You are on page 1of 664

GEOMETRY, PARTICLES AND FIELDS

Based upon lectures given by BJ0RN FELSAGER Odense University, Mathematics Department and The Niels Bohr Institute, Copenhagen

Edited with the help of CARSTEN CLAUSSEN Odense University, P"'ysi.cs Department
.~,~~~

ODENSE UNIVERSITY PRESS 1981 2. edition 1983

PREFACE
The present book is an attempt to present modern field theory in an elementary way. It is written mainly for students and for this reason it presupposes little knowledge in advance except for a standard course in calculus (on the level of multiple integrals) and a standard course in classical physics (including classical mechanics, special relativity and electrodynamics). The main emphasize is laid on the presentation of the central concepts, not on mathematical rigour. Hopefully this textbook will prove useful to high-energy physicists, who want to get acquainted with the basic concepts of difderential geometry. Mathematicians may also have fun reading about the applications of central concepts from differential geometry in theoretical physics. To set the stage I have in the first part included a self-contained introduction to field theory leading up to recent important concepts like solitons and instantons. There are two main themes in part one: One the one hand I discuss the structure of a gauge theory, exemplified by ordinary electromagnetism. This includes a derivation of the Bohm-Aharonov effect and the flux quantization of magnetic vortices in a superconductor. One the other hand I discuss the structure of a non-linear field theory, exemplified by the ~~-model and the sine-Gordon model in (l+l)-dimensional space-time. This includes the construction of a topological charge, the particle interpretation of the kink-solution,and finally the relevance of the kink-solution for the tunnel effect in quantum mechanics is pointed out. Although the present text deals mainly with the classical aspects of field theory I have also touched the quantum mechanical aspects using pathintegral techniques. In part two I have included a self-contained introduction to differential geometry. The main emphasize is laid on the so-called exterior calculus of differential forms, which on the one hand permits the construction of various differential operators - the exterior derivative, the co-differential and the Laplacian - on the other hand the construction of a covariant integral. But I also investigate metrics and various related concepts, especially Christoffel fields, geodesics and conformal mappings.

Apart from the introduction to the basic concepts in differential geometry the second part contains a number of illustrative applications. The Lagrangian formalism is put on covari~nt (i.e. geometrical) form. A detailed discussion of magnetic monopoles, including the Lagrangian formalism for monopoles and the quantization of magnetic charges, continues the investigation of gauge theories initiated in part one. Further examples of non-linear models are presented: The Heisenberg ferromagnet, the exceptional ~~-model and the abelian Higgs' model (including a discussion of the Nambu strings and their relationship with the Nielsen-Olesen vortices). Finally symmetry transformations and their associated conservation laws (i.e. Noether's theorem) are investigated in great detail.

ACKNOWLEDGEMENTS
A project like this would never have been completed were it not for the moral and financial support of a great number of persons and institutions. From mathematics department, Odense University, I would especially like to thank my scientific advisors Erik Kjrer Pedersen and Hans J~r gen Munkholm, who have followed the project through all its various stages. I am also grateful to Ole Hjort Rasmussen for many stimulating discussions about geometry. From the Niels Bohr Institute I would like to thank my scientific advisor Paul Olesen (who originally suggested me to take a closer look at the geometrical and topological structure of gauge theories and who encouraged me to give the lectures upon which the book, is based). I am also grateful for moral support from Torben Huus. Helge Kastrup Olesen look~d over a priliminary version of the manuscript and taught me a lot about english grammar. Carsten Claussen has been of invaluable help to me. He has been reading the whole manuscript in several versions and has suggested innumerable improvements. I would also like to thank the secretaries, Lisbeth Larsen at Odense University and Vera Rothenberg at the Niels Bohr Institute, who, with great patience and professional skill, typed major parts of the manuscript. Finally I would like to thank Odense Universitets publikationskant a for financial support to the printing of the manuscript, and L~ rup and Holck's fonde (at the Niels Bohr Institute) for a generous donation to typing assistance.

CONTENTS

PART I:
Chapter 1:
1.1

BASIC PROPERTIES OF PARTICLES AND FIELDS


ELECTR<l14GNETISM
THE ELECTROMAGNEI'IC FIELD The electromagnetic fieldstrengths. Vector calculus. Maxwell's equations. Charge conservation. THE INTRODUCTION OF GAUGE POTENTIALS The scalar potential and the vector potential. Gauge transformations. Lorenz gauge. MAGNETIC FWX Magnetic flux expressed through a line integral of the vector potential. Magnetic monopoles. Non-existence of a global vector potential generating a monopole field. ILLUSTRATIVE EXAMPLE: THE GAUGE POTENTIAL OF A SOLENOID The vector potential outside a magnetic string. Singular gauge transformations. Dirac strings. The a-fUnction. RELATIVISTIC FORMULATION OF THE THEORY OF ELECTROMAGNEI'ISM Relativistic kinematics. Lorentz invariant equations. The Maxwell field. THE ENERGY-MOMENTUM TENSOR Currents associated with point particles. The energymomentum tenscrassociated with point particles. The energy-momentum tensor associated with the electromagnetic field. Abrahams theorem. Angular momentum. Symmetry of the energy-momentum tensor. SOLUTIONS OF WORKED EXERCISES
3

1.2

1 3

11

1 .4

15

1.5

19

1.6

22

28

Chapter 2: 2.1

INTERACTION OF FIELDS AND PARTICLES


INTRODUCTION Non-trivial gauge potentials and scattering effects outside a solenoid. LAGRANGIAN FORMALISM FOR PARTICLES The Lagrangian. The action principle. The Euler-Lagrange equations. Generalized potentials. The Lagrangian of a charged particle in an electromagnetic field. Gauge invariance of the interaction term. BASIC PRINCIPLES OF QUANTUM MECHANICS Probability amplitudes. The Schrodinger wave fUnction. The Schrodinger equation on integral form. PATH INTEGRALS - THE FEYNMAN PROPAGATOR The Feynman propagator. Path integrals. Feynmans principle of the democratic equality of all histories. The correspondence principle. The Feynman propagator of a quadratic Lagrangian.
31

2.2

32

2.3

39

2.4

42

x
2.5

ILLUSTRATIVE EXAMPLE: THE FREE PARTICLE PROPAGATOR The free particle propagator. Einsteim- de Broglies rule. The wave fUnction of a free particle with specific energy and momentum. BOHM-AHARONOV EFFECT - LORENTZ FORCE The Lorentz force in quantum mechanics. Slit experiments. The Bohm-Aharonov effect. Recovering of the Lorentz force in the classical limit. GAUGE TRANSFORMATION OF THE SCHRODINGER WAVE FUNCTION Gauge transformation of the gauge phase factor, the propagator and the Schrodinger wave fUnction. Gauge invariance in quantum mechanics. QUANTUM MECHANICS OF A CHARGED PARTICLE AS A GAUGE THEORY The electromagnetic gauge group. Gauge vectors, gauge scalars and gauge potentials. The gauge covariant derivative. THE SCHRODINGER EQUATION IN THE PATH INTEGRAL FORMALISM The infinitisemal propagator. Derivation of the Schrodinger equation in the path integral formalism. THE HAMILTONIAN FORMALISM Conjugate momenta. Canonical systems. The Hamiltonian. Hamilton's equations. Identification of the Hamiltonian with the energy. Poisson brackets. CANONICAL QUANTIZATION AND THE SCHRODINGER EQUATION Construction of the Schrodinger equation in the canonical formalism. The Schrodinger equation of a charged particle in an electromagnetic field. The rule of minimal coupling. Gauge covariance of the Schrodinger equation. The Feynman propagator in the canonical formalism. Poisson brackets and commutators. Heisenberg's commutation relations. ILLUSTRATIVE EXAMPLE: SUPERCONDUCTORS AND FLUX The Meisner effect. Superconductivity. Cooper pairs. The super current. Flux quantization. Abrikosov vortex lines. Interaction of charged particles with non-trivial gauge potentials. SOLUTIONS of WORKED EXERCISES

46

2.6

49

2.7

55

2.8 2.9

57

62
65

2.10

2.11

68

2.12

72

79

Chapter 3: 3.1

DYNAMICS OF CLASSICAL FIELDS


ILLUSTRATIVE EXAMPLE: LAGRANGIAN FORMALISM FOR A STRING 83 A discrete model of a string. The continuum limit. The Lagrangian density. Equations of motion. The dispersion relation. LAGRANGIAN FORMALISM FO~ RELATIVISTIC FIELDS The Lagrangian density.' The action principle. The Euler-Lagrange equations. The Hamiltonian density. The canonical energy-momentum tensor. The true energy-momentum tensor. Free field theories. HAMILTONIAN FORMALISM FOR RELATIVISTIC FIELDS Conjugate momenta. Canonical formalism. Hamilton's equations. Functional differentiation. Poisson brackets. Quantization procedures. THE KLEIN-GORDON FIELD The Klein-Gordon equation. The Lagrangian density. The Yukawa potential.

3.2

86

3.3

90

3.4

95

XI

3.5

THE MAXWELL FIELD Maxwell's equations. The Lagrangian density. The energymomentum tensor. Complex wave solutions. Photons. The polarization vector. The degrees of freedom of a photon. SPIN OF THE PHOTON - POLARIZATION OF ELECTROMAGNETIC WAVES Spin. Helicity. Linearly and circulary polarized light. Quantum mechanical description of polarized light. THE MASSIVE VECTOR FIELD The Lagrangian density. Equations of motion. The true energy-momentum tensor. THE CAUCHY PROBLEM The Cauchy problem. The Klein-Gordon field. The Maxwell field. Constraints and gauge conditions. The massive vector field. THE COMPLEX KLEIN-GORDON FIELD Complex fields. Global symmetries. Currents and charge conservation. THE THEORY OF ELECTRICALLY CHARGED FIELDS AS A GAUGE THEORY Electrically charged fields. The rule of minimal coupling. Gauge transformations of the first and the second kind. The electric current. CHARGE CONSERVATION AS A CONSEQUENCE OF GAUGE SYMMETRY Redifinition of the current. Charge conservation. THE EQUIVALENCE OF REAL AND COMPLEX FIELD THEORIES Real fields versus complex fields. Gauge invariance in the real formulation. SOLUTIONS OF WORKED EXERCISES

98

3.6

103

3.7

107

3.8

109

3.9

113

3.10

116

3.11
3.12

119
122

124

Chapter 4:

SOLITONS
NON-LINEAR FIELD THEORIES WITH A DEGENERATE VACUUM The ~~-model. The sine-Gordon model. The mechanical analogue of the sine-Gordon model. Discrete symmetries. Boundary conditions for finite energy configurations. The classical vacua. Fluctuations around a classical vacuum. The mass-formula for a field quantum. TOPOLOGICAL CHARGES The different sectors corresponding to the different possibilities for the asymptotic behaviour. Vacuum sectors. Non-pertubative sectors. Topological charge. Topological conservation laws and topological currents. The topological charge as a winding numher. SOLITARY WAVES Plane waves, travelling waves and solitary waves. The kink and the anti-kink. The energy of the kink. GROUNDS TATES FOR THE NON-PERTUBATIVE SECTORS Ground states versus static solutions. The Bogomolny decomposition. The energy bound for topological non-trivial sectors. The ground state equations. Approximative ground states. Quasi static solutions in a non-linear field theory. SOLITONS The kink as a particle: The soliton. Energy and momentum of a soliton. Exact multi-soliton solutions in the sine-Gordon model: Soliton-soliton scattering. Soliton-ant i-soliton scattering. The breather as a bound state of a soliton and an anti-soliton.

4.1

127

4.2

131

4.3

135

4.4

138

4.5

144

XII

4.6

THE BACKLUND TRANSFORMATION Light-cone coordinates. The Backlund transformation in the sine-Gordon model. Bianchi's permutability theorem. Construction of the two-soliton solution. The most general multi-soliton solution. DYNAMICAL STABILITY OF SOLITONS Topological stability versus dynamical stability. The linearized equation of fluctuations. The eigenvalue problem associated with a stability investigation. The spectrum for the Schrodinger operator. The zero-mode in a non-linear field theory. Derrick's scaling argument. The virial theorem for a static finite energy solution in a (1+1l-dimensional field theory. THE PARTICLE SPECTRUM IN A NON-LINEAR FIELD THEORY The field quantum and the soliton. Composite particles. Bound states of solitons and anti-solitons. The eigenvalue problem associated with the weak-field interaction of solitons and field quanta. The spectrum of the Eckhardt potential. Bound states in the ~-model. SOLUTIONS OF WORKED EXERCISES

4.7

153

4.8

160

164

Chapter 5:

5.1

PATH-INTEGRALS AND INSTANTONS THE FEYNMAN PROPAGATOR REVISITED The Feynman propagator as a path integral. The Feynman propagator as a Green's function for the Schrodinger operator. The position operator and the momentum operator. The Feynman propagator as a matrix-element of the time evolution operator. The energy levels and the trace of the Feynman propagator. The Feynman propagator in momentum space.
ILLUSTRATIVE EXAMPLE: THE HARMONIC OSCILLATOR The classical action for a harmonic oscillator. Caustics. The propagator for a harmonic ascillator. Correspondence with the free particle propagator. The energy levels for the harmonic oscillator. The ground state for the harmonic oscillator. The eigenfunctions, Hermite polynomials and Mehlers'formula. The Feynman-Souriau formula for the propagator of a harmonic oscillator. THE PATH-INTEGRAL REVISITED The path integral as the limit of multiple integrals: Feynman's formula. The p~th integral as a limit of a summation over approximative paths. Fourier paths. The propagator for the harmonic oscillator revisited. ILLUSTRATIVE EXAMPLE: THE TIME-DEPENDENT OSCILLATOR The action for a time-dependent oscillator. Reduction of the path-integral to the action of a free particle. The Jacobiant. The propagator for a time-dependet oscillator. PATH-INTEGRALS AND DETERMINANTS Quadratic path-integrals. The associated eigenValue problem. The path-integral in terms of determinants. The determinantal relation.
172

5.2

5.3

181

5.4

188

194

XIII
5.6 THE BOHR-SOMMERFELD QUANTIZA'l'ION RULE Weak-coupling approximation of the path-cum-trace integral. WKB-approximation. The Bohr-Sommerfeld quantization rule. Analytical mechanics for periodic orbits. Derivation of the Bohr-Sommerfeld quantization rule through the correspondence principle. The harmonic oscillator. Path-integral derivation of the Bohr-Sommerfeld quantization rule. The stationary phase approximation. WKB-quantization of a free particle. WKB-quantization of the breather in the sine-Gordon model. INSTANTONS AND EUCLIDEAN FIELD THEORY Wick rotation. The Euclidean propagator. The Euclidean action. Euclidean path integrals. The Euclidean propagator as a Green's function for the Heat operator. Euclidean finite action configurations. The Euclidean vacuum. Instantons. The correspondence between solitons and instantons. Instantons in quantum-mechanics. Pseudo-particles. INSTANTONS AND THE TUNNEL EFFECT The tunnel effect. The quantum mechanical ground state for a double well. The energy levels in the double square well. The instanton solution as a Euclidean zero-energy solution interpolating between different Euclidean vacua. The Absence of instantons and tunnel effects in non-linear (1+1)scalar field theories. INSTANTON CALCULATION OF THE LOW-LYING ENERGY LEVELS Calculation of the ground state for a single well. Calculation of the Euclidean propagator for a double well. The dilute gas approximation. Calculation of the Euclidean propagator corresponding to a periodic potential. The low lying energy states as a Bloch wave. ILLUSTRATIVE EXAMPLE: CALCULATION OF THE PARAMETER ~ A disaster in the naive approach: The zero mode. Removal of the zero mode. Calculation of the determinants involved. Calculation of the lowest eigen value. The energy split in a double well. SOLUTIONS OF WORKED EXERCISES 198

5.7

212

5.8

219

5.9

225

5.10

236

245

PART II:
Chapter 6: 6.1

BASIC PRINCIPLES AND APPLICATIONS OF DIFFERENTIAL GEOMETRY


DIFFERENTIABLE MANIFOLDS - TENSOR ANALYSIS
COORDINATE SYSTEMS Algebraic and topological properties of Euclidean spaces. Smooth surfaces. Homeomorphisms. Regularity of a smooth map. Coordinate systems. Transition functions. Compatibility of coordinate systems. Atlas. DIFFERENTIABLE MANIFOLDS Differentiable structure. Maximal atlas. Differentiable manifolds. Simple manifolds. Standard coordinates and spherical coordinates on a sphere. The sphere as a nontrivial manifold. PRODUCT MANIFOLDS AND MANIFOLDS DEFINED BY CONSTRAINTS Product manifolds. The cylinder and the torus. The theorem of implicit functions. Manifolds defined by constraints. The configurations space of a double pendulum. 249

6.2

256

6.3

264

XIV

6.4

TANGENT VECTORS Velocity vectors. The tangent space. The canonical frame vectors. Extrinsic and intrinsic coordinates. Transformation rules under a coordinate exchange. Contravariant and covariant quantities. Scalar fields and vector fields. METRICS The inner product in an Euclidean space. Metrics. Metric coefficients. The reciprocal components. Transformation rules under a coordinate exchange. The determinant of a metric. Euclidean metrics and Minkowskian metrics. The light-cone. The canonical frame vectors and the induced metric on a sphere. THE MINKOWSKI SPACE The four-dimensional space-time. Inertial frames of reference. Basic assumptions in special relativity. Poincare transformations. The standard metric in Minkowski space. Lorentz transformations. Spherical symmetry in the Minkowski space. THE ACTION PRINCIPLE FOR A RELATIVISTIC PARTICLE Arc-lengths. Proper time. The four-velocity. The relativistic action. The Christoffel field. The Euler-Lagrange equations. Geodesics. Fictitious forces and the Christoffel field. Galilei's principle. Fictitious and gravitational forces. Tidal forces. COVECTORS Covectors. The dual vector space. Coordinates. Gradient vectors. Coordinate functions. Canonical frame vectors in the dual space. Transformation rules under a coordinate exchange. The canonical isomorphism between a tangent space and the dual space. Contravariant and covariant components of a tangent vector. Raising and lowering of indices. The gauge potential outside a solenoid. Gauge transformations on a covariant form. TENSORS Cotensors as multi-linear maps on the tangent space. Components. Transformation rules under a coordinate exchange. Tensor products. Contractions. Decomposition of cotensors. Mixed tensors of type (k,2). Tangent vectors as tensors of type (1,0). Contractions. Computation rules for mixed tensors. The Kronecker-delta as the components of a mixed tensor field. The canonical identification of mixed tensors of the same rank. Raising and lowering of indices. TENSOR FIELDS IN PHYSICS The electromagnetic field strength as a cotensor field of rank 2. The monopole field in spherical coordinates. The covariant expression for the electromagnetic energy-momentum tensor. The four-momentum as a linear map. The energymomentum tensor as a multilinear map. SOLUTIONS OF WORKED EXERCISES

6.5

276

6.6

284

6.7

291

6.8

300

6.9

308

6.10

318

322

Chapter

7: 7.1

DIFFERENTIAL FORMS AND THE EXTERIOR ALGEBRA


INTRODUCTION The transformation rule for partial derivatives.

325

xv
k-FORMS THE WEDGE PRODUCT Skew symmetric cotensors. k-forms. Skew symmetrizations. The wedge product. Associativity and distributivity of the wedge product. Commutation rules for the wedge product. Decomposition of k-forms. The basic forms in the Euclidean space. Stratifications. Basic forms on an arbitrary manifold.

327

7.3

THE EXTERIOR DERIVATIVE Skew symmetrization of the partial derivatives. Definition of the exterior derivative. Poincares lemma. Leibniz' rule. Closed and exact forms. Simple forms.

336

7.4

THE VOLUME FORM 343 The volume form on a manifold with a metric. Orientability of a manifold. The Maoius strip as an example of a non-orientable manifold. The transformation rule for the determinant of the metric. The Levi-Civita symbol. The Levi-Civita form. The Levi-Civita form as a pseudo-tensor. The Levi-Civita form as a volume form. The cross product in n dimensions. THE DUAL MAP The dual map. Linearity of the dual map. The reciprocal map. Double-dualizations. Dual forms as pseudo-tensors. The scalar product between covectors. The scalar product between k-forms. The dual map as an isometry. Contractions between k-forms and m-forms. THE CO-DIFFERENTIAL AND THE LAPLACIAN Definition of the co-differential. Poincares lemma for the co-differential. The co-differential as a generalized divergence. The Laplacian. Co-closed and co-exact forms. Harmonic forms. Primitively harmonic forms.

7.5

352

360

7.7

EXTERIOR CALCULUS IN 3 AND 4 DIMENSIONS Conventional vector analysis in the Euclidean space R3. The dual map in R3. The scalar product and the wedge product in R3. The exterior derivative in R3. Applications of Poincares lemma and Leibniz' rule. Limitations of the exterior calculus. The dual map in Minkowski space. The exterior derivative in Minkowski space. ELEcrROMAGNETISM AND THE EXTERIOR CALCULUS The electric field strength as a one-form and the magnetic field strength as a two-form. Maxwell's equations in the exterior calculus. The magnetic field outside a wire with an electrical current. The electromagnetic field strength as a two-form. Maxwell's equations on geometrical and covariant form. The monopole field. The gauge potential corresponding to a monopole field. The Dirac string. SOLUTIONS OF WORKED EXERCISES

366

7.8

376

385

Chapter

8: 8. 1

I NTEGRAL CALCULUS ON MAN I FOLDS


INTRODUCTION The 2-dimensional Riemann-integral. The integral of a 2form. Reduction of the integral of a 2-form to an ordinary 2-dimensional Riemann-integral. SUBMANIFOLDS - REGULAR DOMAINS Submanifolds. Regular domains. The interior and the boundary of a regular domain. Cartan's lemma. Orientability of regular domains.

393

8.2

395

mI

8.3

THE INTEGRAL OF DIFFERENTIAL FORMS Definition of the integral of a k-form over a k-dimensional domain. Invariance of the integral under coordinate exchanges. Partitions of unity. Integration on nontrivial domains. Reduction from extrinsic to intrinsic coordinates. ELEMENTARY PROPERTIES OF THE INTEGRAL Linearity. Stokes' theorem. Partial integration. Integrals of scalar fields. Volume integrals. Integrals of vector fields: Line-integrals and flux-integrals. Gauss' theorem in geometrical and covariant form. Integrals of simple forms. THE HILBERT PRODUCT OF TWO DIFFERENTIAL FORMS The Hilbert product. The dual map as a unitary operator. The cO-differential as the adjoint operator of the exterior derivative. The Laplacian as a negative definite Hermitian operator. Characterizations of harmonic forms on a compact manifold. Green's identities.

403

8.4

411

420

8.6

THE LAGRANGIAN FORMALISM AND THE EXTERIOR CALCULUS The covariant action. The covariant equations of motion. The relativistic action of a system consisting of charged particles interacting with the electromagnetic field. Geometrical derivation of the Klein-Gordon equation. Geometrical derivation of the Maxwell equation.

424

INTEGRAL CALCULUS AND ELECTROMAGNETISM 429 The integral representation of the magnetic flUX, the electric flux and the electric charge in 3-space. Space slices in Minkowski space. The integral representation of the magnetic flux, the electric flux and the electric charge in Minkowski space. The Ginzburg-Landau model for superconductivity. The coherence length. The penetration length. The Ginzburg- Landau parameter. Magnetic vortec strings in a type II superconductor.

8.8

THE NAMBU STRING AND THE NIELSEN-OLESEN VORTEX Action of the Nambu string. Equations of motion for the relativistic string. The edge conditions. The total momentum and the angular momentum for the relativistic string. The spinning string. Regge trajectories. The abelian Higgs' model and its relationship with the Ginzburg-Landau model of superconductivity. String-like configurations in the abelian Higgs' model. The Nielsen-Olesen vortex. Confinement of magnetic monopoles in the abelian Higg's model. Quark-confinement. The Nielsen-Olesen vortex as a smooth extended Nambu string. SINGULAR FORMS The naive characterization of singular forms. The a-function on a manifold with metric. The singular current associated with a point particle. The Coulomb field. Singular forms within the framework of distributions. Weak forms represented by linear functionals. Extension of the exterior calculus to weak forms. Weak forms induced by regular domains. The co-differential as a boundary operator. Deficiences of the exterior calculus of weak forms. SOLUTIONS OF WORKED EXERCISES

442

8.9

454

461

XVII

Chapter

9: 9.1

DIRAC MONOPOLES
MAGNETIC CHARGES AND CURRENTS Magnetic currents. The modified Maxwell equations. The singular electric and magnetic currents associated with point particles. Conservation of the energy and momentum. The generalized Lorentz force. Charge rotations. THE DIRAC STRING The Dirac string as a physical string: The modified monopole field in R3. Exactness of the modified monopole field. Dirac strings in Minkowski space. The singular form associated with a Dirac string. Dirac's veto. Dirac's lemma. The modified monopole field in Minkowski space. DIRAC'S LAGRANGIAN FORMALISM FOR MAGNEITC MONOPOLES Dirac's action for a system including monopoles. Dirac's theorem: Derivation of the modified Maxwell's equations and the equation of motion for the electrically charged particles. Derivation of the equation of motion for the monopoles. THE ANGULAR MOMENTUM DUE TO A MONOPOLE FIELD The angular momentum of the electromagnetic field generated from a charge-monopole pair. Quantization of the magnetic charge. The equation of motion for a charged particle moving in a monopole field. Conservation of the speed. Conservation of the total angular momentum. The characteristic cone. QUANTIZATION OF THE ANGULAR MOMENTUM The Hamiltonian of a charged particle in a monopole field. The Dirac formalism versus the Schwinger formalism. The operator for the total angular momentum. The spectrum of the angular momentum. The intrinsic spin. The eigenvalue equations for the angular momentum operators. The bosonic spectrum in the Schwinger formalism. The fermionic and bosonic spectrum in the Dirac formalism. THE GAUGE TRANSFORMATION AS A UNITARY TRANSFORMATION Unitary transformations in quantum mechanics. The phase shift as a unitary transformation. The gauge transformation as a unitary transformation. Gauge equivalence versus unitary equivalence.

9.2

9.3

482

9.4

486

9.5

491

9.6

498

9.1

QUANTIZATION OF THE MAGNETIC CHARGE 500 Charge quantizacion as a consequence of the indetectability of the Dirac string: (a) Unitary equivalence of Hamiltonians based upon different strings. (b) Vanishing of the Bohm-Aharonov effect. Fluxquantization in a monopole field versus fluxquantization in a superconductor. SOLUTIONS OF WORKED EXERCISES

503

Chapter 10:
10.1

SMOOTH MAPS - WINDING NUMBERS


511 LOCAL PROPERTIES OF SMOOTH MAPS The definition of a smooth map from one manifold to another. The induced transport of tangent vectors. Deficiences of the vector transport. Regularity of a smooth map. Regular and critical values. The existence of a local inverse map. Immersions. Embeddings. Submersions. The theorem of Sardo

XVIII
10.2 PULL BACK OF COTENSORS 520 Transport of tangent vectors versus pull back of covectors. Pull back of cotensor fields. Pull back of metrics. Pull back of differential forms. The transport of tensor fields induced by diffeomorphisms. The behaviour of the exterior calculus under pull backs; The wedge product and the exterior derivative. Commutative diagrams. Pull back of integrals. ISOMETRIES AND CONFORMAL MAPS Characterization of conformal maps on manifolds with Euclidean metrics or Minkowski metrics. Dilatations, inversions and special conformal transformations. Isometries as symmetry transformations. The definition of a geodesic on a manifold with a Minkowski metric. The affine parameter on a geodesic. Time-like geodesics, nUll-geodesics and space-like geodesics. Pull back of geodesics. The isometry group on a manifold. The Poincare group. THE CONFORMAL GROUP The inversion. The singularities associated with a conformal transformation. Conformal compactification of the Euclidean q space Rn. Extension of the pseudo-Cartesian space RPxR . Conq formal compactification of the pseudo-Cartesian space RPxR . Conformal transformations generated from pseudo-rotations in the extended space. The conformal group. The conformal group as the smallest extension of the isometry group which includes the inversion. THE DUAL MAP Commutation of isometries with the dual map, the co-differential and the Laplacian. Invariance under isometries of the inner products between differential forms. Behaviour of the dual map under conformal transformations. THE SELF-DUALITY EQUATION The self-duality equation on a differentiable manifold. Conformal invariance of the self-duality equation. The CauchyRiemann equations. Holomorphic functions as conformal maps in the complex plane. Conformal invariance of Maxwell's equations. Conformal maps on the sphere. The sphere as a complex manifold. Holomorphic maps on the sphere. 530

10.3

10.4

540

10.5

551

10.6

554

10.7

WINDING NUMBERS 562 The degree- of a smooth map between compact manifolds. Brouwer's lemma. The winding number of a smooth map. Brouwer's theorem. An integral formula for the winding number. The volume form on a sphere. The winding number of a smooth map into a sphere. The non-compac~ case. THE HEISENBERG FERROMAGNET 570 The local spin direction as an order parameter for the ferromagnet. Heisenberg's model for the ferromagnet. The field equations for the static equilibrium configurations. Boundary conditions and the winding number for a finite energy configuration. The ground states for the topologically non-trivial sectors; The spin waves. The Bogomolny decomposition. The double self-duality equation. Spin waves as conformal maps. Absence of other static finite energy solutions.

10.8

XIX
10.9

THE EXCEPI'IONAL q, 4 -1~ODEL The abelian Higgs' model in (2+1)-dimensional space time. The field equations for the static equilibrium configurations. Boundary conditions and the winding number for a finite energy configuration. Reduction of the second order field equations to a pair of first order equations. The exceptional q,!model. The Bogomolny decomposition. The virial theorem for static finite energy solutions. Spherically symmetric flux tubes. SOLUTIONS OF WORKED EXERCISES

578

Chapter 11:
11. 1

SYf1>1ETRIES AND CONSERVATION LAI,/S


CONSERVATION LAWS Conserved currents. The tube lemma. Conservation of electric charge: The first half of Abraham's theorem. Conservation of energy and momentum: The second half of Abraham's theorem. SYMMETRIES AND CONSERVATION LAWS IN QUANTUM MECHANICS Unitary transformations generated by isometries. The infinitesimal generator associated with a one-parameter group of unitary transformations. The characteristic vector field associated with a one-parameter group of isometries. Symmetry transformations in quantum mecha nics. Conservation laws in quantum mechanics. CONSERVATION OF ENERGY , MOMENTUM AND ANGULAR MOMENTUM IN QUANTUM MECHANICS Conservation of momentum as a consequence of translational
invariance. Conservation of angular momentum as a consequence

11.2

592

11.3

596

of rotational invariance. The intrinsic spin of a vector particle. Angular momentum for a charged particle in a monopole field. The infinitisemal generator necessary to compensate for rotations. The relationship between the infinitisemal generator and the correction term for the orbital angular momentum. 11.4 SYMMETRIES AND CONSERVATION LAWS IN CLASSICAL FIELD THEORY Symmetry transformations in a classical field theory. The characteristic vector field associated with a one-parameter family of diffeomorphisms. Noether's theorem for space-time transformations, internal transformations and generalized symmetry transformations. Noether's theorem for scalar fields. ISOMETRIES AS SYMMETRY TRANSFORMATIONS poincare transformations as symmetry transformations. Conformal transformations as symmetry transformations. Conservation of energy and momentum as a consequence of translational invariance. Conservation of angular momentum as a consequence of Lorentz invariance. Symmetry of the canonical energy-momentum tensor in a scalar field theory. Spin contribution to the canonical energy ~momentum tensor in a vector field theory. THE TRUE ENERGY -MOMENTUM TENSOR FOR VECTOR FIELDS Non-covariance of the canonical energy-momentum tensor for a vector field. Construction of the true energy-momentum tensor. Noether's theorem for vector fields. The conservation laws corresponding to conformal invariance. 600

11 .5

606

11 .6

611

xx
11.7: ENERGY-MOMENTUM CONSERVATION AS A CONSEQUENCE OF COVARIANCE
The metric energy-momentum tensor as the functional derivative of the action with respect to the metric. The metric energymomentum tensor for electromagnetic fields. The metric energymomentum tensor for scalar fields, vector fields and relati~ vistic particles. Covariant conservation of the metric energymomentum tensor.

614

11.8

SCALE INVARIANCE IN CLASSICAL FIELD THEORIES


Scale transformations in Minkowski space. Internal scale transformations. The scaling dimensions for scalar fields and vector fields. Scale invariance of massless field theories in Minkowski space. The conformal energy-momentum tensor for a scalar field.

618

11.9

CONFORMAL TRAl,SFORMATIONS AS SYMMETRY TRANSFORMATIONS


The scale associated with an arbitrary diffeomorphism. Internal scale transformations on an arbitrary manifold with a metric. Conformal invariance of massless scalar field theories in Minkowski space. The conservation laws associated with conformal transformations.

625

SOLUTIONS OF WORKED EXERCISES

632

PARr I

A~ 2(~~

-+

{-Yj
X

>UttttJ]

&It I : ~I I" I I
IIIIIIII

I IIII II

~ B=O

BASIC PROPERrlES OF PARTICI.ES AND FIELDS

GENERAL REFERENCES TO PART I:

R.P.Feynman, R.B.Leighton and M.Sands, "The Feynman Lectures on physics", McGraw-Hill, New York (1965). P.A.M. Dirac,"The Principles of Quantum mechanics", second edition, Oxford (1935). L.I.Schiff,"Quantum Mechanics", McGraw Hill Kogakusha ltd (1968) de Gennes, "Superconductivity of Metals and Alloys", W.A. Benja-

min Inc, New York (1966) A.O. Barut, "Electrodynamics and Classical Theory of Field and particles", MacMillan, New York (1964) Coleman, "Classical Lumps and Their Quantum Descendants" in

Erice Lectures, "New Phenomena in Subnuclear Physics", Plenum Press, New York and London (1975). Coleman, "The Uses of Instantons", in Erice Lectures, " The Whys of Subnuclear Physics", Plenum Press, New York and London (1977) R.P. Feynman and A.R.Hibbs, "Quantum Mechanics and Path Integrals", McGraw-Hill, New York (1964). L.S. Schulman,"Technique and Applications of Path Integration", Wiley, New York (1981).

ehllpter I

ELECTROMAGNETISM
1.1 THE ELECTROMAGNETIC FIELD
-+ -+ -+

The fundamental quantities of the electromagnetic field are the fieZd

strengths:

E(t,x)

-+

and

B(t,x) This vector may depend

The electric field strength is an ordinary vector field, i.e. to each point in space we have attached a vector. on time, too. The magnetic field strength is pseudo-vector field, i.e.

to each point in space we have attached a pseudo-vector which may depend on time, too.
Although we are going to discuss mathematical concepts in greater detail later on, let us clarify the situation a little. An ordinary vector (like is nothing but a directed line segment PQ connecting two points in the Euclidian space. It is defined before we have introduced a coordinate system. But a pseudo-vector can only be specified if we also specify an orientation. Consider two orderei s~ts-+of linea~ilz i~ dependent vectors (UI,U2,U3) and (VI,V 2 ,V3). We can decompose one set on the other set in Q the following way:

E)

Vj = DiAij We say that they are equivalent if det(Ai.O. z I~ this way all possible ordered sets J (U,V,W) are divided into two classes. We p arbitrarily choose one of these classes to represent positive oriented sets, and the other class to represent negative oriented sets. Once we have chosen a specific orientation, the pseudo-vector is represented by an y ordinary vector B = (Bx,By,B z ) but if we exchange the orientation, then the pseudox vector is represented by the opposite vector: Fig. 1 !'= -! = (-Bx,-B ,-B z ) The most well-known example of a pseudovector is the cross product of two ordinary vectors: D x It is sometimes denoted U A V and referred to as the wedge product, but we shall avoid this notation, which we r~serve for differential forms. Once we have chosen a specific orientation, then D x V is specified by the following requirements:

V.

1.

The length of D x V is equal to the a~ea of the parallellogram spanned by U and V

2.
3.

Ux V is
spanned by (D,

pe~pendicular to the parallellograrn


U and V.
x

V,

V)

generates a positive orientation.

(See fig. 2)

z
+

+~
Fig. 2
+ +

3
x

Back to business! ic field. a force locity


(1.1)
V
+

The field strengths

and

can be measured it will experience


+

using the interaction between charged particles and the electromagnetIf the particle carries the charge
F :
+

q,

which depends both on the particle's position

and ve-

F
+ +

q(E + v x

B)
B
+

(Observe that vector!)

is an ordinary vector because

is a pseudo-

This force is known as the Lorentz force.

using the prin-

ciples of special relativity, this means the particle has the following equation of motion:
(1.2)

E dt

p
=

)1- ::

mv

(c

velocity of light)

This relation is of extreme importance and actually it serves to define the electromagnetic field. If the charge is small, we may negUsing lect the influence of the particle on the electromagnetic field. may analyse their motion and hence determine the field strengths: E(t,i) and 13(t,i). The field strengths themselves evolve in space and time hence they, too, obey some equations of motion, the Maxwell equations:

a swarm of test particles moving through the electromagnetic field, we

(1.3)

(No magnetic poles)

(1.4)

at

3B + V x + E

-0
L
EO
~
"t-

(Faraday's Law)

(1.5)

V 3E - c 2v x at

E
B
+

(Gauss' Law)

(1.6)

EO

(Ampere's Law)

Here

is the charge density,

-T

is the current and

EO

is the

permittivity of empty space. Furthenrore we have introduced the vector operator '" 3 3 3 v =(3x'3y'a-z)' In the following table we have collected some useful formulas:

q,(t,~)

is a scalar field; is a vector field; is a pseudo-vector field;

Vq,(t,~)

is a vector field: The gradient. is a scalar field: llie divergence.

E(t,~)
B(t,~)

V'E(t,~l

vxB(t,~)

is a vector field: The curl.

{:"

2 2 3 32 3 = -+ - + -2 2 3y2 3x 3z

Laplacian

(1.

7l
A A x
+ +

+ + + A x B = -B x A

(1.8) (1. 9) (1.10) (1.11) (1.12) (1.13) (1.14)

(BXC) = (AxE) C (EXC) = E(A'C)

(A'E)C

(AxE) V

x C = (A'C)E - A(E'C) (Vq,)

= M = -0 = =
0

Vx
V

(Vq,)
(vxB)

V x (vxB)

V(V'E)

- {:"i3

Gauss' theorem:

f ~'E
(l (l

dV =

f E'~
S

dA

(1.15)

fl

inside region bounded by S Closed surface Unit normal pointing outwards

Stokes' theorem:

f ('7
S

->- ->-

x B)ft dA

,( ->

J Bdr r

->

(1.16)

~
r

r 8

Surface bounded by r Closed loop Unit normal pointing outwards

Theorem of line-integrals:

f 9~ 'd~= ~(Q)-~(P)
r
Q

Smooth curve with end-points P and Q.

let us deduce sane of the irnp:lrtant consequences

of Maxwell' s equations.

If

we take the divergence of (1.6), we get:

(dE)

dt 1

c2

(9 x B)
0

e:

9 9

-T

O
-T

d at (9'E) 1

e: O e: o

.. ..

e: o at

.!Q.

__I_V

-T

where we have used (1.13) and (1.5) Omitting the trivial common factor we have thus shown:

(1.17)

This equation is known as the equation of continuity. portant consequence of charge-conservation. total charge of the system:

It has the im-

To see this, consider the

where we have integrated over all space at a specific time. sity and all currents_ vanish sufficiently far away. tegrals are well-defined. cause the charge-density Now formally Q

We assume,

of course, that the system is confined in space so that the charge-denHence, all our inBut might depend on time be-

p(t,~)

may very well depend on time!

due to the equation of continuity, we get: dQ ; dt

J~ at

3 d x

Jv . j
j

d 3x

Using Gauss' theorem (1.15) this can be rearranged as

~; - J<j.n)dA; 0
surface at infinity

because the system is confined so that

vanishes at infinity.

This

is our first example of a conservation law and we emphasize that it depends strongly on the equations of motion, i.e. the Maxwell equations. Later on we shall discuss examples of conserved quantities which are conserved independently of the equations of motion. They will be conserved for topological reasons rather than for dynamical reasons.

1.2

THE INTRODUCTION OF GAUGE POTENTIALS IN ELECTROMAGNETISM


NOW,

take the Maxwell equation

(1.3).

If we try the following

assumption: (1.18) where

is a vector field, we see that the Maxwell equation (1.3) is

automatically satisfied:

v. B
due to (1. 13) . (1.18).

Hence, we might look for solutions to Maxwell equations on the form However, it is not at all evident that all solutions should To make it more precise, we consider and a pseudo-vector field
-+

automatically be on this form. a specific region in space on

defined

in such a way that

vB

throughout the whole region to find a vector field eXistence of A


+

Q.

Then it is in general not possible Actually, the

with the property (1.18).

depends very strongly on the topological properties


Q

of the space region

Only for very simple space regions (e.g. the

interior of a cube) is it always possible to find a suitable lutions to the Maxwell equations on the form (1.18). Substituting this ansatz into (1.4), we get:

A.

But

let us neglect these difficulties for a moment and just look for so-

aat . (vxA)
which implies that

Vx E
+ E)-

(aA at ait at
=

Using the same trick as before, we try the assumption:

E +
where
<I>

_*" v"
sign~)

is a scalar field.

(Observe the

This obviously solves

(1.4) because:

due to (1.12). Hence, if we look for solutions to the Maxwell equations on the form

(1.19)

v
(1.3) and (1.4)!

it

-v<jl - at

;I:

+ aA

then we have automatically solved the first two of the Maxwell equations:
tials

The new fields

<I>

and

it

are called gauge poten-

and from now on we will be very much concerned with their prosubstitu~e

perties! First, let us equations: (1.19) into the two remaining Maxwell

L
and

e: o
~

.E
+

aA <jl--) V .(- V at
- c 2v
+ xB

-'1<1>

a V - at

,..
o

e:

at

aE

at

(-V<jl

aA) - at
2+

-c 2v

x (vxit)

~1 - ~ -c2~(V.A) + c 2 VA

at

at2

Rearranging these two equations, we finally get the equations of motion

for the gauge potentials:

(1.20a)

(M -

1 tl ;-zat 2 )

-~
EO

-at
-T

[VA +

1 2.1J c 2 at 1 2.1] c 2 at

(1.20b)

1 a 2A (/:;A - -;;zat 2 )

i
EOC 2

+ V [VA +

Observe that we have artifici:ally introduced so impressing.

2 1 a <p c 2 at 2 on both sides of

(1.20a) to make it look like (1.20b). This may at first sight not seem Although we have reduced the number of equations from 4 to 2, they are still complicated. Especially, they are still mixed -+ in <p and A. But now you should remember that we are actually not looking for the gauge potentials <p and A but for the field strengths -+ -+ E and B because they are the only ones you can measure! Now, suppose we have found a solution to the Maxwell equations (1.3) -

(1.6) represented by the gauge potentials <Po and Ao:


-+ Eo

= -

aAo <Po - ~

then we can immediately find other gauge potentials representi:ng the same field strengths Bo and Then: Eo To see this, let x(t,i) be an arbitrary scalar field.

will do the job:

Vx A
by (1.12) and .. ait ;I: ;I: ax aAo a-+ -v<p - at = -v<P o + v at - ~ - at 'Ix

Furthermore, the new gauge potentials <p and A will solve exactly the same equations of motion as <Po and Ao . By substituting <p and

..

-V<P 0

-+ aAo
-~

E0

A
ing

into (1.20a)

and

(l.20bJ, you will easily find that the terms contain-

X drop out.

By representing the electromagnetic field through

the gauge potentials <p and A we, therefore, have discovered a strange symmetry: The field strengths and the equations der the transformations:
(1.21)

of motion are unchanged un-

..
A

A+

Vx

10

Such a transformation is called a gauge transformation and since physics is unchanged under this transformation, we speak of gauge symmetry. Hence, we see that a physical system, like the electromagnetic field, is described not only by a single gauge potential

(<P,A)

but by a whole Con-

family of <gauge pOtentials differing only by a gauge transformation. sequently, we may parametrize the solution in the following way:

(<Px,Ax) =
a specific gauge. If

(<Po -

lx at' A 0

Vx)

By picking a special member of this family, we say that we have chosen This freedom of choosing a gauge is very important. Let us now return to the equations of motion (l.20a) and (l.20b)

(<P,A)

(<Po,A o ) denotes a specific solution, we may then choose


satisfies:

so that

(1.22)

To see this, we substitute

We then get:

o
i.e.

~x - ~ ~ = - [V.A o + c 2 at 2
;!;

1 Ha] c 2 at

This is a wave equation where we know the source term:


-+

[V'A o

+ c 2 at]
We may al-

a<po

We can solve it and hence, we have obtained the appropriate gauge con-

dition

(1.22).

Observe

~hat

the solution is not unique:

ways add a solution to the homogenous wave equation: (1.23)

o
-+

Hence, we are still allowed to make gauge transformations: <P

<p

}t ,

A A + Vx
-+

without spoiling the condition (1.22) prcvided

fies

(1.23) ~

X satisThe gauge condition (1.22) is called the Lorenz*~ondi


In this gauge the

tion, and we say that we work in the Lorenz gauge.

*)

L.V. Lorenz (1829-91). Danish physicist who, independently of Maxwell, constructed a theory of light b~ing electromagnetic waves.

11

equations of motion simplify considerably, because the equations for


~

and

decouple:

(1. 24)

These equations are beautiful wave equations and they can be solved by standard techniques. Passing by, we observe that the Lorenz condition:

(1. 22)

has exactly the same form as the equation of continuity: (1.17) In gauge theories it is a standard "trick" to let the gauge condition resemble the characteristic equation of the source.

1.3

MAGNETIC FLUX
To get some experience with the

~
/

n/-y s
r

B
B
-+

gauge potentials, we will deduce a formula for the magnetic flux. a surface S Let us consider bounded by the closed loop

r
,
n
\

We are interested in the magnetic t

_--------------------_

flux passing through the surface at a specific time

to .

By definition

the flux is given by the formula:


qI

=
~

J B.
S

r;

dA

Fig. 3

Substituting (1.18), we get:


qI

J(VXA) . r;
S

dA

which, by Stokes' theorem (1.16), may be rewritten as:

(1.25)

Hence, we can express the magnetic flux in terms of be static or anything else.

Observe that It need not

we have put no restriction On the electromagnetic field.

This formula has important consequences.

12

Consider a closed surface S. We assume that the electromagnetic field is smooth in a neighbourhood of S but it may be singular at a finite number of pOints inside it. (The electromagnetic field created S by an electron is singular at the actual position of the electron) . Let us now calculate the magnetic flux through

qI=JE.ndA=fA'dr
s
but since S is closed,

r r
is empty. Consequently, the last inte-

gral vanishes and

qI

s
+

+:

SINGULARITIES

Fig.

You may feel I am cheating you! But let me try to convince you in another way: Choose a closed loop r lying on S, which divides S into two regions: Sl and S2 (See fig. 4). The corresponding fluxes through S1 and S2 will be denoted q\l and q\2' We then have: q\1

dA =

n
r

dA

dr

-+

because this time we integrate the other way around get as postulated:

Adding the two fluxes, we

So we have shown:
Theorem 1. vanishes! If the electromagnetic field is generated from a gauge potential

(~,A), then the magnetic fluz through any closed surface automatically

To understand this result, let us consider the static Coulomb field created by a single electron. Let S be the surface of a ball with

13 radius r

and center at the position Then the normal comE

of the electron. easily found:

ponent of the electric field

....

is

1 r2 It is constant on the surface


E

....

. fi

-'L
4'1TEO

and

the electric flux is then easy to calculate:

JE .
s
Fig. 5

fi dA

IdA
S

--.!L. J:.. 4 '1Tr 2


4'1TEO r2

= .3..
EO

All these computations were made just to remind you that the electric flux through a closed surface is pro-

portional to the charge contained inside it.


netic charges, i.e. magnetic monopoles.

Therefore the above re-

sult about the magnetic flux simply excludes the possibility of magOf course, this is not surprising because we started with the Maxwell equations etc. where we have included an electric charge density, but not a magnetic charge density. We are now prepared to consider a space region

with a magnetic

field, which cannot be derived from a gauge potential (a) (b)

A.

Consider:

The electromagnetic field created by a resting electron. The electromagnetic field created by a hypothetical magnetic monopole.

In both cases the fields are singular at the origin clude that point. The remaining part of space is denoted

so we ex-

n:

n
In tions:

= R3 "

(O}

the electromagnetic field solves the following Maxwell equa-

vE

E
n
there is

This can be shown by a direct computation, and we emphasize that they will hold for both case (a) and (b). Observe that in neither electric charges nor magnetic charges. tions. Furthermore, the Cou-

lomb field and the monopole field do not differ by their field equaBut they differ when we try to represent them by gauge potentials!

14

The Coulomb field may be represented by the gauge potential


'i'

"'-~ 471E:

1.
r

A = -0

because this immediately leads to:

B = V x A = -0

and

E = -VIj>
S is dif-

But the monopole field cannot be represented by a gauge potential (1j>,A): This is because the magnetic flux through the unit sphere ferent from zero:
q\

JB
S

fi dA

...9:.. 471

471

We immediately see that this magnetic field contradicts theorem 1. Thus, we have shown that even if
'V

....

B = 0

....

in a space region

, we cannot in general conclude that


x

may be

written on the form:


(1.18)

....

We may summarize the preceding discussion in the following way: The concept of magnetic monopoles and the concept of gauge potentials sean to be in conflict with each other! in peaceful co-existence. Later on we shall spend much time constructing a theory where magnetic monopoles and gauge potentials live

15

1.4

ILLUSTRATIVE EXAMPLE: THE GAUGE POTENTIAL OF ASOLENOID


Up to this point we have discussed gauge potentials in very general

terms only, so let us pause to compute a gauge potential in a special situation. Consider a long coil of wire wound in a tight helix. A current through the wire will produce a very strong magnetic field inside the solenoid, and if the solenoid is very long compared to its diameter, the magnetic field outside the solenoid will be negligible. Let us consider

....

....

the ideal case of an infinitely long solenoid. Then we can safely put

B= 0

outside it.

Inside the sole-

noid the magnetic field direction. Now, let us try to compute the gauge potential potential we need to be concerned about.

will

everywhere be pointing in the same

which is the only gauge

Inside the solenoid we have a constant magnetic field:

This may be represented by


A =
-+

Bn """2

(-y,x,O)

:2

-+

x r

-+

as you can easily verify. This choice of gauge has nice properties: (1) The magnitude of the gauge field is proportional to the distance (2)
-+

from the z-axis.

It is always perpendicular to the z-axis and the radial vector


p.

Outside the solenoid the magnetic field vanishes:


it would be tempting to put a closed loop through

B = o.

Hence,

A=0 ,

but this is wrong!

If we consider

as shown in fig. 8, then the magnetic flux passing

is

JB

ndA

16

but it can also be expressed as the line integral: (1.25)


<l>

fA
A

dr

...
! What can

and this is certainly inconsistent with the choice: we do then?

First, you observe that

the gauge potential inside the solenoid is circulating around the z-axis, but then it is tempting to let the gauge potential outside the solenoid do the same:

a:

(-y,x,O)

Second, we know that the line integral should have the constant value B o na 2 But if we choose r as a circle with the radius Po then it is clear that the line integral is proportional to the magnitude of and proportional to the radius ly proportional to Po:
1

Hence,

IAI

must vary inverse-

IAl
A=
really vanish?

a:

x 2 +y2

Combining these two things, we look for

on the form:

k(- ~ , _ _ x_ ,0)

x 2 +y2

x 2+y2
Does its curl The is The second thing we should

The first thing we should check is the following one: This is easily verified.

check is whether it produces the correct magnetic flux or not. tangential component of k
1 Po

and so we get: dr

...

J:..
Po

211 P

2nk

Hence, we can determine the constant k , because

...

2nk

Bona2i.e. k

=~

B oa 2 .

<lITLD <QII.D
<1-...OJj)

Thus, we have completely solved our problem: ...


A inside

<UTI.P

Fig. 9

IIII1

= ""2
Bo

Bo

(-y,x, 0) ;

...
A outside =

""2.

x 2 +y2

(-y,x,O)

17

Observe that Ainside ary of the solenoid.

and

Aoutside

match continuously on the bound-

You might find it puzzling that we can have a non-vanishing gauge potential A in a space region where the magnetic field vanishes! At first you might think that it occurred only because we worked in a "bad" gauge, so maybe, if we used a gauge transformation, we could "gauge away" have seen that A? A But we have already killed this hope because we

=0

is incompatible with the magnetic flux passing Hence, we cannot gauge away A throughout the

through the solenoid!

exterior of the solenoid. However, it is possible to gauge away the gauge potential a.lm:Jst everywhere! To see this, we consider the function: <P(x,y) Arctan 'I.. x
<p

y
(0,1) 7r <P(O,l)=2:
(x,y)

Observe that

simply produces the


<p

polar angle and that it makes a jump from <P <P :


~ ax

is smooth to
-11

except at the negative x-axis, where +7r First we compute the derivatives of

r<P

+11

<P

-11

(1,0)

<P(l,O)=O (0,-1) 7r <P(O,-l)=-z: Fig. 10

a 'I.. ax Arctan x

2 __ x_

- '1..
x

1+('1..)2

x 2 +y2

E.1
ay They look very much like Ax and Ay
B

a Arctan 'I.. x ay

1 x
1+('1..) 2

X +y

Now consider the gauge transformation with X(t,x,y,z)

= -

o -2<P(x,y)
-~

a2

This produces the equivalent gauge potential:


-~

x 2 +y2

x 2+y2

o o
Q

A'

A + 'ilX

B a2 2

__ x_ x 2 +y2

B0 a 2 -2-

x -2--2 x +y

o
Hence, we have managed to gauge away y A

except for the half plane:

0, x < 0 , where

is singular.

