , J. , ZACHOS ,

695

C .K ., 893

Zassenhaus formula, 202 ZASSENHAUS , G .M., 695 ZAUN , J. , xvi, 1017 ZEE, A., 456, 693, 893, 1161 ZEH , H .D. , 1367 zero-Hamiltonian path integral, 91 zero-modes, 217, 218 of kink fluctuations, 1174, 1177, 1180, 1183, 1187, 1201, 1202, 1245 would-be, of kink fluctuations, 1185, 1187 zero-point energy, 146, 332, 675, 1219 zeta function generalized, 83 Hurwitz, 599 operator, 83 Riemann, 83, 163, 170 ZHANG , B ., 567 ZHU , J. , 1522 ZINN-JUSTIN , J., 566, 568, 694, 1260, 1261, 1367 zone scheme, extended, 575 , 593, 1012 ZUBER, J .-B. , 287, 1260, 1426 ZUMINO , B., 1160, 1163, 1426 ZWERGER, W. , 368, 1260

1580

Index

This page intentionally left blank