Since there is symmetry about

18

the z-axis, we may ignore the z-coordinate for a moment. The problem is then really two-dimensional, and we see that everything is all right except on the negative x-axis. Since X is
discontin~ous

along the nega-

tive x-axis, we conclude that singular - like the negative x-axis.

Vx

becomes

a-function - on the

If we insist on perform-

x
Fig. 11

ing this singular gauge transformation, i.e. insist on using a singular gauge, we see that we have concentrated a singularity in
the gauge potential along the negative x-axis.

SUch a string on which the gauge potential is singular is called a DiraC! string. We will discuss them in greater detail later on. Now what about the fact that

~ =
Dirac string too? where Yes!

fA.
has a

dr

BoTIa

Does it hold in the singular gauge, When we integrate, we a-like singularity necessarily pass the negative x-axis

\
x
A' is

A'

and this gives the correct contribution to the line integral. We may clarify these remarks by recalling the definition of a a-function. It is defined by the property that
+

singular! Fig .. 12

a(x)f(x)dx = f(O)

for any smooth function which vanishes at infinity. Now consider the Heaviside function S(x) defined by:

e (x)
de dx

for

x <!

for x < 0 This function makes a jump of height 1 at x = 0 and we are going to show that:
o(x)
=

In fact. we immediately find:


+
~

-J

de f (x)dx dx

[e(x}f(x)]: : -

J~

e,(x}f' (x)dx

Jf' (x)dx

- [f(x)]~

f(O)

19

where we have integrated by parts and extensively used that f vanishes at infinity. Hence, we have demonstrated in all details that if a function makes a jump, then the derivatives get a-like singularities.

Worked exercise: 1.4.1 Problem: Prove the following formula:


+
~

J f (x) 0 (g

(x) )dx

where

is a monotonous function!

1.5

RELATIVISTIC FORMULATION OF THE THEORY OF ELECTROMAGNETISM


Up to this point we have carefully separated space and tLme, and

we have not used Lorentz invariance. form.

However, it will several times

be more convenient to translate the results into a Lorentz invariant We assume that you are familiar with the basic principles of special relativity, but to fix notation we have collected some useful formulas:

In relativistic formulas we put

c = 11 = EO = 1

(Natural units)

Indices:

Space-time indices: Space indices: Other indices:

a,

a,y, ...

,)l,V

= = =

0, 1, 2, 3

i,j,k,l a,b,c,d

1, 2, 3 1, 2, ... , n

SEace-tLme diagram Coordinates:


P(t,x,y,z)
t

xa : xO=t

Cl

= 0, 1, 2, 3 xl=x x 2=y x 3=z

Parametrization of world line: x


Cl

Xa(T)

T : ProEer time
x

1
~=i (t)

dt = ydT
Y

7i7

20

Metric:

Four vector: ka ka (k ,k i ) naS kS

= =

(w,R)

= (-w,k)
ka k a

w: k :
Tensor:

Scalar part Vector part

-w 2 +

k2

s 00
=

[ ~:O

-:-

I I

S OJ

oj ;i~
Four-velocity:

Velocity:
+

(dX

dt, dt, dt

dz)

va

is a time-like unit vector, pointing towards future

Energy, momentum: E
..

Four-momentum: pa

= my
+
myV

= m dT

dxct

= (E,

p)

p =

-p pa = rn 2
ct

square of rest mass

Charge density, current:


p, J = pv
-t
+

Four current:

Gauge potentials:
<p,

Gauge potential
A
ct

=
=

Maxwell field:

A
Derivatives:

(-<p,A)

Derivatives:

a = __ 3_ =
a axa

(!t'
0

V)
d'Alembertian

21

Field strengths:

Field tensor:
0

-E x
0

-E

-E
-B

z
Y

-+ -+ E,B

Fas

Ex E
Y

Bz z
0

-B

B
0

EZ Lorentz force:

By

-B x

F=

d-+ ~ dt

= q(E+~"B)

Fa

dpa a de = qF SuS

(1.26)

Maxwell eguations:
VoB
->-

-+

0
...Q..

dB + at dE dt

-+

VXE

= -0
-+ B

daFSY + dsFya + dyFas

(1. 27)

VOE =

EO

- c 2V x

"" - ~ EO

d FCl.S S

(1.28)

E~ation

of continuity:
-T

-+ ~+ V

at

d JCI. = 0

CI.

.
dSACI.

(1.29)

Gauge -+ B =

Eotentisl~:

Vx

-+ A ;

E = -V</>

dit -at

Cl.S

dCl.AS

(1.30)

Gauge transformation:
<P

-+

<P

_2.x
at

A -+ A + dCl. X CI. CI

(1.31)

-+ -+ A-+ A +

Vx
...l..ll = 2 at
c

Lorenz gauge:

-+ A +

d ACI. CI.

(1.32)

22

Eq. of motion

for

the gauge potentials:

ll<l> -

.1:..n
2 c 2 at

-E..._.1... (vit at
EO

2.1)
at

(1.33)

llA -

.1:.. ~ 2 2
c at

2-+-

1 -+- -1 -2 -tJ+ V (V A + 2 2.2.) at E c c 0

Observe that the gauge potentials <I> and A have been collected into a single four-vector field ACt = (<I>,A). This four dimensionalgauge potential will be referred to as the Maxwell field.

Worked exercise: 1.5.1 Problem: Use the relativistic formulation to re-examine the introduction of potentials, the equation of continuity and the equation of motion for the gauge potential.

1.6

THE ENERGY-MOMENTUM TENSOR


As a more exciting thing we will discuss the energy-momentum tenAs a preparation we consider a system of N qn0

sor of the Maxwell field. particles with position The charge density

~n(t) and charges p is defined as:


3 -+- -+-

and the current

p(t,~) = L qno (x-x n (t) ) n is defined by:


3(t,it) = L qno (x-xn(t n -tinto the four-vector: J
3 -+- -+-

d-+xn

at

We may collect

and

(For Ct = 0 we remember that xn = t). This is not manifestly a Lorentz invariant but we may rearrange it.

23

To prove the Lorentz invariance of metrization of the the proper time.


JCL :
particle-trajector~es

we use the invariant paraxn


CL

(T)

where

is

Using exercise (1.4.1), we rewrite the expression for

(1.34 )

dt
dT

I t=x~ (T)

The last expression is obviously Lorentz invariant.

Observe that we

may deduce the equation of continuity directly from the definition:

Hence, we have obtained the conservation of charge without using the Maxwell equations. with fixed charges. Each of the P~(t) = (En(t),Pn(t. Hence, the density of energy and momentum is given by: particles carries a four-momentum given by This was entertainment. Now we must go to business. But of course, it is not surprising in this very simple model where we have a finite, but fixed, number of particles

T~~CH (t,~)

=I

p~(t)03(~-~n(t

and the current of energy-momentum is given by:

We may unite them into a single formula:

(1.35)

CLf! ... T (t,x) MECH

24

The important thing is to recognize that this is a tensor. we rewrite the expression for

To see this

T~~CH

CLS -.. T (t,x) MECH

L
n

i.e.
(1. 36)

which immediately shows that it is a tensor.

The components of this

so-called energy-momentum tensor have the following meaning:

(1.37)

[
~~ +

~~E_R~Y_ ~E~~I~~ ~ ~ ~~E~G_Y C~~R~~T 5-1


: _ _ MOMENTUM DENSITY
I

MOMENTUM CURRENT "STRESS-TENSOR"

Now we want the total energy to be conserved.

Hence, we expect the

energy density and energy current to satisfy an equation of continuity:

V.

5 =

~ doToO + djToj = 0 ~ aSTo S

=0
Hence, the

In the same way the total momentum should be conserved. satisfy an equation of corttinuity:

density of the x-component and the x-component of the current should

etc!

Consequently, the conservation of the total energy and momentum

can be expressed by an equation of continuity, satisfied by the energymomentum tensor:

(1.38)

25

Returning to the system of

N particles we get:

't
o

o[ IpCl(t)o
n n

(~-~ (t
n

3 ->-> dx n0 + - . LPCl -dt (x-xn(t

oxl.

o{

i
]

pcl

(t)~
dt

d i

[_0_. 0 (xLxjn oxl.


3

(t

1
J

The last two terms kill each other and we have shown:

(1. 3 9 )

If the particles are free, immediately:

i.e. they are not charged and they do not

experience any forces, etc., then

P~(t)

are constants and we get

(for a system of free particles!) However, if they are charged, they will create an electromagnetic field and then experience forces. In this case we get:

'" 0
and, of course, this is not surprising because we have only counted the
mechanicaZ part of energy and momentum and not the part of it stored in

the electromagnetic field!

Nothing can prevent mechanical energy from

being converted to field energy and vice versa! Now, introducing the energy-momentum tensor of the electromagnetic field

T~~, the total energy-momentum tensor may be decomposed:

and the conservation of energy and momentum ,requires:

26

Since we know

CLS dSTMECH ' we may use this to determine dpCL


n
3
~

CLS EL

=- L
n
CL

0 (X-Xn (t) )

-+

- I -aT
n

dPn dT 3 -+- -+dt 0 (x-xn(t dx S n dT 3(-+- -+- ( S-aT dt 0 x-x n t

- L qn
n

FCL

(due to the Lorentz force)

i.e. we have shown


(1.40)

We have got to rearrange the term on the right hand side.

Worked exeraise: 1.6.1 Problem: Show that Maxwell's equations imply the formula FCL JS = 3 [FCL FYS + !nCLS FYOF } S B Y 4 YO
From this exercise you immediately read off:
(1.41)

Let us summarize the results: We CLS CLS have introduced two energy-momentum tensors: T MECH and TEL' They obey the equations (1.39l and (1.40)

This was a very long computation!

For a system of free particles this immediately implies: CLS dSTMECH because
= 0

pCL (t)
n

are constants.

In a similar way a free electromagnetic CLS dSTEL

field obeys: because JS

for a free field.

Finally, if we have a system of

charged particles interacting with the electromagnetic field, we get:

27

because the total energy-momentum tensor


d T~S

T~S

Exercise 1.6.2
Problem: Show that and are symmetric tensors, i.e.

When we know the energy-momentum tensor of the electromagnetic field, we can especially find the energy density and the momentum density: TOO =-FO F YO cY

E(t,X)

~ (FYOF

yo 1

E2 + 1 4 (2S2_ 2E2) =

1 (E 2 +
2

a2 )

E
i.e.
(1.42)

where we have re-introduced

EO

and

c.

We will finish with some remarks about energy-momentum tensors in general: Let us consider the conservation laws. implies conservation of charge and that tion of energy and momentum. We have seen that

dST~S

a;~ = ~

implies conserva-

Actually, they imply something more:

Theorem 2 {Abraham's theorem} (a) If Ja(x) is a field of four-vectors, then

d~Ja

=0

implies

that: Q

JOd x

is a Lorentz-scalar, i.e. it is independent

xO=tO of the observer. (b) If

T~S (x)
th a: t

is a field of Lorentz tensors, then

aST~S = 0

im-

,. p"~es

p~

--

JT~Od3X

trans f orm

as t h e components of an

ordinary four-vector. This will be proved in chapter 11, section 1.

28

Observe that in the cases we have been discussing


tria (See exercise 1.6.2)

T~6

is symme-

This has an important consequence.


M~6Y(x)

Consider: - x6T~Y

= x~T6Y

(Observe that

M~ey

is not a tensor because

x~

is the coordinate of

point, and not the component of a vector!) Taking the divergence; we find:

M~6Y Y

o~ T 6Y + X~(d T 6Y ) - 06T~Y - x 6 (a T~Y) Y Y Y Y

= T6~

- T~6 is For

Hence, the symmetry of conserved! instance: Now

T~6

implies that

J~6(tO) =

M~6od3~

J~6

is the four-dimensional angular momentum!

2 f1

( j ko xkT jO ld3~ ijk x T -

where

is the momentum density. Since we expect this to be conserved for a closed system, we must demand that the energy-momentum tensor is
symmetria.

... g

SOLUTIONS OF WORKED EXERCISES.


No. 1.4.1

Suppose

g(x)

is a monotonous function and consider:


+<0

f fix)
x=-oo

o(g(xdx

f(x) o(g(x

dg(x)

x=- g' (x)

where g(x o ) = 0 ] and g is increasing


f (g_1 (u
u=-oo

O(u)du g' (x
) > O

g' (g-l (u

f(g-l(O g' (g-l(O g

~;[
-g' (x o)

where g(x o) = 0 ] and g is decreasing g' (x )

If

is increasing, then

0, and if

is decreasing, then

Hence, we may collect the above in the following formula:

<

O.

x=-

fix) o(g(xl)dx

f(xo)
= ----"-

Ig' (x o) I

29

No. 1.5.1 Some time ago we solved the Maxwell equations:

~ B=
with the ansatz:

dB + , dt

~ x E = -0

dt Now we have reformulated these Maxwell equations as

E=

?4J _

aA

o=

daFay

daFya

dyFaa

and we solve them with the ansatz: Faa = daAa - dSAa We can easily check this by explicit computation: daFBY + 0aFya + dyFaa da(daAy-dyAa) + da(dyAa-daAy) + Oy(daAa-daAa)
(dada-dada)Ay +

{ayda-dady)Aa + (dady-dyda)Aa = 0
(dada = dada)

due to the fact that partial derivatives commute From the two remaining Maxwell equations
~

E =L.
0

we deduced the equation of continuity:


J dt Now we have comprised these Maxwell equations into a single equation:

7-

Taking the divergence, we get: but


in
d~dBFaa = 0 because d d is symmetric in aa while Hence, we recover th~ equation of continuity:
d
~

as.

FaS

is antisymmetric

= 0

Finally, we may deduce the equations of motion for the Maxwell field: Aa. Substituting Faa = daAa - dSAa in the Maxwell equation, dSFaB = J~, we get:
d (daAB_dSAa )

= J~ ~

o~(o AS) (ODoS)A~ 6 ~

= J~

If we choose the Lorentz gauge where

daAa = 0

these equations simplify to:

30

No. 1.6.1
Fa Ja = Fa a FaY 8Y 8

(1.28)

a [Fa F8Y ] Y B

[a Fa ]F8Y Y 8

The last term is a mess and we have got to rearrange it separately:


INDEX - GYMNASTIC:

a Y a8 BY [ayF 8]F = [a F ]FBy

~(aYFa8)F
2 1 Y -(a Fa8)F

By

+ ~(aYFa8)FQ 2 ... 1 __ fJy _ _ __

8
Y

~ ~

1" ~ -(a 8Fay)F 8Y 2 1 ..._Y_I!_ _ _ _ Fay = _Fya


. , r - - - - - - - Fy I! = -F 8Y

~(aYFa8)F8Y
2

~(a8FYa)F8Y
y

l[aYFal! + a8Fya]Foy = - l[aaF8Y]F

L-

"

(1.27)---..!

This was the complicated part of the computation.


aoTal! = - FaoJ8 =-a [Fa F8Y] fJ

Now we are almost through:


8

EL

fJ

~ a [naY'F8oF u .]
Y

=~ay[Fa8FBY

n'JYF8oFao]

In the last expression we interchange the dummy indices

and

and get:

31

ch"pter Z INTERACTION OF FIELDS AND PARTICLES


2,1 INTRODUCTION
It is as we have seen possible to introduce a gauge potential describing the electromagnetic field, and this potential

A~ ~ (~o,Ao)
(1.

has the peculiar property that the gauge transformed potential


31)

describes the same electromagnetic field and satisfies the same equation of motion. Now a classical particle interacts with the electromagnetic field in the following way: tion of the particle. irrelevant. Hence if there is a space-time region where The field produces a force and this force deWhat is going on at other places is completely pends only on the value of the field strengths at the momentary pOSl-

... E

and

... B

... B

... B

Bok

vanish, a

classical, charged particle will experience nothing. But we have seen that even if there are no field strengths in a space-time region region.
Q

there may be a non-

trivial gauge potential

A in

this

Here non-triviality For

...=...:::.lL~B a 2 A x 2 +y2

[-y] ~ x IQIT..!)
1111/

(L.!JJ..?
Fig. 13

, I

means tha t the gauge potential cannot be completely gauged away.

0 ~

instance, we have seen that outside a sOlenoid there is no magnetic field but there is a non-trivial gauge potential:

A=~~[-~] 2 x 2 +y2 0
Thus you might ask: Does there exist physically measurable effects Can we distinguish between the situaassociated with the motion of a charged particle completely outside the solenoid? In other words: tion where there is no current in the solenoid and the situation where there is one, just by letting a charged particle go around the solenoid?

32

Clearly, if there is such an effect it cannot be a classical one, because a classical, charged particle will experience no forces in either of the situations. However, there could be a quantum mechanical effect: To study this possibility we are faced with the problem of how to quantize the physical systems above. If the fields are strong so We treat them as that the particle can be regarded as a test particle, then we are allowed to keep the fields on a classical level.
external fields.

However, the motion of the particle should be


~=~(t)

quantized.
~(r,t)

Hence instead of characterizing the particle by its claswe introduce a Schr5dinger wave function At the time t the abis the probability density for finding the
-+

sical trajectory solute square tion of motion:

with the well-known interpretation:

I~(r,t)

12

particle at the position


:;.

r.

Instead of the usual Newtonian equa-

where the dots represent] [ time derivatives the dynamical evolution of the quantum-mechanical system is governed
(2.1)

m x

by the Schrodinger equation:


(2.2)

(fi

= 2~

h = Planck's constant)

~ 11 at

a~

= (-

112

2m

~ + V(r)) W(~,t)

Although you probably already know the Schrodinger equation when there are electromagnetic fields present, it will be rewarding to examine the quantization procedure carefully.

2,2

LAGRANGIAN FORMALISM FOR PARTICLES: THE

NON~RELATIVISTIC

CASE

The first step wheq you want to quantize a system consists of casting the problem into the Lagrangian form. ClaSSically it is a compact way of deriving the equations of motion. Consider for simplicity a one-dimensional motion of a particle, say along the x-axis. We want to consider those motions of the particle where it moves from the space-time point The question is: A: (t1,x 1 ) to the space-time point B: (t 2 ,x 2 ) How can we find the classical path leading from L(x,x) A to B of two variables. we now associate the

A to

B?

First we will introduce a function later on. With each path leading from

It is called the Lagrangian, and the explicit form will be derived

number:

33

(2.3)

called the action associated with


t-axis
SPACE-TIME-DIlIGRAM B (t ,X )

that path.

We want to choose the

path x = x(t) in such a way that the action has an extremum. Suppose x

xo(t)

produces such

an extremum and consider another


t

path: x Fig. 14
x-axis

= xo(t)

+ Ey(t) ; y(t l )
8

= y(t 2 ) = 0
We

We will allow

to vary but

keep y fixed for a moment. know that the action t2 S (8)

L (x

(t) + E'y(t) '*0 +

y)dt

tl which may be considered a function of E E, has an extremum for

0 . Consequently

t2

o = s'

(0)

[~~

yet) +

a~~)

y(t)] dt

tl

aL [ ax
t

y(t) - dt ~

aL

y(t) ] dt

J2 [~Lx
o

_d dt

0(*)

aL]

yet) dt

where we have used the boundary conditions: 0 = y(tl) = y(t 2 ) . However yet) was an arbitrarily chosen function. Therefore the above identity is only consistent if x = xo(t) satisfies the following differential equation known as the EuZer-Lagrange equation

(2.4)

34

Illustrative example: One-dimensional motion in a potential V(x) , Let us try to determine the Lagrangian in such a way that the Euler-Lagrange equation aL d aL (2.4)
reproduces Newton's equation of motion:

d (m:ic) dt You innnediately read off the following relationships aL av L(x,x) (TERMS INVOLVING x) - Vex) ax = - dX ( 2. 1)
- ax
=

av

mx

..

aL = mx ax i.e.

L(x,x) L(x,x)

!m(x)2 + (TERMS INVOLVING !m(x)2 Vex) + CONSTANT

x)

Consequently we have determined the following suitable Lagrangian (2.5) L(x,x) = !m(x)2 - Vex) The above scheme is easily generalized to more complicated
Pl\RI'ICLE 2

systems.

Suppose we have a

system characterized by the folCENTER OF MASS: (x, y, z) = (ql rq2,q


Z

lowing set of generalized coordinates: ql, , qn. If we have two particles for instance, then ql,q2,q3 could be the q 4 ,q 5 ,q 6 could Cartesian coordinates of the center of mass, and be the polar coordinates of the vector connecting particle Fig. 15 1 2. These generalized coordinates take their value in a subset ~ of the and particle
Q

PARTICLE 1

y
x

n-dimensional space qi

R n . We call

the configuration space. We now introduce a A: (tl,q~,

The

dynamical evolution of the system is described by a curve

qi(t)

in the configuration space. L(qi,qi,t)

Lagrangian

which may depend explicitly on time too. , qn)

Next we consider paths connecting the point with the point sociate the action:
S

B: (t2,q~, " ' , q~). With each of these paths we ast2

J L(q
qi

,q ,t)dt

oi

and we want to determine a path Again we suppose that

qo(t)

extremizing the action.

q~(t)

actually extremizes it and con-

sider another nearby path:

35

We will allow

Ei

to vary but keep


t2

y(t)

fixed for a moment.

Con-

sider the action,

J
tl as a fUnction of E1,""",E n
(0,'"",0)

i i i i. L(qo +E y, qo +E y,t)dt

It assumes an extremal value for

-r aE
t t

as

Therefore t z

J [~y aql
yet) dt

I (0,'",0) tl
q
1

aL i aq

Y]

dt

-~ ~] J zr~ La dt a"l
q
1

But this is consistent only if

(2.6)

o ,

ial ,

.. , n

Thus

each of the components satisfies the appropriate Euler-Lagrange Let us return to the case where a single particle moves in the

equation. three-dimensional space. tential If it moves under the influence of a po-

veil

then we can immediately extend the above analysis.

It corresponds to a Lagrangian:
(2.7)

'2 mv
dZi
dt
Z

-+-2

- Vex) Newtonian equation of motion

...

which reproduces the


(2.8)
m

well-~

But we are really interested in the motion of a charged particle in an electromagnetic field. tion
(2.9)
m

This corresponds to the equation of mo-

dZ~

dt

and this equation cannot immediately be reproduced by a Lagrangian

36

of the form (2.7) for two reasons: 1. The force may contain an explicit time dependence because the field strengths may depend explicitly on time. sider for instance the electric field (Con-

-+

in a capacitor
-+ -+

joined to a circuit with an oscillating current.)


2.

The force is velocity-dependent due to the

v B-term.
V(~)

Hence we can try to replace the simple-minded potential by a generalized potential on the form: (2.10) L(x,v,t)
-+ -+

U(x,v,t)

which may depend explicitly So we look for a Langrangian

both on pOSition, velocity and time.


1 2
-+2
-+ -+

mv

- U(x,v,t)

Substituting this Lagrangian into the Euler-Lagrange equations (2.6) we get


0
--::;:

aL

ax

d aL dt --::;: av

Le.

au
ax
-+

d -+ d dt (mv) + dt

au
av
-+

au
ax
-+

+ dt

au
av
-+

To determine the generalized potential we must rearrange the expression for the Lorentz force
(2.9)
m --2

d2~

dt First we introduce the gauge pctentials relations


(1.19)
-+

(<j>,A)

through the well-known


-+

-+

'V x A

...

-+

4-

'V<j>

--at

dA

giving: m --2 dt
-+

d2~

-+ -+ -+ -+ aA q(- 'V<j> - at + v x ('V x A

-+

Clearly, this is a mess and we will have to rearrange the -+ x A)-term. v x USing the standard rules of vector analysis we get

(V

37

-+ v x

V-

dA

-+ -+

ax

substituting this into the equation of motion we now obtain d2 m-2 dt

q(-

dA lP._ at: -+

-+

ax a ax

a ax
-+ -+

(v'A)

-+

-+

_ aA -+ dx -+) ax dt

q(-

-+

[<p

v'A]

-+

at

aA _ aA.d;:) ax dt

q(and we are almost through. A


-+

a ax
-+

-+ -+ dA\ [<P - V'A] - dtJ

Performing the trivial substitution: -(A'v-<p) a~

-:;:(A'V) av

-+-+

-+-+

which is permitted because

<P

does not depend on

-+

we get

-+ -+ ) v'A]

and from this formula we can immediately read off the generalized potential
(2.11)

U(~,~,t)

q[<P - v'A]

-+

-+

Thus we have found the total Lagrangian:

(2.12)

L(x,~,t)

1 2

mv

-+2

- q<p(t,r) + qA(t,r)'v

-+

-+

-+

-+

We may think of the above result in the following way: A: (t l ,Xl) and
B: (t

If we have

a system without electromagnetic fields, then to any path joining


2

,X 2 )
t

there corresponds an action mv


-+2

1 J (2

If we switch on an external electromagnetic field

(<P,A)

then we

have to add another piece of action,the interaction term:

3& t

\
)-------~----~y

The total action now becomes: S

So + Sr' We may rewrite the in-

teraction term in a more elegant

"'-A

\ i
Fig. 16 t2 Sr -q

manner. Let us denote the path joining the space-time points A:(tl,Xl) and B: (tz,xz) by f. We then obtain

tz + q tz

f ~dt
tz

f A.~~

dt
B

q[ f-~dt +
i.e.
(2.13)

Ad;]

f
A

[Aodx" + A.dx i ]
1

In the Lagrangian formalism this formula replaces the Lorentz force! Observe that in this formalism the particle interacts directly with
the gauge pctential and not with the field strengths. Now remember the

gauge symmetry: We are allowed to replace Au by Au + dUX without changing any physics. This is obvious if we look at the equation of motion:

because the field strengths are gauge invariant! But does there exist a quick way to see it in the Lagrangian formalism? What we want to show is that the form of the Lagrangian immediately implies that the motion of a particle is usaffected by a gauge transformation. rt should be possible to see this without explicitly computing the equation of motion. The particle is following a path fo which extremizes the full action S

= So

+ Sr. Performing a gauge transformation we get

a new interaction term:


(2.14)

S'

qfA'udxU qfA dx u +
u f

+ q[ X

(B)

X (A)]

But the last term does not at all depend on the path f !

For any path

the action will therefore be shifted by the same amount. But then fo is still an extremal path and we have explicitly shown the gauge invariance.

39

2.3 BASIC PRTNCIPLES OF QUANTUM MECHANICS


We are now prepared for the quantization procedure! There is actually two alternative ways of quantizing a system: 1. The path integral technique (Feynman's procedure), 2. the canonical quantization. The first one is the procedure which serves our purpose best, so we will start with that following Feynman [1964] . The fundamental concept in quantum mechanics is the probabitity amplitude. Let us assume that we have an electron gun A emitting electrons, a screen B with two holes, and a screen C with a detector that can measure the arrival of electrons

~ ::..-:?'>-------~---------~,-)....~

~ --~.
~,,-

PIx)

-------'--

Fig. 17

~.
P(x) = I~ (x)

An event is the registration of an electron in the detector at a specific position x. With each event tude
~

we associate a probability ampli-

(x), which is a complex number, and the probability P(x) for

the corresponding event is found by squaring the amplitude (2.15)

,2

Furthermore, the electron may have passed through either of the holes 1 and 2. Hence we have two alternative ways in which the event can occur. Each of these is characterized by a probability amplitude
~,

(x) and

~2

(x). The total amplitude is found by adding these two


~

(2.16)

(x) =

~,

(x) +

~2

(x)

When you apply this principle it is important that we have no way of veryfying experimentally, which of the alternatives actually occured. If we make the experiment in such a way that we can decide which of the alternative trajectories the electron followed, then interference is lost, and we see no wave pattern! (See fig. 18).

40

~-----Fig. 18

----r.===
~.

~\V _---1 and 2.

PIx)

Now consider an experiment with polarized light. We have a beam of photons, which has passed through a polarizer O. In the direction of the beam we have put two more polarizers

DErEX:'IDR

=s

0
2
21

Fig. 19

For a given photon there is a probability amplitude there is a probability amplitude

lO

that it

will pass through polarizer 1, and once it has passed through that that it will pass through pola2. This amplitude
in~

rizer 2. We are interested in the situation where a given photon actually hits the detector after passing polarizer
20

is found by multiplying the amplitudes corresponding to the

dividuaZ steps. (2.17)


Observe that this time there are no alternative ways to arrive at the detector. There is through polarizer 2. only one way, which consists of two steps. First the photon has to pass through polarizer 1, and then it has to pass

We have now stated all the fundamental laws of probability amplitudes. They are basic laws, i.e. you cannot derive them from some underlying principles. We simply have to accept them as a starting point for the quantum mechanical description of a system. For later references we have collected them in the following scheme:

41
THE BASIC PRINCIPLES OF QUANTUM MECHANICS:

1.

Each event is associated a probability amplitude event is found by squaring the amplitude:

~.

The probability of the

(2.15)
2.

If an event may classically occur in several alternative ways i = l, . ,n each of which is characterized by the amplitude $i' then the total amplitude of the event is found by addition:
n

(2.16)
3.

$ = l: $.

i=l

If an event occurs in a way that can be decomposed into several individual steps j = l, ,m, each of which is characterized by the amplitude $., then the total amplitude of the event is found by multiplication: J

(2.17)

m $ = IT $.

j=l J

Now we are prepared for the discussion of a system consisting of a single particle. Consider two fixed times t , and t2' At the time tl we have a probability amplitude
-+ -+ ~(r,tl)

of finding the particle


-+ ~(r, -+ ~(r,

at ;. At the time t2 we have another probability amplitude ~(;, t2)


of finding the particle at r. The amplitude
,

t) is of course t) develop in time?

nothing but the Schrodinger wave function. The dynamical problem we must solve is the following one: How does t We will solve it by finding a formula which connects ~(;, t) at times

t,

and t

t2'
-+
-+

Consider ~(;2,t2)' It is the amplitude of finding the particle at r2 at the time t 2 . NoO! let K (!'2; t2 I !'1; t,) denote the p!'obabi lity -+ amplitude fo!' the event, that a pa!'ticle emitted at !', at the time
-+

t, O!i II be 'bb se !'ved at ;2 at the time t2. Then we can argue in the
following way. A particle which arrive at r2 at time t2 must have been somewhere at time t,: Hence there are several alternative paths the particle could have followed, each characterized by the position at time t,. Each of these possibilities is composed of two individual steps: 1. 2. The particle was at the point ;, at time t, The particle was emitted from r, r2 at time t2 ( amplitude
-+ -+

(amplitude = ljJ(~l,t) ).

at time t, , and observed at ).

K(r2;t2Irl;tl)

According to (2.17) each possibility is then characterized by the amplitude: K(r2;t z lr, it, t
2 ,
-+

-+

)~(r"t,).

-+

To find the total amplitude ~(;2,t2) for arriving at;2 at time we must sum op all these amplitudes (cf.

(2.16, i.e. we get!

42

(2.18)

This is the basic dynamical equation of the theory. Although it is an integral equation, we shall see later on that it is completely equivalent to the Sahpodingep equation (2.2).

2.4

PATH INTEGRALS - THE FEYNMAN PROPAGATOR


t-axis
Consider now the amplitude

,
I

7f

K(r,;t,lrl ;tl). It is called the

-+

-+

-,
K(B,A) I

Feynman ppopagator and if we can determine that, we will controll the dynamical evolution of the Schrodinger wave function! Now K(r, ;tllrl ;tl) is the amplitude for the event that a particle which was
-+

I
\

I \

-+

A(r, It,)
Fig. 20

... . r-axu;

emitted at rl
-+

at time tlwill be observed at r, at time t , . TO come from

-+

(rl,tl) to (r2,t2) a classical particle must have followed some path r. Consequently the classical particle could have rroved in several ways, each corresponding to a path leading from A: (;1 ,tl) to B: (;"t,). Let ~r(BIA) denote the amplitude that a particle, emitted at A and observed at B, actually moved along the path r. Then we get:
(2.19 )

K(BIA)

=
A to B.

where we sum over all paths r leading from

Obviously this sum is very complicated, since there is an infinity of paths! It is called a path integpaZ. The precise definition of such a sum is very complicated, and as we are interested only in the basic physical ideas we will continue writing the path integral in the above naive form! We have reduced our problem to that of finding
~r(BIA).

Now this

cannot be deduced from basic principles, so we must simply postulate the value of it. This is where we make contact with the Lagrangian formulation of classical mechanics. With each path r we hava associated a classical action:

43

S(f)

f L(x, a:F' t)dt


tl

t2

-+

d;

Following a suggestion from Dirac, Feynman now used the folloving expression (compare Dirac [1935], 33, p. 125-26)

(2.20)

exp[~
path~epresents

S( r)]

Since each

a possible history and each history

contributes with a phase factor, a complex number of modulus 1, this is often referred to as Feynman's principle of the democratic equali-

ty of all histories!
This prinCiple leads to the following formula for the propagator:

Now, this is a very complicated formula and you might wonder what happened to classical physics? In classical physics you are used to considering only one single path, the path which extremizes the action. In the above formula all paths are on the same footing. There is no special reference to the classical path. But how can the classical path then play such a major role in our everyday experiences? To understand this we must first try to characterize what we expect to be a typical classical problem, and what we expect to be a typical quantum mechanical problem. First you might be tempted to say that a typical classical problem is one where the involved actions are very great compared to h, and a typical quantum mechanical problem is one where the involved actions are comparable to h. However, this is not a good formulation because the action is not at definite number. It is only determined up to a constant and hence the action for at single path has no absolute meaning. But if we take two paths ence
~S

rl

and r 2 , then the differ-

= S(r2)

- S(rl) has a unique meaning. Therefore we are allowed

to say:

44

t-axis

CUlSSICAL

P~1

Fig. 21 x-axis x-axis

A typiaal alassiaal ppoblem is a problem whepe small ohanges in the path oan produae a great ohange l:.s aompared to path only produae a change l:.S aomparable to
Ii. Ii, and a typiaal

quantum meohaniaal problem is one where eVen great ahanges in the


Suppose we consider a classical situation and pick an arbitrary path r,. It contributes to the propagator with a phase factor

exp[~ S(r,)]. But very close to the path r, there will be another
path r2
,

which differs in the action by rrli, but then the correspon-

ding amplitudes will cancel each other, because exp[ Ii S ( r 2


i
= - exp[

)]

Ii S ( r, ) ]

Thus they do not contribute to the general sum due to the destructive interference. Although this is the general situation, i t is not always true. Let rc I be the classical path, then S (r C I ) is extremal and all the nearby paths will have almose the same action. Hence we will have constructive interference from all the paths close to the classical one! Since
~he

main contribution to the propagator co-

mes from paths near the classical path, and since they all have approximately the same action as the classical path, we can put as a first approximation:

(2.22)

L(~C I , dxc I ,t) dt]

crr-

This is known as the alassical approximation.

45
t-axis
B

As we have seen the main contribution to the propagator comes from a "strip" around the classical path where the action varies slowly, and where the changes in the action are smaller than rrh. For a typical classical problem this strip is very "thin", but for a typical quantum mechanical problem the strip is very "broad". ConsequenUy the classical path looses its significance in ty-

REl3ION WITH CONS'l'RUCI'IVE


INl'ERFERENCE

CUlSSlCAL
PATH

Fig. 22

x-axis

pical quantum mechanical situations, sayan electron orbiting around a nucleus. The path of the electron is "smeared" out.

Worked exercise: 2.4.1


Problem: Consider the quadratic lagrangian:

Show starting from (2.21) that the is given by the formula:

Feynman propagator t)dt l

(2.23)

K(X2,t2; XI,ttl =A(t 2 ,ttl exp[~

. t2

t,

L(xCl

,XCI;

(2.24)

Assume now that the coefficients a(t), bit) and c(t) do not depend on time, i.e. they are constants. Show that the Feynman propagator then is given by the formula . t2 K(X2 ,t2 ,Xl ,t,) = A(t2-t,) exp[~ f L(xci ,X" ,t)dt l

tl

In both cases A (.) denotes an unknown function and Xc I denotes the classical path.

The above quadratic Lagrangian includes for instance the following cases:
(a) (b)

A free particle: A forced oscillator:

The harmonic oscillator: L


L

(c)

1sIluF 1snci' -i II1W 2 X2 1smx2 -i mw 2 X2 - '-it) x

Before we use fOnmlla (2.24)

let us point out another feature of

the propagator. By definition K(X2 ;t2lx l ;t l ) is the probability amplitude for finding the particle at the pOint X2 at the time t2 when we know that it was emitted at XI at time t,. Let us keep XI and tl XI ;tl) as a function of Xl fixed for at moment and regard K(X2 ;t21

46 and t2: K(X, ,t,) (x,t) = K(Xitl x, ;t,)

Then K(x, ,t,) (x,t) is simply a Schrodinger wave function since it denotes the amplitude for finding the particle at x! But it is a Schrodinger wave function corresponding to a very special situation, because at the time t, we know exactly where the particle is: It is at x = x,. The amplitude is not smeared out at t = t,! This can also be seen from the integral equation (2.18) wave function
1jJ (X2 ,t 2 )

for an arbitrary Schrodinger

= fK

(X2 ;t2 I x, ;t,) 1jJ (x, ,t l ) dXl

If we put t 2 = t, we simply get


1jJ (X2 ,tl)

= fK (X2 ;t, I x, it,) 1jJ (Xl ,t l ) dXl

But then we can immediately read off that

Hence K(x, ,t,) (x,t) reduces to a 6-function at the time t,: K (x, ,t,) (x,t) = 6 (x-x') Since K(x"t, ) (x,t) itself is a Schrodinger wave function it must furthermore satisfy the integral equation (2.18). We have thus deduced the following important group property of the propagator: (2.25)

2.5

ILLUSTRATIVE EXAMPLE: THE FREE PARTICLE PROPAGATOR


=

We now are in a position where we can calculate the propagator for a free particle. Here L(x;x) ~mX2 and we can use exercise 2.4.1 about quadratic Lagrangians: K(xz ,tzl x, ,t, ) = A(tz-t,) exp

[li- f

. t2

~ x ~L dt]
(xz-x,)z
tz-t,

t,

(2.26)

= A(tz-t,) e vn [i m
'0",

Ii 2"

(The classical path is a straight line, i.e.

is a constant!)

47
To finish we must determine the function A. Substituting (2.26)-into the intergral equation (2.25) we get: 2 A(t3-t , )exp

i. ~
h 2

(X3- X' ) = t3-t,

Using the

alge~raic form~a:

(Xl-X2) + (x,-x,) _ t3-t, [X2 _ Xl (t2 -t, )+Xl (t3 -t,)] + (X3 -x, ) t3 -t, ~ - (t, -t2 ) (t, -t, ) t3 -t, t3 -t,
we may rewrite the last integral:

X2

=+00.

Xl=_CO

[1 m (X3 ':"x,) + exp ft2 t3 -t,

(X2 -x,) 1 dx2 t, -t,

where we have introduced

to get rid of a lot of junk. The total integral formula is thus reduced to: im(x3-x,)2] A(t,-t, ) exp [ ft2~
u=+oo

) exp ( ) ( At3-t2At,-t,

[ i m (x3-x,)']

F2~ exp u=-oo

[i m (t3-t,) 2 ft2(t3-t 2)(t2-t,)U du

Notice that X2 has completely disappeared! The integral does not depend on X3 and x, so both sides have the same dependence on X3 and x" Everything fits beautifully. Using the formula:

x=_
(2.27)

f exp x=_oo

2 [-ax ]dx

=/f
h1lih(t3 -t, )
m ,;",----"",mr.----:--,-

which may be used for complex a too, we finally get: A( t3 -t, ) _ / 211ih (t3 -t2 ) (t2 -t,) _ A(t3-t,)A(t2-t,) m (t3-t,) - /

\r211ih ( t3 -t,)

211ih (t, -t, )

From this we see that,.up to a trivial phase factor exp(iat),which we normalize to 1, the function A(t) is given by A(t)

= ~1T~flt
~ee-partiaZe p~agator:

So now we have computed in all details the

48

E~raise

2.5.1 .. Problem: Show by an expl~c~t calculation that the propagator of a+free particle reduces to a 6-function in the limit where t ~O .

Let us try to get familiar with the free-particle propagator. We know that K(X,tlo,o) is the Schrodinger wave function of a free particle emitted from x = 0 at the time t = 0: (2.29) 1/J(x,t) K(x,t /0,0) = ~ 'exp

.rrll

rimx2]

f!"2 t

Let us focus the attention on a specific point (Xo ,to). If the free particle is observed at x = Xo at time to, we would say that it classically had momentum po= mvo= m Xo to and energy Eo= ?;mvo = ?;m
If
2

x~ t1

we investigate the variation of the phase (Xo; to) we find: 1/J(x,t) = ~ exp r~ ~ ~

mx 211 t

in the viscinity of

rrn-

]
+

a-t(t Ht-to)
Ixo ,to

x2

+ ... ]}

m 21Ti t exp ~

[~

.f!

r X2 ."'to' x
[~

'-- Xo ] .,m tr t

Thus, very close to (Xo,to) the wave function varies like


(2.30)

1/J(x,t) ""

~1T~ht

exp

[Po x - Eot] for (x,t) "" (Xo ,to)

This is closely related to the Einstein-de Broglie ruZe, according to which a particle ~ith momentum p and energy E is associated a wave function with wave lenght A = and frequency V = E/h:

exp [i

(2~X_~)] = exp [~(pX

- Et)]

While 1jI(x,t) represents a particle which is located at x = 0 at the time t = 0, we would also like to have a Schrodinger wave function representing a free particle with specific energy and momentum. At the time t = 0 we suggest the following amplitude: 1jI(x,0) = exp

[k px]

where p is a constant. At a later time t we can compute 1jI(x,t) using the integral equation (2.18):

49

I/I(x,t)

i m (X_X )2] [Ii "2 - - t - exp [~xll w dx l

Using the algebraic identity: pxl+

~ (X_~l

)2

~~ rxl-(x_~l]2

+ px _

we get:
1/I(x,t)

Using the substitution: u 1/I(x,t) = exp

= Xl_ (x-~) m
2

this may be reduced to:


[i m

{fi

[px - L

2m

t]}

V 2~ilit I

~ u=_
u=-oo

exp

11 2" u ldu

. z exp {[ [px- ~ tl}

Thus

the free particle is represented by a solution of the form:


. z exp{ 2:. [px - E- t]} h 2m

which according to Einstein-de Broglie's rule must be interpreted as a particle with the momentum p and the energy:
2

2m Let us summarize: We have played a little with the free propagator and discovered two exact wave functions:
(2.31)
1/1 ( x, t)

= ..; 2lTilit
.

. z exp [[ ~ ~ 1

which represents a free particle, strictly located at x =


(2,32)

at the time t

0, and

l/J(x,t) = exp [

fi [ px -

~tll

which represents a free particle with a specific momentum p and specific energy:
E =

~ 2m

2.6

BOHM - AHARONOV EFFECT

- LORENTZ FORCE
The motion of charged particle Consider a path

Having described the basic principles of the quantization procedure we will return to our original problem: in an external electromagnetic field. A: (xl,tll and

connecting

B: (x ,t ). Before switching on the external field 2 2 this path contributes with the phase factor

~Or(BIA)

= exp

[~ Itz ~
fl
tl

mV 2dtl

50

to the propagator, but after we have switched on the electromagnetic field we must add to the action the interaction term: (2.13 )

=
=
exp[* So(r}]exp [* SI(r}]

Consequently the phase factor is changed into:


<Pr(BIA)

exp[*[So(r} + SI(r)]]

<P~(BIAH>~(B

IA)

SO

you see that the external field produces a change obtained simply

by multiplying the original amplitude with the gauge phase faatop: {2.33}

<P~(BIA) = exP[iq

Ao.dxo.]

This is the quantum meahaniaal law, which replaaes the Lorentz forae!

Now we are ready to give a qualitative discussion of the slit experiment. We first look at a situation where there is no external field. (see fig. 23).

Fig. 23

If the electron gun emits electrons with a characteristic momentum p, then the electron wavefunction has the de Broglie wavelength where A A,

hlp.

Construct~ve

interferences are expected when the dif-

ference between the two pathlengths is an integer multiple of A. If the difference between the pathlengths is a, then the difference of the phases in the electron wavefunctions is
l!.0 =

I . 2'IT
A, where L is the distance between slit

But if we assume that

and screen and d is the distance between the two holes, then simple geometric considerations show that

ta x g = L
(see fig. 24).

an d tg a-ill - d/2

a. =d ~.e.

1 L

g,

51

(1)

2
d

(2 )
Fig. 24 Consequently we may replace difference at the point a by x, thus obtaining for the phase

on the screen:

(2.34)
If When etc!

nO(x)

xd

LA 2~
nO nO

x x

is zero, then

= =
n

0, and we have constructive interference.


,

2d '

AL

then

and we have destructive interference,

Thus we have obtained a qualitative understanding of the wave

pattern on the screen! Now suppose there is a pure static magnetic field between the slit and the screen. We know that this will produce a change of the phase facx (Gee fig. 25).

tor associated with a "classical" electron arriving at the point

o o
G)
0
Fig. 25

0 0

PIx)

0
0

(Strictly speaking we should use the classical trajectories which are circles. If the field is not too strong and the momenta of the electrons are sufficiently high, then the radii of the circles will be great compared with L and we can safely replace the arcs of the circles with straigth lines).
Now the main contribution to the phase comes from the paths and
f2

fl

shown on fig. 25.

Let us look at
phase

rl

Before switching
fl

on the external field, a "classical" electron !lOVing along

wili be associ-

ated with a phase factor

exp[iarl, i.e. a

a1

52 But according to (2.33) switching on the external field will change this phase into

(Observe that phase:

0 , as there is no electric field.)

In a similar way

a "classical" electron moving along

r 2 will

rcw be associated the m::xiified

Therefore the phase difference between the electron wavefunctions is given by: (2.35) Here

nO

is the phase difference, when there is no external field

present, and it was calculated in (2.34). The loop integral is found by integrating forwards along the path But since we are in a static situation magnetic flux enclosed by the are looking for. We will "use" a very special kind of magnetic field. Under certain circumstances iron crystals can grow in the form of very long, microscopically thin filaments called whiskers. These whiskers can be magnetized and they then act like very tiny solenoids! Hence we have a purely static magnetic field concentrated in the whiskers (the field outside is extremely small and we may neglect it), but even though there is no magnetic field outside the whiskers, there is a nontrivial 'gauge potential

r2
loop!

and backwards along dr

r1

p A.

is nothing but the

r2

r1

This is the formula we

it.

This is exactly what we are after.

D
Fig. 26

53

Now we can show that there is a measurable effect when you let electrons move outside the whiskers. Phase factors corresponding to paths lying on one side of the whiskers interfere with phase factors corresponding to paths lying on the other side of the whisker, and at the screen, the whisker (i.e. the gauge potential), has theref=e produced the following change of phase (2.36) Here <I> is the magnetic flux through the whisker. Cbserve, that the phasedifference is independent of the position x on the screen. Hence the total wave pattern On the screen will be shifted a constant amount! This effect is called the
Bohm-Ahcmmov effect. (See fig.

27).

-t-52;;~~~P (x)

----

Classically you can only find out whether or not there is a whisker if you hit i t directly. But quantum mechanically you can detect the whisker because the gauge potential changes the phases of electrons passing by. This remarkable effect was predicted by Bohm and Aharonov (Bohm-Aharonov [1959,1961]) and later verified experimentally (Hollenstedt and Bayh [1962]). D.Bohm and Y.Aharonov, "Significance of electromagnetic potentials in the quantum theory", Phys. Ref., 115 (1959) 485 and - "Further considerations on electromagnetic potentials in the quantum theory", Phys. Rev. 123 (1961) 1511. G. Mollenstedt and W.Bayh, "The Continuous variation of the phase of electron waves in field free space by means of the magnetic vector potential of a solenoid", Phys. Blatt. 18 (1962) 299.
You might still feel uncomfortabJ e about the whole analysis. We have studied the quantum mechanical interaction of charged particles with the Maxwell field and seen that it produces strange effects. We will now derive something wellknown. We will show that this theory reproduces the Lorentz force too. For simplicity we consider only a very simple situation. We consider the slit experiment once more, but this time we put a homogenous magnetic field behind the slit. This homogeneous ,static field is concentrated in a region of width W, where W L (See fig. 28). If we had no magnetic field there would be a phase difference:
(2.34)

54

D
Fig. 28

P(x)

for electron s arrlvlng at the screen . But the presenC e of a magnetic field will produce a phase shift proport ional to the magnetic flux. In our case the magneti c flux is approxim ately BWd. Therefo re the actual phase differen ce is given by: A "

~~

21f +

B Wd

The point of maximum intensit y will be determin ed by:

o " xd 21f LA

+ fiq BWd

i.e.

x = _ EWALq
211'h

classic al limit we would interpre t this in the followin g way: We would see electron s tra velling through the magnetic field shift their directio n of movemen t. Consequ ently they must have experien ced a force. As the electron s have the velocity
v =

I~the

.E

Fig. 29

they experien ce the magneti c field during the time: At =

!i v

If t denotes the time it takes to travel from the slit to the screen, then t is given by:
t

=L
v

As they hit the slit at the point x = _ B~~ the passage through the magneti c field has produce d a velocity in the x-direc tion given by: Av = ~ = - BWq x t

55
Consequently, we would say classically that the electrons experienced a force in the x - direction given by
F

At

= - Bvq

Ap

Using Einstein - de Broglie's rule

A
we finally get:

=~ p =qvB

Fx

=q

(-;;xB)x

and this is exactly the Lorentz force!

2,7

GAUGE TRANSFORMATION OF THE SCHRbDINGER WAVEFUNCTION

Let us now return to the general theory! We have seen that we can represent the electromagnetic field by the gauge potentialAa and also discovered the gauge symmetry: Physics is unchanged if we replace the gauge potential Aa wi th the new potential

Ah = Aa + a(XX Now, how do these ideas fit together with our quantum mechanical
description of a particle ineracting with the Maxwell field Aa? Consider a specific path r
1

in space-time connecting the space-

time points A and B. We have seen that the external field Aa produces a change in the original amplitude corresponding to multiplication with the gauge phase factor
(2.33)

but the interaction term S1 is not gauge invariant! If we perform a gauge transformation, it is changed according to the rule (2.14) S1 (rJ ~s' I (rJ
=

S1 (r) +q [X (B)

-x (A) ]
is not gauge invariant, but

Consequently the gauge phase factor transforms according to the rule:


(2.37)

~r(BIA)

But this has consequences for the propagator! We know that the propagator is given by the path integral formula K(BIA)=

~r(BIA)

D[r]

We can decompose the phase factor <Pr(BIA) into two parts. One, ~\BIA) which describes the system in the abscense of gauge potentials, and one,

~~(BIA), which describes the interaction with the gauge potential. Co~

56

quently we get K(BIA)=

I ~~(BIA)~~(BIA)
=

D[r]

Performing the gauge transformation we then get the new propagator: K' (BIA)

J ~~(BIA)~rI(BIA)

D[r]
J

But the phasefactors exp[ do

o i.9. I ~r(BIA)exp[l1X(B)]~r(BIA)exp

[!SI - flX(A)]

X (B)]
I

exp[-

X (A) ]

not

depend on the path r

so we can pull them outside the

path integral: K' (BjA) = Thus we have shown:


(2.38)

eXP[irX(~)]{f~~(BIA)~~(BIA)D[r]}

exp[- irX(A)]

K(BIA) ... K' (BIA)

eXP[1rX(B)]K(BIA)exp[- irX(A)]

so that the propagator of a charged particle transforms in exactly the same way as the gauge phase factor! This has consequences for the Schrodinger wavefunction! The Schr6dinger wave function ~(r,t) satisfies the integral equation

(2.18)

and since we have seen that the propagator is changed under a gauge transformation we conclude that the Schr6dinger wave function must change too:

~(~,t)

~'(~,t)
~'

To find the gauge transformed wavefunction must satisfy the integral equation:

we observe that it

~'(;2 ;t2) =fK I (;2 ;t2 1;1 ;til ~ I (rlltl) drl

.... -+ + )] I (+ ) + ex P [ - !SI =f exp[ .!9:'" fl x(r2it.2)]K(r2it2Irlittl flx(r1itl ~ rli t 1 drl


The first phase factor exp tegration variable obtaining: exp[-

[iiI X (;2 ,t2)] does

not depend on the in-

1,

so we may pull it outside the integral, thereby

X(i'2it2)]~'

(r2it2)=

57

Comparing this with the integral equation for 1jJ we inunediately obtain: exp[-
i.e. (2.39 )

X(~,t)]1jJcr,t)=1jJ(r,t)

This is really beautiful! Remember that the only physically measurable quantity associated with the wave function 1jJ is the square of the norm: ;1jJ(;,t) I 2, but this is

aompletely unahangedby the gauge


the

transformation. Hence we see ti,at quantum mechanically too, all gauge potentials, differing by a gauge transformation, represent same physical state. You also see that when a charged particle interacts with an electromagnetic field, then the phase of the wave function:1jJ=I1jJiexp[i~] can be chosen completely arbitrarily at any space-time point!

2.8

QUANTUM MECHANICS OF A CHARGED PARTICLE AS A GAUGE THEORY

We have found a lot of formulas which describe the effect of performing gauge transformations. These formulas may be cast into a more transparent fOrm if we introduce some fancy language. Observe that the wave function 1jJ(r,t) is changed simply by multiplying it with a phase factor
1jJ(r,t)~
-+

exp[ 11

.!..9.

x(r,t)]1jJ~,t)

-fiI

-+

If you perform two gauge transformations: 1jJ

eXP [X}1jJ :

exp[x}[exp[~Xl]

1jJ

=exp[~(Xl+X~]1jJ

then this is equivalent to a single gauge transformation, with the combined factor: exp[igx ].exp[igx ]= exp[~ (X +X )]
n2 nl

~21

Thus the phase factors constitute a led the unitary group in one dimension.

group, the group of all comand it is cal(Remember that a unitary

plex numbers of modulus 1. This group is denoted U(l)

58

matrix is a complex matrix with the poperty:

l. At= At A = I
The unitary 2x2 matrices constitute the group U(2), the ,unitary 3x3 matrices constitute the group U(3), etc. Now a lxl
matri~

is nothing

but an ordinary complex nUmber z, and the hermitian conjugate is just the ordinary conjugate: z=x+iy~z=x-iy. Hence U(l) of the phase factors.} We have now attached a group G to our gauge transformations. It is called the consists of all complex numbers with the property: zz=l,i.e. it consists exactly

gauge g:l"Oup,and in this case G=U(l). The group U(l)


for all
1

is an Abelian group, i.e. g.g = g.g


1 2 2

(contrary to U(2)!) and therefore electromagnetism is called an

a belian gauge theory.


If we put x=(r,t) for abbreviation, then to

every spaae-time point X

lJe have attaahed a group element:


(2.40) g(X) exp[iiI xIX) 1

This group element is responsible for the gauge transformation. First we consider
~~e

Schrodinger wave function

~(X).

It transforms

according to the rule: (2.41)


1jJ (X) ~1jJ' (X) = g (X)

w(X)

A quantity transforming in this way is called a gauge vector,

E:x:ercise:2.8.1 Problem: Prove that

aa x = - ~(a g)g-1 1'1 a where g is given by (2.40).


according to the rule:

T~en

there is the gauge potential A" (X). According to (1. 31) and the

above exercise the gauge potential transfoDtlS (2.42)

We shall explore the geometrical significance of this strange formula later on. Finally we have the field strengths FaB(X), which transform according to the rule: (2.43) F aB (X) ~ F ~S (X) = F as (X)

59 and a quanti ty wi th this property is called a gauge saalaP. gauge

As a consequence of the gauge symmetry of our theory we immediately see that any physically observable quantity must be a saalaP. Let us playa little with gauge scalars! Schrodinger wave function
~(X).

First we consider the

As

~(X)

g(X)1jJ(X)

and

we see that ~~ is a gauge scalar. Actually it is observable. It is the

probability density.
Then consider the Maxwell field A a.(X). Here
~le

can form the gauge

scalar:

But this is not the only one. Consider a closed loop f Then the integral

in spacetime.

tji A dxa. r a
is a gauge scalar. To see this we perform a gauge transformation:

mere aa. X does not contribute to the integral because we consider a alosed loop) .Is the above l=p intet-axis

gral
B

~le

? Well, consi-

vo!'l
A

der two space-time points A and B connected by the paths

.rY.,Aa(X )

r l and r 2 Then an external


field will produce a phase
B

r1ri yf2
y-axis

shift between electrons moving from A to B along fl or along

r2

This phase shift is

given by:

Fig. 30
where f=f2-fl' This phase shift is in principle measurable, being nothing but the Bohm-Aharonov effect! Thus we conclude that

exp[i;! tji AadxCl ]


is in principle measurable quantum mechanically.

60
Worked exercise: 2.8.2
Problem: Consider the parallellogram shown on the figure below. Show that the a field :trength FaS is related to the loop integral ~A dx through the followlng formula: a xa+oxCi.
F

as (X)AxSo x a

. 1 I A &c a 11m "E2 !!'


E->0

CI.

xa+D:x.a

Finally we look at the Schrodinger wavefunction O/(X) and the Haxwell field Aa(X) at the same time. What can we make out of them? Consider the derivative aa 0/ of 0/. This is not a gauge vector as

T~e

last term spoils the game, but containing aag(X), it reminds us

of Aa' Consider Aa'V. It transforms in the following way: A 0/ a


+

A' o/'=[A -i~ (a g)g-l]go/= a a q a

g(X)A (X)o/(X)-i~ [a g(X)lo/(X) (l q a Hence the same spoiling term occurs! Consequently: aa o/-i* Aa 0/ = (aa -it Aa) 0/ is a
(2.44)

gauge vector, i.e. it has the correct transformation properties:


)ljJ' "'g(X) (a

_l<! n

A )0/ a

The differential operator Da=aa -"a is called the

gauge aovariant derivative, and it plays a very important

role when you want to construct gauge invariant theories! The discovery of the gauge covariant derivative allows us to write down non-trivial

gauge scalars based upon 0/ (X) and Aa (X). E.g.


liD()o/, (Dao/) (DSo/)'
F

as (Dao/)

(D

0/)

or written out explicitly: i(x) [aao/(X)-i* Aa(X)o/(X)] [aai +

Aai]['aljJ -

ASo/]

etc.

61

Exeraise: 2.8.3 Problem: Let o/(X)

be a gauge-vector. Show that


[D 'D
)1' V

10/= - ~ 0/ 1'1 )1V

So we have walked right into our first gauge theory! We summarize the most important points in the following table: THE

BASIC INGREDIENTS OF A GAUGE THEORY:


G

A gauge theory aonsists of a gauge group

and a gauge potentia l


U(

Electromagnetism is a gauge theory based upon the gauge group Maxwell field Aa(X)

1) and the

To perform a gauge transformation you must attach a group element g(X) to each space time point gauge covariant derivative gauge vectors X.

(2.45)

a = aa o/(X)
Dao/(X)

~Aa
0/' eX) = geX)o/eX)
D' 0/' eX) = geX)Dao/(X)

(2.46)

(2.47)

gauge scalars

Fai3=aaAS-a Aa S
~Aadxa

F' as = Faa
~A~dxa = ~Aadxa
A~eX)

(2.48)

gauge potential

Aa eX )

= Aa eX ) -

i:[

aa g ] g-1

(2.49)

gauge phase factor

tPr(BIA) = exp[-\rJrAa dXa ]

tPreBIA) = g(B)tP r e B1A )g-1(A)

62

2.9

THE SCHRO'DINGER EQUATION IN THE PATH INTEGRAL FORMALISM

At this point we will get in contact with the conventional a]JproQc11 to quantum mechanics based upon the Schrodinger equation. Studying dynamical problems we nave been using the propagator which governs the dynamical evolution through the integral equation:

(2.18)
We now want to show tnat this integral equation actually is equivalent to t:1e Schrodinger equation. For simplicity we will only consider the one-dimensional motion of a particle in a potential V(x). In this case the Lagrangian is given by,
(2.5)

L(x,x,t)= ~mx2- V(x) X(t2)=X2

and the propagator by

(2.21)

K(X;t2Iy;t0=

tl To convert the above integral equation (2.18) into a differential equation we put tl=t and
t2=t+~t:

J exp{ X(td=Xl

K f [; i

t2

2- V(x)>>t} D[x(t)l

~(x;t+~t)= fK(x;t+~tly;t)~(y;t)dy

Since t-axis

~t

is going to be

very small, we expect the phase factor to be approximately equal to: i[m ~2 1 exp{fi 2 (~t) - V(x) J~t} Of course this is not true for a very "wild"

Fig. 31

x-axis

path, but the wild paths kill each other due to the rapid oscillations.

Consequently the "infinitisemal" propagator is given by


i K(xit+AtIYit)~ A exp{fi

[m 2 ~2 At -

V(x)~tJ}

in the limit of very small At. A is a normalization constant, which

arises from summing up all the amplitudes in the path integral and

63

we will determine its value in a moment. Inserting this approximation into the integral equation we get: +'" . [ m ( -x ) 2 1. (2.50) ~(x,t+~t)~ A _~ exp{fi 2 ~ i

V(X)~t]} ~(y,t)dy

A exp[- fi

V(x)~t]

where we have put y=x+n. Observe, that the integral is wildly oscillating for great values of n. Hence the main contribution comes from values of n where n 2 is comparable to ~t. We may therefore expand
~(x+n,t)

to second order in n (second orGer corrections in n


~t

corresponds to first order corrections in factor

because of the phase

Doing this we get


~(x+n,t)~ ~(x,t)+n

3X + %n 2

oljJ

o2ljJ axz

Inserting this into (2.50) we now obtain


~

(x,t+t.t)

A exp[-

V(X)t.tf!

exp[~

i ~~]{~(x,t)+n* +
+""

%~}

dn

~ does not contribute to the integral because it is an The term nax

odd function of n

Thus we have shown:


-too

i ~(x,t+tlt) '" Aexp{- fjV(x) tit} [ ~(x,t)

J exp[2fi n tit] dn +
2

im

\Ji ~

Here the first integral is a Gaussian integral of the standard type, cf. (2.27), while the second is obtained from the formula

f +~2exp[-ax2]
-00

dx

=....!....f!I. 2a a

which follows from respect to

(2.27)

when you perform a differentiation with

a. In this way we finally obtain:

(2.51) ljI(x,t+tlt) '"

A/21T~t exp[- ~V(x)At]ljl(x,t)

Aj21T~t ~t

exp[-

ff(X)At]~:t
+

First we observe that this fixes the value of A. Because when fit

64
the left hand side approches only if: ( 2.52) 1jJ(x,t). But this is consistent

A=! 2TTi~t.t
K(x;tHtly;t) '" V~t exp[fi
~

This should be conpared with the free particle propagator (2.28). In fact the "infinitisemal" propagator can now be decomposed as follows (2 53)
i

2"

m (x-y) 2 i lit lexp[-fiV(x) lItl

KO(x;t+lItly;t)~I(x;t+lItly;t)
where KO(BIA) is the free particle propagator, and cf. (2.33)

~~(BIA) is the phase

factor corresponding to the interaction

~~(BIA) =
Inserting (2.52)

exp[-ifv(x)dtl

into (2.51) we now get a $ 2ffi"axz2

i iht.t 1jJ(x,t+t.t)<>< exp[- h V(x) t.tl1/l(x,t) +


i.e.
1jJ (x,t+t.t) -1jJ (x) ~ ih . t.t - 2m ax

+ exp[ -(l/h)V(x)lItl-1 ,I>(x,t)

t.t

'j'

which in the limit t.t a,I> it ih 2m

0 reduces to:
a2$

ax7 -

KV(x)1jJ(x,t)
.

Finally this may be rearranged as: ih a1/l


h2

(2.54)

at

- 2m

a2'4> axz-

+ V (x) 1/1 (x, t)

So we have captured the Schrodinger equation! By using exactly the same procedure we can find the Schrodinger equation for a charged particle in an external field. However, it becomes technically more complicated, partly because i t is now a three-dimensional motion, and partly because the generalized potential
(2.11)

has a more complicated structure. We will leave this as an exercise: (Details can be found in Schulman (1981.
E=l'aise: 2.9.1

Consider a three-dimensional motion of a particle in a ordinary potential V(x). Use the path integral formalism to derive the Schrodinger equation:

Problem~

ih

at

= - 2m

h2

ai ai 1/1

.... + V(x)ljJ

65
Exercise: 2.9.2
Problem: Consider the three-dimensional motion of a part'.cl p in the genel'alised potential:

Use the path integral formalism to derive the Schrodinger-equation:

ih

a,h if =

h2 - 2m

ai~

11 Ai) a - inA

iq

(i

i)

ljJ +

q<P1/I

2.10

THE HAMI LTON I AN FORMALISr1

The other method you can use when you want to quantize a system is the canonical quantization. Here the starting pOint is the Hamiltonian Consider a Lagrangian by:
(2.55 )

formalism. L(qi,qi,t),i = l, ..... n. With each of the conjugate momentum Pi defined generalized coordinates qi we associate a

i=l, ... ,n

Observe, that Pi is a function of qi,qi,t: Pi= Pi (q .,q ,t); i= l, .. ,n


i Sometimes we may invert tnese equations and express q as a function
i

of q , pi,t. If this is the case, we say that the system under consideration, is a canonical system. Hamiltonian: For a canonical system we then define the
(2.56)

H(q 'P.i,t) =

\"Ie

shall often abbreviate this as H = P . qi - L but you should al1

ways remember that q i should be regarded as a function of the canonical variables p. ,qi. Observe, that we are using the Einstein sum 1 . convention! An expression like Pi q1 implies a summation over i =l, . ,n. The Hamiltonian is very ves: and
us~ful.

We may use it to cast the equations

of motion into a very beautiful form. Consider the partial derivati-

66

Let's work them out:

aE 3Pi

a
3Pi

[po qL L ]=6~ q j+ o.
J
J

aqj

3L

3q j

by the definition (2.55) l'lext we have:

aqj dpi of the momentum conjuc;ate to q j.

J 3p i

.i q

3H =_3_. [p.qj -L]= p. ~_ aq). aq). J J aqi From the Euler-Lagrange equation (2.6) we
d

~~
aqj aqi

3L aq
'i

dt leading to

(2.57 )

and

These equations are referred to as

Hamilton's equations and they

are completely equivalent to the Euler-Lagrange equations.


To check these ideas in a trivial example we consider a one-dimensional motion in a potential V(x). From the Lagrangian (2.5)
) L ( X,x 1 2 2 mx
) - V(x

We see that
P

= 3L = mx

3x
1

so that the momentum conjugate to x in this simple case cone ides with the usual kinematic momentum. The system is canonical and we have:
mP Hence we can find the Hamiltonian:

(2.58)

H(p,x) = px - L = ~ + V(x)

which reduces to the usual formula for the mechanical energy. Finally we get Hamiltons e~uations: x ::; 3H 1
3p =

mp

3H 3x

=_

V'(x) to Newton's
e~uation

which are obviously

e~uivalent

of motion.

For a time-independent system we may identify H with the total energy of the system. TO establish this we consider a system characterizec a Lagrangian, which does not depend explicitly on time:
~y

67

Then the Hamiltonian too does not depend explicitly on time: H = P/li(qi ,P ) _L(qi;qi(qi 'Pj)) j Consider a trajectory, qi= qi(t), which satisfies the equations of motion. Then H is dH dt according to Hami lton' s equations! Hence long the classical total energy follows
pat~. co~stant

along this trajectory:

o
H is a constant of motion
a That this constant of motion is actually the

from the illustrative example above.

Let A(qi;p.) and B(q\p.) be functions of the generalized coordinates ql, ... ,qn and their ~onjugate momnta p , .. . ,p . We define the Poisson bracket of A and B as follows: 1 n dA aB {A; B} (2.59) -. -a-- dB aA aqi OPi aql Pi where a summation over i is implied as usual

Exercise.' 2. 10. 1
Problem: Show that the poissoncrackets satisfy the following simple rules: 1. {A;B}= - {B;A} (Skew symmetry)
i {A ;c}= 0 whenever c is a number, i. e. a constant function of q and Pi'

2.

3.

{Al+A2 ;B} = {AI ;B} + {A 2 {AA;B} = )..{A;B} {A1A2 ;B} {A;{B;C}} +

;B}}. .
Llneanty
0 (Jacoby identity)

4.
5.
Exercise: 2.10.2

{AI ;B} A2+Al {A2 ;B} {B;{C;A}} + {C;VI;B}}

Problem: Consider the i'th coordinate function qi(ql, .... ,qn; pi, ..... ,p) = ql Similarly we can consider the i'th componenf of the conjugate momentum as a special function: Pi(ql, .... ,qn; Pl, .... 'Pn) =Pi i Show that q and Pi :ati.sfy the rule: (2.60 ) {ql;q J} ={Pi ;Pj}

=0

and

{qi ;Pj}

= oi j

Exercise: 2.10.3
Problem: Let F(ql, ... ,qn;Pl," .,p ,t) be a function.of the generalized coordinates, their conjugate momenta and ~he time t. Then '11 and p. themselves depend on time and their time dependence is governed by Hamilton's eqliations (2.57). Show that the total time derivative of F along a trajectory in phasespace is given by:
(2.61) dF dt

= at dF

+ {H.F}

'

(Note that if F does not depend explicitly on time, then F is a constant of motion if and only if the Poisscn bracket {H,F} vanishes).

68

Exercise: 2.10.4
Problem: Let F(ql, ... ,qn) be an analytic function. Show that: { Pi,F (q I , ... ,q) n } = -.aqi aF (q I , ... ,q n) 1. 2. (q\F(ql, ... ,qn)} = q'F(ql, ... ,qn)
n-i - nqi

(Hint = expand F(ql~ ... ,qn) in a powerseries and show that: {Pi,q~} where we exceptionally do not imply a summation over i!)

2.11

CANONICAL QUANTIZATION AND THE SCHRODINGER EQUATION

Now the canonical quantization runs as fOllows: The generalized coordinates (ql, .. ,qn) specify a position in configuration space. Quantum mechanically we characterize the system by its Schrodinger wave function W(ql, ... ,qn;t) which gives us the probability amplitude for finding the system at the pOint (ql, .. ,qn) as a function of the time t. The dynamical evolution of the wave function is governed
~y

the Schrodinger equation which you construct in the folE

lowing way. Consider a Hamiltonian: hand side you substitute, substitute: Pi +


-

ih~t' and on the right hand side you

= H(qi,p.). 1

On the left

iii.

a~i

Then both sides are converted to diffe-

rential operators and the Schrodinger equation reads:


. Ii.

(2.62)

at

-,10 't'

To check these ideas in a trivial example we consider a one-dimensional motion in a potential V(x). We have previously found the Hamiltonian: H(x,p) = p2/2m + V(x) . Thus we get the ordinary Schrodinger equation:
(2.54)
l'li.

~ _ Ii. a ()] ( ) at- -2m l a?"+Vx wx,t


2 2

Let us attack a more interesting problem. Suppose a particle is moving in an external electr6magnetic field, and let us try to quantize this problem in the canonical way: From
(2.12)

we see that the canonical momentum is given by


p =
~

aL

mv +

qA

and hence they do not just reproduce the kinematical momenta! This system is canonical too and v=iiiP-m
~ l'~

S!. A

69

Thus we can construct a Ham1ltonian:

H(x,p,t)

= p.v

- L

= 2k(P

- qA)2+

q~

i'ie may rearrange this slightly:


(2.63)
2

E - q~

=~ 2m

(p -

qA)2 free particle,

Here you see something interesting: TO pass from a where E

=~ ,

to a particle in an external electromagnetic field

you have to make the following substitutions:

(2.64)

-+

E -

q~

p-

qA
ruZe of minimal coupling. It is

and this simple rule is known as the later on.

of very great importance and we shall investigate it more closely For the moment we want to pass to the Schrodinger equation:
(2.65)
[ 1'Ii ~ at

q",

~l ",(~ t)= [-ili~-qAl2 't' r, 2m

Observe that the Schrodinger wave function interacts directly with the gauge potential (~,A) and not with the field strengths! You should also observe that the equation is gauge co-variant. we rewrite it slightly:
'Ii r
1

TO see this

'at

a + i g "'l n
't'

,10 't'

_ -

-Ii 2

[V -

2m

Al2
V -+ V -

,10 't'

Hence to pass from a free particle to a particle in an external field we make the following substitutions:
(2.66)

at

-+

at + 1

These substitutions can be written more compactly as:

*
If

<P

(2.67)

aa

-+

aa

Da

Consequently we have simply converted the ordinary dErivatives to gauge covariant

dErivatives! This is also referred to as the rule of minimal coupling. Using


this we can rewrite the Schrodinger equation as
(2.68)

70

Here you explicitly see that it is a gauge covariant equation, since both sides transforms as gauge vectors. Thus if they are equal to each other in one particular gauge, then they are identical in all gauges! Consider a Hamiltonian operator
H
A

H(q ,-

aql. If we introduce the following inner product for wavefuntions:

in ---.)

<1/J[<P> =

J (x)<P(x)dx

thenAit can be shown under very general assumptions that the operator H is Hermitian i.e.

" 1/Ji<P> <H

where 1/J,<P are arbitrary wavefunctions.


Worked exercise: 2.11.1 Problem: Let ft be a Hermitian operator. Show that the eigenvalues are real(i.e. if ft1/J=El/J then E is a real number) and that the eigenfunctions belonging to two different eigenValues are orthogonal (i.e. if ft1/J1 = E11/J and ftljJz =EZ1/J2 where El *Ez then <1/Jl[1/JZ> = 0).

Furthermore we may generally assume that the set of eigenfunctions {ljJn(x)} is complete i.e. that we may expand an arbitrary wavefunction as a superposition of eigenfunctions: 1/J(x)
=
~

anljJn (x)

Here the coefficient an is determined by the relation am <ljJ m [1/J >

provided the set of eigenfunctions is normalized, i.e.

Worked exercise: 2.11.2 Let {1/Jnix)} be a complet: orthonormal set of eigenfunctions of the Hamil. tonlaJl operator H. Show that.
Pro~lem:

(2.69)

~ 1/J (xd 1/J

n (x,) =0 (Xl- xz)

Consider an eigenfunction of H,i.e.a wavefunction 1/Jn(x) with the property:

This eigenfunction can immediately be extended to a solution of the Schrodinger equation:

71

In accordance with the Einstein-de Broglie rule this is interpreted as tl1e state of a particle with energy En. Observe that the probability distribution is time independent: ln(x,t)1
2

= jlj!n(x)i

and we therefore say that the wave function represents a

stationmy

state.
Let lj!(x) be an arbitrary wave function. Then we can decompose it as a superposition of eigenfunctions: lj!(x) But the Schrodinger equation is linear so we can immediately extend this to the solution (x,t) =

an1/Jn(x) exp [-

Ent]

which reduces to 1/J(x) for t=O. Once we know a complete set of eigenfunctions we therefore ger wave functions. This suggest that the Feynman-propagator it self can be expressed through a complete set of eigenfunctions: controll the dynamical evolution of Schr6din-

Worked exercise: 2.11.:3 Problem: Let {lj! (x)} be a complete orthonormal set of eigenfunctions to the Hermitian operator ft . Show that the Feynman propagator can be expanded as:
(2.70)

Let us make a final comment about the canonical quantization. In the preceding descussion the Schrodinger wavefunction has played a central role. However it is possible to avoid it. The physical quantities are then represented by Hermitian operators that do not necessarily operate on the space of Schrodinger wave functions. These operators cannot be chosen arbitrarily: .In the classical context the physical quantities are represented as functions A(q';p.) of the generalised coordinates and their conjugate momenta. In the quantum cditext these functions are rep raced by Hermitian operators in such a way that their Poisson brc:cket is rep raced by the commutator:
(2.71)

{A;B}+-1-[~,~1
if!

Exercise: 2.11.4 Problem: Show that the commutator [ ~;~ ] ding to exercise 2.10.1.

= ~~ -

~ satisfies the rules correspon-

Especially the generalised coordinates qi and their canonical momenta p. must be replaced by Hermitian operators satisfying the socalled Heisenberg commlltation rures:
(2.72) I\i ; I\j] 1\ 1 [q q = [1\ p.;p. = 0 1 J

ih6~

Compare (2.60).

72

Exercise: 2.11 . .s . P:oblem: Show x~at if we replace the generalised coordinate q" by the multiplicatIon operator q = q and the canonical momentum Pi by the differential operator

~.
I

= _ ih _a_.
aql.

then

and

satisfy the Heisenberg's commutation rules.

Exercise: 2.11.6 . Problem: Show that if ~l. and satisfy Heisenberg's commutation rules then their commutator with other operator satisfies rules corresponding to exercise 2.10.4.

2.12

ILLUSTRATIVE EXAMPLE: SUPERCONDUCTORS AND FLUX QUANTIZATION

We finish this chapter by reviewing another experiment which shows how gauge potentials may produce unexpected quantum mechanical effects~

Consider a suitable piece of metal say aluminum. If we cool


[1964], de Gennes [1966]). Suppose we originally had an ex-

it down it becomes superconducting and interesting things happen. (Feynman ternal magnetic field. This would penetrate into t!1e metal when the temperature is high, but it turns out that when the temperature falls down below a certain critical temperature Tc' then the external field is expelled. This is the famous Meissner effect.

Fig. 32

T>T

T<T

It turns out that there is produced currents in the outher layers of the metal and these currents prevent the magnetic field from penetrating into the metal. This is an interesting situation. We have a magnetic field

B which

stays entirely outside the -lump of metal, but there is a gauge potential

A too,

and it may very well penetrate into the metal!

73

We now change the experiment a little. We take a ring whose width is great compared to the penetration depth of the magneT>T c tic field.
i~e

place it in an ex-

ternal magnetic field

at room

temperature. Nhat happens? The magnetic field is spread throughout the whole space and especially it penetrates into the metal. Then we cool dm'Tn the ring, and when we pass below the critical temperature Tc' the r-ieissner effect occurs. The magnetic T<T

field is expelled due to surface currents in the superconducting ring. Hence some of the field lines pass outside the ring and some of them pass through the hole in the ring. But now we remove the external field. However, this does not mean that the magnetic field disappears completely. Part of the field passing through the hole is trapped by the surface

currents. So the superconducting ring now acts much like a solenoid! Observe, that inside the metal there is no magnetic field a gauge potential A.Since we have trapped a magnetic field through the ring, there is a magnetic flux this flux we are going to examine! Let us try to get a qualitative understanding of the situation. According to the BCS-theory, the electrons in the metal will form pairs in the superconducting state. They are called Cooper pairs. Now the electrons are fermions, but the Cooper pairs act like bosons! This has the important consequence that the Pauli principle no longer applies. Two electrons cannot
occ~py
~

B , only B which pass

through the ring. It is

the same state, but

74

there is nothing to prevent two Cooper pairs from occupying the same state! Actually hosons have a strong tendency to occupy the same state. Suppose ~(r,t) denotes the Schr5dinger wave fUnction for a Cooper pair. We may actually assume that there is a macroscopic number of Cooper pairs all described by the same wave function
~(r,t).

This has important consequences. The absolute square I~(;,t) 12 If N is the total number of Cooper pairs, then
N

is

the probability for finding a specific Cooper pair at the point r .

I~(r,t) 12

is simply the dEnsity q

of Cooper pairs at the pOint r. If furthermore

-+

2e is the charge of a Cooper pair, we conclude that:

is the cha:t'ge dEnsity of the Cooper pairs. Since a solution to the Schrodinger
~tion

is only determined up to a constant we may redefine it:

~(r,t)
single wave function:

-+

~ ~(r,t)

~'(r,t)

So rcw we have a macroscopic number of Cooper pairs described by a

~' (r,t)
and

-I ~

(r, t)

12

simply denotes the charge density at r . Observe


2

that although I~I

in general only has a statistical meaning,

I~'

12

will in this case denote a macroscopic physical quantity! In what follows we will drop the prime and simply denote the wave function by ~(r,t). We may decompose it in the following way:
(2.73)

Here p (r,t) is the charge density, which has a direct physical meaning, and
-+ ~(r,t)

is the phase which has no physical meaning as we

can change it by performing a gauge transformation! In the superconducting state the number of Cooper pairs are conserved and so is the charge. If we denote the Cooper current by we therefore conclude that p and
(1.17)

obey an equation of continuity:

75

We can use this to find an expression for the Cooper current.


p

= -

ljilji

implies that

an __ a,,, at - at lji

lji

at

aij;

Using the Schrodinger equation we can rearrange this as:

which may be rewritten as

at

ap

= 2m v {lji

in

(v +

~ ~

A) lji -

~ lji (v

!g

A) lji }

From this expression we immediately read off the Cooper current:


(2.74)

This show us that the Cooper current is a gauge scalar as it ought to be, and that it is a real quantity. In the following we will always assume that the density p of Cooper pairs is constant thoughout our piece of metal! This is not strictly correct at the boundary where it falls rapidly to zero. Except for that,it is a very reasonable assumption that the Cooper pairs do not "crowd" together but are evenly spaced.

Exercise: 2.12.1
Problem: Show that the expression for the Cooper current can be rearranged as: ~pn ~ qA] (2.75) [ v \1>- 11

J=m

provided pis a constant.

Using exercise 2.12.1 we can now derive the Meissner effect. Since there is no external electric field involved, the Maxwell equations (1.3) and (1.6) reduce to (2.76)

v. B

= 0

J =~

-t

But from (2.75) we then get Vx (VxB)

=~ EoC

Vx.... J

= - Eomc ~

VxA

= - ~2 B Eomc
Vx(V~).

where we have used (1.12) to get rid of the term ranged as

Furthermore using (1.14) and (1.3) the left hand side can be rear-

76

Hence we end up with the simple equation (the London equation) (2.77) where (2.78) is the socalled London-length. That equation (2.77) indeed implies the Meissner effect is left as an exercise, see below. Exercise: 2.12.2
Problem: Consider a semi-infinite superconductor occupying the half-space: z>O. Let us apply a constant field = Eo (1,0,0) parallel to the surface. Show that inside the superconductor the solution to (2.78) is given by: ... [ z B (z) = Bo (exp - ~1,0,0) z > 0

ai3

I"f

it

so that the magnetic field vanishes exponentially for z > AL.(A is typically a L few hundred Angstrom.)

Observe that (2.76) implies that not only

B vanishes

inside the

superconductor, the same thing holds for the Cooper current

3-

Consider now a superconducting ring as shown on fig. 34. Inside the ring at the curve r the Cooper current vanishes, as does the magnetic field! therefore get:
Us~

(2.75) we

~ m

(Vc.p-

9.
Ii.

lA)

But now we can compute the magnetic flux inside the ring beccluse: (2.79)

pit.
r

dr

Ii..! q!ll

V,n' '"

dr

You might be tempted to say that it is zero! But let us look a little closer at the phase lP. The wave function ljJ(r,t)

jl p(r,t) I exp[

ilP(r,t) 1 that it

is only non-trivial in the ring, and all we can demand is

is single valued. But then nothing can prevent lP from making a jump of 2nn. If that is the case, VlP will contain a 6 -like singularity

77

and the integral need not vanish!

(Compare the discussion in sec. 1.4)

TO compute the line integral we assume that V~ makes the jump at the pOint B. If B+ and B- are pOints extremely close to B on each side of B we get: <l>
h

P A d r

h
0<

B+ B
h

q q

L
21Tn

V~

dr

-+

[~(B+) - tp(B - ) 1

+ as B ,B

consequently we get 21Th

(2.80)

<l>

n<1>o

where

<1>0

Iqf

So the trapped flux is

quantized~

This remarkable effect has been

established experimentally (Deaver and Fairbanks [l96l1)~)The quantity <l>o is the fundamental fluxquantum. We have already seen that a vanishing magnetic field inside a superconductor implies a vanishing Cooper current. In the next exercise we will show that the opposite holds too: If t:1e Cooper current vanishes then so does the magnetic field.
Exercise: 2.12.:3 Problem: (a) Show that the expression (2.74) for the Cooper current can be rearranged as: j. = ih iii D.1jJ (2.81 )
1

provided the Cooper density p is constant. (b) Use this expression and exercise 2.8.3to prove directly that a vanishing Cooper current implies a vanishing magnetic field.

In the previous discussion we have assumed that the superconducting state of a metal only depends on the temperature. Actually it also depends on the strength of an external magnetic field. Let us discuss this in some detail for a special kind of superconductor known as a type II superconductor. Consider a superconduction cylinder. Outside the cylinder we have a solenoid. Suppose a weak current flows in the coil. Then it will produce a magnetic field of strength B which is expelled from the cylinder due to the Meissner effect. When we increase the current, B will reach a critical value BCl where the superconducting state starts breaking

~l:--~ ---------------- Fig. 35

~t

*) "Experimental evidence for quantized flux in superconducting cylinders", Phys. Rev.

Lett. 7 (1961) 43

78

down.

Thin vortices are formed where the normal state of the

metal is reestablished. The magnetic field starts penetrating into the metal through these vortices. As the magnetic field strength increases more and more vortices are formed and when we approach another critical field strength BC2 only small superconducting regions are still destributed throughout the cylinder. When we finally pass BC2 the superconducting state breaks down completely and the cylinder is now back in its normal state. If we decrease the current again then the same things happen in the reversed order. Now let us concentrate on a single vortex. Let us inclose it by a great circle

r as shown on the fithus the Coo-

gure. Far away from the vortex the magnetic field and per current vanishes. The magnetic flux through the vortex is given by:

1111

Fig. 36

due to the same arguments as for the superconducting ring. But then we see that the magnetic flux
q~tize~
~

through the vortex is necessarily

The existence of quantized vortices was predicted by Abri-

kosov (Abrikosov [1956 tland they are therefore


~ye

call~d

Abrikosov vortices.

have previously stated that charged particles may interact

with the gauge potential A in a spacetime region Q, even if the field )J strength F)JV vanishes identically throughout this region. This interaction was a pure quantum mechanical effect, the Bohm-Aharonov effect (sec. 2.6). We may throw light on this using our results concerning the flux quantization: Consider two identical rings: A and B. In the following experiment the two rings are placed at room temperature. Inside ring B we also place a tiny solenoid. In
L~is

solenoid we have a current

which produces exactly one half of a flux quantum. Hence in the beginning of the experiment the flux through A is 0, while the flux

through B is

%~o.

*) Soviet Phys. Jetp 5 (1957) 1174

79

Fig.37

Now we cool down the two rings and they become superconducting. !ihat happens? In the first ring nothing happens. But in the second ring the preceeding analysis concerning flux quantization is clearly valid. The superconducting ring will only allow a quantized flux through the ring. comes 0 or
~o

Thus

a Cooper current is produced in the outer

layers and this contributes to the flux, so that the total flux be
L~e

But what is the origin of this Cooper current? If the Cooper pair only interacted with field strength , this would be a mystery because there is only a nonvanishing field strength inside the solenoid. But as we have seen, there is a non-trivial gauge potential outside the solenoid and this penetrates into the ring. This suggests that it is the interaction between the gauge potential and the Cooper pairs that is responsible for the Cooper current. This is also in accordance with the schrodinger equation: 2 (2.65) ih( d~ + ~)~(r,t)= A)2o/(r,t)

ir

:m (V-ir

which shows us that the Schrodinger wave function interacts directly with the gauge potential.

SOLUTIONS OF WORKED EXERCISES:


No. 2.4.1 Now let x = Xcl(t) be the cla~sical path. Then we can write an arbitrary path in the follow1ng form: X=XC1(t) + yet) where y(tll = y(tz) = 0

Let us expand L using Taylor's theorem: dL dL L(xcl +y;x cl +y;t) =L(xcixcht ) + dX y+ax y+
1. '2 [

d1:. . d1:. d1. 21 dxdX y ... 2 dXdX YY + m~ J

This expansion is exact because L is qUadratic! Then we can rewrite the action:

80

=f

t 2 L(x l'x l,t)dt tl c C

=J

~2aL aL (-- y+ - r y)dt tl ax ax

where the term in the miMle vanishes t 2 dL tl

f (3"
x

y +

ar y)dt
x

dL

t 2
=

tl

f (

aL ax -

d aL dt ax )y dt

because x satisfies Euler's equation! cl The last term can easily be computed explicitly:

substituting this into the path integral we get:

D[y(t)J

But Xl and X2 are not at all involved in this last path integral, so it can depend only on tl and t2 ! Consequently we have shown:

If furthermore the coefficients we can easily show that

a(t), b(t) and c(t) do not depend on time, then

Observe that:

t2+At t2+At [ay2 (t )+by 2 (t )+cy:y.Jdt = f [ay2 (t -lIt)+ ... Jdt = f lay ,2 (t)+ ... Jdt t I t I +At t I + At

t2

where we have introduced

y'(t) = y(t - At) . Thus

from which the desired result follows immediately when we put: At

-tl

81
No. 2.8.2

A (X B + f OX B )

A (x) +
<l

<l

2 oA

~ox +

axP

c
B

In the end of the calculation we are going to let + 0 hence we need only to compute the lowest order terms. In this approximation we get:

Interchanging the dummy indices a and

B in

the last term we finally get:

AadX

= 2 [oaAB- "BAal Ax 6x

B a

Taking the limit we obtain the exact result:

No. 2.11.1

(a)

Let E be an eigenvalue and let tion:

be the corresponding normalised eigenfun-

" E =E~I~> = <~I~ = < H~I~>


Using the hermiticity of

Q this

is rearranged as

= ~~~> = ~IE~> = ~1~>E = E


Thus

E = E and therefore E is real

82

(b)

Let E"Ez be different eigenvalues and let eigenfunctions. Then we get:

~"

~z

be the corresponding

No. 2.11.2

Let ~(x) be an arbitrary wavefunction. We can decompose it on the complete set of eigenfunctions:

Using this decomposition we get:

But this clearly shows that:

No. 2.11.3

For fixed (x, ,t I) we know that the Feynman propagator is a solution to the Schrodinger equation. Hence we may decompose it as K'X2; t 40<l;t,J = ~ an(xl;t,) ~n(xz)exp[- fi Ent2l The coefficients can stilldepend on x, and t, . For t2 = t, we know that the Feynman propagator reduces to a 6-function. Thus we get: 6(XI-X2) = l: a (x,;t,)
n n

~.n (x,)exp[- hi

E t,J
n

This forces us to put: an(xl;t,J = an(x,)exp[

Ent,l

since the left hand side is independent of t,. The above formula then reduces to:

but a comparison with exercise 2.11.2 then gives us:

and we are through.

83

ehapter 3 DYNAMICS OF CLASSICAL FIELDS


3,1 ILLUSTRATIVE EXAMPLE: LAGRANGIAN FORMALISM FOR A STRING

We have already discussed the Lagrangian formulation of the dynamics of a system with a finite number of degrees of freedom, say a finite number of particles moving in an external field. Now we want to include the dynamics of ~ields. We will start by considering a one-dimensional string. It has an important property: If you disturb the string at one place, then this disturbance may propagate along the string. We can understand this in an intuitive way. The string consists of "atoms". Each atom interacts with its nearest neighbours. Hence, if you disturb one atom, this disturbance has influence on the neighbour. But this disturbance of the neighbour has influence on the neighbour of the neighbour, etc! In this way a travelling wave is created which propagates along the string!

a
....-"-

x a 2a 3a 4a

- axis
)-

~
Fig. 38 -4a -3a -2a -a
0

xn

na

In our model the string is composed of "atoms", which in the equilibrium state are evenly spaced throughout the x-axis. The important assumption is that the "atoms" are coupled to each other through forces proportional to their relative displacements (Hooke-forces). We will enumerate the "atoms" with an integer n so that the equilibrium position of the n'th "atom" is Xu = an. If we set the "atoms" in motion, then the n'th "atom" will be displaced an amount q from its equilibrium position. When the string is put into vibration~, its dynamical evolution is described by the functions: n = , -2,-1,0,1,2 ... qn = qn(t) The kinetic energy of the n'th "atom" is:

2 mqn
and the potential energy associated with the separation of atom [qn+l - qn 1 Hence the total Lagrangian for the system is: nand n + 1 is:

.2

2 mv

(3.1)

....
L

L =

n=-

Let us determine the equations of motion. Using (2.6) we get 2 (3.2) mqi = mv (qi+l - 2qi + qi-l)

84

It is easy to check that the equations of motion actually allow wave solutions. we put:

If

(3.3)
then this will represent a travelling wave. motion, we get:
2

Inserting this into the equation of

mv Cos (kx -wt) (2Cos(k a) -1) n or

(3.4)
Thus (3.3) is a solution to the equation of motion provided (3.4) is satisfied. Relation (3.4) is called a dispersion relation. Observe that for small k, i.e. in the long wave length limit we may expand the cosine, getting:
(3.5)

which is a linear dispersion relation. Now we want to investigate a continuous string, where there is "atoms" everywhere. We can do this by letting a ~ 0 in our discrete model. We say that we pass to the continuum limit. Now, instead of describing the displacements by the infinite set of numbers: qj(t), we will represent it by a smooth function q(x,t) giving the displacement of the "atom" with equilibrium position x .

Fig. 39
In the discrete model the mass density is: limit, we suppose that it approaches a constant
nuous string, i.e.
m
" .... p

m When we pass to the continuum ~ , the mass density of the contia


~

for

Finally it will be necessary to make an assumption about v. In the disctete model the velocity of a wave in the long wave length limit is (compare (3.3) and (3.5 w k = va We assume that it approaches a constant c, the velocity of a travelling wave in the
continuous string:

va ~ c for a ~ 0 With these preliminaries we can investigate what happens to the Lagrangian in the continuum limit. Let us take a look at the kinetic energy:

!
Xn=-OO

'! [q("n,t)]2axn .... J

x= +1:0

x= -ao We can treat the potential energy in a similar way.

From the observation:

85

we get: !mv [q
n=-oo
Xn=+CO

Xn=+co

n+

1 (t) - q (t)]2

x=-c:o n
x=+ao
-+

l:

xn=-otI

l:

l!! (va)2 [oq (xn,t)] 2 lIx a

ax

Jl x=-c:o

2 [~ ax (X,t) ]2 dx

Thus we get for the total Lagrangian:

(3.6)
showing that in the continuum limit the Lagrangian is expressed as an integraZ over The integrand is called the Lagrangian density.

space.

(3.7)
Observe that it contains not only vative

~ but also ~x come from? o


It came from the term

~ pc 2(~)2 ax

~~

Where did the space deriin the discrete model.

( qn+1-q)2 n

Hence, it reflects the property of ZocaZ interactions. Each point in space interacts with its nearest neighbours. In a similar way we may analyse the equations of motion. In the discrete model we have: We may rearrange the term on the right side: qn+1 - 2qn + qn-1 ax

= [q(x n+l't:}
n

- q(x n , t)] - [q(x n , t) - q(x n, t)]

a[~xq(xn+ ~,t) - 19. (x - g t)]


o ""

2'

Inserting this we get: (3.8) q(x ,t) n


~

a q (x ,t) v 2 a 2 --ax2 n

-+

i. e.

.1....Gt=is 2 2
c at ax2

As before, we may look for a solution representing a travelling wave:

(3.9)

q(x,t)

= A Cos(kx-wt)

If we insert this into the equation of motion (3.8), we get: -w 2A Cos(kx -wt) = -c 2k 2A Cos(kx-wt)
i.e.

(3.10)

ck w satisfies the lin-

Hence, (3.9) is a solution to the equation of motion provided ear dispersion relation (3.10).

86

3.2

LAGRANGIAN FORMALISM FOR RELATIVISTIC FIELDS


You should now be motivated for the abstract field theory. We
~(t,~)

start with a field stretching


q(t,~o)

defined throughout spacetime.

The value

of the field at a particular point field is governed by a Lagrangian preceding example write as:
(3.11)

in the preceding example.

~o ' ~(t,~o)' corresponds to the The dynamics of the L, which we in analogy with the

The Lagrangian density

depends not only on the time derivative

but on the space derivatives, as well:

L
interactions.

= L(~,a\l~)
ai~

The presence of space derivatives

reflects the principle of local If we choose two times tl and

t2 ,we may specify the field at these times. ditions: Any smooth function
~(t,~)

which satisfies the boundary con-

~(tl'~) = ~l(~) and ~(t2,i) = ~2(i)


represents a possible history of the field. To each such history we assot2
S

ciate the action: J Ldt tl t2 3 J JLd 1Cdt tl


(

i.e.
(3.12)
S

n
where and

JL

d"x

n
t=t

is the four-dimensional region between the hyperplanes: t=tl

As usual, we want to determine a history ~(t,~) which 2 extremizes the action. This, of course, leads to the equation of motion for the field. action. Now suppose that
~o(t,~)

really extremizes the

Consider another history:

~(t,~)
where n(t,i}

= ~ o (t,~)

n(t,x}
n(~,i)

satisfies the boundary conditions:

87

Then the action:


SeE)

n
has an extremal value, when

(<I>o+n, dll <l>O+dlln)d x

=0 . Consequently we get

where we have neglected the surface terms due to the boundary conditions on n. But n (t,~) was arbitrarily chosen. 'llierefore the above result is consistent only if <1>0 satisfies the differential equation:

(3.13)

This generalizes the Euler-Lagrange equation for a system with a finite number of degrees of freedom. Observe that all the derivatives occur! This has an important consequence: The equation of motion is Lorentzinvariant provided L is a Lorentz-scaLar.

We would also like to discuss fields with several components: <l>a' a=l, .. , n (like the Maxwell field .Ace l. We leave the deduction of the equations of motions as an exercise.
Exercise 3.2.1 Problem: Suppose the field has several components: <1>, a=l, ,n. Show that each of the components must satisf~ the appropriate EulerLagrange equationl
(3.14 )

a=l, ... ,n

We may also discuss the energy-momentum corresponding to our field <I> a A direct generalization of the Hamiltonian method suggests that the energy density is given by the HamiLtonian density:

(3.15)

(Compare with (2.56.

We expect energy to propagate throughout space,

so we must also have an energy current

associated with the field.

88

Since the total energy: E =

J Hd 3~
H

should be conserved, we must demand that tion of continuity:

and

-+

satisfy an equa-

We can use this to determine an expression for the energy current

-+

s:

dL

doa (do<l>a)

dL dotb a + d(do<l>a) dodo<l>a -

dL d<l>ado<l>a

Using the Euler-Lagrange equations (3.14) we get:

Inserting this you find:

From this we can immediately read off the energy current:

(3.16)

We know that the energy density corresponds to the

TOO-component of

the energy momentum tensor and that the energy current si corresponds to the T Oi component (compare (l.3~), but then the above expressions for Hand s1 suggest that we put

(3.17)

If
-+

is a Lorentz scalar, this is a tensor which reproduces

Hand

s .

Exercise 3.2.2 Problem: Use the equations of motion to show that the above energy-momentum tensor is conserved:

From exercise 3.2.2 we learn that momentum!

~aB

produces a conserved energy-

So everything should be all right:

~aB

is called the

89
canonica~

energy-momentum tensor.

However, there is one point where be symmetric, contrary to our

you should be careful:

~aS

need not

previous demand (sec. 1.6) based upon conservation of angular momentum. Consider the expression (3.17) for the canonical energy-momentum tensor:

Obviously, the last term is symmetric, but the first term need not be
so~

What can we do if the canonical energy-momentum tensor is not We must repair it: Suppose we can find a tensor

symmetric?

eaB

with the following properties: is symmetric (3.18)


2) dSe

aB
od
3

= 0

3)

Je

i =

Then T

TaB

eaB

is a symmetric tensor which is conserved and it.re-

produces the same energy-momentum as

aS = ~aB + eaB

~aS

Hence, we might call There exists a But the

the true energy-momentum tensor.

.systematic method to construct linfante, see for instance:

eaS

(the method of Rosenfeld and Be-

Barut [1964], ch. 3, sec. 4).

method is very complicated and we will not have to use it. Later on (ch. 11) we shall see how it can be const"xucted fran a dEferent point of view. Okay, suppose we want to construct a free field theory. What should we demand about
L

That depends on what we expect of it! it might help to look at the

Although we will not quantize the fields, following very naive discussion: A particle of mass by the wave function: e (where we have put
~

m, energy

and momentum

is represented

~(PX-Et)
= e

iP~xP

1).

When we quantize a field theory of a free


.~efore

field, we expect that the field represents particles (quanta)

we might look for solutions to the equations of motion on the form:

(3.19)

~(x)

=e

ip

x~

where

and we might interpret these solutions as the wave functions of the quanta of the field. For instance, if we quantize the Maxwell field, we expect it to represent massless particles theory of fields. of the form

photons.

(You might be

worried about the fact that we allow comp lex solut'ions in a classical If you are very worried, you may look for solutions Cost p ~XP)I)

90
Let us summarize our expectations: perties:
11 (3. 20) 21 3}
L H

We want to construct a free

field theory based upon a Lagrangian density with the following pro-

is a Lorentz scalar is positive definite

The equations of motion allow solutions on satisfies the the form: e ip xlJ where P]J lJ 2 dispersion relation p plJ =_m
)l

Property 1) quarantees that we are constructing a relativistic theory, 2) guarantees that the energy density is positive, and 3) guarantees that if we quantize the field, it will represent free particles with a mass m, where

3.3

HAMILTONIAN FORMALISM FOR RELATIVISTIC FIELDS

In analogy with the Hamiltonian formalism for a particle we can now associate to each component of the field a conjugate momentum defined by:
a _ aL " - a(ao~a) Here "a should be considered as a function of ~a' ao~a' ai~a' If we can invert tne relation and obtain ao~a as a function of ~a' "a, ai~a we say that the system is canonicaL
(3.21)

Exercise J. 3 1
Problem: Consider a canonical system. to the following HamiLton equations:
(3.22)

Show that the Euler equations are equivalent

~ a"a

a~
0

aH

a~a

_ a "a
=

i~ ~ a

aH

Consider a suitable function space M. A functionaL F is a map, F : M -+ R that maps a function into a real number. If M is the set of all smooth functions f : [a,b]-+R then we can for instance consider the following functiona1s:
b

F1[f] =

Jf(x)dx
a

Fz[f] =

J!(f'(xZdx
a

If the functional F is sufficiently nice, we can introduce a jUnctionaZ derivative. Consider first the case of an ordinary smooth function , f : Rn -+ R .

91

Let

Xo

be given.

Then we can expand f(x O


+ y)

in a neighhourhood of

Xo

To formalise this we consider the map: (xo + Y)[=O we can write it in the form:
+

aE

af

This is a linear map.

Thus

+ y)

a. (x
1

)yi

Here the coefficients rivatives of f:

depend on

xQ

and actually they are the partial de-

at a.(x o) ~ -.(xo) 1 axl This motivates the following definition: consider a functional F. Let fa be given. Then we can expand bourhood of fa F [fa + g] = Ff o ] + ~: (x) g(x}dx + If=fo To formalise it, we consider the map:

in a neigh-

This is a linear map.

g + aE [fa + g] Thus, we canl~~te it in the form:

of

1=0 The function k(x) depends on of F at the function fO :

aa

F[fo+ g]

J k(x)g(x)dx

fa

and we define it to be the funationaL dePivative


= k(x) 5F Oflf=f

N.B. When we perform the variation: g vanishes on the boundary.


Exez>aise 3.3.2
Problem:

fo

fo + g

then we will always assume that

Consider the following functionals: t2 df 2 F} [f] = [jm(dt) - V(f]dt , F2 [f] = f(x O)

J
t}

Show that the functional derivatives are given by:

of 2
Of

- 5(x-xO)

,If =

5F

5'(x-xO)

Exeraise

3.3.3

Problem:a)Show that the Euler-Lagrange equation can be written in the form:

(3.23)
where

5S _ 0 oq,a
=

92
b) Show that Hamilton's equations can be written in the form:
(3.24)

oH

o~a = -

at

alia

6H a<f>a 611a = at"

where

Let

and

G be two functionals of the field components and their conjugate F[<f>a;lI a 1 =

momenta, i.e.

f F(<pa;ai<Pa;lI) a
R3

d 3-+ x

and a similar expression for logy with (2.59):

G.

We may then define their Poisson Bracket in ana-

{F;G}

Exercise J. J. 4
Problem: Show that the Poisson Bracket of two functionals satisfies the properties listed in exercise 2.10.1.

Exercise J. J.5
Problem: Consider a field theory. through the formula: For fixed (t,lt}) we can define a functional <f> a -+
<f>

This functional which depends on t and will simply be denoted as In a similar way we can define a functional: lib -+ lIb(t;~ ) simply denote as 11 (t,x 1 ). Show that the functionals satisfy the rules:
b
+

'S

a (t,~})

~a(t';;:l)

and

{<f>a(t'~l); <f>b(t,X 2 )}
Rearrange the functionals

{1I

(t,x
6 3

-+

);

11 (t,x )} = 0 2

b-+

6~
Hint:

(it} - ~2)
in the form
-+

and (x - Xl) f b -+3-+-+ f 11 (t,x)6 (x - x 2 )


~a(t,x)o
-+3-+

<f>a(t,~})
b .... 11 (t,x ) 2 and use this to show: o<pa(t,xl )
6~b

d3~

d 3;t

ob 63(;t_~) a 1

o<l>a(t,x 1 ) 61!b 6l1 b (t,; )


2

b -+ 611 (t,x) 6<f>a

611
-------

b 3 6 6 a

(x - it2 )

93

Exeraise

3.3.6

Problem: Consider a canonical system. Let F be a functional of the fields and their conjugate momenta. We will also allow F to depend explicitly on time:

The field components $ and their conjugate momenta themselves evolve in time according to Hamilton's e~uations (3.20). Show that the total time derivative along a field history is given by: dF of dt = at + {FIH} Let us make a short comment on how to quantize a field theory. quantum mechanics we can use two different strategies:
(a)

As in the elementary

Path-integraL formaLism:
This is closely connected to the Lagrangian formalism. configuration $a (])(;) at time Consider a field t1 and another field configuration

~ (2)(~)
a a

at time t2 at time

We are interested in the transition ampLitude t2 when we know that it was in the state $ (1) $a (I) the field

<: (2)1$ (1, i.e. the probability amplitude for finding the field in the state at time $a (2) To come from the configuration to $ (2)
a

must have developed according to some history between $ (1) and $ (2):
a a
$ (1) 6~) a

$a(t,~)

which interpolates

~ ( +) "'a t 2 ,x

= "'a ~ (2)(x+)

To each such history we have associated the action: t2 S[$a] = L ($a,oj.l$a}d3~ dt

f J
tl R3

A straightforward generalization of Feynman's principle (2.21) then gives the result: ~a(t2,i)=~~21(~) (3.26) <$a(2)1 <l>a(1)>

exp{~S[~a]} D[~a(t,i)l
$ (1) a and $ (2). a

~a (tloi)~~ 1) (~)
where we sum over all histories interpolating between
(b)

CanoniaaL quantization:
This is closely connected to the Hamiltonian formalism. Consider a canonical field theory. The physical quantities are represented as funationaLs F[$a,~b] of the field components and their conjugate momenta. In the quantum context these functiona1s are replaced by Hermitian operators in such a way that their Poisson Brackets are replaced by commutators: (3.27) {F ,G}
-+

1 [~,G] .." if!

As we have seen the field components ~a(t'~l) and their conjugate momenta n b (t,i2 ) can themselves be interpreted as functiona1s. According to exercise 3.2.7 they must, therefore, be replaced by Hermitian operators satisfying the Heisenberg Commutation RuLes: ['a(t'~l) ~b(t'!2l1 = [wa(t,xll ; h (t,x2 )] - 0

[~a(t'~l}

;b(t'!2)] -

iflo~o3(;2-il)

94
As an illustration of the preceding ideas we take a look at the following Lagrangian density

(3.28)
This is obviously a Lorentz-scalar and if we write out the first term explicitly,
L(</>

, a)l <j

= ~[(l!)2_ at

(l!)2j - U(<j ax

we see that it is a direct generalization of the string-Lagrangian (3.7). The last term can be interpreted as a potential energy density. It can easily be included in the string model too if we simply assume that each atom in the string has a potential energy U(x) arising, e.g., from the gravitational potential. Notice too that the above Lagrangian has the same form as the non-relativistic Lagrangian for a pointparticle (2.5). With this choise of the Lagrangian the conjugate momentum is given by

1T=~
at i.e. it coincides with the kinematical momentum. Furthermore the Hamiltonian density reduces to

(3.29)
This is positive definite, as it ought to be, provided the potential energy-density is positive definite. In the following we will assume that the minimum of U is zero, and that this minimum is attained only when ~ vanishes identically. ~e Lagrangian therefore satisfies the first two conditions in (3.30) which must be valid for any field theory if we are going to make sense out of it. The third condition is specifically related to free-field theories. In our case the equation of motion is given by )l a2~ a2~ (3.30) a~a <j> = U'(<j i.e. at2 - ax 2 = -U'(<j which should be compared with Newton's equation of motion. If we make a Taylor eXpansion of the potential energy density, using that U(O) and U,(O) vanish by assumption, the equation of motion reduces to a a)l<j> = U"(O). + )l
~ U'"(O).2 +

This will only allow solutions of the form <j>(x) = Eexp[ip x)lj provide~the potential is purely quadratic. That follows immediately from the fo~ulas: a a)l<j> _p p)lEexp[ip x)lj )l )l )l U'(<j {U' '(0) + W"'(O)Eexp[ip x)lj +.;.}Eexp[ip x)lj )l )l

In the case of a quadratic potential we get the dispersion relation

_p p)l = U' '(O)l )l which shows that U',(O) is the square of the mass of the particle in the theory. In the general case Eexp[ip x)lj will never be an exact solution, but it will be an approximative solution pro~ided is so small, that we can neglect the higher order terms.In the weak field limit we can therefore treat the field theory with a general potential as a free field theory, where the mass-square of the particle in the theory is still given by U'(O). As the field quanta in the general case are no longer free they must exert forces on each other. We therefore say that they are selfinteracting. Notice that the free field case,where U is quadratic, corresponds to a linear equation of motion, while the self-interacting case corresponds to a non-linear equation of motion.

95

3.4

THE KLEIN-GORDON FIELD.


Now we are in a position where we can construct our first explicit We start by constructing the equations of motion. , this leads to the equation of motion Con-

field theory. stitution

sider the energy momentum relation: p~p~ =_m 2 .


p~ =-ia~

If we perform the sub-

(3.31)
This is known as the invariant.
K~ein-Gordon

equation and it is obviously Lorentz

To check that it has the solutions we want, we observe that

(a~a~ - m2 ) eXp[ip~x~l = (- p~p~- m2)exp[iP~X~1


and this shows us that Gordon equation provided equation of motion </>(x) = exp[ip~x~l is a solution to the Klein2 p satisfies: p p~ =_m
~ ~

Now we must find the Lagrangian density

L.

We know that the

must reproduce From this we immediately read off: a

(~~</ =-a~</>

.... L

=-~ (a~</ (a~</

+ terms involving

</>
a~</>

~=_m2</> .... L = _1 m2 </>2 + terms involving o</> 2 So we have reconstructed the Lagrangian density:

(3.32)
Since </> is a Lorentz scalar, the Lagrangian density is a Lorentz Finally, we should check that the energy density is po-

scalar, too.

sitive definite:

Le.

(3.33)
This is obviously positive definite, so everything is okay! Let us look a little closer at the Lagrangian density. dratic in
</>

It is qua-

and

a~</>

, which is typical for a free fietd theory.

96

Furthermore, it consists of two terms: a) A term,- ~ (a </ (a l1 </, which is quadratic in the derivatives and which acEs ~ike a kinetia energy term. Observe, however, that it includes the space derivatives, too: -1(a </ (a).l</>l = 1 (1.1.) 2 _ 1 (V</ 2 2 ).l 2 at 2 If you compare with the model for the continuous string (sec. 3.1), you see that (v<j 2 should be counted as a potential energy term, due to tfie local interaction. A term, - ~ m2 <j>2 ,which is quadratic in <j>. It is called the mass term, because m gives the mass of the field quanta. Observe the sign! It will be crucial later on.

b)

E:x:eraise

3.4.1

Problem: Determine the conjugate momentum of the Klein-Gordon field and show that the theory of the Klein-Gordon field is a canonical field theory. Determine the Hamiltonian functional

H[</>,n]
and verify by an explicit calculation that Hamilton's equations reproduce the KleinGordon equation.

At this pOint we have only looked for solutions on the form: </>(x)

that had an important significance on the quantum mechanical level. We might look for other solutions. The simplest possible classical solution is a static spherically symmetric solution </> (t,x). '= </> (r) . Using this ansatz, the Klein-Gordon equation reduces to: m </>(r)
2

= exp[ip 11 x).l]

= r dr 2 (r</

1 d

i.e.

From this we find the solution:


(3.34 )

</>(t,i)

= </>(r)

mr

: e

This solution is singular at the origin. Only the decreasing solution is physically acceptable because the contribution to energy from infinity otherwise explodes. How can we interpret the solution (3.34)? In electromagnetism we have a similar solution, the Coulomb solution. There we use the ansatz: <f>(t,~) = </>(r),A(t,x) = 0 where </> and A denote the scalar potential and the vector potential. Using this ansatz the Maxwell equations reduce to

r dr

1: ..i!..2

[rep]

<l>(r) = a

This corresponds to a pure spherically symmetric electric field, the

97

Coulomb field: The field

aA E=-VCP-at
r

b
r2

-+

r
At the point The field

<p<r) =

of singularity, r

!1 =

is interpreted in the following way:

0, we have an electrically charged particle, say

a proton, acting as a source for the electromagnetic field.

itself is interpreted as a potential for the electromagnetic force

E = - Vcp

which other charged particles will experience. interact by exchanging photons.

On the quanThe photons,

tum mechanical level two charged particles will i.e. the quanta of the electromagnetic field, will transfer momentum and when the momentum of particle changes, it experience a force. Let us make the same interpretation of the static spherically symmetric solution of the Klein-Gordon equation. tion between protons. protons. The electromagnetic field Let us assume that the

is responsible for the electromagnetic interacKlein-Gordon field is responsible for the strong interaction between When you quantize the electromagnetic field, you get massIn the same way we assume that you get n-mesons when Two protons (or a proton and a w-mesons: less photons.

you quantize the Klein-Gordon field.

neutron) then interact strongly by exchanging

This exchange produces the strong forces. derived from the potential: cp(r) In other words:

The strong force is then

c
r

-mr

At the position of the singularity, we have a par-

ticle, say a proton, acting as the source of the Klein-Gordon field. It produces the potential cp(r) = ~ e-mr and hence the force which
r

other strongly interacting particles will experience. mr cp(r) = ~ eis called the Yuka~a potentiaL.

The potential

Contrary to the Coulomb potential we see that the Yukawa potential is exponentially damped! Hence, it has only a finite range. Let us estimate this range. If y is a length measured in meters and x is

98

the mass measured in MeV, then in our units where be shown that

=~

it can

x
The observed mass of the range of the force is:
~-meson

is

140 MeV, hence, the typical

2-10- 13 140 meter

= 1,4-10 -15

meter

Now that is exactly the typical distance between protons and neutrons
in the atanic nucleus

am

the Yukawa potential can therefore very well account for the

strong force binding the nucleus together. force was never observed classically.

It also explains why this

In fact, the extranely short range

means that the force operates on spacetime regions so small that the quantum mechanical effects dominate completely.

3.5

THE MAXWELL FIELD.


The next field we attack is the Maxwell field Au Here we know of motion:

the equation
(1.33)

Can we find a simple Lagrangian density which leads to this equation? Since the equation of motion only involves derivatives of,' Au expect that
L

' we

only involves

a~Av

and we should determine it so that:


all

3L _ U v v II II v v II a(aA}-olldA - a (oA)=a (aA - o A ) II v II II

This is a mess, so we will use a trick. gauge invariant,_ ever


All

The equations of motion are This guaranties that when-

Now the Simplest way to construct a gauge-invariant

theory is to use a gauge invariant action.


All + o~X

extremizes the action, then the same thing will be true for Note, however, that it is not the only possibility when

you want to construct a gauge invariant theory (compare the discussion in section 2.2), but let us try it! The starting point is a gauge invariant Lagrangian density Since
L

L.

only involves derivatives of

Aa

we must try to construct But this must be

a gauge invariant quantity out of the derivatives. the field strength:

99

Thus

we expect that

only depends on L = L(P CtB )

F CtB

But the simplest quantity you can construct out of and Lorentz invariant is the square: P So we expect (3.35) where k is an arbitrary constant.
3.5.1

aB

which is gauge

CtB

p CtB

Worked Exercise Problem: Show that

4F\l\l

Using exercise 3.4.1, we see that the ansatz (3.35) leads to the correct equations of motion:

To determine CtB

we investigate the energy-momentum tensor:

a(~~A\I)

aCtA\I

Ct nCtBL =-4k[p B\la Av -

~nCtBpyOpYO}

But here something is wrong: should make the replacement: aCtA\I

It is not symmetric, and in fact it is aCtA \I This suggests that we

not gauge invariant, due to the term:

aCtA\I - avACt = pet

because the new term is gauge invariant. Therefore we try to repair the canonical energy momentum tensor a CtB = Is this legal? Remember that properties (3.18). CtB with the correction term:

4kFB\l a ACt

aCtB

should possess three characteristic CtB+aCtB is symmetric:

Pirst we observe that

~CtB + aCtB =_4k[pBVpCt


Secondly we observe that a CtB

_ ~ netBp pYa} v 4 YO is conserved:

4ka a{ pBVavACt]

..

4k{(a pav)a ACt \I

pBva

B v

ACt}

100

because the field's equation of motion kills the first term, and the antisymmetry of the field strength kills the symmetric tensor Pinally we observe that

0Sov Aa !

e aB

does not contribute to the energy-

momentum of the Maxwell field:

4kJ povovACld3x 4kJ (oiFoi)Aad3X


vanishes at infinity. equation of motion.

4kJ F

Oi

ai Aa d 3 x

4kJ (opFo~}Aad3X = 0

where we have neglected the surface terms because the field strength The last integral vanishes due to the field's

So we have succeeded in repairing the canonical energy-momentum tensor. Compairing the energy-momentum tensor:

a B +
(1.38)

aBp pYa} e aB =-4k[p BV p Cl V _ :1 411 Yo

with the true energy-momentum tensor for the Maxwell field:

we finally obtain the result

~!

Thus we have constructed the Lagrangian density for the Maxwell field:

(3.36)

EM

Exercise

3.5.2 and show that

Problem: Determine the conjugate momenta of the Maxwell field A~ the theory of the Maxwell field is nota canonical field theory.

As we have seen

LEM

is Lorentz and gauge invariant, and it pro-

duces a positive definite energy density:

(Compare with (1.41)). motion:


(1.33)

It remains to investigate the equations of

We look for solutions of the form:


(3.37)

Here

is a Lorentz vector, called the poZarization vector.

If we

insert the wave function into the equations of motion, we get:

101

o=
Hence A)J vided:

(-P)Jp)Jv + pv(p)J)J

exp[ip)Jx)J]

)Jexp[ip)Jx)J] is a solution of the equations of motion pro(P)Jp)J)v = (p)J)J)Pv .

This algebraic condition has two kinds of solutions: and Let us examine the condition al. If b) p )J p)J ~ 0 , we conclude:

(p)J)Jl (p)JP)J)

Pv

Consequently v is proportional to Pv and we see that the equations of motion allow solutions of the form: (3.38)

But these solutions are trivial. can gauge them away! AV' We therefore Choosing Av + avx
X

They are not observable because we


i.cL exp[ip)Jx~],we get:

disregard them! b). If P)JP)J

Then we are left with the case

o ,

we conclude

but this is only consistent if p)Jj1 = 0 Thus , we conclude that the equations of motion allow solutions of the form:

(3.39)

with

and

Since

P)JP)J

= 0,

the photons are massZess!

These solutions are However, it

non-trivial, i.e. we cannot completely gauge them away. is not all the components mation. X

)J,)J = 0,1,2,3 , which carry physical infor-

To see this, we make a gauge transformation choosing This leads to the transformed potential:

iaexp[ipjJx)J]

102

+ a x = E exp[ip xll] - ap expUp xll] = ( - ap lexp[ip xll] v v v II v II v v II So if we perform a gauge transformation, we can change the polarization vector according to the rule:
A' v
= A

(3.40l

To investigate the consequences of this gauge freedom, we consider a wave travelling along the z-axis. pll where pllEll w Then the four-momentum is given by (w,O,O,kl Due to the condition

= -k

because the photon is massless.

=0

, we conclude:
o
=
3

Thus , the wave function of the photon has the form:


All

[E 3 ,E 1 '2'3] exp[iw(z-tl]

But we are still allowed to make gauge transformations:

Hence, i f we choose z-axis


A = eiw(z-t)

3/W

we

have gaugEd away pletely!

EO and 3 comIf we split the polariza-

pf

A =~
q,

II

II

longitudinal

eiw(z-tl

r ...

part
+

tion vector into its different components, then: 0 1 2 3 is called the scalar part

=Oeiw(z-t)

11

ll= (o,~ l

0 ,+

.1.

is called the transverse part is called the longitudinal part

~=~
Fig. 41

y-axJ.S

The above result then shows that we can gauge away the scalar photons and the longitudinal photons, but we cannot change the transverse part. Consequently all the physical infoonation is con-

tained in the transverse parts

and

2 !

E:r:eraise

3.4.3

Problem: Compute the complex field strengths corresponding to the above travelling wave and show explicitly that they only depend on 1 and 2

103

3.6 SPIN OF THE PHOTON - POLARIZATION OF ELECTROMAGNETIC WAVES


Let us investigate the physical meaning of closer. ing way: [0'1'2'3] where = b( 2 1 0[1,0,0,0] + +[O,l,i,O] + _[O,l,-i,O] + 3[0,0,0,1] Each polarization vector 11 1 and 2 a little may be decomposed in the follow-

+ i2 l

We may decompose the wave function All ll exp[ iw (z-t) ] We want to show that this decomposition is To see this we must investigate what happens

in exactly the same way. closely related to spin!

if we perform a rotation about an axis, say the z-axis. Now a rotation is nothing but a special kind of Lorentz-transformation:
X,ll

= aY

XV

represented by the matrix (allvl where


1

o
Cos 8 -Sin 8

o
Sin 8 Cos 8

o o
a
1

all

o o o

Here

is the angle of rotation.

In the new coordinate system we

have a new coordinate representation of the wave:

with

'=aY v II v

and (~llvl the reciprocal matrix of (all v}. Hence, the rotation only attacks the polarization vector. We now get,

[1,0,0,0] ... [1,0,0,0]

: 0 [0

o
Cos 8 Sin 8

o
-Sin 8 Cos 8

~1

[1,0,0,0]

104

and similarly for [O,O,O,lJ. Finally [O,l,i,OJ transforms into

o
Cos 8 Sin 8

o
-Sin 8 Cos 8

o
o
[0,Cos8iSine,-Sin8iCose,OJ

o
[O,l,i,OJ

o o

a
1

o
RS

= exp(ie).[O,l,i,O]
Thus, if we let denote the rotation operator, then: [ 0,0,0,1] exp[ihl (z-t) ] [1, 0 , 0 , a ] exp[ ihl (z-t)] and

are eigenfunctions with eigenvalue 1, and [0,1, i, 0 J exp[ihl (z-t)] and [0,1, -i, 0 Jexp[ihl (z-t)]

are eigenfunctions with eigenvalues:

is is You then see that

But the rotation operator is given by: the operator of angular momentum around the z-axis. projection

the wave function for a scalar or a longitudinal photon carries spin 0 , while the wave functions for the transverse photons You should, however, remember that only the transverse photons are observable! Thus, a photon is a spin l-particle, and i t has carry spin projection l.

spinpro-

------..
jecti~ I

spinp')ojection +1 either spin projection 1 in the direction of the momentum or it has spin projection -1 ~ p photon in the direction of momentum, but i t never

...

has spin projection O!

We therefore say

that the photon has heZicity l.

The preceding discussion may seem to bear little resemblance to what you have previously learnt about electromagnetism. Maybe the following remarks will clarify this;. We have looked for solutions on the form:
(3.37)

~(x) = ~ exp[ip~x~]

The above solution is complex valued and hence, it can have relevance only on the quantum mechanical level. Rut clearly the real and imaginary part will solve the Maxwell equations too, and they are real solutions so that they may have relevance On the classical level. We want to find out how we can interpret the above solution classically and quantum mechanically. Let us discuss the classical interpretation first. We know that Maxwell's equations allow plane waves as solutions If we introduce a coordinate system, where the 3-dimensional wave vector along the z-axis, this formula reduces to:
A~(X)

points

= ~Cos(wz-wt)

105
We may decompose it into a scalar part and a vector part: AP (~,it) ~ gOCos(wz-wt)

it

;:Cos (wz-wt)

But as we have already seen, we can gauge gO and g3 away! If we work in the special gauge, where gU = g3 = 0, the scalar potential drops out, and the vector potential is represented by:
(3.41)

A=

;Cos(wz-wt)

where ;: = [,1,,2,0] is orthogonal to the wave vector We can also calculate the field ,strengths. Observe that since it is time dependent, this wave will also represent an electric field:
(3.42)

k.

R= g x it - -(;Xk) Sin(wz-wt) E = -~~ =;w Sin(wz-wt)

Here the electric field is pointing along the polarization vector and the magnetic field is orthogonal to bot~ the wale vector and the polarization vector. As B and E are pointing in constant directions in space, we speak about linearly E polarized light. ' The density of momentum g is pointing along E Poynting's vector (C=EO=l): 2 ~~~~~~~~~~~~ g = Sin (wz-wt) Hence, there is a simple connection between the momentum density g and the wave vector k which represents the momentum of a single photon (P--hk). Similarly, the energy density is given by (c='o=l) ~ 2 ~2 dx,t) = i(l> +B )=w2 Sin 2 (wz-wt) Let us consider a big box at a certain time, say t=O. Assume that the box contains N photOns. Each has an energy given by ~w and a momentum given by ~k. i.e. in relativistic units, where 'o=c=~ = I, we see that the total energy and momentum is given by: E = Nw , P =Nk tot tot Now let us compare this with the similar classical calculation. Here the total energy is represented by: w2V E = Sin 2wzdxdydz w2 ,(to)d 3x -2tot

...

EXR kW

15tot

Jg(~,t)d3x

J wk J S'in2wzdxdydz
~

wkV -2-

where V is the volume of the box. Thus, you see that the classical and quantum mechanical calculations are consistent provided you put: N = wV 2 We have also found waVes representing spin: [O,l,i,O]exp[i(wz-wt)] We extract the real part and get the classical wave:
(3.43)

A- [

;:i~:::l ]

106

which solves the Maxwell equations. In this case the vector potential is circulating around the wave vector with a frequency w Let us compute the field strengths:

(3.44)

....
E -

- -

at

aA

....
= w

l;!~i~:::~

A=

-cos(wz-wt)] w [

Sin(~z-wt)

Clearly they are rotating too, and we speak about circularly polarized light. Then we turn to the quantum mechanical interpretation of the complex~wave solution: A]:I(x) = Pexp[i(~-wt)] which we interpret as the quantum mechanical wave function for a photon with the momentum =~k and energy E= ~w Again we work in the special gauge where 0 and E3 disappear:

A' (xl -

[l~

"",(i (ki-ll

We may decompose this wave function along the x-axis and the y-axis:

A' (x l

"

[i 1

,,,,wki_ll'"

[! ],""

(i

(ki_) l

As the solutions are only determined up to a factor, we may normalize the solution so that: (1)2 + (2)2 = 1 i.e. 1 = Cos a ;2= Sin e, where 9 is the angle between and the x-axis. Since all states can be obtained as a linear combination of A~ll) and AP(2) , these states represent a frame for the linearly polarized state~. AP (1 ) repr~sents a photon polarized along the x-axis, and AP(2) represents a photon polarized along the y-axis. In the general state, A).! (x) = cose A).! (1) (x) + Sin a A).! (2) (x), Cos e will give the probability amplitude for finding the photon polarized along the x-axis and Sin e will give the probability amplitude for finding the photon polarized along the y-axis. An experiment may he performed in the following way: If we have a beam of photons, a light beam, which passes through a polarizer then some of the photons will be going through the polarizerwhile some of them will be absorbed. After the passage, the light is polarized. All photons are now in the same state, they are polarized along the characteristic axis of thepolarize~ Now, suppose we insert a second polarizer which is rotated an angle a relative to the first polarizer. What is the intensity I of the beam after the passage of the second polarizer in terms of the intensity I of the polarized beam? Let us introduce a coordinate system whePe the x-axis points in the direction of the axis of the second polarizer: Then the polarized beam is characterized by a polarization vector:

; =

I :::: I
e
AP(I) + Sin

and the wave is decomposed according to

A~ - Cos

AP (2)

107

Detector

I~I
Fig. 43 Polarizer 2 Polarizer 1

Unp:>1arized ligth beam

(Observe that we have a macroscopic number of photons all occupying the same state. This is possible because photons are bosons. Compare with the discussion of superconductivity). Consider a single photon: The probability amplitude that it is polarized along the x-axis is Cos a. Thus, there is a probability cos 2 a that a given photon goes through the second polarizerl But as there is a macroscopic number of photons in the same state, we may actually interpret cos 2 a as a macroscopic physical quantity. If Ni is the total number of photons before the passage of the second polarize~ and Nf is the number of photons after the passage of the secondpolarize~ then: Nf ~ Ni cos 2a . As the number of photons are proportional to the classical intensity of the light beam, we may rearrange this as,
(3.45)

and that is a formula which you Can test very simply in a classical experiment.

3.7

THE MASSIVE VECTOR FIELD


We have seen that the Lagrangian density of the Maxwell field only Correspondingly, the

contains a kinetic energy term and no mass term. photons are massless!

Now, suppose we add a mass term to this Lagran-

gian density and consider the following field theory:

(3.46)

The first thing we observe is that we have spoiled the gauge invariance.Therefore ACt and this field ACt is not a gauge field. So in what follows: Let us find ACt + 0Ct X represent different physical situations. the equations of motion. Using (3.14) we get:

m2 A f!
These equations imply that:

m2aSAa = aadetFetS= 0

108

Hence, the Lorenz condition, for the massive field. siderably: (3.47) (a)Ja)J m2 )Aa = 0

aaAa

0, is automatically fulfilled

But then the equations of motion simplify conand aaAa = 0

Each of the components Ail thus satisfies the Klein-Gordon equation, and the field Aa will clearly represent free particles with mass m. the components the Loren<: pendent. Ao We call Aa All a massive vector fieZd. Observe, however, that are not independent, as they are connected through

condition. Therefore only three of the components are indeAs we shall see in a moment, this means that the scalar part The reamining three degrees of freeUsing exactly the same method as we did in The projection of the spin onto the and this time the longitudinal

can be eliminated completely.

dom correspond to spin. particle is a spin

connection with the Maxwell field, we can show that the massive vector 1 particle. z-axis assumes the values -1, 0, +1 part also contributes!

The massive vector field is not a gauge field.

Worked exercise

3.7.1

Problem: (a) Calculate the canonical energy-momentum tensor for the massive vector field and show that it is not symmetric. (b) Show that we can repair the canonical energy-momentum tensor by adding the following correction term: aaS =-0 (FSPAa )
P

leading to the true energy momentum-tensor


(3.48)

(c) Show that the energy-density momentum tensor is positive definite.


'lb conclude

TOO

corresponding to the true energy-

we have presented

very naive semi-classical arguments to mo-

tivate three kinds of fields:

Name (3.49) (3.50) Klein-Gordon field Maxwell field Massive vector field

Lagrangian density - ~ (0)Jq,) (allq,) _ ~ m2q,2

Equation of motion (O)J all - m2 )q,

- '4
1

Ila

F aS FaS _ 1 m2AaAa

(alldll)Av- dv(OpAll) all - m2 )A Il = 0


Il

=0

(3.51)

- 4

as

'2

(0

II

aaA

=0

and we have seen that they represent spin 0 particles, massless spin 1 particles (photons) and massive spin 1 particles (vector particles).

109

3;8

THE CAUCHY PROBLEM


In what follows we will treat our fields as classical fields, i.e.

we will only admit reaL solutions to the field equations. The first thing we will study is the Cauchy problem. If we have a single particle moving in a one-dimensional space, it has the equation of motion: Now, fix a time change of x tl
d 2x m dt 2

= -v

(xl x , and the rate of

and prescribe the value of dx dtlt

at this particular moment: x(tll

l What can we say about the dynamical evolution of the system once we have prescribed these initial data? This is the Cauchy problem. The above differential ex and the rate of change x In this case there is a simple answer. solution once we have specified the value of of x at a given moment.

xl

VI

quation is an ordinary second order equation, hence, it has a unique It is because of this that we say that

is a dynamical quantity and that the system in question is uniquely characterized by this single dynamical t-axis
SPACE-TIME
DL~l

quantity. Now let us look at the Klein-Gordon field. Again we fix a time tl and we prescribe the value of the field and the rate of change at that moment: '

y-axis

)r
Fig. 44

What can we say about the dynamical evolution? Well, let us write out the equa tions
0

f motion :

(3.49)

or

110

This

is a second order equation anl thus it has a unique solution once $ and

we have prescribed

O<P at

at a given moment.

In fact, it is easy

to see that we may construct the dynamical evolution directly by using an iteration procedure. We know the value of $ and a~ at the time at t) We may then find them at time t = t) + at using the formulas: + + a$ ..$ (t) + at,x) "" $ (tj,x) + at at (t),x) a$(t + at )

at,~)

"" aa$t(tj,X) +

at~(t ,x) at2 )

+ a$ 2 at (t),x) + ot[M (t),x) _m $ (t),x) 1 a$ But when we know $ and at at the time t = t) + at , we may repeat a$ at the time t the procedure to find $ and t) + 2at, etc. Thus at we can reconstruct the dynamical evolution of the Klein-Gordon field.

This shows that the field variable ly characterizing the system.

is a dynamical quantity completeAu

Then we look at the Maxwell field and prescribe the values of Au and

we say about the dynamical evolution of the Maxwell field? Nothing! And here is the reason why: We know that if we find a solution Au to the equations of motion, then the gauge transformation:
Act
+

We select a time t = tj at this time. Now, what can

Act

auX

produces another solution.

Thus, if we choose

X so that it is zero

in a neighbourhood of t t ) , this will not disturb the boundary conditiond and you see that there is infinitely many solutions to the equations of motion, all satisfying the boundary conditions. On the other hand we have seen that the equations of motion are second order differential equations (avaV)A V - aV(avAV) so you might wonder what is Wrong? There is a subtle reason for this. Remember that the full equations for a Maxwell field with a source look as follows
(1. 33)

These equations automatically guaranteed the conservation of charge:

But this conservation has a price. rator

It means that the differential ope-

111

o beys a special identity: (3.52) Suppose we isolate

o
00
on the left side:

o [(0

all)AO II

This equation is interesting. left! Removing

On the right we have time derivatives

of maximum order two, but then the same thing must be true for the

ao

we see that:

(allall)AO- aO(oIlA ll )
can only contain time derivatives of maximum order one! sider the equation of motion for the scalar part
(ollall)AO AO

So if we con-

0 (oIlAll) = 0 ,

we observe that there is something wrong. equation. (3.53)

It is not a second order

By explicit computation you can easily find that it reduces to

Le.

VE =

This is not an equation of motion at all! the scalar part even prescribe
~

This shows us clearly that In fact, we cannot t


+ l ,

is not a dynamical quantity. and


~a

tt

at the initial moment and


at (t l ,x) 0'" at (V
7

because once

we have prescribed:

+ oA

then

has to obey:
I'l~ =
-

A)

For this reason we say that equation (3.53)


straint.

is an equation of con-

This leaves us with the three remaining field variables

as dynamical variables.
Le.

We may rewrite their equations of motion as:

(0 all)Ai - ai(a All) = 0


II II

(3.54)

aA = flA ot 2

v[v . A +

~]

so the Maxwell equations reduce to one equation of constraint and three equations of motion. blems. But we still have not solved our dynamical proTo get rid of this we We are still allowed to make gauge transformations and thus the It is tempting to use the Lorenz gauge,

dynamical evolution is completely indetermined. must choose a specific gauge.

112

(1.22)

not only because the Lorenz condition is Lorentz Using the Lorenz condition we can now eliminate

invariant, but also

because the massive vector field automatically obeys this condition.

too.

That solves

our dynamical problems:

... A

The only true dynamical quantities are the three space components They are governed by the equations of motion:
(3.55)

a21 at 2
~

The scalar part

is eliminated by the equation of constraint

and the Lorenz condition:


(3.56)
We should, however, be a little careful with the iteration procedure. There are no problems with the A-variables. But consider the scalar part. It should satisfy (3.56) at all times, if our method is to be consistent.

Worked exercise

3.8.1

Problem: Use the iteration procedure and the equation of motion to show that the equation of constraint and the Lorenz condition are automatically preserved at all times. From this exercise it follows that the method is in fact consistent. You might also be worried about the fact that we can still perform gauge transformations, provided
(1. 23)

X satisfies:

Thus , you might think that we could repeat the argument from before to show that there are infinitely many solutions to the equations of motion, all satisfying the boundary condition. But this is not so. The only solution to the equation (1.23), which satisfies the initial data~ x(it,O) = 0 is the and

x-function which vanishes identically!

In the case of the massive vector field you can show by a similar analysis that only the space components Ai are true dynamical quantities. This time the Lorenz condition is automatically satisfied and therefore the scalar part AO is automatically eliminated. Observe that when we constructed the Lagrangian density for the massive vector field:
(3.46)
L

we directly generalized the Lagrangian density for the gauge potential.

113

Especially we kept the gauge invariant kinetic energy term.


L

= - ~ (a~Aa

aSA~)

a (aa AS - aa A ) -

~ m2A~Aa

But the massive vector field is not a gauge field, so you might wonder if it would not have been much easier to start with the following Lagrangian density:

(3.57)

This would be a direct generalization of the Klein-Gordon field. puting the equations of motion you easily find:

Com-

(3.58)

Consequently each of the components satisfies the Klein-Gordon equation. Thus, in this case, all four components are true dynamical variables. But why did we not choose that Lagrangian density? have seen, the scalar part the vector part Ai AO represents a spin I-particle. represents a spin Because, as we a-particle, while Hence, if we 0and spin

adopted the Lagrangian density I-particles!

L' , we would construct a field theory

which, when quantized, would represent a mixture of spin gy-momentum tensor.


Exercise 3.8'.2

Furthermore, this would lead to troubles with the enerThis is the content of the following exercise:

Problem: Calculate the canonical energy-momentum tensor corresponding to the Lagrangian (3.57). Show that it is symmetric but that the energy density is indefinite.

3,9

THE COMPLEX KLEIN-GORDON FIELD


Now suppose that we have two real classical fields:
~1

and

~2

We will base their dynamics upon the Lagrangian density:

(3.59)
Clearly there are no interactions between the two Klein-Gordon fields as their equations of motion decouple:
(3.60) (dllo ll 2

m )<pl = a and

(dlla ll - m2 )<p2 = 0 We fuse the two real

Now we perform a trick which is very useful. fields


<PI

and

<P2

into a single compZex field:

114

(3.61) We can then rearrange the Lagrangian density as follows:

(3.62)

This is only a fancy way of writing the same thing, but what about the equations of motion? vary
~l

To derive the Euler-Lagrange equations we should

and

~2

independently, getting:

~=

oL

HI

)J a (o)J </11)

and

n;

aL

aL )J 0 (a)J </1 2 )

But due to the formulas:


~

~l + i</l2
~1

4i
of motion: (3.63)

and

~1 =
~2 =

~ (</I + 4')

H2

~i(~-q;)

these equations of motion are equivalent to the following equations

aq;

oL

oL = o)Jo(o)J~)
~

and as independent variables:

where we formally treat


L

and

L(~,~,a)J~,o)J~)'

For instance, we get in the above case (3.62)

o
or

aq;

1 - 2

m 2 ", 'I'

+ 12 a )J a)J", 'I'

o =

~[o)Jo)J

m2]~
reproduces the Klein-Gordon equation for So the complex field obeys the Klein-

which, if you decompose Gordon equation too:

~!

the real and imaginary part.

(3.64)

0,

</I = ~ 1 + i~ 2

Now, returning to the Lagrangian density: (3.62) you observe that i t is invariant under the substitution: (3.65)
~ (xl'" e

-io.-

~ (x)

115

Thus, we have discovered asymmetry. But when there is a symmetry, there should also be a conservation law. (This is known as Noether's theorem). Let us examine this a little closer. Let

be an arbi-

trary space-time region, and consider the corresponding action: (3.66) 5ince
L

5 =

J L(<p,~,a)J<p,a].lq;)d4x
n
<P
+

is invariant under the substitutions

-ia<P

we conclude that:
5(0.)

I
n

L(eia<p,
a.

e-ia~, eiao)J<P' e- ia a)J)d 4 x


Hence, differentiating with respect to

is a constant function of

we get:

5 '(0)

But since

J ~L
.[

r r (<P ~ aL
oL

+ a,j, OL ) _ (" o~ + 0 aL )] d 4 ].1"'0 (0)J<P) '" 0<P )J<P a (a].lf) x

was chosen arbitrarily, this is consistent only if:

~ (<Pa;p+ d)J<Pa(a)J<p

oL

(<p af + d].l<Po(a].liP

dL

oL

Using the equations of motion we can rearrange this as:

o
i aL a].ll<p~ )J

<P 0 (0)J~)

oL

Consequently we conclude that because of the symmetry and the equations of motion, the following four-vector, (3.67)
J].I

i[ <Pa_a_L _ (a)J<pl
J].I

_ - __ 0 __ ] <Po (a)J~l

will obey the equation of continuity


(1.29)

For this reason we will call anything to do with to


Q =

a current, although it need not have In the same spirit we will refeI

electromagnetism~

JJ d x
O 3

t=t O as a charge, although it is not necessarily an electric charge!

116

A symmetry like:

~(x)

eia~(x)

is called an internal symmetry,

in contrast to a space-time symmetry where the Lagrangian density is invariant under a coordinate transformation:
yl.l = al.l
XV

But internal symmetries and space-time symmetries have one thing in common:
JIJ :

They produce conservation laws!

For instance, we have just leads to a conserved current

seen that the symmetry

~(x)

eia~(x)

Exercise 3.9.1
Notation: Let Wp = A\.l (1) + iA\.l (2) be a complex vector field. Let G\.lV denote the corresponding complex field strengths: G\.lV = dpW V - dVWp The Lagrangian for the massive vector field is immediately generalised to 2 L = - ! G GPv - ! m wWll 4 \.lV 2 \.l Problem: (a) Show that the Lagrangian is invariant under the internal symmetry: Wp + e iCl W W)J + e -ia w\.l ll (b) Show that the corresponding conserved current is given by:
JV =

i(w
2

\.l

C;)Jv -

W G\.lv) \.l

3.10 THE THEORY OF ELECTRICALLY CHARGED FIELDS AS AGAUGE THEORY


Let us now examine the complex Klein-Gordon field:
(3.62)

In this case the conserved current


(3.68)

J\.l

is:

Can we give a physical interpretation of this current?

Suppose

you want to construct a theory of an eZectrioally charged field, then you would certainly expect that we could define a reasonable fourcurrent
JI.l , and this electric current should satisfy the equation

of continuity
(1.29)

a]J J\.l =
~

But then the complex Klein-Gordon field is an obvious candidate. ever, the complex field cal fields
~1

How-

= ~1 + i~2

actually consists of two classi-

and

~2

Hence, if we quantize it, we expect that

117

it represents two kinds of particles! Can we understand this in a simple way? Yes! In a relativistic For theory we may have production of a particle-antiparticle pair. instance, a photon may, split into an electron-positron pair: y-+e gative charge. +e +

and if the particle has positive charge, then the antiparticle has neThus , if you want charge conservation, the theory of a charged field must represent both of these particles, one with positive charge and one with negative charge! This makes the complex Klein-Gordon field an even more obvious candidate. But if q, is going to be a charged field, it must interact with the Maxwell field. In fact the current J~ must act as a source for the Maxwell field. The question is then: Can we construct a suitable interaction term for the total system consisting of the charged field and the electromagnetic field:
L

LA, + LI + L
'I'

Aa

=-!(~) (o\lq,) I m2ijiq, + 2 ~ - 2:

LI -

! 4

as F

etS

To answer this we may take advantage of the gauge symmetry!


(1.3J.)

We know

that the theory has to be invariant under the gauge transformation

but then we can use what we have learnt studying quantum-mechanical systems. The combined theory is gauge invariant provided we exchange the usual derivatives (2.44) Therefore we suggest the following total Lagrangian:
(3.69) L

0a

with the gauge covariant derivatives:

= ~[(o~ -

ieAjl)q,][

(o~ - ieA~)q,] - I2

m $<1> -

2-

Fas F

as

This is invariant under the combined gauge transformation:


(3.70)

Using this total Lagrangian, we have, in fact, gained something. In the original Lagrangian we were only allowed to make the transformation, (a) where a was a constant. to make the transformation, (b) where a(x) q,(x)
-+

eiaq,(x)

But in the combined theory we are allowed q, (x) ... eia(x) q, (x)

is space-time dependent.

It is customary to refer to (a) as a gauge transformation of the

U8

first kind and to (b) as a gauge transformation of the second kind. Hence, incorporating the second kind. We have seen that there is a natural way to incorporate the interaction between the complex Klein-Gordon field and the Maxwell field. In what follows we shall refer to
~axwell

field, we have extended the symmetry

from a gauge symmetry of the first kind to a gauge symmetry of the

cp.

= cp 1 +

iCP'2

as a charged Klein-

Gordon field and we shall base the theory upon the Lagrangian density:

(3.69)

Now, what about the current? involving All:

Since we have extended the Lagrangian, It may contain terms All

it is not necessarily equal to (3.68) any more.

+ terms involving

In fact, the old expression must necessarily be wrong because it is not gauge invariant! derivatives
(3.71)

This suggests that we simply exchange the usual

all

with the gauge covariant derivatives

Dll

giving

Jll

~ !t$Dllcp - ~DllcpJ
2

This is a real, gauge invariant quantity and in the absence of the Maxwell field it reduces to the old expression (3.68). From Is it conserved? The easiest way to see this is to calculate the Maxwell equations.

we get

and

aL aAa
which immediately shows that:

ie [~Da~ _ ~Da$J
'l' 'l'

[ II a - 3 a (dllA)J= II ie -ollF lla ;-(dlld)A 2

[~Da~
'l'

- a 'l'-CPDcpJ

Hence, we read off the conserved electromagnetic current:

(3.72)
which up to a constant factor reproduces (3.6.8).

119

Exeraise

3.10.1

Problem: Consider the charged Klein-Gordon field. can be rearranged as


L

Show that the interaction term

where

J'1

JaAa - e 2A A~$$ ~ is the electromagnetic current (3.72).


I
=

Exeraise

3.10.2

Problem: Consider a charged particle moving in an electromagnetic field. the interaction term (2.14) can be rearranged as (3.73) s = fJaA d 4x
I

Show that

where

Ja

is the electromagnetic current (1.34)


3.10.3
w~

Exeraise

Notation: Consider a charged massive vector field coupling corresponds to the substitution: with
v~wv = a~wv

Here the rule of minimal

ieA~Wv

Problem:

Show that the electromagnetic current is given by:

J~

= ie [w (VUWv - VVwlJ) -

W (V~Wv ~

- vV;ju)

3,11

CHARGE CONSERVATION AS ACONSEQUENCE OF GAUGE SYMMETRY,


<P

We conclude the discussion of charged fields with sane general remarks. We saw

above that it was easy to couple a aomp ~ex field well field Aa If
<p

<P 1

+ i <p 2

to the Max-

was described by the free Lagrangian density:

Lo

Lu(<p,a]J<p,~,alJ~)

then the total Lagrangian density was obtained simply by replacing the ordinary derivatives with the gauge co-variant derivatives and adding the usual piece for the Maxwell field:
L

= Lu(<P,D]J<p,~,DlJ<P)

FaS FaS
We can now give

This Lagrangian density is obviously gauge invariant. Consider the action for the combined system:

a general definition of the current associated with the charged field:

= So

+ S

+ SA

We have decomposed it into the actions corresponding to the free fields


<p

and

Aa , and the interaction

term.

120

Consider the interaction term: Sr = Sr(<i>,A Ct ) rf we produce a small change in the Maxwell field

this will produce a small change in the interaction term: Sr Here OSr
->-

Sr + oSr oACt , so we can write it as

will depend linearly on

(3.74)

J J Ct (x)oACt (x)d 4x
n
JCt which measures the response of the interaction

The coefficient

will be defined to be the current! Carrpare with exercise 3.10.2 in the case of charged particles. We nrust show that this is in agreanent with our previous ideas: First, we observe that,

(3.75)

Sr =

J Lr(<i>,a~<i>,A~)d4X
Q
a~Av

where the interaction term does not depend on we replace ACt by ACt + soACt Sr(E) = then dS oSr = from which we read off that aLr a ACt

since the presence rf

of the Maxwell field comes from the gauge covariant derivatives. and consider the displaced action, +

J Lr(<i>,a~<i>,A~
Q

EOA~)d4x
oA~ d x
4

dsls=o

J n

aLI
aA~

(3.76)

JCt(x)

This is in agreement with the Maxwell equations = - a F~Ct Ct ~ since the interaction term is the only term which depends explicitly

a~ a(a~ACt)

aL

aL aA Ct

i.e.

aA

aLr

on

ACt . We can now give another argument for the conservation of electric We will only use the part of the action containing <i> :

charge.

121

Sq, = So + Sr This is a gauge invariant action. q,(x) we get: S(E)


+

Performing a gauge transformation Av(x)


+

eie[EX(x)]~(x)

Av(x) + EdVX(X)

ILo
n

(e ie [EX(x)]<P(x) ... )d 4x + ILr(eie[E:X(x)]<P(X) .. ,Av(X)+E:dvX(Xd4X

n
o_
dS" - dE: IE:=O $ solves the vanishes on the

which is constant. Hence, we conclude as usual:

Let us make some assumptions.

First we assume that

equation of motion (3.63). Next we assume that X(x) boundary n This has the following consequence: eie[EX(X)] ~(xl where
~(x)

q,(x) + E~(X)

is a smooth function vanishing on the boundary of

n.

(We are not interested in an explicit expression of ~(x) .) NOw, observe that since q, solves the equation of motion, it will not contribute to we get:

~I dE: =0
o

when we actually perform the differentiation! Thus

de: Ie:=o

r~ aLr dvX(x)d 4 x J
n
V

I
n

J V (X)d v X(x)d x
4

Performing a partial integration, we finally obtain

o
But as x(x)

=-

4 (d vJ V)X(X)d x

was arbitrarily chosen, we deduce

So we caught our equation of continuity! The above argument was very general and abstract! So maybe it is good to summarize the conclusions: We are studying the interaction between a charged field ~ = ~l + i$z and the Maxwell field A~ :
L = Lo(<P,d~$) + LI(~,a~~'IA~)+- LA~(F~S) . '

From the total Lagrangian density we get the equations of motion:


d~aCl~ =

( aea ~ )A B -

( aSd B) ACl = ....

Usually we use Maxwell's equations to identify the current the right-hand side of the Maxwell equation:

JS.

It is equal to

122

(dSdU)A S - (dSdS)A

= ~[~;d~~;~l

The conservation law then follows automatically from the form of the Maxwell equation. We do not use the equations of motion of the charged field at all! The new argument shows that the dynamics of the Maxwell field is superfluous. We need only bother about the ~-field:

Lip = Lo (<P;d):l<jY) +

LI(<Pld~~;Aa)

From this Lagrangian density we get the equations of motion of the charged field (3.63). The current is then identified through the interaction term (3.76) and the conservation law "BJ~ = 0 follows automatically from the equations of motion of the charged field. We do not use the dynamics of the Maxwell field at alZ.! What we use, however, is the gauge symmetry, so again we see that gauge in-

variance implies a Worked exercise

conse~ed e~ent.

3.11.1

Problem: Consider the complex Klein-Gordon field <p coupled to the Maxwell field. Write down the equations of motion for ip and write down an expression for the current J]J Show by explicit calculation that the equations of motions for <p imply that JP is conserved.

3.12

THE EQUIVALENCE OF REAL AND COMPLEX FIELD THEORIES.

As we have seen, it is very easy to couple a complex field 4> = ~ + i4>2 to the Maxwell field Aa. All you have to do is to exchange the ordinary derivat1ves da with the gauge co-variant derivatives Da = "a - ieAa . Of course there is nothing mysterious about the fact that we use complex fields. It is nothing but a trick which makes life easier, and we could easily have avoided it. To see this, let us return to the problem of constructing a charged field. The starting point is the free Lagrangian

(3.59)
(we have got to have two classical fields because the theory should incorporate both particles and anti-particles). If we collect the two components into a single vector,
(3.77)

we may rearrange the Lagrangian as follows:

(3.78)
But then it is obvious that it possesses the symmetry,

(3.79)
where a is a constant.
4> + e

[~21) + [C~sa-Sina ][4>1] S1na Cos a 4>2


~

Of course, this is completely equivalent to the formula:

(3.70)

ia

4>

123

In the same way we can construct a gauge co-variant derivative:


(3.80)

D~

au - eA~

r 0 -11 II 0 J

Operating On the combined field, we find,

D; uI

a r~l]
U l~2

- eA lr ~2 1 ~lJ
U

which is completely equivalent to the formula:


Du~ (all - ieAu)(~l + i~2) = aU(~.l + i~2) - eAj1(-cp.2 + i~l) Hence, we can write down a gauge invariant Lagrangian density:

(3.82)

In this case "gauge invariance" means that it is invariant under the combined transformation:
(3.83)

~l] [ ~2

[cos(ex(x Sin(ex(x Aa(x)


+

-Sin(ex(x].[~l]
Cos(ex(x
~2

Aa(x) + "aX(x)

Here is a dictionary which allows you to pass from the "complex" formulation to the ureal II formulation: Gauge group Group element Gauge vector Gauge
co-variant

SO(2)

U(l) exp[ia]

[c~sa -Sina ]
S1na Cosa

;1

[:d
[0 -1] 0
~2
'i.

~ = ~1 +

i~2

derivative Gauge scalar Gauge phase factor

Du = au - eA u 1
= += ~I ~ 1
~ 2 +
1

Du = au - ieA u

$~

= (~l

- i~i(~l + i~2)

exp{[~

Cl -6]eJAa dx }

a exp[ieJ Aadx ]

Concerning the gauge phase factor, you should observe that:

[~-~r -[~~]
From the formula you therefore get

124

exp{x[~ -~]}

[1 0. 1 Cx - x 3 + xS~ -... )[0 L x + 1 x-"')OlJ+ 1-11 OJ -I} = [C~s x -Sin x J = Cos x [1 0 01 lJ + Sin x[ 0 1 0 x Cos x
= Cl 2

4~

3~

S~n

Thus the gauge phase factor lives in the gauge group as any decent gauge phase factor ought to do~ So you see that there is nothing sacred about complex numbers~

SOLUTIONS eF WORKED EXERCISES:


No. 3.5. , FasFyo) aF 0 1 aF a_S_ F So [ __ + F ~ = 11 aY 11 a callA,,) yO as a COilA,,) J a COilA,,) dF aB Thus , we must first compute a callAv) aCF a
11 11

aS F ) as a callA,,)

C ay So

aF a (aaAs-asAa) aB a(aIlA,,) = a (aIlA,,) inserting this, we get: naYnSo[Co)J\) - o)J')F + F (0)J\) - 0 )J~] as yo of

o.S

Sa yo

(0)J\) - 0 )J')F aS + Fyo (0)J\) - 0)J')

as

Sa

yo

oy

F Il " _ F"1l + FIl " _ F VIl = 2(F Il " _ F"Il) = 4F UV (due to the anti symmetry of FIJ "), so we have shown once and for all:

No. 3.7.' Ca)

Here the first term is not symmetric, while the last two terms are trivially symmetric. Cb) Observe first that, using the equations of motion, we may rearrange the expression for eaS eaS =-C3 FSp)Aa _ FSP a Aa P P a 2 a B SP =_m A A _ F aA
p

Here the first term is born symmetric, but its pre$ence is very important for the verification of the various properties eOS ought to have:

125
<lS + e<lS ~[_F<l FPS _ In<lSF FyoJ1 + {m2A<lAS _ 1 n<l2 A AY] P 4 yo 2 Y

(1)

This is clearly symmetric. Observe that the first piece is identical to the energy-momentum tensor for the electromagnetic field.

(2)
(3)

dSe<lS =-dSdp(FSPA<l) = 0

since

aBd p is symmetric in SP .
ai(FoiA<l)d3x

J a<ld 3x =-J

ap(FoPA<l)d3x

=-J

(c)

But the last integral can be converted to a surface integral which vanishes automatically provided the fields A<l vanish sufficiently fast. =[_FO FPO _ ! F FYO} + [m 2AOAO + 1 2A AY l P 4 yo 2 m Y J

~(ii2

+ 132) +

m2{ (A O)2 + (AI)2 + (A2)2 +

(A3)~]

which is clearly positive definite,


No. 3.8.1

0
immed~ately

From the equation of constraint and the Lorenz condition you

find:

If you prescr1be

->-(

->A t1,x) ->-

and
dA

~ (tl,X) dt

from the equations: M(tl,x) = ->-

V'at (tl,x)
~

->-

and you can then use iteration to determine

and

2i at

at

2i at

(t

+ at,x) '"

at

~~

(t ,;:) +
I

We must then show that the equations of constraint and the Lorenz condition are still valid at t = tl + ot

Equation of constraint:

using the equation of constraint and the Lorenz condition at time tl this is rearranged as ->->+[aA ->.7->->= - 9. dA (t ,~) + ot ~(-9 ->A(tl,x = -\I :it (tl,x) + ot 8A(t l ,X) }
dt
I

using the equation of motion this is rearranged as


->-r

\I -

aA lat

(t
I'

->-

x) + ot -

dtZ

126

Lorenz condition:
~( +) = ~(+) +) at tl+ct,x at tl,x + ct ~$ ( tl,x

Using the equation of constraint and the Lorenz condition at the time t is rearranged as =-VA(tl,X) - ct at(V'A(tl,X)) =
No. 3.11.1
-+-+

1,

this

-+

a"-+-+

-9[it(tl,*)+Ot.~~(tl'*)]

= -9'A(t 1+Ot,*)

The starting point is the free Lagrangian:

LO
L$($,all$,A a )

=-

2112

~~)(a~$) - ~m2~~

We make it gauge invariant by using the rule of minimal coupling:

= LO($,D
= [-

$)
~

..!.rf";)(a l1 $) 2

~m2~$1 + ~A [$al1~_~all$l 2211


11 11

From this we get the equations of motion (3.63) (a a l1 )$ = m2$ + ie[A a l1 $+a (A~$)] + e 2A A~~
11 11

and we can immediately read off the interaction term:


L
I
=

ie A [$al1~_$al1<p] _ 2e 2A AI1$$
2 11 2 11

From this we get the current:


J~

Of course this is identical with the previously found current (3.72): Then we compute a J~:
1.1

1.1

Jll = ~[$(a all )$ - $(a a~)$] _ e 2 a [AI1 $$]


21111
1.1

From the equations of motion we now get:

-~(m2$+ie(A al.l$+a (AI.I$))_e 2A AI.I$)] 1.1 11 11 e 2 a [A~;P~]


~

Inserting this, we get:

a11 JI.I

e 2 a [A~~$] - e 2 a [A I1 $$]
~

11

=0

Consequently we have reproduced the equation of motion without using Maxwell's equations!

127

chapter 4 SOLITONS
4.1 NON-LINEAR FIELD THEORIES WITH A DEGENERATE VACUUM
In this chapter we will discuss in some details a few important models in classical field theory which have recently attracked great attention due to their so-called "topological" properties. To keep the discussion as simple as possible we will only consider classical field theories in (1+1)-space-time dimensions. generalizations are by no means trivial.) The theories we are going to consider will be non-linear field theories, i.e. in the non-relativistic case, with which we shall be mostly conserned, they are based upon a Lagrangian density of the form, (3.28) where the potential energy density U is no longer quadratic in given by (3.30)
U ' (<1
~

(Although some of the

results have suitable generalizations to higher dimensions, these

(cf.

the discussion in section 3.3). The associated equation of motion is

and to ensure a positive energy density we shall as usual assume that the potential is positive definite:
U(<I
~ 0

To obtain non-trivial results we shall however furthermore assume that it has more than one (g~oba~) minimum. Two examples are of particular interest: Illustrative example 1:

The ~~model.
~:

In this model the potential energy density is given by the following fourth-order polynomial in
(4.1) U(<I

}(<j>2_ f)2

Notice the sign of the quadratic term which is opposite of the usual mass-term (cf. the discussion in section 3.4). The potential energy

128 U[cp] Klein-Gordon , potential


I

;1
Fig. 45a

it

1>-axis

+~ A
Fig. 45b

Cp-axis

density has two distinct minima, (4 . 2 ) 1> + = and and has the characteristic shape indicated on fig.45a (known as a double well). Illustrative example 2: The sine-Gordon
2

'It

mode~.

In this case the potential energy density is given by


(4.3)

U(CP)

11 IT(l

COSACP)

which clearly is periodic in cp. The potential energy density, which has the characteristic shape shown on fig.45b, thus has an infinite series of minima: (4.4)

,. = 'l'n
a all</>
)J

nT
2

ZTT

.. ,-2,-1 ,0, 1 ,2, ..

Consider the equation of motion (4.5)


=

LSinA</> A
Icpl
1, we can safely replace

In the weak field limit, where with


(3.31)

SinAcp

A</> whereby we recover the Klein-Gordon equation:

Now, because the equation of motion reduces to the Klein-Gordon equation in the weak field limit and because the exact equation involves a sine-function it has become customary to refer to equation' (4.5)

as

the sine-Gordon equation. So that is the origin of the funny name for our mOdel.*)

*) According to Coleman the name was invented by Finkelstein. In Coleman [1975] there is a quotation from a letter from Finkelstein: 1 am sorry that I ever called it the sine-Gordon equation. It was a private joke between me and Julio Rubinstein, and I never used it in print. By the time he used it as the title of a paper he had earned his Ph.D. and was beyond the reach of justice."

129

Remark: This model has a famous analogue in cJassical mechanics which allow you to "visualize" the basic properties of the model: Imagine a fixed string on the x-axis. To this stri.ng we attach an infinite equidistant series of pendulums. These pendulums are only allowed to move perpendicular to the string, i.e. they can rotate in th~ y-z-plane. The position of a single pendulum at xl' = an is then completely characterized by its angle Ij>(xn,t) relative to the y-ax1S. (For convenience we imagine that the y-axis is pointing downwards).

Y-axis

Fig. 46

Finally we introduce a small coupling between neighbouring pendulums, e.g. by connecting them with small strings. We can then write down the energy for this system. A single pendulum contributes with a) Its kinetic energy
1

2[dlj>(X n ,t)]2

2mr

at

b) Its interaction energy coming from its coupling to the neighbouring pendulums

c) Its potential energy due to the gravitational field mgr[l - Coslj>(xn,t)] The total energy is therefore given by H" n;...."
+00

~mr2 ~(Xn't)

[]2 + ~k2[cjl(Xn+l't)-cjl(Xn't)] 2+ mgrll-Cos</>(xn,t)]


2

Furthermore the dYnamics is controlled by the Lagrangian L" T-V

"E~mr2[~]

~k2[Ij>(Xn+l,t)-Ij>(xn,t)]

- mgr[l-Coslj>(xn,t)]

(Compare with the Lliscussi<Jn in section 3.1). We then pass to the continuum limit where a+O while at the same time m/a + rand ka + a . In this way we obtain the following Lagrangian L"

J::_oo(~pr2[~]2

~a[~]2

- pgr(l-Cos</]dx

We therefore See that for a suitable choice of parameters the infinite system of pendulums is equivalent to the classical field theory in (l+l)-dimensions characterized by the potential energy density (4.3).

Notice that in both of the above examples the potential U(cjl) possesses a discrete symmetry which transform one minimum into another. In the cjl4-model it is the reflection Ij> + - Ij> ,

130

while in the sine-Gordon model it is the translation


cP
+

<P

+ 2rrA

The existence of such a discrete symmetry is a typical feature for the kind of models we are going to examine. We proceed to investigate various configurations in a non-linear field theory. In a classical theory we will always restrict ourselves to configurations with a finite total energy. Since the total energy is given by (4.6)
H

the assumption of finite ditions:


(4.7)

~gy

naturally leads to the boundary con-

0 . It + 0 ; U(cp) + 0 as I xl+ 00 ' ax At infinity the field is consequently static and approach a constant value
at
+

lim

-+~OQ

CP(x,t)

due to the first two conditions. The third condition states that the asynptotic value must in fact be one of the global minima for the potential, i.e.

Among the configurations with finite energy we have especially the vacuum configurations. The classical vacuum is characterized by having the lowest possible energy, i.e. it is characterized by a vanishing energy density. It follows from (4.6) that a classical vacuum must satisfy the equations:
(4.8)

o
<p(x,t)
=

and

Thus a classical vacuum is represented by a constant field

CPo

where

CPo is one of the global minima for the potential, i.e.

Notice that the assumption about several global minima for the potential implies that there are several classical vacua. A non-trivial non-linear field theory is thus characterized by a degenerate classical vacuum! Next we consider small fluctuations around a c 1 aSSlca . 1 vacuum
~ ~o:

<P(x,t) = <Po

+ n(x,t)

131
In the weak field limit, where interaction of the field, thus get! L(n,a"n) ...

In(x,t) 11 , we can neglect the selfn(x,t) we

i.e. we need only keep the quadratic terms

in the Lagrangian density. In terms of the shifted field

!a nalln 2 II

!U'~. 0 )n 2
(3.32). In the quan-

But this is exactly a free field Lagrangian, cf. cuum

tized version of the theory small oscillations around a classical va-

.0

consequently represent free particles with the mass-square

U"(.o). If the fluctuations are large we can of course no longer neglect the self-interaction, i.e. when the density of field quanta is high the field quanta no longer act as free particles. If the various classical vacua are all connected by a symmetry transformation then U"(.o) will be the same for all the classical vacua, i.e. the field quanta associated with two different classical vacua will have the same mass. In the .'-model, e.g., the mass-square of the field quanta is given by
(4.9)

while in the sine-Gordon model it is given by


(4.10)

4,2

TOPOLOGICAL CHARGES
As we have noticed finiteness of the energy implies that the asymp-

totic values

.+

- co

lim

~~(X)

.(x,t)

are independent of time. These asymptotic values furthermore correspond to classical vacua. Let

denote the space conSisting af all smooth finite energy con-

figurations. As a consequence of the degeneracy of the classical vacuum this space breaks up into different sectors characterized by the different possibilities for the asymptotic behavior. Consider, e.g. the theory. Here the boundary conditions are given by

.4_

lim
++00

.(x,t)

= :!:

This gives four different possibilities for the asymptotic behavior: I III
_....l!..

-00 =
. - oc

If

;
;

.+- = .+co: =

-it
_Jl.

II IV

cp-oo =

-.t
Jl.

; ;

CP+CI> =
cP+<

= +If

-1!.

If

-'" = +If

+* +*

132 Thus
E++. E

breaks up into four sectors, which we label

E __ , E_+, E+_ and

The vacuum solutions

$+ belong to the sectors E++ and E __ . These

sectors are consequently referred to as vacuum sectors. What can we say about the remaining non-trivial sectors? Consider a given smooth finite energy configuration $ A smooth deformation of $ is a one-parameter family of finite energy configurations $A(X), such that $A (x) depends smoothly upon:\ and that $A (x) reduces to the given configuration $(x) when A=O. Clearly a smooth deformation cannot "break" the boundary conditions, i.e. lim $A (x) $oo

-+co

is independent of A. A smooth deformation thus necessarily stays within a single sector. You may think of $:\ (x) of the given configuration $ as representing a pertubation . It follows that a smooth pertubation of

a classical vacuum necessarily stays within the corresponding vacuum sector. On the contrary a configUration from a non-trivial sector can not be obtained by pertubing a classical vacuum. The non-trivial sectors are therefore called non-pertubative sectors. Notice that if the model possesses a discrete symmetry operation which relate the different classical vacua, then this symmetry operation will also relate the different sectors to each other. E.g. in the $4-model the inversion will interchange the vacuum sectors, i.e. it will map E Similarly it interchanges the non-pertubative sectors E_+ onto E++. and E+_.

Thus the structure of the two vacuum sectors and similarly the structure of the two non-pertubative sectors are completely equivalent. As another example we consider the sine-Gordon model. It possesses the translational symmetry p integer. The different sectors in the 'sine-Gordon model can be labelled by a pair of integers (n,m) such that a field configuration belonging to
E

n,m

satisfies the boundary conditions lim $(x,t) = 2;n


X

lim
++00

$(x,t)

+-00

21Tm A

The translation operator then maps the sector En,m onto the sector En+p,m+p. In the sine-Gordon model there is therefore a discrete series of qualitative different types of sectors labelled by a single integer q=m-n (where q=O corresponds to the vacuum sectors)

133

Next we look for conserved quantities in our non-linear models. Due to our boundary conditions the asymptotic values lim q,(x,t) x -+oo are conserved, i.e. independent of time. Let us especially focus upon their difference
(4.11)
Q

This can be interpreted as a conserved charge with the corresponding charge density given by
(4.12)

p[q,]
Q

Le. (4.13 )

q, (+oo,t) - q, (-oo,t)

We will now compare this charge with the charge associated with the complex Klein-Gordon field, cf. the discussion in section 3.9.
Exercise 4.2.1 Problem: Let q, = q,1 + iq,2 be a complex Klein-Gordon field. Show that the charge corresponding to the conserved current (3.67) is given by
(4.14)

Q=

R3 ab

q,a~ld3x
dt

The main difference between the stJ:ucture of (4.13) and (4.14) is that the ordinary charge (4.14) depends upon the time derivative of the field as well. Notice that the conservation of the ordinary charge presupposes that q,a obeys the equations of motion: Differentiating (4.14) with respect to time we get

the above integral can be rearranged as follows

But this can be converted into a surface integral at infinity. Provided the gradients vanish sufficiently rapid at infinity the surface integral vanishes and the the charge is conserved. The conservation of an ordinary charge is thus dependent not only upon the boundary conditions at infinity but also upon the dynamics of the field.

134

On the contrary the conservation of the new charge (4.13) is independent of dynamics. It only depends upon the boundary conditions. vation law is called a topological conservation law. Now, whenever we have a conserved charge we expect it to be associated a conserved current, i.e. a current which together with the charge density obeys the equation of continuity (1.17). In the present case the topological charge density (4.12) is associated the relativistic current (4.15) Notice that this current is trivially conserved:

EDr this

reason it is called a topological charge, and the corresponding conser-

a~ J~ =

E~va

a\! ~ =

due to the symmetry of

a~av'

I emphasize once again that the dynamics is thus "automatically"

of the field has not been evoked in the above derivation of the equation of continuity. Since the current (4.15) conserved it is called a topological current. Let us investigate the topological charge (4.13) a little closer. Notice first that it is conserved due to the requirement of finite energy, but that the structure of the potential energy density is irrelevant in this connection. Thus the topological charge (4.13) will be well-defined and conserved for any relativistic scalar field theory in (l+l)-dimensional space time. In an ordinary scalar field theory we will have a potential energy density
U[~J

with a unique minimum at

~=O.

But in that case the boun-

dary conditions (4.7) reduce to lim

-+oo

~(x,t)

=0

Thus the topological charge is completely trivial in an ordinary scalar field theory, since it automaically vanishes. As another example we may consider a model where the potential energy density is absent, i.e. it is based upon the Lagrangian density
L

=-

~a cpa\Jcp
~

In that case the classical vacuum is characterized by

being an arbi-

trary constant. Thus the topological charge can take any value. Between these two extremes we have the non-linear models in which we are particularly interested. In these models we have a potential energy density with a discrete set of global minima. The topological charge can therefore only take a discrete set of values. In the
~'-model,

e.g.

135 the topological charge is on the form (4.16)


Q

+2J<. = O'-If'
211 = -r q,

while in the sine-Gordon model it is given by (4.17)


Q

q integer.

Since the topological charges in this kind of model can only take a discrete set of values we say that it is quantized. But notice that it is quantized already on a classical level! We speak about topological quantization. Observe that the topological charge vanishes for a classical vacuum.

Non-pertubative configurations thus correspond to the topo~ogica~ charge.


In its mechanical analogue the classical vacuum corresponds to the configuration where all the "pendulums" are pointing downwards. In general finite energy implies that sufficiently far away the "pendulums" are static and point downwards. When we go along the x-axis from
-~

non-trivia~

va~ues

of

We end this section with a remark about the sine-Gordon model.

to

+~

the pendulums will rotate around the x-axis, but they will necessarily rotate an
integra~

number of times. Fig. 47

Thus each finite energy configuration is characterized by a winding number.

Clearly the winding number coincides, apart from a constant, with the topological charge in the sine-Gordon model.

4.3

SOLITARY WAVES
We proceed to look for wave-like solutions. In a free field theory
p~ane

we would naturally look for the form


~(x,t)

wave solutions, i.e. solutions on

ACos(kx - wt) represent solutions with finite energy.

Although they have infinite energy themselves, we can construct superpositions of plane waves, which (This corresponds simple to a Fourier analysis of such a solution). In a non-linear field theory such plane waves can never be exact solutions.

136

In stead we will look for solutions on the form (4.18)


~(x,t)

= f(x - vt) .

Such a solution will be called a travelling wave. Especially we will be interested in travelling waves with finite energy, i.e. if we introduce the variable, ; = x - vt then f must satisfy the boundary conditions (4.19) t; ...: ...
lim
f(O

As f is essentially constant sufficiently far away we say that such a wave is localized. A solution on the form (4.18) which satisfies the boundary conditions (4.19) is called a solitary wave. Let us write out the equation of motion (3.30) explicitly

a2cb - at2

a2 cb ax 2 =

U'[~]

Inserting the ansatz (4.18) this reduces to the ordinary differential equation
U' [f]

If we multiply this equation with ranged as

df/dt;

on both sides it can be rear-

But that can be immediately integrated whereby we obtain

The integration constant must however vanish since both df dE; and U[f] vanish at infinity. Thus the equation of motion reduces to
(4.20)

with

= v'1=V2

This first order differential equation can be integrated too:


(4.21)

This formula gives t; as a function of f, i.e. we have actually determined the inverse function. Notice that the boundary conditions are satisfied. Close to a zero, as follows
~o'

of the potential, we can expand the potential

U[f] ~ IU'~~ o ](f-~ 0 )2 Thereby the solution (4.21) reduces to

137 Thus
~

goes to infinity when f approaches a classical vacuum. The

asymptotic behavior of f is therefore given by (4.22) If(O-<J>ol "" exp {

~ v1JIifQ

y(s-so)}

as

lsi'"

co

Clearly we get two types of solitary waves: The first type is in-

creasing,

(corresponding to the plus-sign in equation (4.20) ). It inter-

polates between two neighbouring classical vacua, and it is known as a

kink. The other one is decreasing and similarly interpolates between two adjacent classical vacua.It is known as an anti-kink.

:o..='~-----

----...:-:....- - - anti-kink

- - - -

-<J>~:

- - - -X-a
5

X-axis

----------~----~---=~~
~v

-----:-:-= - - - Okay, let us look at some specific examples: In the <J>'-model the kink is given by
f

-----~----

Fig. 48

dt =V2 If Y( .~-~o) , '" _-{f2J ~ ---z-_t 2 :""Artanh[-f] ~ ~ oA


i.e.
(4.23) In the sine-Gordon model the kink is similarly given by
f
Y(s-~O)

=~f

dt 12 (1-C05>"1') TriA

~Af

Iil

If

J~(1-C052t)

dt

~Jst!f
~Tr

1>.

!IT

i.e.
(4.24)

We can also easily determine the energyof the kink. Inserting the ansatz (4.18) in the energy functional (4.6) this reduces to +co

J(~~2(~~)2+

U[fj)dt;

~ing

the equation of motion (4.20) this reduces to

+co (4.25)

138

+co

<j>+co y J /2U[f] df <j> -co

Thus the kink energy can be computed directly from the pctential. We need not know the explicit form of the kink at all! Once again we look at some specific examples: In the <j>"-model we get (4.26)

+]1/1X
H[<j>kink] = yJ I~(~ - i')df -]1/ IX
271 / A ,-,,.-_ __
2

Y-3- A

2/1 1!...3

while in the sine-Gordon model we obtain (4.27) H[<j>kink1

= yI

~r~(I-COSAf)

df

Y~

o
Notice that the solitary waves constructed above interpclate between different vacua. Consequently they belong to the non-pertubative sectors Since they actually interpolate between neighbouring vacua, they carry the smallest possible non-trivial topological charge. The corresponding sectors are called the kink sea tors respectively the anti-kink seators.

4.4

GROUND STATES FOR THE NON-PERTUBATlVE SECTORS

Now that we have gained some familiarity with the solitary waves we will rederive them from a somewhat different point of view. As w~ have seen the space of all finite energy configurations, E, breaks ~p into disconnected sectors characterized by the different possibilities for the asymptotic behaviour of the finite energy configurations. In each sector we will now look for a ground state, i.e. a configuration <j>o(x) which has the lowest po~sible static energy in that sector. Since the ground state minimizes the static energy Hstatic[<j>] =

J~(~!)2

+ U[<j>(x)] dx

it must satisfy the associated Euler-Lagrange equation

Notice that the ground state extends to a static solution <j>(x,t)=<j>o(x) of the full field equations (3.30). Actually the following holds: The ground state for a given seator aorresponds to the solution

Of the field equations whiah has the lowest possible energy.

139 Proof: To see this let


~(x,t)

be a non-static solution in the corresponding

sector. Then there is a space time point (xo,t ) such that o

~(Xo,to) f'
and

Consider then another solution specified by the initial data

It has the same asymptotic behaviour at infinity as to the same sector.

~.

Thus it belongs
~

Furthermore it has the same static energy as

but

it misses the kinetic energy! Consequently a non-static solution cannot minimize the energy in a given sector.

Okay, let us look for possible ground states. In a vacuum sector the ground state is simply given by the corresponding classical vacuum. But what about the non-pertubative sectors? To look for possible ground states in these sectors we will first examine the static solutions in our model (since a ground state is necessarily a static solution although the converse need not be true). The field equation for a static solution reduces to
(4.28 )

2 d " = U' :::.....z 2 dx

[~l

This can be integrated once (using the same trick as before), whereby we get

(~) 2 = 2U[~l
dx Thus we get two first order equations:
(4.29)
~ dx

v'zU[~l

<1>-2

<1>-1

.-

dx -

~- -v'zU[~l

~.
~,

<1>0 monotone

~-axis
~

~2:constant

They have two kind of solutions:

solution

solution

(a) On the one hand there are constant solutions, corresponding to the zeroes of U, i.e. the constant solutions reproduce the classical vacua. (b) On the other hand there are monotone solutions, which interpolate between two adjacent vacua:

(4.30)

x - Xo

~Jv2at<l>l

They correspond precisely to the static kink and the static anti-kink. This suggests that the kink is the groundstate in the kink sector!

140

Notice that there are no static solutions which pass through a classical vacuum. If the model possesses more than two classical vacua there will consequently be sectors (consisting of configurations which interpolate between non-neighbouring vacua) which has no ground state! Observe that if
~(x,t)

is any solution to the field equations, then

so is the Boosted configuration: (4.31)


~(x,t)

~[y(x-vt) ;y(t+vx)]

due to the Lorentz-invariance. If we apply this to the static solution


~(x)

we thus produce a solitary wave:


~(x,t)

~[y(x-vt)]

We can therefore easily get back our solitary waves once we have determined the static solutions! Next we want to tackle the problem of whether the static kink really
is a ground state configuration for the kink sector. This will be shown

using a beautiful trick going back to Bogomolny.

He showed that in many

interesting models the static energy can be decomposed as follows: (4.32)

Here

P(~;~)

is a first order differential operator acting on

~ while

the topological term only depends upon the asymptotic behaviour 01 the field at infinity, i.e. it is constant through out each sector. Provided the topological term is positive we therefore get the following bound for the static energy (4.33) Hstatic

{TOPOlOgiCal term}
~

Furthermore this bound is only saturated provided order differential equation ; (4.34 )

solves the first

Any solution to this first order differential equation is thus a ground

state. A decomposition of the type (4.32) is known as a Bogomolny decomposition. The associated first order differential equation (4.34) is known as the ground state equation. Let us try to apply this to our (l+l)-dimensional models. Guided by the kink equations (4.29) we try the following decomposition

141

The rest term is given by


<11+ co

I
-

/2U[<j>] d<j>
co

Thus it is a topological term only dependent upon the asymptotic behaviour at infinity! Let us specifically look at a model with a finite or infinite number of classical vacua

where the different classical vacua are related by a symmetry transformation. In the sector En consisting of all configurations which interpolate between <j>o and <j>n' we can now apparently rearrange the above decomposition as follows:
(4.35)

Hstatic[<j>]

~f{~

:;:

hU[<j>]}2 dx

Thus we have shown:


(a)
The energy in the sector E

<j>1

is bounded below by

(4.36)

Hstatic'"

InII~
<j>o

d<j>

(b)

A configuration in the sector En is a ground state i f and only if it satisfies the first order differential equation:

(4.29)

+: n positive { -: n negative

This shows especially that the kink is the ground state for the kink sector. It also shows that the energy of the static Kink is given by
(4.37)

<j>1 H[<j>kink]

= IhU[<j>] d<j>

which of course is in agree~~nt with our earlier result (4.25). The Bogomolny decomposition can also be used to examine the higher charged sectors. Consider as a specific example the sine-Gordon model. We know in advance that there are no exact ground states in the higher charged sectors. We know also that the static energy is bounded below by Hstatic[<j>] ~ InIH[<j>kink] Now let X1,X2, ... ,X n be widely separated consecutive points on the xaxis and let furthermore <j>+(x) denote the static kink solution centered

142

at

X~O

(i.e. we put lim


X
+-00

v~so=O

in (4.24) ). It satisfies the boundary con27f

dition s

Consider now the superposition

~[ ](x) = ~+(X-Xl) + ~+(X-X2) + ... + ~+(x-xn) .n;x1""'Xn


It clearly satisfies the boundary conditions
lim
X
+-00

~(x)

lim

$'(x)

2 n )..7f

++c:o

so that it belongs to the

Fi . 49

811/)..611/1..
47f/)..

hi A
X2
x

X-axis

Notice furthermore that the kink solution is strongly localized, since it approaches the classical vacuum exponentially, cf. (4.22). Outside a small region of width perposition: (a) Except when we are close to the centers configuration ~(x) cuum. X1 ,X2, .. ,X the n is exponentially close to a classical val/~the

kink solution essentially reduces to the

classical vacuum. This has the following consequence for the abg.ve su-

(b) Close to a center, x k ' it behaves like a kink solution. Especially the difference

:i!
(4.38)

"'1 v'2U

[~]

is exponentially small. As a consequence the above configuration has a static energy given by

Hstatic[~] = nH[~kinkl

+ {exponentiallY small error}

It is thus extremely close to the lower bound given by the Bogomolny decomposition. furthermore we clearly have k=l, .. ,n-l Thus we can find approximate ground ted kinks.
sta~es

conSisting of widely separa-

143

Apart from a configuration consisting of widely separated kinks we may also consider strings of widely separated kinks and anti-kinks. Let us this time look at the ~4-model, and let us even specialize to one of the vacuum sectors, say E__ . As before centered at x=O, while
~_(x) ~+(x)

denotes the static kink

denotes the static anti-kink centered at

x=O. We now consider the following configuration in E __ :

~(x) = ~+(X-Xl) + ~_(X-X2) + ... + ~+(x-x2n_l) + ~_(x-x2n)


Notice that this time we must let the kinks and anti-kinks alternate and furthermore, since we are in a vacuum sector, there must be an even number of centers.

]lIlA
X-axis

As before we see that 2U[J +

Fig. 50
{exponential small error}

When the kinks and anti-kinks are widely separated the configuration (x) is thus almost a static solution. Since the static energy is thus not truly stationary under deformations we call such a configuration a
~ionary

quasi-s~a

configuration.

Although a quasi-stationary configuration need

not be an approximative ground state it will almost be a local minimum for the static energy. This is clear since any deformation of the quasistationary configuration (except for a pure displacement of the centers) will either destroy one of the claSSical vacua involved, or it will destroy one of the kinks/anti-kinks involved. Thus the static energy will increase (or at least stay constant).

3H 1

SINE-GORDON MODEL

vacuum vacuum

o
kink

~~~~__~~~~~__~~__~-+~s~e~ctors

n=2

n=3

n=4

sector sector

sector sector

Higher charged sectors

144

4. 5

SOLITOr~S

Observe that asymptotically the kink solution approach the Yukawapotential (3.34). But it is a solution of quite a different type! The Yukawa-potential (3.)4) is a singular solution to the Klein-Gordon equation and (just like the Coulomb field) it signals the presence of an external point source, i.e. a foreign particle. But the kink is a smooth solution everywhere, and consequently it is not associated with any external sources of the sine-Gordon field. Notice furthermore that they are localized solutions, where the energy is concentrated within a small region, i.e. they represent a small extended object. In accordance with this kink-solutions in non-linear field theories are generally interpreted as particles -an interpretation which in the case of the sine-Gordon model goes back to Perring and skyrme*). Since such particles are associated with solitary wave solutions they are referred to as solitons. We will now look at some of the properties of the kink-solutions which justify this interpretation: The kink-solution ~+(x-xo) represents a soliton at rest centered at x = Xo . In the same way the boosted solution (4.31) represents a soliton moving with the velocity v! We can also consider the energy and momentum of a single soliton. It is given by the total energy and momentum contained in the field, Le. by +OO plJ =

-00

where the energy-momentum tensor is given by


a We know that plJ transforms as a Lorentz vector and in this example it is furthermore reasonable to localize it, since the energy is concentrated in a very small region! Thus we can attach the Lorentz vector to the center of the soliton.
TlJV

alJ~aV~ + nlJv[~a ~aa~ + U(~)]

Soliton in (2+1)dimensional space-

For a soliton at rest the energy E is equal to the rest mass M M = C:TOOdX

time. Fig. 51

C] ~(*)2

+ U[x] }dX

and the momentum P vanishes because the solution is statiC. Due to the Lorentz covariance of plJ, we therefore see that a soliton moving
*) A model unified field equation, Nuel. Phys. 31, 1962 (550)

145

with velocity
(4.39)
P

has the energy and momentum


E

= Myv

= My

(in units where

c = 1

The kink-solution therefore represents a free particle with rest mass M , while the anti-kink solution represents the anti-particle corresponding to the soliton. We now specialize to the sine-Gordon model. In the higher charged sectors, where the winding number is numerically greater than one, it is tempting to see if we can find solutions which represent several solitons. 'iliere are no static solutions. When several solitons are present, they consequently no longer behave like free particles. This means that two solitons exert forces on each other. This phenomenon is closely associated to the non-linearity of the sine-Gordon equation: The superposition of two kink-solutions is no longer a solution. We have seen, however, that if we choose the superposition carefully, then it is "almost" a solution.
Worked exercise 4.5.1
Problem: (u) Consider the :;t.>'I,'t superpo"it.ion
~(x,t)

= ~+[y(x+vt+xo)]

211 + ~+[y(x-vt-xo)] - )l

Show that it satisfies


T (~] - Sinh(~yx) g 4 - Cosh[~Y(vt+xo)]

(b) Show that it has the asymptotic expansion: (4.40)


e-~Yxo Sinh(~Yx]

t->>

Cosh(~Yvt]

Based on computer simulations Perring and Skyrme guessed the following analytic expression for an exact time-dependent solution (Cf. the asymptotic expansion (4.40) :) (4.41)

T9(A~s4s]= vSinh(~yx]

Cosh[ llyvt]

Remark: It is definitely not "trivial" to verify that this is in fact a solution. In the end of this section we shall present a general method to solve the equations of motion whereby we can derive (4.41 ) "relatively" easy.

Exeraise 4.5.2
(4.42a) PrOblem: Show that the above solution (4.41) can be expanded asymptotically as 211 ~ss(t,x) ~+[Y(x+vt+xo)] + <p+(y(x-vt-xo )] - T
t+->
~

" (4. 42b)

ss (t,x)

t++<">

14.6

has a very simple According to exercise 4.5.2 the solution (4.41) interpretation: In the remote past it represents two far separated solitons moving towards each other and in the remote future it represents two far separated solitons moving away from each other. Thus it describes a scattering process between solitons!

Scattering between two solitons.

(Sketch based upon a computer simulation)

Fig. 52

Observe that the two solitons are identical, so it is not possible to say if the two solitons move through each other or if they are scattered elastically: All we can say is that there are two particles going into the collision center and two particles going out. Each of these particles carries an energy and momentum Pl~ and P2~ which can be calculated by integrating the energy momentum tensor over the central region occupied by the particle (Outside this region, the energy momentum tensors dies off expoX-axis nentially and we can safely neglect it). If you look back on Abrahams theorem (Sect. 1.6)' it is now clear that we have decomposed the total energy momentum as Fig. S3 p~ = Pl~ + P2~ where
Pl~ , P2~

themselves transform like Lorentz vectors. (Of course

this decomposition only works in the remote past and the remote future, where the solitons are far separated. During the collision it has no meaning to speak of individual solitons!). But the total energy-momentum stored in the field is conserved. Thus we get

147 (4.43)

Pl~(-~) + P2~(-~) = Pl~(+~) + P2~(+~)

i.e. the total energy and momentum of the solitons are conserved during the scattering. We can also compute the energy of the 2-soliton solution: This is most easily done using the asymptotiC expressions v we simply get:
M[ 4>]
(4.4~

Since they consist of two far separated solitons moving with velocity

(4.44)

2~ly

where

is the rest mass of a single soliton.

Perring and Skyrme found two other interesting analytic solutions. They can be obtained from
(4.41)

using various "analytic continuations". (x,t)


~

First we apply the "symmetry-transformation" the singular complex valued solution:

(it,ix)

and obtain

Sinh[~-1-'
Tg[AP(t,x)]= v 4

t]

COSh[~--V,I1_v 2

(f:V2

x]
v

)
whereby we obtain the regular

Then we perform the SUbstitution solution:


(4.45)

v ~1

-(t,x) Tg[ A4> S 5 ] = I Sinh["yvt] ,,_ 4 v Cosh[~yxl

Exercise 4.5.3

Problem: (a) Show that (4.45) represents a scattering process between a soliton and an anti-soliton by investigating the asymptotic form of the solution as t+-oo and t->+oo. (b) Show that its energy is the same as the 2-soliton solution: (4.46)

Performing the 'analytic continuation"

-+

iv

(4.45)

can further-

more be transformed into the periodic solution:


(4.47)

with

,1l+v 2

This solution is strictly localized at all times within the region ixl < (~r)-l since it falls off exponentially due to the denominator. 0, so it belongs to the vacuum sector. It reIt has winding number

presents clearly a periodic oscillating configuration, called a bpeat-

148
her. The energy of the breather can also be obtained by "analytic conti-

nuation"
2M

Slowly moving breather.

Il+v2 Observe that it is less than the total energy of a far separated solitonanti-soliton pair. The breather is interpreted as a bound state of a soliton and an anti-soliton. For this reason the particle it represents is called a bion. (Bion is an abbreviation for bi-soliton bound state). When investigation the breather it is in fact more natural to introduce the cyclic frequency wand the "amplitude" n: (4.49)
W

= Ilrv

In terms of these variables the breather takes the form: (4.50) Sin(wt) nCos h (nwx) with
n

Notice that when t=nX (n integer) the breather momentarily degenerates to a classical vacuum. At these times the energy of the breather is thus purely kinetic.
Exercise 4.5.4
Problem: (4.51 ) Show, by explicit calculation, that the energy of the breather at the time t~O is given by
E = 2M(l - ~2)2

w2

where M is the soliton mass, cf. the earlier obtained result (4.48).

Now suppose that

Il

Let us furthermore assume that Sin[wt) is positive, say get the following asymptotic expansions:

O<t<X' We then

1T

A<j>b Tg[-4- 1

""

l~
2w

eXP{Il(X + 1l-11nf2Bs:nwt)}

when xO when x""O when X0

eXP{Il(X _ 1l_ 1ln [2US:TIWt)) }

149
On the other hand we get from (4.24) that vely anti-kink, is given by "<j>+

static kink, respecti-

(4.52)

Tg[~]

When t

is not too close to 0 or IlSinwt


w

1T ~

we know by assumption that

is a large positive number. Consequently we can interprete the asymptotic behavior as fOllows: We have (a) a soliton to the far left with the center x (t) ~ _1l-11n[2BSinwt] o w (b) an anti-soliton to the far rigth with the center xo(t) ~ (c)

1l-'1n[2Bs~nwt]

a classical vacuum ( <j> '" 2"Tf ) in the central region between the soliton and the anti-soliton.

This thus confirms our interpretation of the breather as a bound state of a soliton and an anti-soliton. The breather belongs to the vacuum sector. Thus it is not a nonpertubative configuration. Notice the following two extreme limits of the breather solution: (a) When

0 the mass of the breather approaches twice the

soliton mass and we further get "<j>b t Tg [ =--!}l"T---.. 4 ] ~ Cosh[llx]

as

But this is the same as the limit of the soliton-anti-soliton solution

4.45) when we let v


thus becomes unbound. (b) When

O.In this limit the soliton-anti-soliton pair


+

11 the mass approaches 0, while <j>b

O. In this

limit the breather is thus just a small pertubation of the vacuum.

4.6

THE BACKLUND TRANSFORMATION


The existence of exact multi-soliton solutions is a peculiar

property of the sine-Gordon modeL E.g. there are no exact multi-soliton solutions in the <j>4-model. Interestingly enough it turns out that there exists a systematic method for solving the sine-Gordon equation known as the inverse spectral transformation. Due to its complexity we shall however confine ourselves to a discussion of the Backlund transformation, which is a related powerfull technique that allows a systematic computation of

150

all the multi-soliton solutions. To simplify the analysis we introduce the so-called light-cone
coordinates:

(4.53)

x+ = ~(x+t)

= ~(x-t)
1... a ax - at

The derivatives with respect to the light-cone coordinates are given by


d , 2

In light-cone coordinates the sine-Gordon equation thus reduces to (4.54)

X Sin[A<p]

for simplicity we put A = ~ = 1 in the following. The crucial idea is to reduce the solution of this second-order differential equation to a pair of first order differential equations, cf. the philosophy behind the Bogomolny decomposition. In the last century Backlund discovered the following remarkable pair of first order differential equations:
(4.55)

"[h::h] '" + 2

aSinr~] 2
L

Differentiating the first Backlund equation with respect to x- we get

tsing the second Backlund transformation the right hand side can be further reduced to (4.56)

"_"+[~]

COs[~]Sin[~]

Similarly we get from the second Backlund equation


(4.57)

"+"- [~]
2

Consequently we get by adding and subtracting (4.56) and (4.57)


(4.58)

"+"_<P1 = Sin[<P1] <PO,<P1


which satisfy the Backlund equations

We have therefore shown:


Any two functions (4.55) must necessarily solve the sine-Gordon equation too!

In other words: The sine-Gordon equation is the integrability condition for the Backlund equations. We can now reformulate the above observation in the following way: Suppose Then we respect we have we have been given a solution <Po to the sine-Gordon equation. can insert it into the Backlund equations and solve them with to <P1.Since the new solution depends upon the old solution <Po in this way constructed an operator Ba , which transforms a

151 I given solution of the sine-Gordon equation into another solution. The operator Ba is known as the Back lund transfo1'mation (with the scale parameter a). As an example we apply the Backlund transformation to the classical vacuum. The Backlund transformed vacuum ~ solves the first order differential equation: a .. = 2aSin 2 a ~ = ~Sin2
+'" 2

Introducing the new variables f,; they reduce to


=

ax+ + !xa

1 ax + - a -x

at;~
(4.59) Tg[2] 4

2Sin~

with the obvious solution

= exp[t;-f,; 0 ] =

exp[y(x-vt-s )] o

But that is precisely the kink-solution (4.24)! By applying the Backlund transformation to the classical vacuum we therefore create a single soliton. In principle we can now obtain the two-solution (4.41) by Backlund transforming the one-soliton solution, etc. Remarkable enough it turns out however that further integration of the Backlund equations can be reduced to pure algebra. This important observation is due to Bianchi, who showed that successive Backlund transformations commute and that furthermore the following non-linear superposition principle holds for ~o,
~l
~3

Bal [~o], ~2 Ba2 [~l]

=
1

Ba2[~o]

and

= Ba

[~2]

(4.60)

(this is known as Bianchi's permutability theorem). We leave the details in the proof as an exercise:
worked exeraise 4.6.1 Introduction: Consider the following Backlund transforms of a given solution
~l=Bal[~o];~2=Ba2[~o];~~=Ba2[~l];~~~Bal[~2]

~o:

We are going to show that the integration constants leading from ~l to ~~ , respectively from ~2 to ~~' , can be chosen in such a way that ~l coincides with ~~' .

152 Problem: (a) Assume for a moment that the Backlund transformations B and B a2 mute and denote the Common value of ~l and ~l' by ~3' al Show that ~3 has to satisfy the consistency relation
(4.61 )

I
COffi-

with (b) Return to the general situation where-the Backlund transformations need not apriori commute. From ~O'~I and ~2 we now construct a new function ~3 by imposing the condition (4.61). Show that ~3 is the Backlund transform of ~l with scale parameter a2, i.e. show that ~3 satisfies the Backlund eQuations:

(Similarly it follows that ~3 is the Backlund transform of ~2 with the scale parameter al' This shows precisely that the integration constants can be chosen such that ~l and ~l' coincides with the function ~3 defined by (4.61).) (c) Show that the consistency relation (4.61) is eQuivalent to the Bianchiidentity (4.60). In the Bianchi identity (4.60) we now substitute for vacuum and for We then obtain: (4.62)
~l
~o

the classical

and

~2

two kink-solutions with different velocities.

Tg[~]

1 +Tg [ ~ 2 ] Tg [ ~ I

a2+al exp[S2-S!']-exp[Sl-SO] a2-al 1+exptS2+S1-s!~s!1

For simplicity we further put a = al = - 1 (i.e. V2=-Vl) and sA = Sl' =0. a2 Then (4.62) precisely reduces to (4.41). This constitutes the promised derivation of the two-soliton solution. Remark: If one has patience enough one can work out 3-soliton solutions, 4-soliton solutions etc. In fact it is possible to write out the general N-soliton/anti-soliton solution explicitly. It is given by the following simple expression:
(4.63)

Cos~

1 -

2a\.la\.l~{Det[M]}
1 1 1

Mij

a.:a.
1

i j COSh[6 :6 ]

For sOliton/anti-SOli ton-solutions we put I-v.


a~
1

= __ 1
I+vi

6.
1

= ::y. (x-v.t-t;.)
o~cur

Bions are included by allowing the a!s to be complex. In that case they must pairs of Hermitian conjugate numbers7

in

The formula for 8i

is then modified to

153

4.7

DYNAMICAL STABILITY OF SOLITONS


Let us now turn to the question of the stability of solitons. Ac-

tually this has already been solved, since we have shown that the static kink is the groundstate for the kink sector and as such it is clearly stable. However the argument was based upon the Bogomolny decomposition. It is thus essentially a topological argument, i.e. we have shown that the soliton is topologically stable. It is not always possible to give topological arguments for the stability of a static solution. It will therefore be useful to rederive the stability of the soliton from a conventional dynamical point of view. By analYSing the dynamical stability we will furthepmore be able to extract useful information about the particle spectrum in the model. Okay, we want to investigate the stability of a static solution
~o(x}. This is done by studying small fluctuations around the configuration:

(x,t) "" ~o (x) + n (x,t)

In(x,t) I

Inserting this in the equation of motion we can expand the potential energy density to the lowest order U'[<j>(x,t) ""
U'[~o(x)l

+ U"[<I>o(x)n(x,t)

Thus the linearized equation of fluctuations becomes:


(4.64)

This is not an ordinary Klein-Gordon equation because U "[~o (x) 1 is a function of x. Since it is not explicitly time-dependent we can,however, decompose n on normal modes:
(4.65)

Inserting this we get

i.e. each mode is a solution to the eigenvalue problem:


(4.66)

Notice that this is just an ordinary Schrodinger equation for a "particle moving in the potential W(x) = U"[<I>o(x)". It can be solved (in principle) and we should especially look for the eigenvalues. Observe especially that if there is a negative eigenvalue then the

154

corresponding cyclic frequency w must be purely imaginary. Thus the k "oscillation" exp[iwkt] in fact grows exponentially in the future (or in the past). Clearly this contradicts the stability, according to which a small fluctuation must stay small. Thus we have deduced the following criterium for dynamical stability
A static solution ~o(x)

is dynamically stable provide1 the ei-

genvalue problem

(4.66)

{-

..: + U'Ho(X)J}~k(X) ax2

Wk~k(X)

has no negative eigenvalues.

We can also derive this criterium by looking at the static energy functional. A stable static solution must at least be a local minimum of the static energy functional. Consider now a small pertubation:
~(x)

= ~o(x)

+ E:1jJ(x)

Expanding the potential energy density to second order we get

In the last two terms we perform a partial integration whereby we get

H[~ o J +E:Jr{_~o+ U'[~ 0 J}l/J(X)dX , dX2

!E:2Il/J(X){-~ +U"[~ 0 J}l/J(X)dX dX2

But here the middle term'vanishes due to the equation of motion, cf. (4.28). Thus the change in the static energy is of the second order in E: (4.67) oH[l/J]
=

~E:2I

l/J(X){-

dX2

+ U'H (x) J ~l/J(X) dx


0

We must now demand that the quadratic form OH[l/JJ is positive (semi-) definite. If we solve the eigenvalue problem {- ::2
+

U"[~o(X)J}~k(X)

wk 4>k(X)

we can now decompose the fluctuation l/J(x) l/J(x)

on the eigenfunctions:

ak~k(x)

Inserting this the quadratic form (4.67) reduces to


(4.68)

oH[l/J)

Thus we see again that a negative eigenvalue would be catastrophic, Since fluctuations along the corresponding mode ~k(x) would lower the static energy.

155

Okay, having motivated the eigenvalue problem (4.66) we now proceed to solve it. Notice that the "potential" W(x) = U"[<j>o(x)] is strongly localized and satisfies the boundary conditions:
(4.69)

lim x

W(x)

= U"[<j>+co]

~ut

here the right hand side reproduces precisely the mass-square of the field quantum, cf. the discussion in section 4.1. Making a trivial shift in the eigenvalue we can further reduce the eigenvalue problem (4.66} to the case where the "potential" vanishes at infinity.

ILLUSTRATIVE EXAMPLE: THE SPECTRUM FOR THE SCHRODINGER OPERATOR


Let us look at foliowing equivalent problem: Find the eigenvalues
for the second order differential operator

(4.70)

dt 2

W(t)<j>

A<j>(t)

where W(t) vanishes exponentially at infinity.

(Notice that we have momentarily replaced the position variable x with the time-variable t. This is because we can now appeal to our intuitive knowledge of classical mechanics. This is a standard strategy in physics: When you want to discuss the qualitative behaviour of a solution to a differential equation you try to find a mechanical analogue.) If we rearrange the above differential equation as follows
d " ~
2

dt 2

= -A<j>(t)

+ W(t)<j>(t)

we recognize it as Newtons equation of motion for a particle of unit mass, which is under influence of partly a constant force, -A, and partlya time-dependent force W(t). Notice that the time-dependent force only acts for a very short time. For simplicity we assume in the following that W(t) is a negative function, i.e. it corresponds to an attractive force. When discussing the particle motion, x between two cases:
(a)
Positive eigenvalues, A = w2 (i.e.
=

<j>(t), we now distuinguish


the continuous spectrum).

When ), is positive the constant force is attractive, i.e. it drags the particle toward the center x = O. In the remote past it will therefore oscillate xCt) ~ a -<X'Coswt + b -ex:Sinwt
t
+- ac

for a short time it comes also under the influence of the time dependent force W(t), but is then left again to the sole influence of the constant force, i.e. in the remote future it oscillates again

156

xCt)

t "'+0<

a +00Coswt

b +00Sinwt

Except for a change in amplitude and a phase shift it thus has the same qualitative behaviour in the past and the future.

X-axis

X-axis

t-axis

CONI'INUOUS SPECTRUM
(b)

DISCRETE SPECTRUM

Fig. 55
the discrete spectrum).

Negative eigenvalues, A =_w 2 (i.e.

In this case the constant force is repulsive. We shall investigate the motion of a particle, which in the remote past is "almost" at rest at the origin, i.e. the particle motion is subject to the following asymptotic behaviour
~Ct) ~

exp[ wtl

-+-00

What will happen with this particle? Especially we will be interested in the following question: lnder what circumstances does it return to (This corresponds to an eigenstate) . the origin in the remote future.

Okay, were it not for the attractive force W(t) the particle would just fly off to infinity. Thus for A less than the minimum of W(t) the total force is always a repulsive force and the particle flies off. When A passes the minimum of W(t) there will be a short time where the particle experiences an attractive force too. For a value Ao sufficiently high above the minimum
th~

attractive force will just be sufficient

to decelerate the particle, send it back to the origin, so'that when W(t) dies off, the particle has just the reverse of the "escape-velocity" (i.e. it approaches the origin exponentially), cf. fig.55. When A is slightly higher than Ao the attractive force will become greater. hus the velocity obtained will be too great and the particle pass through the origin and escapes to infinity. For an even higher value Al the attractive force will now be great enough to decelerate the particle, send it back through the origin, decelerate it again and leave it with the reverse of the "escape-velocity", so that once again it approaches the origin exponentially, cf. fig. 55 . And that is how the game proceeds!

157

Equipped with the experiences from the mechanlcal analogue the following classical theorem
Theorem 1 The spectrum of the second order differential operator
shoul~

not come as a great surprice:

d2
dx 2

we x)

with
x

lim+ Wex) = 0
-+
-00

consists of a continuous spectrum, A=k2, with positive eigenvalues extending from zero to infinity, and a finite discrete spectrum
'" <

Aa

<

AI

<

A2

< .. <

An

< 0

The eigenfunctions of the continuous spectrum can be characterized by the following asymptotic behaviour (A=k 2 ):

exp[ikx]
(4.71)

when

<Pk ex)

{ aek)exp[ikx]+Sek)exp[-ikx]
The eigenfunctions <PKex)

when

of the discrete spectrum can be cha-

racterized by the following asymptotic behaviour (AK=-K ' ):


(4.72)

<PKex)

_ {c_.ex)eXP[+KX] C+",ex)exp[-Kx]

when when

x x

0 0

The ground state <poex) has no nodes, i.e. it vanishes nowhere. The lowest non-trivial eigenstate <pleX) has precisely one node, etc.

This concludes our illustrative example.

Okay, let us return to the eigenvalue problem (4.66). To show the dynamical stability of the soliton we must show that if <Po(x) is a static kink or static anti-kink, then the associated eigenvalue problem has no negative eigen values. The proof is surprisingly simple. It is based upon the observation that we can in fact immediately write down the ground state! Notice first that if <Po(x) is a static solution with finite energy, then it satisfies the field equation (4.28) dx'

~o=

U'[<P
0

(x)]

Differentiating this once with respect to x we get


3 d A. = U"[<P (x) P:'.:I: dA. O :::.....:t o o dx

This we rearrange as

{ _ d' dx '

+ U"[<Poe x )

]}~o=

0
d<Po/dx is an eigen

A comparison with (4.66) then immediately shows that function with the eigenvalue

A=O. Recall then that the kink, respectively

158

the anti-kink, are monotonous functions. Thus the assiciated zero-mode has a constant sign throughout the whole x-axis. But then the zero-mode
d~+

/dx

has no nodes, i.e. it is precisely the ground state for the

eigenvalue problem (4.66): This shows that there are no negative eigenvalues. In the above argument the precense of a zero-mode ated with the finite-energy static solution lational symmetry, i.e. the configurations
~o(x)
~o(x) d~o/dx

associ-

is very central. Its

existence can be understood in the following way: The model has transand
~o(x+a)

has the same static energy. configuration as

For a small

we can expand the displaced

~o(x+a) ~ ~ (x)+a dPO (x+a)


o Clearly dPOCx+a) da1a=O Thus we see that a pertubation along

dala=O

d~o/dx

leaves the static energy

invariant. According to (4.68) it must then be a zero-mode, i.e. an eigenrnode with the eigenvalue O. To conclude: The zero-mode comes from
translational symmetry.

Before we leave the stability considerations we want to mention that it is some times possible to prove the instability of a static solution by examining a particular simple deformation of the given configuration. The most important example is Derrick's scaling argument: Consider a scalar field theory in D space dimensions based upon the Lagrangian density
L

= -~(d ~a) (d)J~a)

)J

_ U[~al

The corresponding static enerBY functional is given by


=

~f (~) (~)dDX
ddx dX

fU[~aldDX
d

Hl[~al

H2[~al

R
~A (x)

R
a
= ~

Suppose ~a(x) is a static solution and consider the scaled configuration

(Ax)

The scaled configuration has the static energy

159 If ~a

is stable this must be stationary at A=l i.e. we must demand

o
This implies the following stability condition (4.74 )
If

D>2 H2

the coefficients

(2-D)

and

has opposite signs. Since

Hl

and

are non-negative they must both vanish. Thus we conclude:

A non-trivial static solution in a scalar fie ld theory is unstable if the dimension of the space exceeds 2.

If

D=2

there is a small loop-hole. The condition (4.74) this


H2[~1

time reduces to

and this can be fulfilled by a non-trivial configuration provided the potential energy term is absent. This is known as the exceptional case in two dimensions and we shall return to it in section 10.8. If D=l the condition (4.74) finally reduces to

i.e. a static configuration corresponding to a single scalar field can only be stable provided it satisfies the virial theorem; (4.75)

~J(~)2dX = JU[~(X)ldX

Interestingly enough the kink and the anti-kink in fact satisfies this identity point-wise, i.e. not only are the integrals identical but so are the integrands too! This follows immediately from (4.29). Moral: With a single exception Derrick's scaling argument shows that there are no stable solitons in higher dimensional scalar field theories. To stabilize the static solutions we must therefore extend the model somehow. This can, e.g., be done by coupling the scalar field to a gauge field. We shall return to this in section 8.7,8.8 and 10.9.

160

4.8

THE PARTICLE SPECTRUM IN NON-LINEAR FIELD THEORIES


Finally we will give a brief (and naive) discussion of the par(l+l)-dimensional scalar field theory with a dege-

ticle spectrum in a

nerate vacuum. When we include quantum mechanical considerations we see that there are two basic types of elementary particles: (a) On the one hand there are the field quanta represented by small oscillations around a classical vacuum. A field quantum has the mass (4.76) where
~

m=~ o
is a classical vacuum.
c~~ssical

(b) On the other hand there are the solitons/anti-solitons, which are present already on the
(4.77)

level. A soliton has the mass

M
~l

fV2U[~l
~1

d~
(Strictly speaking

where

and

~2

are consecutive classical vacua.

there are quantum mechanical corrections to this mass formula but we shall neglect them.) Apart from these elementary particles there may also be composite
particles, Which can be interpreted as bound states of the elementary

particles, i.e. the field quanta and the solitons/anti-solitons. E.g. we have seen in the sine-<;ordon model that a soliton and an anti-soliton can form a bound state - the bion. Such a bion must then also be included in the particle spectrum. In the classical version its mass can be anywhere between zero and two soliton masses. Quantum mechanically the mass of the bion is quantized, i.e. it can only take a discrete set of values. We shall return to this in section 5.6. There is also the possibility of having bound states consisting of a field quantum bound to a soliton. Let us consider the interaction between field quanta and a single soliton in the weak field limit, i.e. we consider small "osci1lations" around a kink-solution
~(x,t)

~+(x)

+ n(x,t)

As we have seen the "oscillation" n(x,t) can be decomposed on normal modes


~(x,t)

~ ~k(x)exp[-iwktl

where

~k(x)

is an eigenstate for the eigenvalue problem


{_ d_2 }

(4.66), i.e.

+ UII[~+(X) 1 ~k(x) = wk~k(x) 2 dx Let us introduce the "potential"

161 (4.78 ) and the "shifted" eigenvalue (4.79)

A
~_ d L dx2
2

= w~

- m

Then the eigenvalue problem reduces to

~ W(X)}~k(X)

A~k(x)

with
x

lim W(x)

-+oo

Thus we can take over the results from theorem 1: (a) There is a continuous spectrum consisting of non-negative eigenvalues lity Asymptotically the fluctuation is given by (4.80) n(x,t) = ~k(x)exp[-iwkt] ~ exp{i[x/w~-m2 - wkt]}
xO
0<1,<00. They correspond to

frequencies w satisfying the inequak

This corresponds to a free particle with momentum P and energy E given by


P =

/w~

m2

and

Notice that E2- p2= m2 ,i.e. the rest mass of the particle is precisely m. Thus it is interpreted as a field quantum which is scattered on the
soliton. When the field quantum is far away from the soliton it beha-

ves like a free particle, but when it is close to the soliton they interact in a complicated way and the approximation (4.80) correspondingly breaks down. (b) There is also a finite discrete spectrum AO < Al < . < An < 0 As we have seen the groundstate AO corresponds to a zero-mode, i.e. AO

_m 2

The groumd state arises from the translational symmetry of the model. It represents no increase in the energy relative to the soliton, but simply reflects the fact that the
kink-solution,~+(x),is

degenerate, i.e. it kink-

is just one particular member of a continuous

family,~+(x-Xo),of

solutions. The other discrete eigenstates, if they are actually present in the model, are more interesting. In that case the fluctuation dies off exponentially when we go away from the soliton. They are interpreted as bound states of a field quantum and a soliton. Notice that the total mass of the bound state lies between M and M+m. The question then arises if there are models with non-trivial discrete eigenstates.

162
Exer>cise 4.8. 1

Problem: (a) (4.81 ) (b) (4.82)

Consider the ~!model. Show that the potential (4.78) is given by


W(x) = - Cosh 2 [axl with a =

7z
1S

Consider the sine-Gordon model. Show that the potential (4.78) given by
W(x) Cosh2[axl with a
II

Fran exercise 4.8.1 we see that in our most popular models, the ~' model and the sine-Gordon model, we must determine the discrete spectrum for the Eokhardt potential:
w(

2 a s (s+l) - Cosh 2 [axl

ILLUSTRATIVE EXAMPLE: THE SPECTRUM FOR THE ECKHARDT POTENTIAL.


The eigenvalue equation is given by (4.83) -

~
dx
2

- a s(s+1) Cosh2 [axl

A~

By a suitable transformation we can eliminate the transcendental function in front of


~.

Consider first the substitution:

~(x) = Cosh-s[axlw(x)
By a straigth forward computation we get the following equivalent equation for w (x) : (4.84 )

- --- + 2asTanh[axldx - a 2 s 2 w(x)


dx 2

d 2w

dw

AW(X)

Since the Eckhardt potential is an even function, i.e. it posseses the symmetry, x
~(x).
~

-x, we need only look for even and odd eigen functions is an even function, so we need also only
O~x<oo.

Furthermore Cosh[axl

look for even and odd solutions to (4.84). Especially we need only solve equation (4.84) on the positiye semi-axis, i.e. for The odd eigenfunctions then correspond to the boundary condition w(O) = 0 while the even eigenfunctions correspond to the boundary condition dw (0) = 0 dx We can now make the further variable substitution
~

Sinh2[axl

The eigenvalue equation (4.84) then reduces to (4.85)

163

The even solutions can be expanded in a power-series like

~a I;n n=O n while the odd solutions can be expanded in a power-series like
(4.86a) w+(I;) = (4.86b) w_ (I;)

00

I; 2 l: an <:n n=O

substituting (4.86a) into (4.85) we get the following recursion relation: 1 A n(n-s)+ij"(az+s 2 ) a n +l = (n+\) (n+l)
A discrete eigenvalue occurs when this power-series terminates after a finite number of terms, i.e. when

4a 2 (n

Since

~+(x)

(l+I;)->;sw+(I;)

~
I; -+ 00

(l+I;)->;sl;n

we must furthermore demand 2n < s in order to get a norrnalisable wave function. In the odd case we get the recursion relation
A (n+>;) (n+>;-s) + 1 ij"(az+ s2)

Ie :
(n

25 3 n

+ :2

+ :2)

This generates a polynomial (corresponding to an eigenfunction) when

Since

(x)

(1+1;)

-\sw

(1;)

we must furthermore demand 2 (n+l.;) < s The even and odd case may be comprised in the following way: The eigenvalues are given by (4.87) Ak = Even
K

a 2 (s_k)2

k integer, k<s

then correspond to even eigenfunctions, while odd k correspond

to odd eigengunctions. The associated eigenfunction is given by when k is even (4.88)


~k(X)

when k is odd

164 where

P denotes a polynomial of degree n, which is obtained from n the above recurrence relations. Notice that ~k(x) has precisely k nodes

in accordance with theorem 1:

Okay, we can then apply the results obtained in this example to the ~'-model and the sine-Gordon model. (a) In the ~4-model we get from (4.81) that s=2. Consequently there are precisely two discrete eigenstates. The corresponding eigenvalues are given by and
Al = - a' =

~ll'

According to (4.9) and (4.79) they correspond to the frequencies i.e.

and

wf =

The first one is of course just the zero-mode, while the second one represents a genuine bound state! We can also determine the corresponding eigenfuction where
~1

(x). It is odd and corresponds to the case

w_(~) reduces to ~~. Thus ~, (x) is proportional to


Sgn[x] (1+0 -l~~ =

Sinh[~x]
Cosh' [~xl

(b) In the sine-Gordon model we get from (4.82) that s

1. Con-

sequently there is preciely one discrete eigenstate (corresponding of course to the zero-mode). In the sine-Gordon model there is consequently no bound state of a field quantum and the soliton.

SOLUTIONS OF WORKED EXERICSES:


No . 4.5.1(a) Using the trigonometric identity

Tg(a+s-i) ; Tg(a)Tg(S)-l Tg(a)+TgtSJ


and the expression for the static kink solution
Tg(~)

A~+

exp(llx)

exp(2\lYx) - 1 exp[lly(x+vt+Xo )l+exp[lly(x-vt-x0 )1


(b) To find the asymptotic behaviour as
t~~

we use that
RI

lim Cosh(x+a) Coshx X ~ a:

exp(a)

i.e.

Cosh(x+a)
x

ea.cosh(x)
'"

165

No.4 .6.1
(a) If we introduce the notation

a
(4.89a)

exp(,,)

we can write down the BacklUd equations in the following condenced form:

d+(,/ll+<jlo)

(4.89b) d+(<P3+'!>l)
(4.89cJ

2exp[::"11Sin[1 + 2exp[::"21Sin[~1 2exp[::"21Sin[P2; PO l + 2exp[::"11Sin[<P3; P2 1 ::Zexp[::"11Sin[Pl;P01+ Zexp[::"21Sin[~1 + + P1.:.P.1 -Zexp [+ -~2 1Sin [P.d.! Z 1+ Zexp[-"llSin[ 2 1
+
+ +

d+(<P2:;<PO)

(4.89d) d+(<P3:;<P2)

By adding and subtracting the first two we get

d+(<P3+<PO)

Similarly we obtain from the last two +

- ) d+ (<P3+<PO

AS the left hand sides coincide we therefore deduce tpe consistency relation

CXP[::"11{Sin[~1~Sin[Pl;P01}
+

==

eXP[::"21{Sin[Pi;Pll:;Sin[~1}
-

Using the well-known trigonometric identity

Sinx :: Siny

ZSin[~lCos[~l

the consistency relation reduces to

Using the condenced notation introduced in the exercise this can be written as

(4.61)
(b)

exp[::"llSin[A+l == exp[::"21Sin[A_l
+

Differentiating the consistency relation (4.61) we get

exp[::"llCos[A+ld+A+
Writing out

exp[::"21Cos[A_ld_A_
+
+ +

A+ this leads to

{eXP[::"llCoS[A::l- eXP[::"21COS[A+l}d::<P3 {exp[::"llCOS[A;:l- eXP[::"21COS[A:;1}d;:<Po + {eXP[::"llCOS[A;:l-

eXP[::"21COS[A~1}d;:<Pl

:; {eXP[::"llCOS[A::l- eXP[::"21COS[A:;1}d::<P2
The middle term is okay, but the first and the last term must be rearranged using (4. a) and (4. c) so that we can get rid of d+<PO and d+<P2. After some algebra we then end up with

166

(*)

{exp[!AllCOS[A!l-

eXP[!A21COS[A;1}a!~3==
+

!{eXP[!AllCOS[A!1-eXP[!A21COS[A;1}a!~1
+ -4exp [+ -AI 1exp [+ -AzlCos [A_1Sin[ ~ 2 1

;2{eXP[~AllCOS[A!1+ eXP[~A21COS[A;1}eXP[~AzlSin[PZ;POl
The first term on the right hand side is just what we want, but the remaining two terms need a little massage! Introducing. the abbreviation

B ; (p3+PO)~(pZ+Pl)
+

we have

<h:!h
2

!B ; A + +

~; ~B+ +A 2 +

P3-~1

; B

+-

If we introduce these abbreviations the last terns in (*) reduces to

exp[:AlleXP[~A21{2COS[A;lSin[B!lCOS[A~1+ 2COs[A+1Sin[A;lCos[B+l ;
+

-4COS[A_1Sill[A+1COS[B+l}
-

'- - - - - - - - - - - - - - - ~- - j

+ + + + Using the consistency relation (4.61) at the combination indicated by the broken line this is further reduced to:

exp[~A21eXP[~A21{2COS[A_1Sin[A_1COS[B+1-2COS[A_1Sin[B+1COS[A_l}

2exp[~Allexp[~A21rCoslA lSin[B+ lCos[A+ 1+ CostA+ lSin[A- lCos[B+ l} 1 .. + + - 2eXP[~AzleXP[~A21{COS[A_1Sin[A_1CoS[B+l+ COS[A_1Sin[B+1COS[A_l}


+ + + -

2eXP[~A21{exp[!AllCOS[A!1- eXP[~A21COS[A+l}{Sin[B!lCOS[A+l+Sin[4;lCoS[B~1==

2eXP[!Azl{eXP[~AllCOS[A!1- eXP[~AzlCOS[A+l}Sin[P3;~11
Inserting this back into (*) this fin~lY reduces to

a~~3 ; !a~~l + 2exp[~A21Sin[P3;~11


(c)
Writing out the consistency refation (4.61) explicitly as

aISin[(p3-PO)~(pZ-PI)1
it can be rearranged as

azSin[(P3-PO)~(Pz-PI)1

aISin[~lCos[~l + aICos[~lSin[PZ~Pll

== a Sin[~lCos[PZ-~ll - a Cos[~lSin[~l 4 4 4 4
This immediately leads to

(az+al)Sin[~lCos[~l == (az-al)Sin[~lCoS[~l
from which the Bianchi identity (4.60) follows trivially.

167

ehapler 5
PATSINT.BGRAI,S AND INSTANrONS
5,1 THE FEYNMAN-PROPAGATOR REVISITED
In the remaining chapter we would like to include a few aspects of the quantum theory of fields and particles. single particle in one space dimension. In ordinary quantum mechanics the central concept is the Feynmanpropagator K(X lt lx ;ta ). As we have seen it denotes the probb b a ability amplitude for a particle to move, i.e. to propagate, from the space-time pOint (xa,t a ) to the space-time point (xb,t ). b The propagator governs the dynamical evolution of the Schr6dinger wave function according to the rule (2.18) W(xb;t b ) To simplify the discussion we begin our considerations with quantum mechanics of a

JK(Xb;t b I xa;ta)W(xaita)dxa Furthermore it satisfies

and as a consequence all information about the quantum behavior of the particle is stored in the propagator. the group property (2.25) K(xc;tclxa;t a )

= JK(Xc;tclxb;tb)K(Xb;tbIXa;ta)dXb

It will be useful to recall what we have learned so far about the propagator. For simplicity we consider a physical system characterdx 2 ized by the Lagrangian

tm(lt)
Then we have previously shown that

- V(x)

1)

The propagator can be

~ep~esented

as a

path-integra~

x(tt)=2b
(2.21)
=

ifbm dx 2-v ( x) 1dt } D[x(t) exp{fi [Z(iIT)


t

where we sum over all paths oonneoting the spaoe-time points


(xa;t a )
and

(xbitb).
so~ution

2)

The propagator is the unique

to the Sohrodinger equation,

satisfying the initial oondition


K(xb;talxa;ta) o(x b - xa)

(Cf. the disoussion in seotion 2.4.

In the mathematioal terminology the

propagator is thus a Greens function for the Schrodinger equation). ;;) The propagator has the following Moomposition on a complete set of eigenfUnotions for the Hamiltonian operator:
(2.70)

1< (xb;t b Ixa;t a ) =

n=o

<l>n (x a ) <Pn (x b ) exp[-

En (t b - tal 1

cf. the worked exercise 2.11.;;.


Of these characterizations only the last two have an unambigious meaning. The first one, involving the path-integral, is only a formal description since we have not yet indicated how one should actually perform such a summation over paths! Before we discuss the properties of the propagator further we will indicate yet another characterization of the propagator. that in quantum mechanics a physical quantity a Hermitian operator eigenvalues An T Recall is represented by

and that this operator can only have real In general a

which represent the possible values of the quantity

T, when we try to measure it in an actual experiment. for the operator

wave function can be decomposed on a complete set of eigenfunctions

T:
an is then interpreted as the probability An' i.e. in an actual experlinent we will

The coefficient measure the value operator


(5.1)

amplitude for measuring the value An with the Xo E.g., the position

~ility

Pn ~ 1~12. is represented by the multiplication

The corresponding eigenfunctions are given by


(5.2)

Wxo (x)

o(x-x o )
This is re-

Notice that these eigenfunctions are not normalizable. trum, i.e. a continuum of eigenvalues.

lated to the fact that the position operator has a continuous specThe eigenfunctions, however,

169 still satisfy the completeness relation (2.69):

WXo (x 1 ) WXo (x 2 )dx o = JO(XO-X1)O(XO-X2)dXo = o(x 1 -x 2 ) Similar the momentum Po is represented by the differential

operator (5.3) PoW (x) = - Hi.

1:. ax
with p

The corresponding eigenfunctions are given by the plane waves


(5.4)

Wk (x) = __ 1 __ exp[ik x]

nrr

o = ilk 0
t~e

(Notice that we parametrize the momentum eigenfunctions by number ko' rather than the momentum

wave

Po = ilk )' The momentum o eigenfunctions are not normalizable, but, like the position eigenfunctions, they satisfy the completeness relation (2.69):

Here we have used that the Fourier transform of Dirac's delta function is simply 1, i.e.

Jo(X)eXP[iYX]dX
(5.5)

= exp[iyx] lx=o

By Fourier's inversion formula we thus get

(xl

= 2111

exp[ -iyx] dy

Exercise 5.1.1 Problem: Compute the free-particle propagator by means of the formula (2.70), which in this case reduces to
+00

-+2
:;---r::--"T"

K(xb,tblxa;ta)

f ~k\Xa/Wk(xb)

exp[-

E-.

2m (tb-ta']dk

where

~k

(x)

is given by (5.4).
i exp [ - E HT].
It has

Consider now the time-evoZution operator the matrix element

where

Ix > is the position eigenfunction (5.2). Using a complete a set of eigenfunctions for the Hamiltonian H we can now expand the

position eigenstate Ix a > =


n

<wnlxa>lw n >

If

we furthermore use that

170

then the matrix element reduces to <xblexp[- fi HTl IXa> =


i -

L -Wn(x a )

<xb1wn> exp[- fi En T ]

Wn(xalWn(xb) exp[-

%En T ]

According to the third characterization we have thus shown:

4)

The propagator can be represented as a matrix element of the time evolution operator exp[-

BTl , i.e.

(5.6)

K(~;Tlxa;O)

<~Iexp[- ~

Hrllxa>

As an especially important example of how one can extract infor-

mation from the propagator, we will now show how one can in principle calculate the energy spectrum of a particle. For this purpose we consider the trace of the time-evolution operator
(5.7)

G(T)

Tr{exp[-

HT]}

= I<xo1exp[-

HT]lx o > dx o

jK(Xo;T1Xo;0) dx o Using the third characterzation of the propagator this reduces to


(5.8)
~n(x),

where we have used the normalization of energy eigenfunctions

Thus the trace of the propagator decomposes very simple in terms of

harmonic phase factors on ly depending upon the energy leve 1-s!


Introducing a "Fourier transform"
(5.9)

G(E) =

% J
o o

G(T)exp[i ET]dT

we in fact immediately get:


(5.10)

G(E)

i Y J n n=o

exp[-

~(En-E)TldT

Consequently the energy levels show up as poles in the transformed

trace of the propagator.

171

In the above discussion we have been focusing upon propagation in the "position space". We could equally as well use other physical properties, such as the momentum, as our starting point. Thus we define K(Fb;tblp;t) as the probability amplitude for a particle, which is released with moment~ a p at the time t . . l ) denotes a to be observed w1th momentum Pb at the t1me t . Thus K( Pb;t a p;t b the probability amplitude for a particle to "proBagate in momentum ~pa~e" from the point (p;t) to the point (pt,;t ). According to the basic rules of quantum b mechanicsath~ momentum propagator is now related to the position propagator as follows K(pt,;tbl Pa;t a ) = H<pt,IXb)K(Xb;tblxa;ta)<xal Pa)dxadxb where Ip) is the momentum eigenfunction (5.4). interpret~d as follows: Here the various terms can be

a) <x Ip) is the probability amplitude for a particle with momentum Pa to be at theapo~ition x. b) K(~;tblxa;ta)a is the p,obability amplitude for a particle to propagate from xa to x b c) <Pbl~) is the probability amplitude for a particle at the position xb to be observed with the momentum Pb' As usual we then sum over all the alternative ways the particle can propagate from Pa to Pb ' i.e. we integrate over the intermediary positions. Using that and The above relations can now be rearranged as

The momentum propagator is thus obtained from the position propagator by a "double" Fourier transformation. Consider e.g. the free-particle propagator 2 I m im (x b -x a ) (2.28) VZTI1b(t b -t a ) exp[zt t b -t a ] In momentum space this is converted to
K(ptiTIPaiO)

r-m- 1 II i m (~-Xa) = V2:ilir 2rr exp{K['2 --r'2'I' [xa


m

P~

This is just a Gaussian integral, cf. (2.27), which can be calculated in the usual way by "completing the square". Using the identity,

'2 - r

(~-Xa)2

- ptxb

+ Paxa =

T nor.. 2 + iii(Pa - ""r')] -

2iii T + (P a -Pt)~ ,

the x -integration cancels the factor formu1a: 2 K(Pb;T IPa;O) = exp[i Pa K 2m

;f2TI~~T

in front and we are left with the - Pb)xb]d~ We there-

T] 2rr

J exp~(Pa i

But according to (5.5) the last integration just produces a a-function! fore finally get: 2 Pb - P a i Pa (5.12) a(--~-) exp[- fi' 2m T]

This simple result has an intuitive explanation: Since the momentum of a free particle is conserved, it follows that once we release a free particle with

172 I momentum p it will necessarily still have momentum p once we decide to measure it. a This explains the a-factor. Furthermore a a free particle with momentum p will have the definite energy E = p2/2m. Thus the propagator has the charact~ristic phase factor indicated by the d~ Broglie rule, cf. (2.32),
exp[-

i 1i

ETl

= exp[- i fl

Pa Tl 2m

Evidently the free-particle propagator is much simpler in momentum space. This is a general feature of quantum theory and therefore it is the momentum propagator which is most often displayed in the literature.

5.2

ILLUSTRATIVE EXAMPLE: THE HARMONIC OSCILLATOR


As an example of how to use the preceding techniques we now take

a closer look at the harmonic oscillator, which is characterized by the Lagrangian:


(5.13 )

The corresponding equation of motion is given by 2 d x 2 dt Notice that w


2 w x

is the cyclic frequency of the oscillator. Sinde the Lagrangian is

First we must calculate the propagator. gator, cf. the worked exercise 2.4.1. reduces to the form, (5.14 ) where (xb;T) x
t

quadratic, we can use the same trick as for the free-particle propaThe propagator consequently dx

A(T)exp[~
= xcI

f L(xcll

d~l)dtl
(xa;O) and

o
(t) is the classical path connecting

To determine the classical action we notice that the general solution of the equation of motion is given by xcl(t)

=a

Coswt + b Sinwt a and b so that the boundary

we shall then adjust the parameters conditions

= xa
are satisfied. unique solution
(5.15)

and this is easily obtained and we get the Xb - xa CoswT Sinwt

If

wT

nIT

and

173 If on the contrary wT = nrr (corresponding to either a half period This is because a classical

or a full period) we are in trouble.

particle, after a half-period, necessarily is in the opposite pOint. After a full period it is similarly necessarily in the same point.

X-axis

Classical trajectories for the harmonic oscillator.

wT-axis

Fig. 56

For

wT = nrr

we can therefore only find a classical path if

xb = (-l)n Xa , and, if that is the case, any classical path passing through (xa;O) will in fact also pass through (Xb;T)! In analogy with optics we say that the classical path has a caustic nrr. At the caustics the propagator thus becomes singular! Okay, neglecting the caustics for a moment we can complete the calculation. The classical action is given by S[X cl ] mw =:r 2 2. 2 [(b -a )CoswT S~nwT - 2ab Sin wT] a and b we now get after a when wT =

and substiting the values (5.15) for lengthy but trivial calculation: (5.16 )

As expected this is highly singular at the caustics, i.e. when

wT

nrr. We proceed to calculate the amplitude A(T). It can be deterBy translational mined by the same trick as the one we used for the free-particle propagator, i.e. by using the group property (2.25). invariance in time the group property reduces to

Substituting the expression (5.14) we now get

174

This integral looks pretty messy, but after all it is just a harmless Gaussian integral. Using the identity

imw (X +X )C05WT Z - ZX3xZ ~[ SinwT + Z imw Sinw(T Z+T,) zh SinwTZSinwT, [xZ


we can immediately integrate out the x -dependence whereby the right 2 hand side reduces to

3Z

Most reassuring the x 3 ,x 1 -dependence is the same as on the left hand side and we can therefore factor out the phase factor completely. We then obtain the following functional equation for the amplitude A(T):

A(TZ+T,) A(TZ)A(T,)

ZninSinwTZSinwT, mwSinw(TZ+T,)

mw

It has the general solution

but as in the case of the free-particle propagator we shall neglect the additional phase exp[iaT], i.e. put

a=O.

(This corresponds to A(T) reduces

a normalization of the energy,and has no physical consequences). Notice that, in the limit where to the free-particle amplitude A(T) as we expect it'to do! Consequently we have now determined completely the propagator for

0, the function

= J2nlhT

the harmonic oscillator. (5.17)


K(XbIT!Xa,O)

In its full glory it is given by

=.; 2TIif:WSinwT e:xp[

2fi ~l~wT{

(X; +

x!)QOSWT - 2X b Xa }]

175

Notice too that it also reduces to the free-particle propagator in the limit of small we in fact get im til 2 2 } exp[2fl SinwT {(X b + xa)COStilT - 2xb x a ] ... T. Expanding the phase to the next lowest order

2 2 2 T im 2 imtil exp[ 2fiT (x b - xa) ]exp[- ~ {Xb + xa + xbx a } 3] Here the first phase factor is the free-particle phase factor, while the second phase factor reproduces the interaction phase factor, exp[ -

t JV(x)dt]
o

integrated along the free-particle path, i.e. the straight line x

= xa

(xb-Xa)t/T.

Thus the "infinitesimal propagator" is in Observe also that if we had

accordance with our previous ansatz (2.53), which we used in our derivation of the Schrodinger equation. included the phase factor exp[iaT] To be honest, the above formula, this would no longer be correct. (5.17), for the propagator of This is due to the

the harmonic oscillator is not quite correct. integrals of the type (5.18)

fact that we have been rather careless when performing Gaussian


+""

J e- iax2

dx =

The above formula is obtained from the rigorous formula, (2.27)


[ A

>

by performing an analytic continuation. a complex number.

But this analytic continua-

tion is actually double-valued, since it involves the square root of This gives rise to a phase ambiguity in the GaussAs a consequence the propagator has a phase
~1

ian integral (5.18).

ambiguity (corresponding to a phase

or

~i).

Below the first caust-

ic, i.e. when T < ~ , the formula (5.17) is correct simply because til it is in correspondance with the free-particle propagator in either of the limits til
+

or

O.

But once we have passed the first We shall return

caustic, we no longer control the phase ambiguity. to this problem in the end of this section.

- 176 easily calculate the energy spectrum. is given by


G(T)

Now that we have the Feynman-propagator at our disposal, we can The trace of the propagator

V2(C05WT - 1)
But this is easily rearranged as (5.19)
G(T)

2iSin!wT

- e -iwT

~ e-i(n+!)wT
n=Q

A comparison with (5.8) then immediately permits us to read off the energy spectrum (5.20) In the case of the harmonic oscillator the possible values of the energy are thus evenly spaced and, perhaps a little surprising, the ground state energy is not zero as in the classical case, but is instead given by (5.21)
The zero-point energy is often explained by referring to Heisenberg's uncertainty principZe, according to which the indeterminacies of the momentum and the position are bounded below by (5.22)

<x><p>

fi

Unlike the classical case, the particle therefoce cannot be at rest in the bottom of the potential well, since this would cause both <x> and <p> to vanish. It must necessarily "vibrate a little". A rough estimate of the energy is given by

E
~/<x>

=~ 2m

+ tmw2<x>2

We want to minimize this subject to the constraint (5.22). Replacing <p> and minimizing the resultin~ expression with respect to variation in produces the minimum

by <x>

Except for the factor !, which we cannot account for by such a primitive argument, this is the same as the zero-point energy (5.21).

Once we have the propagator at our disposal we also control the dynamical evolution of the Schrodinger wave function. solution to the Schrodinger equation. the normalized wave function At the time t=O As in the we consider free-particle case we can now find an exact and especially simple

177

(5.23)

1jJ(x,O)

y'~

'rmw

mw 2 exp[ - 2fl(x-a) ]

This corresponds to a Gaussian probability distribution centered at the point x=a. At a later time t the wave function will, according to (2.18) evolve into
1jJty,t)
=

IK(y;tIX;O)1jJ(X,O)dx
4/iii;; 2/ lItJJ V ~ V21TlflSinwt

illtJJ 2 2 exp[2li.Sinwt[(y +X )Coswt - 2yx} -

IlliJ

2 2li.(x-a)] dx

As usual this is just a Gaussian integral. Using the identity

illtJJ 2 2 2fISinwt [(y +x )Coswt - 2yxJ -

lItJJ 2 m(x-a)

lItJJie iwt . t 2 2 1W (aSinwt - iy)] - ~[y 2li.Sinwt [x - iethe integration can be immediately performed, and the wave function at the time t reduces to

(5.24)

~(y,t)

WK exp[- ~(y-aCoswt) 2 ]exp[- l~t]exp{_ ~[2aySinwt -

~ Sin2wt]}

Apart from a complicated phase factor this is of the same form as (5.23). P(x,t) Consequently the corresponding probability distribution, 11jJ(x,t) 12 =

/I

exp[ -

~w

(x - a Coswt)2]

is still a Gaussian distribution, but this time it is centered at x = a Coswt This is highly interesting because this means that the wave packet

oscillates forth and back following exactly the same path as the classical particle!
If a=O the probability distribution reduces to the stationary distribution

corresponding to a classical particle sitting in the bottom of the potential well. state. function for (5.25) Thus it is closely related to the classical ground It should then not come as a great surprise that the wave a=O, i.e.

178

represents the quantum mechanical ground state. the simple time dependence exp[- ~ Eotl. with !hw. equal to the ground state energy

Notice that it has Eo precisely being !fiw.

It must therefore in fact

be the eigenfunction of the Hamiltonian with the eigenvalue

Exermse 5.2.1
Introduction: Consider a free particle, where the corresponding wave fUnction at the time t=O is given by ~ x2 1jJ(x,O) = ~2 exp[-1j"(?"l Problem: a) Show that its wave fUnction at a later time t is given by

(Unlike a particle which is at rest in the bottom of a potential well, this wave packet thus spreads out in time!) b) Show that the corresponding probability amplitude in 'momenttun space is given by

~(k,t) =

i2 an

2(t)

HIt 2 2 exp[ - k a (t)l exp[8ma2cr2(t)lexp[ <p><x> =


'!\ "2

and as a consequence

Also the rema~nlng eigenfUnctions can easily be extracted from the propagator. With the third characterization of the propagator in mind we decompose it as follows K(y;Tlx;O) = e- iwT / 2

vi

mw 2' T exp{ mw 2'wT [~(y2+x2)(1+e-2iwT)_ 2xye- iwT l} 11fl(1-e- lW ) 11(1-e- 1 )


z=eiwT this can be rewritten as follows _ ~[ x +Y1_~2(2 2) 2
2

If we introduce the variable K(y;Tlx;O)


J..

r-mw= z-~ exp[- ~(X2+y2)l~ exp{


,'"'
fJ. n

xyzl}

This should be compared with n=O n

E ~~ (y) exp[_i E Tl = z- E ~ (x)~ (y)zn


n n=O jn n

If we put

~n(x)

= ~ exp[-

~x2J(2nn! )-!Hn(~)
and

and introduce the rescaled variables

u=~
it follows that
(5.26)

v=~,
has the generating formula
( 1 -z 2
)-.

Hn(u)Hn(v)

exp

2 2 2) 2 [uvz - (u +v Z 1 1_Z2

179

Since the left hand side is a Tay~or series in z we can actually find 2-n H (u)H (v) by differentiating the right hand side n times and thereafter put z o. I~ fol~ows trivially that H (u) is a polynomial of degree n in u. In fact it is a Hermite polynomial and thg above formula is the so-called Mehler's formula. We leave the details as an exercise:

Worked exercise 5.2.2


Introduction: mula The Hermite polynomials Hn(x) n=o Problem: a) Deduce Rodriques' formula 2 d n _x2 Hn(x) = (_On eX e n dx b) Show that e -x 2
1 _t,2

are defined by the generating for-

liT f

+ 2it,x dt,

and consequently H (x) = (_2i)n e x2


n

liT

Jt,n e-t,2 + 2iE,x dE,


[l-z2j

c) Prove Mehler's formula

(5.28)

n=o

-! exp [ 2xyz -

x 2+ Y - z

(2

2) z2

Using Mehler's formula we can in fact recalculate the propagator. must then determine the eigenvalues En and the eigenfunctions ~n(x) Hamiltonian: 2 2 h d 2 2 (- + jlllL1l X )$ (x) = E ~ (x) 2m dx2 n n n

First we for the

(This eigenvalue problem is treated in any standard text on quantum mechanics!). Once we have the eigenvalues and the eigenfunctions at our disposal we can finally explicitly perform the summation in (2.70) by means of Mehler's formula. Notice that the phase ambiguity is still present. The Taylor series in Mehler's formula has convergence radius 1, due t~.&e poles at z = 1. We are especially interested in the behaviour at z = e ~ We are thus actually working directly on the boundary of the convergence domain! This boundary, i.e. the circle Izl= 1, is decomposed into two disjoint arcs by the poles z = lJ Furthermore the poles correspond precisely to the caustics wT = nTI. At two times t and t , h separated by a caustic, we have thus no direct relation between the pftases.

180

Let us finally tackle the problem of the phase ambiguity in the propagator for the harmonic oscillator. Below the first caustic we know that the exact propagator is given by
(5.17)

K( ~,TI Xa' 0) - e-i1T/4 v'ZnMInwT rrrr;;- exp {imu 2 2) COSW:r - Z~Xa]} mSinwT [( ~+Xa T

1T T <w

When

=~

this reduces to the particular simple expression 1T.

K(~;Zwlxa'O)

=e

-i1T/4 !iii;

vin% exp[-

lh~xal

.1Illl

From the group property (2.25) and the translational invariance in time we know that

K(Xb;~IXa;O)

= =

!K(Xb;!wIX;O)K(X;z:IXa;O)dX i1T Z e- / i1TJexp[- i~(Xb+Xa)xl~dx

which by (5.5) reduces to

This takes care of the first caustic. We now proceed in the same way with the second caustic 1T K(xbl : lx a ;O) = J K(x b ; ~IX;O)K(X: ~IXaIO)dX

Continuing in this way we finally obtain the propagator corresponding to an arbitrary caustic: (5.29 ) But the propagator on a caustic serves as the initial condition for the propagator on the subsequent segment, i.e. we must demand lim
T ...
+

K(xb;T!xa;O)

=e

-in.~

!!!
w

n o[x b - (-1) Xa1

Consider the expression e


- i 1T /4 / mw { imw [( 2 2) T 1} V21T~ISinwTlexp ZnSinwT Xb+Xa Cosw - ZXbXa

181

(Notice that it differs from Feyruman's expression (5.17) by an inclusion of the absolut value of SinwT under the square root). If wT
=

nrr

+ WE:

0<1

we can approximate the above expression by

e-i1T/41z~exp{(-1)n
e -i1T/4

lR-[(X +X!)(_1)n - 2x x a ]} b n ]2}

v'~exP2nxb-(-1)Xa

rm--

rim [

Compairing this with the "infinitesimal" free particle propagator we see that it has the limit
O(x b -

(-1)n Xa )

All we need to get the correct boundary condition is therefore just an inclusion of the additional phase factor -i ~ Ent [ WT 1
e

1T

where Ent[xl is an abbreviation of the entire part of X, i.e. the "greatest integer below x. The correct propagator for the harmonic oscillator is consequently given by the following formula:

(5.30)

This is the famous Feynman-Soriau formura which takes into account the proper behaviour of the propagator at the caustics. It was discovered by Soriau in 1975 and has an interesting geometrical-topological interpretation, which we unfortunately cannot explain within this simple framework. It has also been derived by adding a small anharmonic term, say x4, to the potential, and then studying the limit of the anharmonic propagator as +0.

5,3

THE PATH-INTEGRAL REVISITED

Up to this point we have been treating the path integrals in a handwaving way. Although we are definitely not going to turn it into a rigorous concept, we shall now try to make it a bit more precise. The basic idea is to construct a limiting procedure, which allows us to compute the path integral as a limit of ordinary multiple inte-

182 grals (in much the same way as an ordinary integral can be obtained as the limit of Riemann sums).

There are several different ways in which this procedure can be deduced. Let us first concentrate on the same reasoning as we used when we derived the Schr5dinger equation (section 2.9). To compute

we slice up the time interval from

ta

to

tb

in

equal pieces:

with By repeated use of the group property (2.25) we then get


x(\)=~

J exp{~S[x(t)l} D[x(t)l
x(ta)=Xa
(TO simplify the notation we have put xa = Xo and xb = X ). At N this point the formula is exact, but it presupposes the knowledge of the propagator we want to calculate! propagators. In the limit where N -+co

we

can however use that the propagators involved become infinitesimal We can therefore replace them by the approximate (2.53),
V~ exp[fi

expression, cf. K(y;t+lx;t) ~

r-rn-

i m (x-y) 2 {i x-+-v} 2 1 exp - fi V(~)

Here the first two terms comes from the free-particle propagator while the last term comes from the interaction amplitude. arrive at the following limiting procedure: We thus

. [Vz=Ik:"l rm N . expLr;: i E2 m = 11m


N -+CD

J f

1T

"1<.=1

(_)2 + 4< 4<-1 4< ~-1 - V(-2-)ldx1~_1

with

183

In fact this is the original formula given by Feynman. We can rederive the same formula from a slightly different point of view: Consider an arbitrary (continuous) path leading from (xa;t ) to a (xb;t b ). Using the time-slicing we can approximate this path by the piece-wise straight line connecting the intermediary pOints (xo;t )' o .. , (xN;t N). Rather than summing over all paths we now restrict ourselves to a summation over piece-wise linear paths. The action of a piece-wise linear path is given by

Piece-wise linear path

t-axis

Fourier
path

t-axis

Fig. 57a

Fig. 57b

V(x)dt

The

~entiated

action is therefore approximately given by {i N

. exp Cfi:

s [x (t)]} '" exp i

k=l

i -=---=-...!-

(xi-x _,)

-V(~

x.

x } 2 i 1 )

But since the piece-wise linear path is completely characterized by its vertex-points (x 1 , . ,x _,), we can now sum up the contributions N from the "N'th order" piece-wise linear paths simply by integrating over the vertex coordinates. In the limit where N + 00 the piecewise linear paths "fill out" the whole space of paths and we therefore propose the following limiting procedure:

184

x(1)=~

(5.32)

J exp{iS[xCt)l} D[x(t)l

x(ta)=xa

Comparing this with (5.30), we see that it is given by almost the same expression, except that we have "forgotten" the integration measure:

(/21T~fi) N
For various reasons, we will however prefer to neglect the integration measure. Rather than calculate a single path-integral, we will therefore calculate the ratio of two path-integrals (both involving a particle of mass m). Then the measure drops out and we end up with the formula

x(tb)=~

exp{k S[x(t)l} D[x(t)] xCta)=xa


(5.33)
xCtb)=~

!M.

J exp{k So[x(t)l} D[x(t)l


x(ta)=xa
def.

lim
N+ex>

Usually So is the free-particle action so that we are actually calculating a propagator relative to a free-particle propagator. When we calculate a path integral we need not necessarily sum over piece-wise straight lines. Other "complete" families of paths can be used as well. Suppose e.g. we want to calculate a propagator of the form

185
K(D;TIO;O)

By slicing up the time-interval we again break up an arbitrary path, x = x(t ); 1 1 ... ;x _ 1 X(t _ ). This time we will approximate the given path N N 1 x(t) by a Fourier-path, i.e. a path of the form x(t), in segments connecting the intermediary pOints

N-1
x(t) =
If we choose the coefficients (5.34)

a
k

k=1 a

kt k Sinn T
so that

N-1

k=1

Sin [ n.]!] N

j=1, ,N-1

we evidently obtain that the Fourier path passes through the same intermediary points as the given path, cf. figure 57b. Again the approximate path is completely parametrized by the vertex coordinates nates. between (x 1 ' ,x _ ) and we can therefore sum up the N 1 contributions from such paths by integrating over the vertex coordiIn practice it is however more convenient to integrate over and (a1, . ,a _ ). Since the relationship (5.34) N 1 (a , .. ,a _ ) is one-to-one and smooth 1 N 1

the Fourier components (x , .. ,x _ ) 1 N 1 we evidently have:

J. .. I
is independent of
x(T)=O

J...

But the transformation (5.34) is in fact linear, so that the Jacohiant (a , ... a _ )! We can therefore forget about it N 1 1 and we finally arrive at the formula:

Jexp{~[x(t)l}D[x(t)l
Jexp{~o[x(t)l}D[X(t)]

x(O)=O

(5.35)

x(T)-O

x(O)=O

To see how it works, we shall now return to our favorite example:

TEST CASE: THE HARMONIC OSCILLATOR


According to exercise 2.4.1, the path-integral can be expanded around the classical path, whereby we get

186

x(T)=O

K(X ;Tlx ;0) = exp{ks[x ]}Jexp{if[1(t)2 -1 ix2]dt} D[x(t)] Z 1 C1 x(O)=O 0

According to (5.35) this can be further rearranged as


. K(x Z;Tlx 1;0) = exp{w>[xclnKo(O;TIO;O) 11m .
1 .

N + '"

.- exp{~[ l: ,\Sm-r-l }da1 ~1 J J ----.-.-...:""N'"'--1-------J.- J . kITt } da1-~-1 exp {1. KSO[k~,aksm-Tl


~1

N-1

. kTIt

where

Ko

is the free-particle propagator, which is given by Ko(O;TIO;O)

=.j2rr~'tJ.T

The remaining problem is therefore the computation of the multiple integral. Using the orthogonality relations for the sine-function it is easy to calculate the action which reduces to S{ N-1
k=1

a k Sin

k t}

mT N-1
k=1

2 k 2 TI2 2 a k (-:2 - w )
T

But then it is trivial to perform the integration over Fourier components, since the exponent is not only quadratic but even diagonal in (a , , a _ ) !
1 N 1

J J

N-l . kTIt . exp{fi S[ l: akSln~l}da ... da N_ 1 k=1

wf4TIin)N-1 N-1 k 2 rr2 2-l n (-2 -w ) mT k=1 T

To calculate the same multiple integral for the free particle action we simply put w=O i~ the above formula. Thus we finally get:

But using Euler's famous product formula for the sine-function, (5.36) this precisely reduces to sin(TIx)= rrx[
k=1

(1

2 x )]

187

j 27Tifi

m SinwT

exp{~
11

s[xcll}

in accordance with (5.17)! What about the phase ambiguity? This can also be resolved if we are a little more careful! Consider the analytic extension of the Gaussian integral. If we carefully separate the phase from the modulus we get 7T i 4 when A > a e +'" 2 iAx dx Tn . 1T Je -~'4 e when A < a

=/ =
i

jU 1m

Now, going back to the action of the Fourier path, (5.37) exp lii S 1 N-1 = exp {imT 4n I
A

2 (k 2 7T an --:2 - w ) k=1 T

2 2}
k < wT/1T, and "negaThe

we observe that the analogue of

is negative when

positive when k > WT/1T. Consequently there are Ent[wT/1Tl ti ve terms" (where Ent [x 1 denotes the entire part of x). multiple integral is therefore actually given by

For the free-particle propagator this ambiguity does not occur. corrected formula for the propagator thus comes out as follows:

The

(5.38)

But that is precisely the Feynman-Soriau formula!


Now that we have seen how the limiting procedure based upon Fourier paths work, let us return for a moment to the problem of the integration measure. If you want to calculate a path-integral directly, i.e. you do not calculate a ratio any longer, you must necessarily incorporate an integration measure in the limiting procedure. Feynman suggested that one could use the same integration measure as for the piece-wise linear path. This assumption leads to the formula:

188

x(T)=O
(5.39)

J x(O)=O

exp{i}[~(dX)2_ TIO 2 dt
lim
(V21iThi)
a>

V(x)]dt} D[x(t)] J . N-1 k t ... exp{i' S[ L akSirr.f-]}dx ... dx N -1 k=1

r;;- N J

N ...

lim
N ...
a>

But here the right hand side diverges as you can easily see, when you try to calculate it in the case of a free particle, cf. exercise 5.3.1 below:

Exercise 5.3.1 Introduction: Consider the matrix A involved in the linear transformation (5.34), i.e. 'k Sin [,T I ] j,k 1, ... ,N-l Ajk
N

a) Show that AXT

N-1

=2

and

Det LA.]=

[~]-Z

b) Show that the series involved in the limiting procedure (5.39) is given by lim
N-""

[N]

r (N)

TI

N-1~
21TihT

and use Stirlings' formula for the r-function to verify that it diverges. In fact, by considering the case of the free particle propagator, it is not difficult to construct the correct integration measure, which (when you integrate over the Fourier components) turns out to be given by:

XjT) =0 i exp{[ S[x(t)]}


x(O)=O

D[x(t)]

Nl!~ (vrz)

1T N-1

r(N)[V~]

rm

NJ

_jexp{~[k:1akSln.y- d

. N-1

k1Tt]} N-1

(As a consistency check you can u~ this formula to rederive the propagator of the harmonic oscillator). So now you 'see why we prefer to neglect the integration measure: Every time we introduce a new limiting procedure, i.e . a new denumerable complete set of paths, we would have to introduce a new integration measure!

5,4

ILLUSTRATIVE EXAMPLE: THE TIME-DEPENDENT OSCILLATOR

As another very important example of an exact calculation of a propagator we now look at the quadratic Lagrangian: L dx 2 = tm(at) - tmW(t)x 2

189

Since it is quadratic we can as usual expand around a classical solution, so that we need only bother about the calculation of the following path-integral
x(tb)~O

x(t }~O
a

exp{211 [(CIt)

t imfb dx 2 ta
Performing

Let us first rewrite the action in a more suitable form. a partial integration we get tb
S [x (t)]

=- ~ J [
ta
boundar~

Here we have neglected the tions x(t a )

terms due to the boundary condi-

= x(t b ) =

O.

Inserting this we see immediately that

the above path-integral is actually an infinite dimensional generalization of the usual Gaussian integral since it now takes the form:

x(tb)=O

im b dZ exp{- 2fJ. fx (t) [dt Z

W(t)]x(t)dt}D[X(t)]

x(ta)=O

ta

To'compute it we ought therefore to diagonalize the Hermitian operator


-

2
2
+ WIt)

dt

At the moment we shall however proceed a little different. the free particle action! differential equation
(5.40)

By a

beautiful transformation of variables we can change the action to Let f be a solution the second order

d2 {-2 + W(t)} fIt) dt


f

i.e.

belongs to the kernel of the above differential operator. The only thing we will assume is

The solution can be chosen almost completely arbitrarily within the two-dimensional solution space. that f does not vanish at the initial endpoint:
f (t a ) " O.

(Notice that

fIt)

is not an admissible path, since it breaks the Using f we then construct the following

boundary conditions!).

linear transformation, where x(t) is replaced by tha path y(t):

190
t

(5.41 )

x(t)

f (t)

J
ta

~ f (s)

ds

Here we assume that the transformed function y(t)also satisfies the boundary condition y(t ) = 0 a Differentiating (5.41) we obtain
t

x' (t) = f' (t)

J Yf'(~sl

ds + y' (t)

"""f'(t) x(t) + y' (t)

f' (t)

ta so that the inverse transformation is given by


t

(5.42)

y(t) = x(t) -

fTs)

f' (s) x (s) ds

ta We can now show that the above transformation has the desired effect. Using that
t

x"

(t)

f"

(t)

J Lill f(s)

+ ft (t)f (t) fIt

+ y" (t)

we obtain:

{;i$

+ W(t)}x(t) ={f"(t) + W(t)f(t)}

I[~'(~S)]dS

f'n~r(t)

+y"(t)

ta But here the first term vanishes on account of (5.40). the action reduces to: S[x(t)] = Consequently
t

-1 J

~
[F(t)f' (t)y' (t) + F(t)f(t)y"(t)] dt with F(t)

= J L...!& f(s) ds

ta Performing a partial
integrat~on

on the second term this can be

further rearranged as:

s[x(t)

- 1 [X(t)Y'(t) Jt
x(t).

tb

But here the boundary terms vanish due to the boundary conditions satisfied by Thus we precisely end up with the free-particle y(t)!

action in terms of the transformed path

191

There is only one complication associated with the above transformation, and that is the boundary condition associated with the final end-point path y(t): tb tb' The boundary conditions satisfied by x(t) are transformed into the following conditions on the transformed

J
ta handle directly. the identity, Ii (x(t )) = -,}; b cf. end-point: x(1)=O

~dS

The second boundary condition is non-local and therefore not easy to We shall therefore introduce another trick! Using

J exp

{-ia x(t ) }da

(5.5), we can formally introduce an integration over the final

J exp{iS[X(t)]}D[X(t)]
x(ta)=O

2~

x( t ) arbitrary b

Jf~eXP[-iaX(~)]eXP{kS[x(t)]}da D[x(t)]
a

x(ta)~O

This is because the integration over

now produces a Ii-function (Notice that if we x ). N Changing

which picks up the correct boundary condition! must now also integrate over the final end-point variables we then get

attempt to calculate the path integral by a limiting procedure we

Here the infinite dimensional generalization of the Jacobi-determinant is independent of linear. y(t) because the transformation (5.41) is "Completing the square" The remaining integral is Gaussian.
~

the whole formula therefore reduces to

OX = #ed liy ]

J
-'"

exp{ -

2iif

ill Z 2

y(tb ) arbitrary b dt im bdy 2 f (\)f -z-}da exp{crifCat) dt}D[y(t)] t f (t) t a yet )=0 a a t

with yet)

yet) -

~f(~)ff~)
ta

192 At this point we get a pleasant surprise: the a-integration! In fact we get: We can actually carry out

Furthermore the remaining path-integral is with-

in our reach, since it only involves the free particle propagator.

y(1)

arbitraIb

Jexp{~ {(~~)2dt}D[y(t)]
y(ta)=O a

== f
-0>

K o (x;1b IO ;ta )dx


2

==

lz C 1b
7Tih

-t )

LeXP[~

0>.

(1, ~tidx ==

This should hardly come as a surprise since, by construction, the above path-integral represents the probability amplitude for finding the particle anywhere at the time tb. The total path integral thus collapses into the simple expression:

(5.43)

- W(t)x 2]dt}D[X(t)]

It remains to calculate the Jacobiant!


using a very naive approach. dure!). The paths points

We shall calculate it

(The following argument is only includ-

ed for illustrative purposes, it is certainly not a rigorous proceAs in the approximation procedure for path-integrals we x(t) and y(t) are then replaced by the multidimensional discretize the linear transformation by introducing a time-slicing.

The linear transformation (5.42) can then be approximated by T

- N

k~ 1 ~ ---'2"---:":"--:""

f' (t k )

(x k " x k - 1 )

(This is actually the delicate pOint, since the discrete approximation of the integral is by no means unique, and the Jacobi-determinant turns out to be very sensitive to the choice of the approximation procedure). Okay, so the Jacobi matrix has now been replaced The determinant thus comes exclusively by a lower diagonal matrix. from the diagonal, i.e.

Taking the limit

-T

we then find:
N

lim
N +
a>

exp[log n (1 k= 1

lim
N +
a>

exp[ E log(1 - ! ( t )
k= 1

'(t k ) T
k

N) 1

tb

f'J11 exp[-! T(t)dtl

f ta

Consequently

Inserting this into

(5.43)

our formula for the path-integral fi-

nally boils down to the following remarkable simple result:

The path-integraZ corresponding to the quadratic action


(5.44)
t t m b dx 2 2 m b d2 S[x(t)l == 2 f[(dt) - W(t)x ldt == - 2 fx(t) [dt2 ta ta
+

W(t)lx(t)dt

is given by
x(~)=O

(5.45)

f
where

exp{~[x(t)l}D[x(t)l

x(ta)=O
f(t) is an almost azobitrary soZution to the differentiaZ equation

{~2

W(t)}(t)

0
~

the onZy constraint being that

f(t a )

O.

Notice that the above formula in fact includes the free-particle propagator (with W(t)=O) and the harmonic oscillator (with W(t)~2). You can easily check that (5.45) reproduces our previous findings f(t)=Cosw(t-t ). a in these two cases by putting f(t):1, respectively

194

5.5

PATH-INTEGRALS AND DETERMINANTS


Now that we have the formula for the propagator corresponding to

a quadratic Lagrangian at our disposal, we will look at it from a somewhat different point of view. tb Sex (t) Since the action is quadratic,

;:J

x(t)

{~+

W(t)} x(t)dt

dt For this purpose we consider the Hermitian

we can "diagonalize" it. operator

- -:-2 - WIt)
dt which acts upon the space of paths, x(t ) = x(t ) = O. b a normalized eigenfunctions n(t): ary conditions: x(t), all satisfying the boundIt possesses a complete set of

A given path,

x (t), N

can now be approximated by a linear combination tb ann (t) with an

xN(t)

n=1

ta

n(t)x(t)dt

Notice that the corresponding action of the approximative path reduces to

The summation over approximative path's can therefore immediately be carried out since it is just a product of ordinary Gaussian integrals:

I. f exp{~s
In the limit where by

[XN (t)

1 }d3., ... da N
the approximative paths fill out the whole

space of paths and the path-integral is therefore essentially given

By analogy with the finite-dimensional case we define the determinant


of the Hermitian operator eigenvalues.

_a t2

- WIt)

to be the infinite product of

Of course the determinant will in general be highly

195

divergent but we can "regularize" it in the usual way by calculating the ratio of two determinants. From the above calculation we then learn the following important lesson:

The path integral corresponding to a quadratic Lagrangian is essentially given by the determinant of the associated differential Operator, i.eX(s,)=o ~
(5.46)

exp{~fX(t)[-:tr2ta

W(t)]x(t)dt}D[x(t)] =

MDet[-~2- W(t)J}-~

x(t )=0 a

hlhere the right hand side should actually be interpreted as a limiting procedure (i.e. it is a short hand version of the follOhling expression:
(5.47)
N _1
2

lim
N ->- co

L',(N)[ IT;')
n=1

lim L',(N)f.N ->- co

fexp[~ l: Ana~l(vj~~/dNa
N n"'1
11

Notice that we have included a factor

(12 ~ft f in

the integration

over the generalized Fourier components, i.e. the proper integration

measure for evaluating the determinant is given by


(5.48)

k=1

[/2:Hi da k ]
L',(N)

Apart from that we still need a further integration measure

since the determinant is only proportional to the path-integral. Coleman [1977] .

The

above characterization of the path-integral is the one used by e.g. Since we already know how to compute the path integral we can now extract a relation for the determinant. is given by To avoid divergence problems According to (5.45) it we calculate the ratio of two determinants.

Det[-a~- W(t)]

(5.49)

Det[-a~- V(t)]

In this formula it is presupposed that at tao singular limit, where fW and fV

fW

and

fV

does not vanish tal Let us de-

It can however be simplified considerably by going to the

does vanish at

196 note by fO

{- a~ -

the unique solution to the differential equation,

W(t)}f(t)

0,

which satisfies the boundary conditions

2.. fO (t ) dt w a
Similarly we denote by conditions:

= 1

f~

the solution which satisfies the boundary

o
In the above identity (5.48), we can then put fO It follows that and Finally the limit of the integral tb

W+ E W

f1

f = fO + d 1

f
bourhood" of verges like

dt ta [f (t)]2

diverges, due to the vanishing of

fO at tao But since almost all W the contribution to the integral comes from an "infinitesimal neighta (in the limit where s+O), it follows that i t di-

Consequently
lim
{

a Thus the identity

tr, f -w
t

dt

f2 (t) }

I{

ftr, e
t

dt
(t)}

(5.49)

,V a collapses into the extremely simple deter-

minantal relation:
Det[ -a~-W(t) ]

(5.50)

DetH~-V(t) ]

If, eg., we put V(t)=Q

(and consequently

f~(t)=t-ta) it follows that

Det[-a~-w(t)l

DetH~l

197

This can be used to calculate a propagator (corresponding to a quadratic Lagrangian) relative to the free-particle prop~gator Ko: . (5.51)

K(~;tblxa;ta)

==

K(O;tbIO;t a ) exp{~[xcll}

Det{-d~-W(t)}]-2 KO(O;tbIO;t ) Det{-d~} a


m

-I

2TIiflf~ ( tb )

exp{~s[xCl]}

E.g. in the case of the harmonic oscillator we put W(t)=w 2 and consequently 1 fW(t); ;;Finw( t-t ). In this way we recover the by now well-known result (5.11). a
/

Rema:r>k: Incidenti.lly the

investigation of determinants corresponding E.g. the

to linear operators has a long tradition in mathematics. twenties*).

basic determinantal relation (5.50) has been known at least since the This is very fortituous, since in our derivation some The passage from quotients of is actually independent dirt has been swept under the rug. gration measure n(N)

path-integrals to quotients of determinants only works if the inteintroduced in (5.47) WIt). of the potential function Since the result we have deduced,

i.e. the determinantal relation (5.50) is known to be correct, we have thus now justified this assumption. In fact the determinantal relation (5.50) beautiful reasoning which emphasizes its basic who are acquainted with more advanced analysis argument: Consider the expressions 2 Det(- a g(A)

can be proven by a very general and position. For the benefit of those we include the main line of the

WIt) - A) t 2 Det(- a - V(t) - A) t

and
h (A)

It

is important for the following that A is treated as a complex variable. can be proven quite generally that a differential operator of the form,

It

- a2 t

- WIt)

has a discrete spectrum of real eigenvalues 1.. 1 ,1.. 2 ' , An'

*) J.H. Van Vleck,

Proc.Nat.Akad.Sci.

li,

178 (1928).

198

which are bounded below and which tend to infinity as n~. Each of these eigenvalues has multiplicity one, i.e. the associate eigenspace is one-dimensional. When A coincides with one of these eigenvalues, say A = An' it follows that the "shifted" operator 2

{-at - wit) - An}

has the eigenvalue zero, so that its determinant vanishes. As a consequence the function g(A) becomes a meromorphic function with simple zerOs at the eigenvalues A~ and simple poles at the eigenvalues A~. Consider next the function to the differential equation

f~+A(t).

By definition it is the unique solution


-

{- at2

w(t)

A}f (t) = 0

which satisfies the boundary conditions

flta)
It fo llows that

=0
WIt)} if and only if

is an eigenvalue of the operator


f IW+A) It b ) = 0

and when that happens f(W+A)(t) is in fact the associated eigenfunction (except for a normalization factor). As a consequence the function heAl V becomes a meromorphic function with simple zeros at AW and simple poles at A k The quotient function n' g (A) /h IA) is thus an analytic function without zeroes and poles! (Thus it has a behaviour similar to e.g. exp[A]). Furthermore, from general properties of determinants (respectively solutions to differential equations) it can be shown that g(A) (respectively heAl) tends to 1 as A tends to infinity except along the real axis. The same then holds for the quotient, which as a consequence must be a bounded analytic function. But then a famous criterium of Liouville guarantees that it is constant, i.e.

Specializing to

A=O

we precisely recover the determinantal relation (5.50).

**=

5,6

THE BOHR-SOMMERFELD QUANTIZATION RULE

we shall now encounter the important problem of computing quantum corrections to classical quantities. Especially we shall consider quantum corrections to the classical energy. In the case of a onedimensional particle we know that it can be found by studying the trace of the propagator. Inserting the path-integral expression for the propagator, this trace can be reexpressed as

199

(5.52)

G(T)

j J eX~t~S[X(t)]}
-00

x(T)=x x(O)=x

D[X(t)]dx o

== J exp{~S[X(t)]} D[X(t)]
x(O)=x(T)
where we consequently sum over all path's which return to the same pOint. This expression is known as the path-aum-traae integral. The There are now in principle two different approximation procedures available for the evaluation of this path-cum-trace integral. first method is the hleak-aoupling approximation. the potential has the general shape indicated on fig. 58. the potential well, we can approximate the potential by
V(x)
RJ

Let us assume that To calcu-

late the low-lying energy states we notice that near the bottom of
2

W"

(0)x

iV"(O)x 2

E3

E"

E2
El

Eo

X-axis Fig. 58

Thus the problem is essentially reduced to the calculation of the path-cum-trace integral for the harmonic oscillator! already solved, cf. consequently given by
(5.53)

This we have

(5.19-20) and the low-lying energy states are


2

with

v"(O)

mw

The second method is the so-called WKB-approximation which can be used to find the high-lying energy states. non-perturbative method. It is thus essentially a It leads to the so-called Bohr-Sommerfeld

quantization rule, which in its original form states that

200

(5.54) where q

f pdq

= n

h p the conjugate momentum and

is the position coordinate,

we integrate along a periodic orbit.

Notice that the left hand side

is the area bounded by the closed orbit, cf. fig. 59, so that the quantization rule states that the area enclosed in phase-space has
to be an integral multiple of

h.

The original WKB method which was But before we jump out

based upon the construction of approximative solutions to the schrodinger equation, will not be discussed here. in the path integral version of the WKB method it will be useful to take a closer look on classical mechanics and the original derivation of the quantization rule (5.54).

p-axis Phase-space
diagram.

--+--"'------""'t'---f-"'--..... q-axis

Fig. 59

ILLUSTRATIVE EXAMPLE:

THE BOHR-SOMMERFELD QUANTIZATION RULE

Let us first collect a few useful results concerning the action. Suppose (xa;t a ) and (xb;t b ) are given space-time points. Then

S(Xb;tb[xa;t a ) will denote the action along the classical path connecting the two space-time points. The
res,ult~ng

function of the initial and final (If

space-time point is known as Hamilton's principal funation.

there is more than one classical path connecting the space-time points (xa;t a ) and (xb;t b ) it will be a multivalued function). All the partial derivatives of Hamilton's principal function have direct physical significance: (5.55) Here Pa

as

Pa

as
ata
xb and E

at b

as

is the conjugate momentum at the initial point

the conjugate momentum at

xa ' P b is the energy (which is con-

served along the classical path).

201
Proof: A change, 6x b , in the Epatlal position of the final space-time point will cause a change, 6x(t), in the classical path. The corresponding change in the action is given by tb 6S =

J [~~

6x(t) + aL Qx(t) J dt ax

ta tb
t

J
a

CaL

-~~J
dt " aX

~
6x(t)dt + [aL " aX 6x(t) ]

But here the integrand vanishes due to the equations of motion and we end up with: 6S
It

= a~(tb)QX(tb)
ax follows that

aL(t )6x(t ) a a ax

aL ) 6X b = -;-(t b ax

Pb 6X b

6S 6X b

Pb Similarly we may consider a change,

6t , in the temporal position of the b final space-time pOint. Notice first that

The corresponding change in the action is also slightly more complicated,

[~ 6x(t)
ax ta

~ Qx(t)Jdt
ax

where the first term is caused by the change in the upper limit of the integration domain.

As before we can now make a partial inte-

gration leaving us with the formula 6S

aL L(t b )6t b + -;-(t )6x(t b ) b dX b ax b

L(t b )6t b -

~(tb)x(tb)6tb

aL
dX

aL' ) J 6tb = [ L(t ) - -;(tb)x(t But the energy is precisely given by

202 E =
p~

- L

aL
a~

X- L

so that we end up with the identity

6S =

Eot b

Le.

as
3t

Notice that the above considerations were in fact anticipated in the discussion of the Einstein-de Broglie rules, cf. the discussion in section 2.5. After these general considerations we now return to the discussion of a one-dimensional particle in a potential well like the one sketched on fig. mental period T: x = xT(t) Each periodic orbit will furthermore be characterized by its energy E, which thus becomes a function of the period From (5.55) we learn that (5.56 ) where S(T) dS = _ E xT(t). We can now S(T). T, i.e. E
59.

In the general case there will exist a one

parameter family of periodic orbits whi9h we can label by the funda-

= E(T).

OT

is the action of the periodic orbit

introduce the Legendre transformation of the action function It is defined by the relation: dS W(E) = S(T) - dTT=S(T) + E'T where T should be considered a function of E.

(Notice the simiIn

larity with the passage from the Lagrangian to the Hamiltonian). analogy with (5.56) we then get: (5.57) dW _ dS dT dT dT dT E dE + T + E CiE = T

dE - OT OE + T + E dE

= -

Finally we can get back the actfon function by a Legendre trans formation of
(5.58)

W(E), S(T)

= W(E)

- T'E

= W(E)

dW - dE E T. Notice that

where

should now be considered a function of

the Legendre transformation is in fact given by the phase integral in (5.54) W(E) S(T) + E'T

f [m(E)
o

T
- V(x)]dt +

f [m(~~}
o

2
+ V(x}]dt

i.e.

203

(5.59) since p

WIE) dx = m at W(E)

I o

2
m(dx) dt dt

for this simple type of theory. which we are going to quantizel

It is thus the

quantity

From (5.57) we now get, dE where w

1 = -T

dw = ~ dw 2~

with

2~ =T

is the cyclic frequence of the periodic orbit.

Following

Bohr we then assume that only a discrete subset of the periodic orbits are actually allowed and that the transition from one such periodic orbit to the next results in the emission of a quantum with energy ple.
2~

nw, where

is the frequency of the classical orbit.

The

last part of the assumption is Bohr's famous correspondence princiIt follows that W can only take a discrete set of values and W can be stated as +
2~ftc

that the difference between two neighbouring values is given by

= n.
c

Thus the quantization rule for W(E ) = n


2~(n+c)ft

2~n

where

is a constant which we cannot determine from the correYou can look upon this formula as the first W(En) where we expand in Thus the correct

spondence principle. powers of 1/n.

two terms in a perturbation expansion for

Bohr and Sommerfeld simply assumed it was zero, but

using the WKB-method it was actually found to be ,. quantization rule for the above case is given by (5.60)

As an example of its application we shall as usual consider the harmonic oscillator. features: This example turns out to have two remarkable T SIT) breaks down completely. As a 1) The assumption of a one-parameter family of periodic cannot be defined for a har-

solutions labelled by the period consequence the action function monic oscillator. W(E)

2) The quantization rule (5.60) is exact. A general periodic orbit

Since the action function cannot be defined, we shall define through the phase integral (5.59). x(t)= ACoswt + The associated energy is consequently is given by BSinwt

mW 2 (A2 + B2)

while the phase integral turns out to be

Thus
W(E) = 2rr E
w

and the Bohr-Sommerfeld quantization rule reduces to the well-known result (5.20)

Now that we understand the Bohr-Sommerfeld quantization rule we return to the path-cum-trace integral (5.52). This time we are going to expand around a classical solution x(t) = xcI (t) + nIt) The potential energy is expanded to second order: V[Xcl(t)] + V'[xcl(t)]n(t) + Pl"[x c1 (t)]n (t) Inserting this, and using the classical equation of motion, the action then decomposes as follows
Rl

V[x(t)]

S[X(t)]

Rl

S[Xcl(t)] +

b{I(~)2-!V"[XCl(t)]n2(t)}dt
T

+ 1bn(t){-~- ffiV"[Xcl(t)]}n(t)dt

d2

Consequently the path-cum-trace integral reduces to (5.61)


G(T)

Rl

7
-00

l:

exp{~[xcl(t)]}
cl

n(T)=O

J expHKn{-~- ~Vff[xc1(t)]}ndt}D[n(t)]dxo

n(O)=O

where xcl(O) = xcl(T) =xo' But the remaining path integral we know precisely how to handle. According to (5.45) it is given by

j __ m_-J""-Tdt
2rrifJ.f(O)(T)

with

2 (t) j

o
In fact we can relate f to the classical path: solves the Newtonian equations of motion:
d XcI m~

The classical path

- V'[x Cl (t)]

Differenting once mOre, we thus find


m

~
dt

3 d x

= -

V"[x lIt)]
C

205 Consequently we can put

Thereby the path-cum-trace integral reduces to (5.62 )


dx
0

IlxdP o
+00

T dt

where we sum over all classical paths satisfying the constraint xcl(O) = xcl(T) = xo. Despite its complicated structure it is essentially a one-dimensional integral of the type

exp{if(x)} g(x)dx

Such an integral can be calculated approximatively by using the


stationary phase approximation.

Thus we look for a point f' (x ) = O. o

Xo

where

the phase is stationary, i.e. around this point:

We can then expand


2

fIx) "'" f (x o ) + ,f" (x o ) (x-x o ) g(x) Rl g(x ) + g' (x o ) (x-x o ) o

In this approximation the integral then reduces to a Gaussian integral:


+co

(5.63)

exp{if(x)}g(x)dx

Rl

g(xo)eXp{if(Xo)}vI,,(:~)

Using this on the path-cum-trace integral (5.62) we see that we need only include contributions from the classical paths which produce a stationary phase, i.e.

o = x a ao

But according to (5.55) this gives

o
i.e. the momenta at odic solutions! od T. t=O and

P b - Pa
t=T are also identical. Thus the

stationary phase approximation selects for us only the purely periThey can be parametrized by their fundamental periT we should therefore only inCorresponding to a given

clude the contributions from the periodic orbits:

206
XT / 2 (t)

x T/ 3 (t) r .. rXT/n (t) I

Before we actually apply the stationary phase approximation we can therefore restrict ourselves to a summation over these periodic orbits:

N::>tice that the parametrization of the periodic orbit is only determined up t, i.e. we can choose the starting point quite arbitrarily on the closed path. p-axis Each of the x-values between the turning points x 1 and x 2 occurs furthermore twice on the path, cf. fig. 60. We are now ready to perform the x -integration. The x x 0 ~~____________-+____+-__-+~2~.. action of the periodic orbits does X-axis not depend upon the choice of the starting point, and neither does the expression to a translation in
T

Fig. 60

dt
(~) 2

Thus the only contribution to the integral comes from x2 +00 dx o 2 dt = ~ I~T(O) I x 1 Cit

Having taking care of the xo-integration we must then compute the integral over the reciprocal square of the velocity. Here we get
T

~2
(x T / n )

J o

dt

2n

(2[E(T/n) - V(xB-~) dx
m

The same integral can however be obtained by a different reasoning! Consider the phase integral W(E), which is given by x WeE)

= 2n

J /2[E
x1

- vex) 1 dx

for the periodic orbit in question. respect to E we obtain:

Differentiating twice with

207

dT dE consequently

-3/2 (2[E - V(x)]) dx

Jo
reduces to:
(5.64)

dt (x T / n )

_ m3/2 dT dE

Putting all this together, the path-cum-trace integral thus finally

G(T)

Rl

1.

_1_

m I2nifl n=1

exp{i nS - } ft

-[T] T~dEI n n dT

T
T~

This was the hard part of the calculation! energy levels.

Now we can extract the The transformed

As usual this is done by gOing to the transformed


(5.9)

path-cum-trace integral and looking for the poles. path-cum-trace integral is given by, cf. G(E)
i

G(T) exp{fiET} dE

(Notice that latter as obtain

now both represent an integration variable and the To avoid confusion we shall write the G(T) we Inserting the above expression for

energy of the periodic orbit. Ecl I).

G(E)

Rl

iiiKv'21Til'i

n~,Jo
~

""

Here it will be preferable making an exchange of variables, i.e.


~

= Tin,

is the fundamental period of the periodic orbit in question:

G(E)

Rl

. 1 ~v'21Ti~

n=1

""""f ~ dT E exp{1f(ET+S[T])}T/I~Iv'n
o
as
T

The T-integration is then performed by use of the stationary phase approximation. A stationary phase requires
+ S[T])

a o = 1T(EI

=E

- ~

=E

- Ecl

Thus for a given value of

we get the main contribution from the E!


give~:

periodic orbit which precisely has the energy


(5.63) the stationary phase approximation now

According to

208

G(E)

Rl

2- _,_

~ jZTTift

mI'lv'ZTTifl n='

fd2S cr:rr
W(E)

_,_

Using that dE

dT

and

S (T)

+ ET (E)

the formula finally reduces to

(5.65)

G(E)

Rl

T(E)

n='

exp{%nW(E)}

i mt

(E)

This clearly has poles when (5.66) and thus we have precisely recovered the old Bohr-Sommerfeld quantization condition (without the half-integral correction term). simply: The reason we missed the half-integral correction term is, however, very It is due to the fact that we have not taken into account precisely as for the The exThe additional the phase ambiguity of the path integral.

harmonic oscillator we should pick up phase corrections. pression (5.62) is completely analogous to (5.38). phases then come fran the singularities in the integrand

-,

U f2 ~:) ] o
These singularities correspond to the zero's of to the turning pOints turning
~ts

f(t)

x(t), i.e.

In the present calculation the x and 1 are thus analogous to the caustics. Each time we pass e

a turning pOint we therefore phase-factor

iTT/2

e~pect

, cf. xT/n

(5.29).

that we pick up an additional 2n turning

Since there are

points in the orbit

we consequently gain an additional phase

e- iTTn = (_1)n
which should be included in the formula for the propagator. The transformed path-cum-trace integral (5.65) is thereby changed into

(5.67)

G(E)

Rl

T(E)

n=1

(-1)nexp {%nW(E)}

209

This causes a shift in the poles, whichare now given by


(5.68)

and that is precisely the correct quantization rule according to the original WKB-calculation. Notice too that near a pole we get the expansion

Using the approximation 1 + exp4i W(E)}


Rl

1 + exp{k W(E

ll

}exp{~ T(E) (E-En )} = 1 - {1

~ T(E) (E-En )}

it follows that

G(E)

behaves like G(E)


Rl

1 E--=-E
n

when

Rl

En'

This should be compared with (5.10) and it clearly

shows that the approximative path-cum-trace integral has the correct asymptotic behaviour close to a pole. we have been working hard to derive a result which, as we have seen, can in fact almost be deduced directly from the correspondance principle! When we are going to use the path-integral technique in quantizing field theories we will actually have to work still harder Before we enter into these dreadful technicalities I want to give a few examples of almost trivial applications of the preceding machinery. The first example is concerned with the free particle.
L

To find

the energy levels we enclose the free particle in a box of length and use periodic boundary conditions. Thus we have effectively replaced the line by a closed curve of length effect that the energy spectrum is discretized. uous free-particle spectrum back by letting and around the closed circumference: given by xT(t) =
L

L.

This has the We get the contin-

go to infinity. x is

The free particle can now execute periodic motions by going around The periodic orbit T

Lt T

It has the following energy and action: m dx 2 _ mL 2 E(T) = Z(dE) SIT)

-;;Z

ET

210 Consequently W(E) ; SIT) + ET ; LI2mE

Since there are no turning pOints in this problem the quantization , rule is given by (5.66), i.e. Ll2mEn i.e.
(5.69 )

21Tnfl

These are the correct energy levels in the discrete version. this we notice that the free particle Hamiltonian is given by
H=----

To see

fi2

d2

2m dx2

Since the SchI1.5dinger wave function must be periodic, the eigenfunctions are given by 21Tnx] o/n(x) = exp [ 1. -Land the corresponding eigenvalues precisely reproduce the above result
(5.69)

The second example is concerned with a field theory. of periodic solutions: are given by (cf. (5.70)
~(x,t)

In the

discussion of the sine-Gordon model we found an interesting family The bions (or breathers). In a slightly w, they changed notation, where we emphasize the cyclic frequency (4.50): n Sinwt ] with Arctan [ Cosh(nw~

4 =~

= ---w--

/].l_w 2

and

O<w<

The corresponding energy is given by (cf. 4.51):


(5.71) E = 2M!j1 _

w~
\j

where

M is the mass of a single soliton.

As we have seen it re-

presents a classical bound state of a soliton and an anti-soliton. We therefore expect that it will generate quantum bound states which should be recognizable in the energy spectrum. be labelled by the period the form Since (5.70) is a one-parameter family of periodic solutions (which can easily T), we can in fact use the naive BohrFor a classical field theory it takes Sommerfeld quantization rule.

211

T
(5.72)

+00

WeE)

21Tnh

with

WeE)

JJ
o
-00

1T(X,t)

dt dx

(where

1T(X,t) WeE)

is the conjugate momentum density).

Rather than

calculate

directly from the phase integral we shall use the dW dE =T =


W

basic relation (5.57 ) From (5.71) we see that (5.73 ) dw dE


21TW

-1
E

can be replaced by
21T)1-

and we get

(1

--.,-)

E2

-'

4r.1"

This can trivially be integrated WeE) But using that state, where then gives: (5.74)
E

= 41T~-1M ArCsin(~)
W=)1,

+ C

E=O, i.e.

corresponds to the classical ground

W obviously must vanish too, we see that

C=O.

trivial application of the naive Bohr-Sommerfeld quantization rule

o<

< :~

Notice the upper limit of crete set of energy states. When n and (5.73) is given by T(n) But when T n

n, i.e. there is only a finite disThis comes about in the following way: T, which by (5.57)

grows so does the fundamental period dW dE 21T

the corresponding soliton-antisoliton pair becomes above the treshold, the classical state thus

unbound and for soliton.

becomes unstable and breaks up into a separate soliton and antiIn the above derivation we have been using the naive Bohr-Sommerfeld quantization rule. As in the case of the single particle we In the can improve this uSing the path-integral version of WKB.

multidimensional case the WKB-formula, however, becomes more complicated than in the onedimensional case, where we just had to include a "half-integral" correction term. For the special case of the sine-Gordon model it can nevertheless be shown that the energy spectrum of the bion-states is still given by (5.74), with a suitable redefinition of the parameters M and
)1.

212

5,7

INSTANTONS AND EUCLIDEAN FIELD THEORY


At this point we introduce a very important trick, which will

enable us to extract non-trivial information about the quantum mechanical ground state of a system. space with its indefinite metric ds 2 Let us consider the Minkowski 2

=-

dt

+ dx

+ dy2 + dz

Formally we can turn it into a positive definite metric by using Weyl's unitary trick, i.e. by replacing the Minkowski time the imaginary time
1

with

= it:

Obviously we can think of this as a kind of analytic continuation if we regard the 4-dimensional Minkowski space as being a 4-dimensional real subspace of an 8-dimensional complex space. The above trick has been widely used in special relativity, where it allows you to avoid the distinction between upper and lower indices. But it is also widely used in quantum field theory, where it In quantum field theory, it has further been given a The analytic continuation to imaginary time is by has had profound implications for our understanding of a variety of phenomena. special name:

quantum field theoreticians known as performing a Wick rotation. To illustrate the applications of the Wick rotation we consider for simplicity ordinary quantum mechanics in one space dimension.

As we have seen, the most important ingredient in ordinary quantum


mechanics is the Feynman propagator this to be an analytic function of K(Xb:tblxalta)' tb and t If we assume , we can perform a

a wick rotation whereby we produce the so-called Euclidean propagator: KE(XbITbIXa;Ta)' E.g. the free particle propagator is Wick rotated m (X b - Xa) } exp { - 2fi T - T b a Similarly the propagator for the harmonic oscillator
(5.17
2

into the following Euclidean propagator:


(5.75)

cf.

(2.28).

is Wick rotated into (cf.

where the trigonometric functions have been replaced by hyperbolic

f,"..,t-; ons.

Notice that the caustics have disappeared and so have

213

the phase ambiguities - the Euclicean propagator is only singular when Tb = Ta! In the Minkowski space the Feynman propagator can be reexpressed as a path integral. The same is true in the Euclidean case. To Wick rotate the path integral we notice that the substitution, t ~ - iT, generates the following change in the action Tb

sex (t)]

- V (x)} dt

J
1a

V(x)}dT

consequently the Wick-rotated path integral is given by

x~Tb)=~

J
XCTa)=xa

Tb exp[iJjc*)2 + VCX)}dT] D[X(T)] Ta

where we sum over paths in the Euclidean space. Let us introduce the abbreviation Tb (5.77) SE[X(T)] <kf

J{jC*) 2
Ta

+ Vex) }dT

The quantity SE[X(T)] is known as the Euclidean action. Notice that the Euclidean action is positive definite. The Euclidean path-integral is now given by
X(Tb)=~

exp{-iSE[X(T)]} D[X(T)]

X(Ta)=Xa This is one of the main achievements of the Wick rotation. Rather than being an oscillatory integral, which can never be absolutely convergent, we have now turned the path integral into a direct generalization of a nice decent Gaussian integral where the exponent is negative definite and furthermore it is at least quadratic. This puts the path integral on a much more sound footing, where we avoid completely the divergencies and phase ambiguities which plague the Minkowski version. The Euclidean path integral can be defined through limiting procedures like (5.31) or (5.33). If they are quadratic, i.e. of the Gaussian type they can also be diagonalized, so that we can use relations like (5.45) (suitable modified). In the Euclidean case it is in fact even possible to define it in terms of a rigorous integration theory on an infinite dimensional measure space using the so-called Wiener measures. Returning to the Euclidean propagator we now get from (2.21) that it can be reexpressed in terms of an Euclidean path integral:

214
X(Tb)=~

(5.78)

exp{-k SE[x(T)l} D[x(T)l

X(Ta)=Xa The other characterizations of the Feynman propagator can also be carried over to the Euclidean version. Notice that the Wick rotation turns the Schrodinger equation into the Heat equation: (5.79) = 2m
t\.2 2 ~

a ax

$(X,T)

- V(X)$(X,T)

As a consequence the Eualidean propagator is the unique solution to

the Heat equation:

satisfying the boundary aondition


KE(xbiTa!XaiTa) = 6(x b - xa) Next we observe that the Hamiltonian operator is unaffected by a Wick rotation (since it does not contain the time). This evidently leads to the characterization: The Eualidean propagator has the following deaomposition on a aomplete set of eigenfunations for the Hamiltonian operator: (5.80)
cf.

(2.70). Finally it can be reexpressed as a


1 1

matri~
~

element of the Eualid-

ean time evolution operator


(5.81)

exp [ - fi HTl , i.e.


~

KE(XbiT!Xa;O) = <xblexp[ - fi HTl !xa>

As we have seen, the Euclidean formalism offers a chance of constructing in quantum mechanics a rigorous path integral which is not plagued by singularities and phase ambiguities. To some extent this is also true in quantum field theory. Consider the Lagrangian density

(For technical reasons U[~] should be at most a fourth order polynomial in <p. This guarantees that the theory is renormalizable,

215

i.e. that one can get rid of various divergent expressions in quantum field theory using a so-called renormalization procedure). In the Euclidean formalism this Lagrangian density leads to the following Euclidean action
(5.82)

SE[~l

J{!(~)

2
+

!(:~)

2
+

U[~l}d~dT

J{!6~Va~~av~

U[~l}d~dT

Notice that apart from a change of sign, the Minkowski metric has been replaced by the Euclidean metric. As in quantum mechanics, the Euclidean action is positive definite and the Euclidean path integrals

are of the nice infinite dimensional Gaussian type with a negative definite quadratic exponent. As a consequence it has been possible to construct a rigorous integration theory for such path integrals in one and two space dimensions. In three space dimensions (corresponding to our actual world) there are, however, still unsolved technical problems. In quantum mechanics the Euclidean formalism with its associated Euclidean path integrals is a funny trick, which is not really necessary (after all, it is easy to solve the original Schr5dinger equation). In quantum field theory it is on the contrary the only way one knows to construct a rigorous theory: First one must construct the Euclidean version and thereafter one wick rotates back to the Minkowski space. we now turn to a very interesting aspect of the Euclidean field theory. In a Minko.wskian field theory we have seen that the path integral is dominated by classical solutions with finite energy, and this we have been using to calculate a semi-classical approximation of the path integral. In the Euclidean version we can do the same, i.e. we expect the Euclidean path integral to be dominated by classical solutions to the Euclidean equations of motion. To investigate Euclidean configurations a little closer we look at the space of smooth configurations with a finite Euclidean action, Le. In analogy with the Minkowski case a Eualidean vaauum is defined as a configuration with vanishing action. Since the Euclidean action is given by

216

(5.82)

SE[$] =

f{i(~~)

2
+

i(l!)

ar

U[$] }drd,

this implies that

11 = 11 = U[$] aT ar

i.e. a Euclidean vacuum (like the Minkowski vacuum) corresponds to a zero point of the potential energy density:

with Next we observe that finiteness of the Euclidean action implies that the field asymptotically approaches a Euclidean vacuum. We then look for solutions of the Euclidean equations of motion. Obviously a Euclidean vacuum is such a solution, but there may also be non-trivial solutions with the following two properties a) It has a finite Euclidean action; b) It is a local minimu m of the Euclidean action (i.e. it is stable) Since such a non-trivial solution is localized not only in space, but also in (Euclidean) time, 't Hooft has proposed to call it an instanton. The basic role played by the instantons is then that we expect it to dominate the Euclidean path integral and thereby help us with the calculation of the Euclidean propagator. At this pOint you have probably recognized a similarity with our discussion of solitons in chapter 4. This is no accident since there is an extremely close connection between instantons and solitons which we can formalize in the following way: Consider a Minkowskian field theory in D space dimensions. It is characterized by an action of the form

When studying solitons we look for a static configuration $(r,t) = $(r) with the following two properties: a) It has a finite static energy; b) It is a local minimum of the static energy (i.e. it is stable). Here the static energy is given by the expression:
2

E = f{!(li) static ar Next we consider a field theory in D spacetime dimensions, based upon the same potential energy density U[$]. The Euclidean version (5.83)

217

of this theory is based upon the Euclidean action

which is identical to (5.83)! As a consequence we therefore derive the following extremely important principle:
A static soliton in to an instanton in

space dimensions is completely equivalent

spacetime dimensions.

We shall apply this principle to solitons in one space dimension. They correspond to instantons in one spacetime dimension, i.e. in zero space dimensions! This example is therefore somewhat degener~. Now a field in zero space dimensions can only depend upon time, i.e. it is on theforrn $a(t). Consequently a field theory in zero space dimensions corresponds to a system with a finite number of degrees of freedom, i.e. to ordinary mechanics. If there is only one field component $(t), we can identify it with the position of a particle moving in a single space dimension, i.e. put x = $(t). Let us contretize this: A particle moving in a one-dimensional potential V(x) is characterized by the Lagrangian
L

= !m(~)

- V(x)

In the Euclidean version this corresponds to the Euclidean action


+oe

SE[$ h)] =

Om (~~)

+ V ($) ]dT

where we have denoted the Euclidean path by x = $(,). Since , is now a space like variable, we can finally denote it by x whereby the Euclidean action becomes
+ V[ $ (x))} dx

which is precisely the same as the expression for the static energy in a (1+1)-dimensional field theory. Notice too a very pleasant surprise: Although we have started out with a non-relativistic theory for a particle moving in a potential V(x), we see that the corresponding Euclidean formalism corresponds to the static version of a relativistic theory in (1+1)-spacetime dimension! We can now take over the results obtained in the sections 4.3 4.4 concerning the kinks in (1+1)-dimensional field theories.

218

First of all a kink in a (1+1)-dimensional field theory corresponds to an instanton in ordinary quantum mechanics. E.g., a particle moving in a double well
V(x)

(x

2 2 - a )

is associated an instanton given by


(5.84)

$(T) = a Tanh {a::f (T - 'oj} Similarly a particle moving in the periodic potential
V(x)

cf. (4.23).

)l4 f1 T l - Cos (IX)] 11 $

is associated an instanton given by


(5.85)

$(T) = ~Arctg

{ )lIT-To)}

cf.

(4.24).

Furthermore we have seen that a theory with a degenerate vacuum is characterized by a topological charge Q and that the static energy is bounded below by the topological charge due to the Bogomolny decomposition. In the Euclidean version this corresponds to a decomposition of the Euclidean action
(5.86)

+00

SE[$(-r)]

{~ +" PV(*) d, m

$+
+ [Q]

with

Q =

J $-

~)d$

where $(T) is a configuration interpolating between the classical and $+. This leads to the following simple picture: vacua L The space of all smooth configurations with finite Euclidean action breaks up into disconnected sectors characterized by different values of the topological charge Q. In each of these sectors the Euclidean action is bounded below by the topological charge

we expect the contribution from a given sector to the Euclidean path integral to be dominated by a configuration which saturates the lower bound, i.e. from a configuration satisfying the differential equation (5.87) d$ = + flV($) err - m

In the lowest non-trivial sectors this leads precisely to the oneinstantons or anti-instantons (like (5.84) and (5.85). In the

219

higher sectors wekocw that we cannot saturate the bound exactly, but we can still corne arbitrarily close by considering multiinstanton configurations consisting of widely separated instantons and anti-instantons, cf. the discussion in section 4.4. Due to the close relation between solitons and instantons, Polyakov has suggested to refer to instantons as pseudo-particles, a label which is also currently used in the literature.

5.8

INSTANTONS AND THE TUNNEL EFFECT

Having understood some of the formal properties of instantons we now turn to the question of their physical significance. We know that solitons correspond to a particle like excitation of the field. In a similar way we will now show that instantons are associated with the so-called tunnel effect in quantum mechanics. Let us first recall that instantons in ordinary mechanics are associated with a particle moving in a potential well which has a degenerate minimum (like the double well or a periodic potential). As a consequence the classical vacuum is degenerate and formally the instantons represent Euclidean solutions which interpolate between two different classical vacua. We now proceed to examine the quantum mechanical ground state of such a system. To be specific, let us concentrate for a moment upon the motion in a double well (cf. fig. 63 page 231). Classically there are two ground states represented by the configurations and
x_It) := -a

corresponding to a particle sitting in the bottom of either of the wells. Quantum mechanically we must solve the Schrodinger equation and thereby determine the energy spectrum to see which eigenfunction has the lowest eigenvalue:
_ fl2 d 2 ,,, + ;;;....x. V(x) ljI(x) '" E ljAx) 2m dx2

Based upon the classical analysis we can give the following qualitative description of the low lying energy states: Each of the two wells correspond to a harmonic oscillator potential. Thus the two lowest eigenstates will be related to the ground states of these two oscillators. Furthermore the potential V(x) is invariant under the inversion, x ~ -x, and consequently the eigenfunctions of the Hamiltonian can be chosen to be eigenfunctions of the parity operator pljI (x) = ljI (-x)

220

i.e. they can be chosen as either even or odd functions. In the case of the double well, the two lowest eigenstates are therefore of the form 1jJ+(x)
1 { $ (x-a) - $ $ (x) = --

12

1 (x+a~

where $0 is the ground state wave function for a harmonic oscillator (i.e. it is given by (5.25) where we put t=O). There are now two possibilities: Either the quantum mechanical vacuum is degenerate (like the classical vacuum), i.e. $+ and $_ has the same energy, or the Classical degeneracy has been lifted, i.e. $+ and have different energies. To find out which possibility is actually realized, we must compute the associated eigenvalues, i.e. solve the Schrodinger equation. We will not do this for a general potential, but to get a feeling for what is going on we look at a particular simple example:

w_

ILLUSTRATIVE EXAMPLE:

THE DOUBLE SQUARE WELL

Consider the double square well ~ shown on fig. 61. Because the Vex) potential is constant "piece by piece" it is trivial to solve the V Schrodinger equation. Due to the 0 symmetry of the wave functions we need furthermore only solve it in the regions I and II. Since the II I IV / III I potential at x=b is infinitely high, the wave function must :/ / / / / / / / / I vanish at x=b. At x=a we must -b -a a b Fig. 6-1 furthermore demand that the wave function and its derivative is continuous. This can be simplified by demanding that the logarithmic derivatives match,i.e.

~ ~ ~ ~ /;

~
~

~
~
~
~ :;
x

~ ~

(5.88)

Okay, now for the solutions. In region I, the solution to the Schrodinger equation is given by $I(x) = A

sin{vI~

(X-b)}

221

(where A is a normalization constant which will soon drop out of the calculation). In region II, the solution depends on whether the total wave function is even or odd. For an even wave function we get the solution $II(x)

Cosh

{!

2m (V0 - E)
2

fi.

x} x}

while in the odd case we get

In the even case the boundary condition (5.88) now reduces to (5.89a) Tan {v~ (b-a)

/2niE

= -

~ Coth {/2m(vfi.2 v~
o

E)

a}
a}

while in the odd case it reduces to (5.89b)

Tan{~~

(b-a)}

V~ Tanh 1 / L

;--p;o

f j2_m_(V....::o~_E_)
fl

In the limit where Vo + 00, i.e. where the potential barrier is infinitely high, the two conditions collapse into the single cond~ tion (5.90 )

Tan{vI~

(b-a)}

(Notice that both Tanh x and Coth x tend to 1 at infinity!). In this case the spectrum is therefore degenerate, i.e. both the even and odd wave functions have the same energy which is given by (5.91)
E -

fl2

2 2
11

- 2m (b-a) 2

n = 1,2,3,

(cf. the free particle spectrum (5.69. For finite Vo the spectra are however different in the two cases due to the difference between the hyperbolic tangens and cotangens functions. We can furthermore estimate the two lowest lying energies. If Vo is high, we expect the correct energy to be close to the "degenerate" value
2m (b-a)

i.e. we can safely put E = Eo + oE

222

Using the approximation

j
We then get

2mE

Tan{j:~E

(b-a)}

oE Tan { 2Eo

I~ ~

(b-a)

(Notice that = 0 and use the addition theorem for the tangens TI function!). As loEI Eo' we can furthermore forget about the tangens function itself so that the boundary conditions (5.89a) and (5.89b) finally reduce to:

Tan{~2m~0 (b~a)}

7(b~--a~)~v~2~m~V-o Coth{V'[2.a}
(5.92)

2EoTI

!iiiiV

(even case) (odd case)

- (b-a)V2mV

2E o11

rhus the effect of including an additional potential well is one the one hand to produce a general shift in the ground state energy ;riven by (b-a) I2mV0 In the other hand it causes a small split between the energies of the even and odd wave functions. Using the asymptotic relations Coth x '"
x+oo

1 + e

-2x

and

Tanh x

1 _ e- 2x

He in fact get that the even wave function has the lowest energy and that the energy split between the even and odd case is given by (5.93)
E_ E+ "'"

4Eoll (b-a) 12mvo

exp

[1 -!i

(IZmvo ) 2a

Returning to the case of a general double well we thus see that the classical degeneracy has actually been lifted. The symmetric ground state wave function suggests that there is an equal probability of finding it somewhere between the wells. This is related to the tunnel effect. Classically a particle with energy E+ cannot penetrate through the potential barrier, so it is totally unaffected by the presence of the second well. Quantum mechanically there is,

223

however, a small chance for the particle to be in the region between the wells. Semiclassically we can thus describe the ground state in the following way: Most of the time the particle is vibrating close to the bottom of either of the wells, but from time to time it leaks through the barrier and moves from one well to another, i.e. it tunnels forth and back. Notice that the instanton solution (5.84) precisely interpolates between the two classical vacua and we can therefore think of it as representing the tunneling between the two vacua. Similarly we may think of the multiinstanton solutions as describing the tunneling forth and back between the two vacua. Let us further consider the Euclidean equations of motion and the associated Euclidean energy: Minkowski version (real time) dZx dt 2

(5.94)

Euclidean version (imaginary time) d x m -2 d.


2

Equation of motion Energy

- V' (x)
2
+

V' (x)
2
+

dx im(dt)

V(x)

!m(~~)

V(x)

The Euclidean energy (which is the Wick rotated version of the Minkowskian energy) is indefinite, but it is still a constant of motion. In the Minkowski case, the demand of vanishing energy leads to the classical vacua, i.e. x(t) ; + a But in the Euclidean case it leads to the instanton equation (5.87),
i.e.

dx
d1:

I/V(x)
21m

Thus the instantons are zero energy solutions which interpolate between the classical vacua. (The same is "almost" true for the multiinstantons) . This is now a general aspect of the tunneling phenomena: The existence of a Euclidean zero-energy solution which interpolates between two different classical vacua signals the presence of tunneling between these vacua, and this tunneling lifts the degeneracy of the classical vacuum.

224

Let us in the light of the above interpretation of the instanton return for a moment to the double square well. Here we have shown that the energy split is given by 4Eo11 {1 ~ (b-a)~ exp - ~ 2a /2mV0 o
}

This can be related to instantons in the following way: to (5.86) the action of a single instanton is given by +a

According

f
-a

~d$

Za/ 2mvo

Thus the energy split is precisely proportional to exp { This strongly suggests that the energy split is related to some Euclidean path integral of the form
x(oo)=+a

So} .

exp{- iSE[x(T)]} D[X(T)]

x(-oo)=-a

where we have evaluated the path integral by expanding around the instanton solution in the usual fashion. We shall make this precise in a moment. As a final remark in this section, we consider a scalar field theory in (1+1)-spacetirne dimensions based upon a potential density U[$] , which has a degenerate minimum. This includes our favourite examples: The $4-model and the Sine-Gordon model. what happens when we quantize such theories - does the vacuum degeneracy survive the quantization or is it lifted by instanton effects? The vacuum degeneracy persists! An instanton in this case would correspond to a soliton in two space dimensions, but the existence of such a soliton would violate Derrick's theorem (see sec. 4.7). Thus a scalar field theory in (1+1)-dimensiohs cannot support instantons. It can also be argued that two different classical vacua are separated by an infinite potential barrier. To be specific, let us consider the $ 4-mode l. To be able to "visualize" the argument let us furthermore use the string-analogy of the field theory, cf. section 3.1. The potential energy U[$] corresponds to a double valley separated by a ridge of height ~4/1X, see fig. 62. A classical vacuum corresponds to a string which lies in the bottom of one of the valleys. Now if we want to convert one vacuum into another, we would have to lift the string a=ss the ridge. It requires the energy

225

U-axis U[<I>] <I>-axis

X-axis
Fig. 62

4 LL A

to lift the string across a ridge of length infinite amount of energy to lift the string long ridge. The same type of argument shows quantum mechanically stable, i.e. that it is into one of the vacua.

L.

Thus it requires an across an infinitely that the soliton is not allowed to decay

5,9

INSTANTON CALCULATION OF THE LOW LYIMl

ENE~Y

LEVELS

Now, that we know the general aspects of the instantons, we shall try to make an explicit calculation of the energy-split in the double well. As a byproduct of this calculation we will also be able to verify the previous given description of the ground state. Since we are not interested in the high-lying energy states we need not evaluate the total path-cum-trace integral. It suffices to calculate the asymptotic behavior of the propagator as T goes to infinity. This follows from the decomposition (2.70), which in the Euclidean formalism becomes: (5.80) KE(xb;TblxalTa) = nL <l>n(X a )<l>n(xb)exp{ -

En T }

In the large T-limit the right-hand-side will clearly be dominated by the ground state. Also the "amplitude" will be related to the wave function.

226 Consider e.g. a single potential well as the one sketched on figure 58. The classical ground state is clearly given by

which in fact is the absolute minimum of Euclidean action, so this path will dominate the path-integral. Consider a small fluctuation around the classical ground state
x(T)

n (T)

Expanding the potential to second order, we get 2 2 2 V[X(T)] R1 W"(0)n (T) = tmw n (T) with The Euclidean propagator from given by 0 to 0

mw 2 = V"(O)

is therefore approximately

But the a~tion corresponding to the classical ground state is zero and the remaining path integral is just the Euclidean propagator for the harmonic oscillator. Consequently we get (cf. (5.76)):

we_then go to the large T-limit where we can replace the hyperbolic Sine by an exponential function, so that we end up with (5.95)
KE(O;"! 0; - 2) '"
T+oo

TI

,tmW - ~'T VTiK e

On the other hand we get from (5.80) the following asymptotic behavior: (5.96) A comparison therefore reveals that (5.97)
$0 (0) ""

4!IDw VTiK

and

This is in precise agreement with the previous obtained results concerning the harmonic oscillator (see especially (5.25)) respectively the weak coupling limit of the low-lying energy states in a single potential well (cf. (5.53)).

227

Having "warmed up" we now turn our attention to the double well, cf. fig. 63. In this case the classical ground state is degenerate
X. (,) '= :!: a

and we shall correspondingly take a closer look at the following four propagators: K(-a;il- a ; -~) K(+a;il- a ; -~)

K(+a;il+a;-~)
K(-a;il +a; -~)

The first two connect a classical ground state with itself, while the last two connect two different classical ground states. The calculation of their asymptotic behavior will be performed in two steps: In the first step we will deduce the general structure of the asymptotic behavior leaving a single parameter uncalculated. This will especially put us in a position where we can deduce the qualitative behavior of the quantum mechanical ground state. In the second step, which will be performed in an illustrative example, we will then calculate in all details the missing parameter. Then this will finally give us an eXplicit formula for the energy split. Okay, to be specific, we concentrate upon the propagator
(5.98)

In this case the path integral is in particular dominated by the instanton solution (in the large T limit). Besides that we will also get contributions from strings of instantons and antiinstantons. Let us denote such a path by
X

[n;'1"""n]

(t)

where

n is the total number of instantons and antiinstantons, while

'1"""n are the consecutive positions of the consecutive centers. The centers consequently obey the constraint
(5.99) -

i < '1

< '2 < . <

'n <

In order to get a string of widely separated instantons and antiinstantons the difference between two consecutive centers should actually be at least of the order of ~ (the width of the instanton). In the large T limit the subset of configurations where some of the instantons and antiinstantons overlap, will, however, be very small compared to the total set of configurations satisfying the constraint (5.99) and therefore we will neglect them. Notice too

228

that since we propagate from -a to +a the integer n must necessarily be odd. Okay, according to our general strategy for semiclassical approximations we must now expand around each such quasi-stationary path

and approximate the potential by 2 V[X(T)] '" V[X(T)] + W" [X(T)] n (T) The corresponding contribution to the path integral then decomposes as

where SE is the Euclidean action of a particle moving in the time dependent potential
U(n)
=

!V"[x

[nn 1'. ,Tn1

(T) 1n

The remainingpath integral corresponds to the Euclidean propagator of a particle moving in the above potential. Let us for notational implicity denote this propagator by
K (0;2 iO ;

2)

where we have suppressed the dependence upon the centers (T1' . ' Tn). Notice that the Euclidean action of such a string is approximately given by
S

[~[ n;T1' . "n1 1 "" ns0

where So is the Euclidean action of a single instanton, cf. (4.38). Finally we must sum up the cpntributions from all the quasistationary paths, i.e. we must sum over n and integrate over T , ... ,T . The total approximation of the propagator (5.98) is thus n 1 given by

(5.100)

229

We then turn our attention to the funny propagator first the case of a single instanton centered at T=O. the potential
U[n]

Kn. Consider Notice that

!V"[x(T)]n

degenerates to a harmonic oscillator potential except when we pass right through the instanton, i.e. outside the small interval [-~,~] the potential is given by

This suggests that we should compare the distorted propagator K with the propagator of the harmonic oscillator, which we will denote by Kw. We therefore put
(5.101) K (0!10 w '2 '

where ~ is supposed to be small. Using the known expression for the Euclidean propagator of the harmonic oscillator (5.76) we then get
K(01-2 I01 - .".) ,t;

if mw = ~2lT'fi Sinh[wT]

Since we are going to investigate the asymptotic behavior in the large T limit, we can safely approximate the hyperbolic Sine by an exponential function and we end up with
K(01j101 -

i)

~ ~e-WT/2 ~

Before we proceed to the case of multiinstantons, we want to make two important remarks about the behavior of the correction term ~ in the large T limit:
In the large T (1)
(2)
~
~

limit

is independent of T, and is independent of the center of the instanton.

Proof: Suppose ~ values of T:

depended upon

and consider two different large

IZlFZ7l\
-(1 +(1

230 According to the group property (2.25) we then get


~ T2 T2 K(O;TIO; - T)

T2 T1 ~ T1 = ff~ K(O;TIY;T)K(Y;Tlx;

T1 ~ - T)K(x;

Outside

[- T1/2;T1/2] x

the curly propagator reduces to the oscillaK(y;T1/2Ix; - T1/2) is exponentially and yare close to the "classical" vacuum

tor propagator.

Furthermore

small, except when

= 0,

cf. the behavior of the ground state wave function (5.25). Using the group

In the above integral we can therefore safely replace K(y;T1/2Ix; - T1/2) by ~(T1)Kw(Y;T1/2Ix; - T1/2). property once more we therefore get

T2 T2 T2 T2 K(O;TIO; - T) ~ ~(T1)Kw(0;T!0; - T) i.e.


~(T2)

= ~(T1)'

The second statement is proven by a similar

argument.

Notice too that the correction term corresponding to an antiinstanton is the same as for an instanton, since they generate the same potential

Urn].
Here the

We now turn our attention to the multi-instanton case. "particle" moves in the potential

For large

and widely separated instantons and instantons:

- T/2

T/2

we then get exactly as before that the potential reduces to the oscillator potential except when we pass through one'of the instantons or anti-instantons. From the group property (2.25) we get

As in the above proof we

c~n

safely replace

K(xk;Tk!xk_1;Tk_1) by

Using the group property once more we therefore deduce TI TnT T) ,n -wT/2 .~ Kn (0;-,: 0; - Z) "" ~ Kw (0;'11 0; - Z """ e Vifi

231 Inserting this our approximate formula for the full propagator (5.100) now reduces to

But neither So nor ~ depend upon the intermediary times '1""'n' Consequently the time integration can be trivially executed. Using a symmetry argument we quickly get

Inserting this we can finish our calculation: (5.102)


=

rmw e -wT/2 v"iTh

Had we considered the propagator from -a to -a itself everything would have been the same except that this time n should be even. We would therefore end up with the analogous formula (5.103) KE(-a;il-a; -

i)

~ ~
T...oo

e-

wT 2 / Cosh

{~T

exp { -

So}}

Finally the propagators are invariant under the sUbstitution a ... -a -a ... a

putting everything together we have thus shown:

(5.104a)

(5.104b)

232

Hx)

-a

+a

Fig. 63a

Fig. 63b

Comparing this with the general formula (5.80), we see that the ground state has split up into two low-lying energy states $0 and $1' The energy split is given by (5.105) E1 - Eo = 2fi.tI exp [ and
~1

So]

and the wave functions $0

satisfy:

Thus the ground state wave function $0 together with its close companion ~1 peaks at the classical vacua ~a, but the peaks are only half as great as the corresponding peak for the oscillator ground state, cf. (5.25)

~o (+a) = $0 (-a) = 12. ~ = $1 (+a) = -

$, (-a)

Clearly the true quantum mechanical ground state $0 corresponds to an even combination of the corresponding oscillator ground states $_ and ~+, while $, corresponds to an odd combination. Notice that in accordance with the node theorem in section 4.7 the even combination has no node while the odd combination has precisely one node. We have thus in all details verified the qualitative description outlined in the preceding section. Let us make some general comments about the method we have used: If we imagine that the instantons are some kind of particles (which they are not!), then we can think of an arbitrary distribution of instantons and anti-instantons as a kind of "instanton gas". To calculate the Euclidean propagator we are then summing up various types of configurations in this "instanton gas", but the crucial

233 point is that the calculation is dominated by configurations consisting of widely separated instantons and anti-instantons. For this reason the above method is often referred to as the dilute gas
approximation.

We can now check if the method is self-consistent, i.e. if the main contribution to the approximation (5.104) really does corne from widely separated instantons and anti-instantons. Consider the intermediary result (5.102):

KE(a;il- a ; -

i)

T~oo

~~

e-

wT/2

n odd

1
n!

{LIT exp [ - ~ Sol}

Now, in a series like


n n!

xn

the main contribution comes from the terms where


n
~

(As long as n < x, the factorial is beaten by the power and vice versa). In our case the main contribution thus comes from terms where

i.e.

But niT is the mean density of the instantons and anti-instantons, and since the right hand side is supposed to be small, this mean density is very small. Thus the main contribution really comes from "dilute gas" configurations. As another application of the dilute gas approximation we now consider a particle moving in a periodic potential, sayan electron moving in a one-dimensional crystal. Thus the potential is very analogous to the potential in the Sine-Gordon model, and you should keep the Sine-Gordon model in the back of your mind when reading the following discussion. In the case of a periodic potential the classical ground state is infinitely degenerate (cf. fig. 63b)
Xj (T)

ja

= ...

,-2,-1,0,1,2, ...

We shall correspondingly try to estimate the propagator which connects two such classical ground states which we label by the integers Land J' +

234

The corresponding path integral is dominated by multi-instanton configurations, i.e. by strings of widely separated instantons and/or anti-instantons. But notice that this time there is no restriction on the positions Qfthe anti-instantons relative to the instantons. If we label such a multi-instanton configuration by

(where

is the number of instantons,

the number of anti-inn, m

stantons etc.) then the only restrictions to be placed upon instantons, are accordingly given by n - m

and the centerpositions of the instantons, respectively the anti-

j+ -

T - 2

< ~1 <"'<~m < 1: 2

Following the same chain of arguments as in the case of the double well we therefore obtain the following asymptotic estimate:

Using that 211

~ vab = 211

e i6 (a-b) de
n

o
we can in fact get rid of the restriction on the summation over and m and we thereby get

But then we can explicitly perform the double summation which, in fact, decomposes into the product of two separate summations:

_ ,/mW -V'rt

-wT/2

2rr

211

f
o

-26T

exp

~
u

S }cose
0

-ie(j+ -

j_)

de

So the calculation has come to an end, and after a little rearrangement we get the final result:

235

(5.107)

Now, how are we going to interpret this asymptotic behavior? If we compare it with the general expression for the asymptotic behavior of an Euclidean propagator (5.80), we see that this time the classical degenerate vacuum has in fact been split up into a contin-

uous band of low-lying energy states


e. ground state (5.108) Notice also that
'"
'!'

~e(x)

labelled by the angle


~e

Furthermore the energy of the eigenstate e=O) is given by E(EI) "" 26exp {-

(relative to the

so} (Cose - 1)

(.

EI

_ Ja ) -

4~w -

-1- e ije 11ft I21i


~e

This suggests that the eigenstates form (5.109) where


~

is a superposition of the

~(x-ja)
(x)

ije

is the oscillator ground state wave function, cf.

(5.25).

Notice that the periodic potential possesses a discrete symmetry, since it is invariant under the translation

x ... x + a
If we neglect tunneling effects, and correspondingly try to represent the quantum mechanical ground state by the discrete series of wave functions <P j (x) "" <P (x-ja) then we break this translational symmetry. symmetry: The correct eigenstates The effect of the instantons is thus, precisely as for the double well, to restore this <Pe (x) are given by (5.109) which
is

are now eigenfunctions of the translation operator: <pe(x+a) = e


~e(x)

All the preceding results are in fact well-known from solid state physics, where wave functions, which like <Pe are eigenfunctions of
~aves.

the translation operator, are known as Bloch

236

5.10,

ILLUSTRATIVE EXAMPLE: CALCULATION OF THE PARAMETER

So far our discussion of the double well has been dominated by qualitative arguments. Now we will finish the calculation, so that we in principle can compute the actual figure representing the difference in energy between the two lowest lying states. We have already deduced the formula: (5.105) In this formula So is the action of a single instanton, which
+a

according to (5.86) is given by (5.110) So

=J

12mV(x) dx

-a
Thus it is a computable number once we have specified the potential. The other parameter 6 is, according to (5.101), given by
K(O;~I 0; - ~)
T T

Here

is the Euclidean propagator corresponding to the motion of


U[x]

a particle in the potential

2 lV"[x (1:)]x o

where

X (1) is the one-instanton solution (5.84), and Kw is the o propagator of a harmonic oscillator with the frequency w. In

T-+oo

x(+I)=o

f
thus get

_f +J2 dx 2 1 eAll-~ f(d1) + w2x 2}d1]D[X(1)]


-2
According to (5.46) we

x(-z)=O

One way of calculating this ratio is to use the relationship between path integrals and determinants.

lim

T"''''

237 But in the limit where operator, (5.111) has a zero eigenmode given by (5.112 ) i.e. by construction
~o T~

we know in fact that the differential


V" [~(-rll

is an eigenfunction of the differential This is the old story which

operator (5.111) with the eigenvalue O.

we encounter every time we have a symmetry in the Lagrangian which is broken by the particular solution in question, cf. the discussion of the translational mode in connection with the solitons (sec. 4.7). But the other differential operator (5.113) has no zero eigenmode. - 0 2 , - w2

It's lowest eigenstate is given by: 7[2 Le. ~1 (-r) = Sin[;: 1 +-::2
T

As a consequence only the first determinant goes to zero in the large T limit. Evidently this implies that as But this is a disaster!
T ...
00

The whole discussion preceding this illu6 T is a very for large TI

strative example was based upon the assumption that small quantity, which is furthermore independent of So where did we go wrong?

Fortunately it turns out that all the conclusions in the preceding discussion are in fact correct, but that we have been a little sloppy in the introduction of subtle but serious error. To understand that, we must reconsider the whole philosophy behind our approach: To calculate a path integral, we proceed in two steps: First we find the quasistationary paths, i.e. in the present case the multiinstantons

6, where we have in fact committed

x[

ni

-r 1'

,'f n

1(-r)

Next we consider the fluctuations around each such configuration:

x(-rl = x[
T~e

n;'1''''''n

1 (-r)

+ n (,)

path integral is then evaluated as a sum of contributions coming

from each quasistationary paths, i.e.

238

(where the sum over the instanton centers integration!). evaluated in the Gaussian approximation.

'1' ... "n

is actually an

Each of the "shifted" path integrals is thereafter The problem now arises To study It comes

because the set of quasistationary paths is labelled not only by a discrete parameter, but also by continuous parameters. this in detail consider the one-instanton contribution. from a curve in path space, the instanton.

we perform the Gaussian approximation corresponding to fluctuations around we in fact make a "doublecounting". the instanton curve in path space include contributions from the nearby one-instanton paths

x, o (,)

where

'0

labels the center of

'0

The fluctuations along

but they have already been included when we "sum" over the positions

'0 Thus we see that when performing the Gaussian approximation we ought only include fluctuations

which are perpendicular to the instanton curve in the path space! Notice that
(5.114 ) Consequently the fluctuations along the instanton curve are precisely generated by the zero eigenmode, i.e. these fluctuations are of the form

This indicates a connection with our determinantal problem:

On the

one hand it is the zero mode which generates fluctuations along the instant on curve (and thereby causes a "double counting" in our naive approach). On the other hand it produces a zero-eigenvalue in the T + "') determinant (as

We can now resolve the problem in the following way: When calculating the shifted path integral in the Gaussian approximation, we

239

must only integrate over fluctuations which are perpendicular to the instanton curve. Let us denote the resulting path integral by:

Let us furthermore consider the associated differential operator (5.111). It possesses a complete set of normalized eigenfunctions

2 { -a T

!V "[ X0 (T) m

l}

<I>

n (T)

where we know that where T

<l>O(T)

is proportional to

dXo/dT

(in the limit

goes to infinity).

Normally we would integrate over.all

the "fourier components" in the shifted path, nIT)

ao<l>O(T) +

n"'1 but in the above path integral we have excluded integration over a ' o since this is to be replaced by an integration over the instanton center To' This causes, however, a slight normalization problem. In the standard evaluation of the path integral (leading to the determinant) we use the integration measure

an<l>n(T)

cf.

(5.48).

We must now reexpress this in terms of

dTo'

According

to (5. 114), we get

and consequently

The correct integration measure associated with the integration over the instanton center is therefore given by (5.115 )

j~dTo
dTo' This will now be includTo sum up the correct defini-

When we originally performed the integration over the instanton center, we missed the factor in front of ed in the corrected definition of 6.

240

tion of

is therefore given by

(5.11b)

lim
T
+
00

=~ T+oo lim

!'Jet{-a~

- lVII[XO(T))}]-l

Det{-a~ -

w2

(where the stroke indicates that we shall omit the lowest eigenvalue in the evaluation of the determinant!). computed using the following strategy. [- 2; + 7 1
T T

This expression can now be On a finite interval Ao(T) will only be

the differential operator (5.111) will not have an exact T


+

zero eigenmode (i.e. the smallest eigenvalue vanishing in the limit where tion (5.50) we therefore get:
00).

Using the determinantal rela-

(5.117)

AO(T)Sinh[WT1]lz Wfo(!;T)

/5

Ao(T)eXP[WT1]!;

lim
+
00

[ 2wf (!;T) o

It remains to calculate
a)

fo(~)!
f
o

and

Ao(T):

Calculation of

(7):
is the unique solution to the differential

By definition equation (5.118)

fO(T)

{- a2 T

1 m

V"[ 2 (T)l} fO(T)


0

which satisfies the boundary condition

241

cf. the discussion in section 5.5. above differential equation is


S

One particular solution to the

-,

dx0
d1:

In the large

limit this is properly normalized and must there-

fore have the asymptotic behavior:

since (5.111) reduces to (5.113) for large behavior of the one-instanton solution deL
(5.119 )

1:.

Here

Co

is a con-

stant, which by construction, is extracted from the asymptotic

lim 1:-+0> is given by


(5.110) .Thus Co can be computed once

Here.the parameter So follows that $0

we have specified the potential.

Since

Xo

is an odd function, it

is an even function.

Consequently we may summa-

rize its asymptotic behavior as follows:


(5.120)

$0(1:)

The second order differential equation (5.118) has two independent solutions. $0 as follows: The second one can be constructed explicitly from Put 1:
(5.121 )

CO e

-W1:

as as

+00

Co

eW 1:

1"..... T ...,..

J o $~
$0 has no nodes and that

dS
(S)
~o

(Notice that $0

is not proportional to

since the integral is 1:-dependent). 1:


I

Differentiating the above

identity we now get


~o (1:)
$~(1:)

J 0 $~ (S)
1:

dS

+~

and
~~ (1:)
~~ (1:)

J $~ (S)
0

dS

They imply the following two identities:


(5.122)
W[$o;~o]
= $0

o err -

d1/J

~o

<rr

d$o

242 and (5.123) (The second one suffices to show that Since
~o
~o

is another solution).
~o

is an even solution to (5.118), it follows that

is From

an odd solution. We are also gOing to need the asymptotic behavior of (5.120) we get
T
~o'

C e -WT o

JC2 d~2WT e
o

The asymptotic behavior may thus be summarized as follows

(5.124)

~O(T)

R;

1 WT 2C w e o

as T ... as T
-+

+00

{-

-WT 1 e 2C w o

We proceed to investigate the particular solution the one we are really interested in. combination of
~o

fo

which is

We know that it is a linear

and

~o:

fo(T) = A~O(T) + B~o(T) Clearly A and B can be expressed in terms of Wronskians:

and

conditions satisfied by
W[~o;fol

But these Wronskians may also be computed directly using the boundary f . o

=[<11

dfo_ f dT 0 df

df]~ =_ f
d~

d~

~o(- 1)

and
W[~o;fol

=[~ o ~dT

dTO]IT

=-

~o( - 1)

"2

Consequently (5.125) From this expression, which is exact, we can extract the asymptotic behavior of fo:

243

2w

Especially we get: (5.126 ) in the limit of large T.

b) Calculation of Ao(T): To compute the lowest eigenvalue we must first determine the solution (5.127 ) which satisfies the boundary conditions fAIt) to the differential equation

T T f A(- 2)=3,f A (- 2)
The eigenvalues tion

A are then determined from the subsidiary condi-

For small values of (5.128)

A we can now use the approximation: fA (,) ~ f o (') + A dfA (B\I

A=O
If we introduce the function
<P(,)

= cn:-Ih)
A=O

dfA

we get from (5.127) that it is the unique solution to the differential equation: (5.129) which satisfies the boundary conditions T d T <P(- 2) = d,<P(- 2) The general solution to the inhomogeneous equation (5.129) is given by: atP o (') + 131/10(') +

J
T

[1/Io(s)tP o (') - tP o (s)1/Io(')] fo(s)ds

- 2

244

But from the above boundary conditions it follows that both S vanishes, i.e. ~ is given by

and

(5.130)

~(,)

J
- 2"
T

[1/Io(s)tP o (') - tP o (s)1/Io(')] fo(S)ds

Inserting this in (5.128) we thus end up with the following approximation: fAt,)

f o (') + l

J
T

[1/Io(s)tP o (') - tP o (s)1/Io(')] fo(s)dS

- 2"
But fo is given by (5.125). tPo and
T
T

Inserting this and using the symmetry T limit:

properties of

1/1

we now get in the large

2"
fo(i) +l
T.,.."

f l ('2") ...

J {tP~(f)1/I~(s)
T

1/I~(i)<I>~(s)}

ds

- 2"
But in the same limit

2"
T

+00

J <I>!(S)dS ... J
2"
while
T

.~(S)ds

2"
T

J 1/I!(S)dS ...

wT

- '2
2 (.'!'.) goes like Consequently the first term is bounded (since tP o 2 WT e- ) while the second term grows exponentially. Inserting the asymptotic behavior of fo and' 1/1 we thus end up with:
0

It fOllows that the lowest eigenvalue is given by (5.131 )


lo(T) '"

4C~

we-

wT

245 Now that we have calculated both return to the formula for (5.117) we obtain: (5.132 )
C
~.

T f o (2)

and

Ao(T)

we can

Inserting (5.126) and (5.131) into


SW

rn ~
__ 0_

Thus the energy split (5.105) is given by (5.133) E1 - Eo ::: 2C

VmS~flW

exp { -

So}

This ends our discussion of the double well. As an example of the application of the above machinery you can now work out the detailed formula for the energy split in the double well specified by the fourth order potential *) v(x)

,4

(x

JT)

2 2

*) Details can be found in: E. Gildener and A. Patrascioiu, Phys. Rev. D16, 423
(1977) .

SOLUTIONS OF WORKED EXERCISES:


No. 5.2.2

(a) From Taylor's formula we get


Hn(x) =.a::.... az n exp[2xz_z 2 ]=e X2 n -"- exp[_(x_z)2] "Zn

Iz:O
axn

Iz-O

2 n a {exp[_(x_z)2]} = (-1) ne x ---

Iz=O

(b) By completing the square


_u 2+2iux = _(u_ix)2_x 2

we can immediately perform the Gaussian integral:


e J

'"-u2+2iux

du

246

Using Rodriques' formul~ we now get 2 n (_2i)n x 2J n -u 2+2iux i=lln x a J -u +2iux r= e - - e du du fiT e u e .11 axn
(c)

n=O

n (-2i )2n X2+y2J n n 2+2 2+ 2 . L 2.... e JU v e-U lUX-V lvy dudv n=O 2n n! 11

We can now easily perform the Gaussian integrals whereby we obtain

PART II

THE MOBIUS STRIP

PARITY VIOLATION!

BASIC PRINCIPLES AND APPLICATIONS OF DIFFERENTIAL GEOMETRY

248

II

GENERAL REFERENCES TO PART II: M. Spivak, "A Comprehensive Introduction to Differential Geometry", Publish or Perish (1970). S.1. Go IdberC], "Curvature and Homology", Academic Press, New York (1962). G. deRham, "Varietes Differentiables. Harmoniques", Hermann, Paris 1960 V. Guillemin and A. Pollack, "Differential Topology", PrenticeHall, Englewood Cliffs (1974). 1.M. Gelfand and G.E. mic Press, New York L. I. Schiff, "Quantum
~,
~'ormes,

Courants, Formes

"Generalized Functions", Acade-

(1964).
~Iechanics",

McGraw-Hill Kogakusha ltd (1968)

A. Jaffe and C.H. Taubes, "Vortices and Monopoles", Birckhauser (1981)

249

II

ehapter 6
DIFFERENTIABLE MANIFOLDS TENSOR ANALYSIS
6.1 COORDINATE SYSTEMS
To simplify our discussion we shall work entirely with subsets of Euclidian spaces! l-le will assume the reader to be familiar with Euclidian spaces, but to fix notation some useful properties and definitions are collected in the following table:

I I
i

n The Euclidian spaceR consist of all n -tuples (xl, ... ,xn) n where xl, ... x are arbitrary real numbers.

Rnis a vector space with an inner product


['inea:t' s truature :

Addition:

x+y

Multiplication with a scalar:

Ax=

, [Xx"'nl ]

[~Xx"'nl]
A

Algebraic structure

Inner product:
<xiY> = xlyt+ ... +xnyn The inner product has the following characteristic properties:
1. Symmetry: <xlp = <ylx> 2. Bi-linearity: <Ax+).Jy z> = A<xl z> + ).J<Yiz> 3. positive definiteness, i.e. <xix> > 0 if x o.

The inner product generates the Euclidian


norm:

250

II

Metric: The distance between two points x and


y is defined as d(x;y) =

II

x-yll = .; (xl_yl)2+ ... +(xn_yn)2

Balls: The open ball with center Xo and radius


<: is given by

B(x ;E) = { x Rnl d(X;\) <<: } o . "", ... ,


,

--

~,

' ..... _-,,,

r---: ,
0,

,\

n Open sets: A subset A c R is called open if

ITopological
I structure

each point in A can be surrounded by a ball lying entirely in A.

Neighbola'hood: A is a neighbourhood of Xo if

+ ' ...... _~I

,....... -...,

it contains a ball centered at Xo

Convergence : A sequence xn converges towards


Xoo

if:

Continuity: A map, f :

A~B,is called continuous if it has one of the following two , equivalent propert ies :

1. It transforms 'every convergent sequence into a convergent sequence, i. e. x ..... xCI) implies n that f(X ) -+ f (xc), n 2. The preimage f- 1 (U) of every open set U in B is itself an open set in A.

251

II

n Observe, that once we have equipped the Euclidian space R with a metric, the rest of the definitions are standard definitions that are common for all metric spaces. In what follows we shall especially be interested in subsets of Rn which might be referred to as "smooth surfaces". Consider, for instance, the unit sphere 52 in the three-dimensional Euclidian space R': 5 2 = { xE R'I (XI)2+(X2)2+(X')Z= 1 } If you prefer you may think of 52 as a model of the surface of the Earth. "ow we can clearly characterize a point on the surface through its Cartesian coordinates (X I ,X 2 ,X'). This on the other hand is rather clumsy because one of the Cartesian coordinates is superfluous: We can use the equation (X I )2+(X 2 )2+(X')2= 1 to eliminate e.g. x'. Because it suffices to use two coordinates when we want to characterize points on S2 we say that 52 is a smooth two-dimensinal surface in R'. In geography it is customary to use lattitude and longitude as coordinates. In our mathematical model they are referred to as polar coordinates (6,lp) 6 being the polar angle while the azimutal angle.
~

is

Let us try to sharpen these ideas. Let M be a subset of the Euclidian space RN . We want to give an exact meaning to the statement:" M is a n-dimensional smooth surface". The key concept is that of a aool'dinate system: Let U be an open subset of Rn and let ~: U ~ M be an injective map. AS ~ is injective, i t establishes a one-to-one correspondence between pOints in the domain U and points in the image ~(U). Thus each point P in ~(U) is repren sented by exactly one set of coordinates (xl, ... ,x ) (see fig65 ),

We may therefore characterize the points in ~(U) through their coordinates. But that is not enough! To be of any interest ~ must respect the topology on M. Being a subset of RN we know that M is equipped with a metria. Obviously we must demand that points in M lie close

252

II

to each other if and only if their coordinates lie close to each other! We can formalize this in the following way: Let P be an arbitrary pOint in
~(U).

The coordinates are expected to

give information about the properties of the surface around P. But then it is necessary that the coordinates covers a neighbourhooe of P. We can achieve that by demanding that of M.
~(U)

should be an open subset

(x

,~,
I'

,x)

(x +e:

I'

IX

"" +e; )
Fig. 65

Hence if Q is sufficiently close to P, then Q lies in represent it by a set of coordinates: (XI+EI


I '

~(U)
).

and we can

xn+e:

TI

Now, let(P i )i=1, ... be a sequence in ~(U) convergirg to anoint P in ~(U) and look at the corresponding coordinates: (Xi""'X~) and only if (xi,'"
~-I

(X!"",X~). We want the following to hold: Pi converges to P oo if and IX~) converges to (x!, .. ,X~) I but this v!e can achieve by demanding that 4> is a homeomorphism, i. e. that 4> and the inverse map
are continous maps! n By demanding that 4> is a homeomorphism from an open subset UC R to

an open subset 4>(U) c M we have thus guaranteed that the coordinates respect the topology of M. But it would be also be nice if we could express t:1e smoothness of M in terms of the coordinate map 4>. First we observe that ~ can be regarded as a map, U ~ RN,
Y 1 = ~ 1 (x 1,
,

xn)

But then we must obviously demand that if> is a smooth map, i.e. that the components

~l, ,~

have partial derivatives

.2. j
ax

~
'

ax ax

J"

k'

J" k I' etc. ax ax ax


~

a3~i

of arbitrarily high order! If

is not smooth we could easily be in

trouble as you can convince yourself by considering the example:

253

II

<P(x,y)

(x,y,1 xl)

(See figure 66 )
<P(x,y)
~

(x,y, Ix!)

y-axisj

<P

~aXiSx-axis
FIG. 66

But even that is not enough! Given a smooth map <p:U ~ RN, we may form the Jacobi-matrix:

(6.1)

We say that trix

r11 I l
ax
j
<p

<p

is regular at a point xo~(x~, ... ,x~) if the Jacobi ma-

has maximal rank at that pOint. The maximal rank is n x


o

and that

is regular simply means that the n-colum vectors

~ ax

, ... ,

~ ax

~ ax
in U. This is

~ ax
<P

are linearily independent. We will demand

to be regular everywhere

rather technical assumption, but we will give a sim-

ple geoQetrical interpretation in a moment. Let us first look at an example which shows that the assumption is necessary:

~iS
\ x-aus

x-axis
FIG.
Ip.

67

this case

<p

is a smooth map, but the Jacobi-matrix:

254

Il
~

is not of maximal rank when x = O.

Thus

is singular along the

y-axis. The corresponding points in M form the sharp edge! Let us sillMilarize the preceding discussion:

Definition 1 n Let M be a subset of~. A coordinate system on M is a pair (<P,U) where Uc:R is an open subset and ~:U '" M is an injective map with the following two properties: (a) (b)
~
(~ ~

is a homeomorphism trom the open set U onto the open set <prU) respects the topology of M)
(~respects

is a smooth regular map.

the smoothness of M).


as a parametrization of

You may consider a coordinate system

(~,V)

a "smooth surface", but this parametrization need not cover the whole surface. Actually there need not exist a single coordinate system covering a qiven "smooth surface". The sphere,e .g., cannot be covered by a single coordinate system (this will be demonstrated in section 6.2). Let us look again at the Euclidian subset MCR N . Suppose that we have a family of coordinate systems
(~i,Ui)iI

all of the same dimen-

sion n, and suppose furthermore that this family covers M, i.e.

Mc

In this case some of the coordinate systems may of course overlap, which means that the same point P is represented by different coordin nates, say (x1, ,x ) with respect to the coordinate system (~ ,V )
1 1

and (yl, ... ,yn) with respect to the coordinate system (<P2,U2), i.e. P = ~dx', ... ,Xn) = <P2(yl, ... ,yn) In the overlap region, <PI (U I )n<p2(U 2 ) , we have thus the possibility of exchanging coordinates!

255

R~
To investigate
t~is

FIG.

69

closer we now intrOduce the sets (See fi0. 69)


-1 -1

U12~

<PI

(<Pdu!ln<pdU 2 (<PdU!ln<P2 (U 2

U21 ~ <P2

Observe, that the overlap region,CPl (U I )n<P2(U2) ,is an open subset of


M and since
<PI

and CP2 are continuous maps we conlude that Ul' and D21 (See fig. 70)
_1

are open sets. We now introcuce the transition functions


CP21
(6.2)
cP
12

~ <P

2 0

,Pt
'1>2

-1 <t> 1 0

two sets of coordinates,

Suppose P is a pOint in the overlap region. Then P is represented by n (Xl, .. ,x ) and (y 1, .. ,yn) ,where (x 1 , ... ,x n ) ~ CP-:(P) arrl (yl, ... ,yn) = <t>-~(p)

(xl ~,xn)

gjJ
(Y+,",Y

FIG. 70

But then it is evident from the diagram, fig. 70,that: (yl, ... ,yn) = CP21 (xl, ... ,x n ) Thus the transition function CPu has a simple meaning: It
e~resses

the

new coordinates (yl, ... yn) in temzs of the oZd ones


n (x 1 , ... ,x ) = CP12(yl, ,yn)

(x1, ...

:xr).

In a similar way:

256

II

i.e. the second transition function $12 expresses the old coordinates in terms of t:1e new ones! As it follows
t~at
-I

and

<P 12 '"

<P

<I.

the transition functions are homeomorphisms. It will,

however, be important later on that

~~e new coordinates (yl, ... ,yn)

not only depend continuously but even smoothly upon the old coordinan tes (xl, ... ,x ). Therefore we demand that the transition functions are smooth functions. vfuen the transition functions are smooth we will also say that the two coordinate systems are smoothZy re Uzt:ed. The preceding discussion motivates the fOllowing definition:

Definition 2 Let M be a subset of RN. An atZas on M is a fcurrUy of coordinate systems ($i'Ui ) i EI with the following properties: (a) The famiZy (<Pi,Ui)ifI covers M, i.e. M ci~I
(b)

<Pi (Ui )

Any two coordinate systems in the fami Zy are smooUy rdated.

6.2

DIFFERENTIABLE MANIFOLDS
N
equipped with an atlas (<Pi,Oi)iEI' Sup-

Let us look at subset MCR pose f

M ... R is a map from M to R. We know what it means to say that

"f is continuous" because M is a metric space. Let

Po

be an arbitrary

pOint in M. Now we also want to assign a meaning to the statement: "f is differentiable at the,point PO" Let us choose a coordinate system (<PI,OIl which covers Po

$tixL .. ,x~).

Then we can represent f by an ordinary Euclidian function nate system 0


n

1 defi-

ned on the domain of the coordiI ,

y=1 (x', ,x ) with


(Xl, ,X ) eOI

'Q~ I";) i _____ ~ ~o)


f_=_f_.<P_l_ _ _

)~
Fig. 71

f(P

where

I =

fO<p

But as f itself is defined on a neighbourhood of Po we see that the Euclidian representative 1 is defined on a neighbourhood of (x~, . ,x~

257

II

Consequently we define Definition 3 f is differentiab le at the point Po if and only it. the Euclidian representative is differentiable in the usual sense at (xL ... ,xo). You might be afraid
t~at

the definition is meaningless, because you


(~2,U2)

could choose another coordinate system going to do, if it turns out that

covering PO. Then

would be represented by another Euclidian function i'and what are we

is differentiable at (x5, ... ,x~),

while i'is not differentiable at (y!, ... ,y~)? But do not be afraid! i' ven by the expression is gi-

"f' = fO~2t which you can rearrange in the


following way: i'= fO~2= fO(~10~ I)O~2 (fO<l> Jl
0
-1
-1

(<I> 1 O~2) = iO<l>12

As the two coordinate systems are smoothly related, we know that the transition function ever
~lZ

is a smooth function. Thus

we see that when-

is differentiable, so is i ' !
+

Because an atlas makes it possible for us to introduce the concept of a differentiable map, f: V differentiable structure on M. Now suppose we have been and
(~j,Vj)jJ

R, we say that an atlas generates a two different atlasses (<I>i,Ui)iI

~~ven

which cover the same subset M. If all coordinate sy-

stems in the first atlas are smoothly related to all t,1e coordinatesystems in the second atlas, then they will obviously generate the same differentiable structure, i.e. a continuous map, f:M tiable relative to
t~e ~

R,

will be differen-

first atlas exactly when it is differentiable

relative to the second atlas. This motivates the following concept Definition 4
An atlas (~i,Vi)il:;[ on a subset M is called maximal if it has the following property: Whenever (~,V) is a coordinate system which is smoothly related to all the coordinates systems (~.,V.) in our atlas, then (~, V) itself belongs to our atlas, i.e. N, V) = (<I> .,V.) for ~om~ jQ;.X

Thus

a maximal atlas comprises all possible smooth exchanges of the

coordinates on M.

258

II

Given an atlas, say ($i,Di)iEI' you may easily generate a maximal atlas. You simply supply the given atlas with all possible coordinate systems that are smoothly related to the given atlas. The use of maximal atlasses will be important for us when we later on are going to study the principle of general covariance. We can now formalize the definition of a
differ~iable

structure

Definition 5 A differentiable struetu:ro on an Euelidian subset M is a maximal atlas on M.


Finally we can give a precise definition of what we understand by a "smooth surface", which we from now on will refer to as a differentiable

manifold: Definition 6 A differentiable manifold M is an Euelidian subset M equipped with a differentiab le struetu:ro.
Let M be a differentiable manifold. If we can cover it with a single coordinate system, we call it a simple manifoZd. In that case a map, f:
M~R,can

be investigated through a single coordinate representative:

(Xl,

,x n )

In general, however, we will need several coordinate systems to cover

M. A map,f: M--R,must then be represented by several coordinate representatives:


-

fi (x , ... ,x )
(~i,Di).

corresponding to the various coordinate systems lapping regions


Q

In the oVer-

ij = $i (Di)1) $j (D j )

the coordinate representatiJes are tpan patched together through the relations
-

fi (x , .. ,x )

fj (y , . ,y ) (See fig. 72 )

where

Finally we mention that a n-dimensional differentiable manifold M n is frequently denoted M , where the dimension of M is incorporated explicitly.

259

II

l'i g. 72

Illustrative example:

The sphere.

We are now armed with heavy artillery! Let us return to the sphere S2 and see how the machinery works. We want to construct an atlas on the sphere, and thus may define a map, ~: U~S2,in the following way: we must construct some coordinate systems. If U denotes the unit ball in R2, we

~(XI,X2) = (xl,x', ';1-(X ' )'_(X 2 )2)


Clearly this map has a simple geometric interpretation: If we identify R' with the x ' -x 2-plane in R 3 , then ~ is nothing but the projection along the x 3 -axis! (See fig.73). Let us check that (q"U) is acI 2 3 (X ,X ,X )

tually a coordinate system: U is obviously open and q,(U) is the northern hemisphere, (U) ={xES 2 [x 3 > O}

<p

which is open too (when regarI


1<1>

ded as a subset of the topological space S2!) As (Xl ,X2) "',; 1- (Xl) 2_ (x2) 2 x
2

is a continuous map it follows that


<I>

is continuous!

-1
<I>

is

(x ,x )

FIG. 73

nothing but the projection on the first two coordinates

260

II

-1

Hence

<p

of S2. Then we must

is continuous. T.hus we have shown that <p respects the topology check that <p is a smooth map:

We know that the square-root function 'it is smooth when t>O. (1\t t=O it is not smooth due to the fact that:

<it

(It)

2~

-> '"

when

->

0+ ).

But we have restricted U to values of (Xl,X2) where (Xl)2+ (x2)2<1. Thus there are no troubles and <P is a smooth map. We must check the Jacobi matrix too, but that is easily done:

[~] ax6

~ ax

a<p l
aJ{7

0 0 -Xl 1-(xl) 2_(X2) 2 _x 2

axr
~ ax

d<p

~ ax
~ ax

1- (Xl)

2_

(x 2 ) 2

~ and we see that ax] always have rank two due to the upper square unit matrix. Consequently we have also shown that <p respects the smoothness of S2.
Clearly we may construct six coordinate systems in this way and they obviously cover t~e sphere, (see figure 74)

To see that they form an atlas we must check that they are smoothly related, (s ee figure 75).

261

II

FIG. 75

From the diagram you read off:

(XI,X2)~(XI,X2,

A_(XI)2_

(X 2 )2)

~(XI,A_(XI)2_

-r

(X 2 )2)

Hence the transition function


yl= Xl

~21

is given by,

and the corresponding domain

U1 2 by
(X 2 )2 <

UJ2 = [

(X I ,X 2 )

I x2

> 0,

(XI)2+

1}

As the square root never obtains the value 0, the transition function
~12

is obviously smooth: In a similar way you can check that all the

other transition functions are smooth and thus we have shown in all details how to construct an atlas on S2. Consequently S2 is a differentiable manifold and we shall refer to the coordinates constructed above as standard coordinates. What about the polar coordinates correspond to a coordinate system
~(8,~)
(8,~)?

In our new context they

(~,U)

where:

(Sin8Cos~,

Sin8Sin~,

Cos8)
~-axis

But what can we choose as our domian U? There are two kinds of troubles: The first trouble is not so serious. Due to the periodicity of the trigonometric functions several values of

8-axis

the coordinates actually describe the same point:


~(8, ~ ~) ~)

~(8,

+ 2IT) + IT)

(8,

~(-8,~

262

II

But the second trouble is more serious. At the northpole and the southpole the coordinate pute the Jacobi matrix:
~

is completely indeterminate! Let us com-

2Jtl
as

~ll
a~

CosSCos~

-SinSSin~

[a~~] ax]

~2

.t.2
a~

ae
~3

CosSSin~

SinSCos~

.t.3
a~

ae

- SinS

At the northpole (S=O) or the southpole (S=n) we see that sinS =0, and therefore we find

[~
which shows that
~

o
o

o
Thus the coordinate

is singular when S=O or S=n.

system breaks down at the northpole and the southpole and we will have to exclude these from the range of $ lar pOints. because they are singu(Let me emphasize once more: The sphere itself is per-

fect and smooth at the poles. The singularities we have chosen a "bad" coordinate system!).

onZy arise because

The best thing we can do is therefore to put U equal to something like

= {

(S,~)

[0 <S<n

O<~<

2n }

You see that we have not only missed the poles but also the arc r joining the poles. The arc "singularity" is due to the periodicity in
~.

Clearly we may construct an atlas using two coordinate sys(see figure 77) .

tems of this kind

FIG. 77

11

263

II

Now we have constructed two different atlasses on the sphere. Does it mean that we have constructed two different types of differentiable structures on the sphere? No, because if we try to express the polar coordinates in terms of the standard coordinates, we find that the polar coordinates depend smoothly upon the standard coordinates and vice versa:
8

Arcsin[ Arctg [

(Xl) 2

+ (x')' 1

x 2= Sin8' Simp

xr

x2

Thus all the polar coordinate systems are smoothly related to the standard coordinate systems! So they are obviously equivalent, i.e. if a function,f : S2~ R,is a differentiable function when expressed in polar coordinates it will also be differentiable when expressed in standard coordinates. Thus it is irrelevant whether you use standard coordinates or polar coordinates on the sphere, since they will generate the same maximal atlas and consequently the same differentiable structure.
~fuen discussing the sphere S2 you might have wondered why it was

necessary to use several coordinate systems to cover the sphere. It required 6 standard coordinate systems, respectively 2 polar coordinate systems, to cover the sphere! If we were smart enough, might we. then hope to find a single coordinate system covering the whole of the sphere? However, it is impossible to do that:

Any atlas on the sphere con-

sists of at least tuJo coordinate systems.


The reason for this peculiar fact is purely topological. It is intimately connected with the fact that the sphere single coordinate system (CP,U) covering then
cp

~s

compact

topo-

logical space, since it: is a i:Jour:ied closed subset of R


~he

If we had a

whole of the sphere,

would be an homeomorphism of U cR '" '" U homeomortpic onto 52

onto the sphere 52:

But this forces U to be compact itself. Therefore U is both

compact and

open. But the only subset of an Euclidian space which is both open
and compact is the empty set. This is a contradiction! Hence the sphere 52 is not a simple manifold. However, if you cut out just one point from the sphere, then the rest is not compact and nothing can prevent us from covering the sphere minus one pOint with a single coordinate system! For simplicity we cut out the northpole. Then stereographic projection defines a nice

264

II

coordinate system (see fig. 78)

onto

....

6.3

PRODUCTMANIFOLDS AND MANIFOLDS DEFINED BY CONSTRAINTS

Now consider two Euclidean manifolds ~e RP , Nne Rq where ~ is an m-dimensional and Nn is an n-dimensional manifold. We can form the product set M X N consisting of all pairs (xiY) with

xM and yEN. Clearly M x N is a


subset of the Euclidean space RPx Rq Let
FIG.

79

RP+q.We want to con-

struct an atlas covering M x N.


(~,U)

"
M N
~III\I!(UXV)

be a coordinate system
(~,V)

on M and let

be a coordi-

nate system on N:

1 - - -.......- - - .
-x I
M
~(U)

FIG. 80

Then we define a coordinate system on the product set M x N as follows:


(6.3)
~19~:UxV"'MxN

&

~ (Xl, ,xmiyl, ... ,yn)

Clearly U x V is an open set in Rn x Rm

Rn+m and it can be

265 checked in all details that a regular map.


(~i8 ~j,Ui8Vj)
~

II

is a homeomorphism and that it is


(~j'Vj)jEJ is an atlas for

Thus it is a nice respectable coordinate system. If

(~i,Ui)iEI is an atlas for

M and

N, then

is an atlas for M x N. "Ie call M x N the proJu.ctmanim+n dimensional ma-

foZd of M and N. By construction M x N becomes a


nifold.

Some examples may throw light on this rather abstract procedure:

The eyZinder: SlxR, fig. 81.


We have used polar coordinates on Sl. The cylinder is supposed to be extended to infinity in both directions.

, "..---- .... ....

The. torus: SlxSI


In figure 82 the two circles :.ave been drawn with very different radii a > b.

(Observe that strictly speaking SlxSl should be constructed as a subset of


2 2 ,

R x R = R , but that is not so easy to visualize. Observe also that if a


~

b then

FIG.

82

the torus constructed on figure 82 becomes a smooth surface with self-intersections. such a surface is not a Euclidean manifold.) We will now sketch a very general method to construct Euclidean manifolds. First we remind you about the following problem from classical calculus: Let f: Rm+ n ~ Rn be a smooth function and consider the equation: m ,x ,y1, ,yn) =0

:: . 1 (x 1, f(x,y)=o i.e. {

fn (Xl, ,xm ,y1, ,yn)=O

Suppose (xo,Yo) is a solution of the equation f(x,Y)=O. Under what circumstances can we find a

neig~UrhOOd

U of (xo,Yo) in which we

can solve the equation f(x,y)

=0 with respect to y and obtain y as

a smooth function g of x.i.e.y=g (x)?

266

II

n R
y

Rmt-n

';;o,Yo)

f(;;:,y)

FIG. 83

This problem is wellknown. To motivate the solution we study the linearized problem, where
f(xo+~x;yo+~Y)

is replaced by

f(xo;yo) + Dx f~~I+ Dy f~YI


Thus the equation f (x,y) = 0

is replaced by the equation


,

Dxf ~~I + Dyf. ~YI = 0

where we have taken into account the assumption that (xo,yo)

is a

solution, i.e. f(xo,yo)=O. Now the linearized equation can be sol= provided = af i ved with respect to ~YI Dyf=(-a--.) is a regular square matrix. In that case we :furt:heJ::nore get: yJ
~YI

= = - (Dyf) = -1 . (Dxf) = ~xl =

With this in our mind we now state the sOlution to our problem:

Theorem 1 (Theorem of impHdt functions) mrn ~ R n be a smooth function. If U f is a regylar squaro matrix, when Let f: R it is evaluated at a_solution (XO,Yo) to th~ equation f(x,y)=O, there exists a neighbourhood U of (xo,yo) and a smooth function g:Ff' ~ ~ such that:
1)

f(x,y)=

iff y = g(x)
=_1

for aU (x,y) in the neighbourhood U.

2}

D~

=-

(DI)

{ljxf}

(i.e. the partial derivatives of y with respect to x are found by imp licit differentiation:

ai +:d..
a)

al a)

o)

Using this cornerstone of classical analysis, we can now show: Theorem 2 Let f: Rq .... rf be a smooth function (q > p). Let M be the set of solutions to the equation f (iii) = o. If Mis nonempty and if f has maximal rank throughout M, then M is differentiable manifold of c1:imension (q - p).

267

II

We will not go through the proof in all details, but let us Rketch how to construc t coordina te systems on M. Suppose M is not empty and suppose Zo belongs to M, i.e. f (;,) = O. We know that:

Df-Z

Z=Zo

l~
dZ I

~fP

~j
dZ

dZ I

af P q
dZ

has rank p,Which is the highest possible rank. Hence we can find p columns which taken togethe r constitu te a regular square matrix. To Simplify let us assume that the last p columns are independ ent and let us introduc e the notation : = G,y) i.e. (zl , ... ,zq) = (xl, ... ,Xq-P,yl , ... Then D f is a regular square matrix. Accordin g to the theorem of implicit functions y We can now find a neighbou rhood U of Zo = (xo ,Yo) of the form U = { (x,y) I x~-< xi< x~ + e , y~ -e< yj< y~ + e } and a smooth map g:R q- P ~ RP such that

?)

(x,y) EM n U iff
TI(U)
= { x

I x~

- e< x<

x~

g(x) when x belongs to TI(U), where + e }

i.e. TI(U) is the pr~jection of U into Rq - P , cf. fig. 84.

Fig. 84

x:

.)
-ax~s

. x 1-ax~s

Rq-P

But this means that restrict ing ourselve s to U we can define a coordina te system ($,TI(U)) , $:TI(U) ~ M,in the followin g way:

$ (Il) = (iC ,g(iC) ) Such a coordina te system will be called a standard coordina te system. We must now check that $ really defines a coordian te system: a) Since U is open, M n U automat ically becomes an open subset of M. As the projection map IT = f1is triviall y continuo us, we see that $ is a homeomorphism of the open subset TI(U) onto the open subset MnU.
b)

is a smooth function since g is. Furtherm ore

$ is triviall y regular since

D $ = x

' [

, '. 0

.,

where the left part of the Jacobi matrix is the unit (q-p) x (q-p) matrix which is triviall y regular.

~"'l
dyl

268

II

We can clearly cover M by standard coordinate systems and it is not too difficult to check that they are all smoothly related. Thus we have defined an atlas, and this shows that M is a differentiable manifold of dimension (q-p), see fig. 85.

If M is generated as the set of solutions of an equation f(z)=O, we say that M is defined by an equation of constraint. Clearly the sphere S4 is defined by the following equation of constraint:
(Xl)2+ (X 4 )2+ (X 3 )4=

If you review the illustrative example of the sphere, you will in fact find that most of it was an exemplification of the above abstract discussion. The assumption that f has maximal rank is essential. Consider for instance the following trivial example: ven by Let f: R4 ~ R be the smooth map gif (x,y) = X 4_ y 2 af Of Then Df= (ax; 3y) = (2x,-2y) is singular at the point (x,y) longs to (0,0) u But the pOint (0,0) clearly be-

= {(x,y)

Ix 2 -y2= a}. In accor-

Fig. 86
It is impossible to

dance with this we easily find that M is not a I-dimensional differentiable manifold.

find a smooth map ~ which maps an open interval bijectively onto an open neighbourhood of (0,0) in M. (See figure 86). As another illuminative example we consider the smooth map, given by
f:R3~R,

269

II

which is singular at all points on


M={(x,y,z) Ir<x,y,z)= 0 }= {(x,y,z) [x2+y2=1, z = O}. The subset

z- axis
y-axis

M is,nevertheless, clearly a
I-dimensional differentiable manifold (it is isomorphic to the circle S 1). IVhat goes wrong

; 0

is the dimensionality. If f had maximal rank, then M would have been a

2-dimensionaZ manifold!
As a final example of differentiable manifolds defined by con-

straints,we look at the configuration space for a system consisting of a finite number of particles in classical mechanics. Consider for instance a double penduZwn: It consists of two particles where the positions are constrained through the equations
(ql)2+(q2)2+(q3)2= I (q'_ql)2+(q5_ q 2)2+(q6_ q 3) 2= I

Since the smooth map,f: R6~ R2, given by


(ql)2+ (q2)2+ (q3)2_ I (q6_ q 3) 2_ I

[ (q'_ql)2+

(q5_ q 2)2+

is regular on the configuration space M, we conclude that M is a 4dimensional differentiable manifOld. In fact M is isomorphic to
S2xS~

6.4

TANGENT

VECTORS

\'le are now in a position to introduce tangent vectors. Consider an Euclidian manifold M c RN (figure 89) Let P be a point in M. Then we have a lot of vectors at P in the surrounding space RN. Observe that a vector is defined as a directed line segment PQ Thus a vector is characterized not only by its direction and length but also by its base point P!
R
N

2.70
Hence if two vectors have different base pOints P and p', then they are different vectors, even if they have the same length and direction. Let us return to the vectors with the base point P. Not all of them are tangent vectors to the smooth surface M! Let A:
p, the velocity vector,
(6.4)

II

~M

be a

smooth curve passing through P. Then A generates a tangent vector at .... dA vP=dt

Ji

=a
aU possible smooth curves in M passing

The set of vectors generated by denoted by:


Tp(M)

through P forms a set of vectors called the tangent space and it is

To investigate the structure of this tangent space we introduce a coordinate system


(~,U)

which

covers a neighbourhood of P. To simplify we choose the coordinate system so that P has the coordinates (0, ... ,0) (see figure 91). Consider the curves: (6.5) AJlt) =
~(t,O,O,

I~
; .. ; An(t) = et>(O,O, ... ,O,t)

... ,O)

They are called the coordinate lines. By differentiation we now obtain the tangent vectors
(6.6)

..
e

-n

dA dt
2

Jt ,

a
1

They are called the canonical frame vectOl's. To investigate 1 them a little closer we will write out the parametrization of the coordinate lines in their full glory:

FIG. 91

Adt)

~l(t'O'.'.'O)l = . , .. ,
N et> (t,O, ... ,a)

~~ (0, ,a ,t)
et> (0, ...

l.

,a ,t)

By differentiation we get

271

II
1

... e

a<p 1
l

= dt I t=O

dl'l

ai{T

, ...

... e

dAn n

axn a<P J
a$N

axr

a$N

dt1t=0

I (0, .. ,0)

"3XTI.

I (0, . ,0)

Here the coordinates are the Cartesian coordinates in RN and we will refer to them as the extrinsic coordinates of the vector. Now take a look at the Jacobi matrix a<p l
(-ax-r)

a<p

3i{T

~ ax
~ n
ax

a$N
ai{T

It is obvious that the n collDTlYl vectors in the Jacobi matrix are nothing but

the extrinsic coordinates of the canonical frame vectors. But we have previously demanded
(

~ .)

ax] to be of maximal rank n. Therefore the column vectors are linearly independent! Thus, we have shown that the canonical frame vectors are

linearly independent.
H"ext we want to show that the canonical frame vectors span the whole tangent space T (M). Let ~ be an arbitrary tangent vector. ... p p Then v is generated by a smooth curve A. Let us write out the parap

metrization of A: First we observe that A has the coordinate representation:


I(t)

with respect to a coordinate system (cp,U) , cf. ::ig. 92.

Fig. 92
We can then write down the parametrization of A with respect to the Cartesian coordinates in the surrounding space

RN.

272
A(t)

II

[q,l (~l (t) , .. ,x

n (t

q,N (Xl (t) , ... ,x n (t By differentiation we get

(6.7)

... vp

dA dt! t=O

n l ~ dx + . + ~ dx ax ~ ax dt aq,N dx n ~dXI ax ~ + + axrr dt

dt el

dx l

...

+ +

qxn ... at en

t ly we see that the vectors ;;1' ... ';;n really span the tangent vector vp. On the other hand, if ~ is a vector spanned by ~l' ... '~n
i.e.

conseque:

u=a1el+ .. + an~

then ~ is a tangent vector, since it is obvious that it is generated by the curve A(t) with the coordinate representation
A (t).

As e l , , en are linearly independent we therefore conclude that the tangent spaae Tp (M) is a n-dimensional veatorspaae. The vectors -+ -+ e l , .. , en constitute a frame for this vectorspace. Formlla (6.7) now motivates the following definition
DEFINITION 7

-+

-+

Let

tp be

a tangent veator generated by A and let A have the ooordinate repre"A(t)


(x1(t) . ,:rr(t))

sentation:

Then the numbers are aaUed the intrinsia speat to the aoordinate system

. di a1- = at

aoordinates of vp with re-

(q,.U).

Observe that the intrinsic coordinates a i are simply the compo-

...

nents of;j with respect to the canonical frame (e 1 , ,en) . p . The numbers al. are also referred to as the aontravariant components of v P We shall explain this mysterious name in a moment! Now, let us see what happens if we exchange coordinates, say from (x1, .. ,x n ) coordinates to (yl, ... ,yn)_ coordinates. Then the same tangent vector vp is described by two set of numbers, (al, .. ,a n ) with respect to the x-coordinates

...

III

(1)

n { (al, ... ,a ) with respect to the y-coordinates


121 (2)

We want to find the transformation rule which allows us to pass from x-coordinates to y-coordinates.

II

It will be convenient to introduce the notation 0 21 for the Jacobi matrix of the transition function, i.e. ax J Okay, we are ready: Let + vp be generated by the smooth curve A. Then A has the coordinate representation, xi

021

(~)

= xi (t)

in x-coordinates, and it has the coordinate representation, yi = yi (x j (t) ) in y-coordinates. Consequently we obtain
(6 .8)

~I

= di

di

a i dxt fu (it =

ai ~

(II

a1

In a matrix formulation we may collect the intrinsic coordinates into a column vector

Then we can rearrange the transformation rule in the following way:


(6.9)

This is the most impontant of all the formulas involving the tangentvectors. Let us also investigate the connection between the old canonical frame vectors (Il ]. and the new canonical frame vectors \I ~ ].,.NOte that , , ~p- has the old coordinates &)J = = (0, ... ,1, ... ,0), where the number 1 is in the i' th posi tion. But OCJW we may carplte the new coordinates from the above formula:

e,

oI

aj =
(21

~ ak
axk
(J)

~
axk

ok

~
axi

ConsequenUy the vector ~Ii has the following decanposition along the nfM eanonical

f,,-~me

vectors:

-+

(6.10)

~Ii

e"a j
(21J (21

-+

e,

(21J

274

II

Let us fix the notations: ~21 and ~12 are the transition functions with the corresponding Jacobi matrices 0 21 and 0 12 ,
y = ~21{X)
X=12(Y)

(6.11)

5 21 =

[5]

As ~21 and ~12 are inverse maps we conclude that 0 21 and 012 are reciai . a ~ procal matrices! Thus ~ and ~ are components of reciprocal maJ ayJ ax trices and we may therefore rearrange formula (6.1U in the following way (6.12) A quantity which transforms in the same way as the canonical frame vectors is said to transform ao-v~antly (co-varians = transform in the same way). Hence a covariant quantity is a quantity Wj which

transforms according to the rule axi (6.13) w. = w. --. {2jJ 11I~ ayJ when we exchange the coordinates. Observe that we have in the lower position: wi' It is referred to as covariant that a covariant index is transformed with the matrix right. put the index i index. Observe axi from the ayj
i

. dx The canonical frame vectors transformed with the matr~x .. A quan~ ayJ tity which transforms with the reciprocal matrix axJ is said to transform aontra-variantly (Contra-variantly = transforms in the opposite way). Hence a contravariant quantity is a quantity a i which transforms according to the rule

(6.14 )

a
(21

~ axJ

a
(I)

when we exchange 'the coordfnates. This time we have put the index i ~n the upper position: a i It is referred to as a contravariant index. Obs~!'r a"i ve that a contravariant index is transforI'led ",ith the matrix .::...L...ax j from the left. As we have already seen the components of a tangent vector transforms contravariantly! In the preceeding sections we have only discussed the local aspects of functions and tangent vectors, i.e. we have focused upon the i"l\'!rediate neighbour-hood of a specific IXlint P. We will row briefly discuss how we can introduce scalar fields and vector fields gn a manifold M. Let us look at a scalar field first. In our geometrical interpretation, a scalar field is simply a smooth function ~:M ~ R.

275

II

If we introduce two coordinate systems, say x-coordinates and y-coordinates, we may now repr.:sent q, by the ordinary Euclidean functions and and these are related through the
transfo~ation

rule

~ (xi)
(I)

~ (yi)
(2)

where xi and yi are coordinates of the same space time point p. When we investigate such a scalar field p .... q,(P), we will always introduce coordinates and use the Euclidean representative. :')e then "forget" the bar, and write the corresponding Euclidean function as

Next we have vector fields. To construct a vector field A on an arbitrary manifold M, we attach a tangent vector, Ap,to each pOint P in our manifold, (see figure 94 ).

...

Fig. 94 If we introduce coordinates (x1, .. ,x ) on M, then the tangent vector P(xl, ... ,x n ) can be decomposed along the canonical fran

at the point

me vectors el, .. ,e n as follows:

....

....

a (x , ... ,x lei
We say that the vector field is smooth, if its components ai(xl, ... ,x n ) are smooth functions of the coordinates. In general we want to investigate a vector field A through its components, ai(x), but one should be careful. If the manifold M is a non-trivial manifold, then several coordinate systems will be needed to cover it! (Compare the d~sion of 8 2 in section 6.2) x2

-+

[T"
(~
i

2 U2

--..

xl

EJ

Fig. 95

276

If we have two such coordinate regions S"h and the region


(II

Q2

which overlap in

Q 12,

then you will have to use t\qO coordinate expressions

ai(x) and ai(y) to describe the vector field


(2)

A.

In the overlapping re-

gion

Q 12

these two coordinate expressions are row connected through the


ai, n ~ a.f(xl, ... ,x )

tr ansf orma tion formula (6.8):

ax]

(I)

6.5

METRICS
<xly>

n In the ordinary Euclidian space R we have an inner product <I>,

xlyl+ ... + xnyn

which is symmetric, bilinear and positive definite. We want to extend this concept to the tangent vectors on a manifold M. To each tangent space on the manifold M we therefore associate an inner product. This family of inner products is called a metPia and it is denoted with the letter g. If vp and ~p are tangent vectors belonqing to the same tangent space Tp(M) we denote the value of their inner product by:g(;p:~p)

R
i1Q

FIG. 96

Observe that we do not define the inner product of tangent vectors vp and

-+

U Q

with different base pOints.

As in the Euclidian case we shall assume that g is symmetPia,

(6.15)
and bi linear ,
(6.16)

277

II

But we shall not demand that the lCletriCS is positive definite. We only demand that it is non-degener>ate, Le. the only vector which is orthogonal to all other vectors in Tp(M) is the zero vector:
(6.17)

g(VpiUp)~O= for all ~p in Tp(M) implies Vp=

Let us introduce coordinates and let canonical frame vectors. The nunbers
(6.18)

e , ... ,en
1

be the corresponding

are referred to as the metric coeficients. Let vp and up be two tangent vectors with the coordinates aiand b i . T'le can then express the inner product of vp and,u p ent~rely in terms of the metric coeficients and L the coordinates a and b L :
(6.19)

At each pOint P the metric is therefore completely characterized by its components gij(P), Thus we may regard the metric coefficients as functions defined on the range of the coordinate system. If ~Ie identify P with its coordinates (x!, .. ,xn) l.,re may furthermore regard gij as ordinary Euclidean functions In gij= gij (x , ... ,x ) t'le say that the metric is smooth if g ij depends smoothly upon the coordinates. In ""hat follows we will only work with smooth metrics! Let'us summarize the discussion:

Definition 8 Let M be a differentiable manifold. A smooth metric g is a family Of inner products, defined on each of the tangent spaces, with the following properties:
d) g is a bilinear map, which is symmetric and nonderenerate.

b) The metric coefficients gij depend smoothly upon the coordinates

(Xl,."

n ,x ).

The metric coefficients are often collected in a matrix G = (gij)' which by construction is symnetric. Let us denote 17.f ~I and I), the coordinates of two tangent vectors vp and up' Then we may rearrange the formula for the inner product in the following way: j g(vp;u p )= gijaib = aigijbj= Next. we want to show that the matrix is a column vector with the property

a7 GE,
regula:!>, Let us assume that

G is

E,

278

II

We have finished the proof as soon as we can show that

h,

is necessari-

ly the zero vector! Let E, be the components of a fixed tangentvector up and denote by a,the components of an arbitrary tangentvector vp' From GE,= cr,we conclude:
-+ -(GO I )

a,

for all for all

i.e.

g(vp;~)

E,=

Consequently vanishes because g is non-degenerate. It follows that 0 as we wanted it to be.

... up

=0

AS G = (gij) is a regular matrix we knolll that it has a reciprocal matrix! We will denote the elements of the reciprocal matrix with gij. By construction we get

(6.20)

Then we want to see how the metric coefficients transform .,hen we pass from one coordinate system to another. But that is easily done. Let there be given two coordinate systems, x-coordinates and y-coordinates.lf ~afl are the old metric coefficients corresponding to x-coordinates and 8~fl are the new metric coefficients corresponding to y-coordinates, we now get

9:nD
'4~ "

=
-+

gee jeD)
(2\a (2l"

= gee

lIlY

ax Y aya '

(11 0

e ax O)
aye
axo ayfl

g(eYieo)-(1)
(I)

-+

ax Y axO
aya

(~yo aya

axY

where we have used (6.12). So the metric coefficients transform covariantly:


(6.21)

Worked exercise: 5.5.1


Problem:
L:~

gij be the covariant components of the metric. Show that the components

glJ of the reciprocal matrix transform contra-variantly.

Exercise: 6.5.2
Problem: Consider the Euclidean space R3 with the usual metric. In Cartesian coordinates the metric coefficients reduce to the unit matrix

GCartesian '"

[ ~ ~ ~]
001

279

II
(r,e,~)

Let us introduce spherical coordinates

through the relations:

X] y [z X] [:

[r sin e ] r sin e Slntp r cos e


(p,~,z)

C~stp

and cylindrical coordiantes

through the relations:

[P costp 1
P

s~ntp

Show that the metric coefficients in spherical respec:ively cylindrical coordinates are given by (6.22)

ITspherlcal .

[~
=

o 0 0 ] r2 o r 2 sin 2 e

ITcyllndrlcal . .

(Hint: It is preferable to work in the matrix langmage. Let D12 the Jacobi matrix corresponding to a coordinate transformation. Then the transformation rule(6.21) reduces to (6.23)

~j be ay

(~) = Dl~?P12

'* =

We will need some elementary ?roperties of the metric. F~rm the regularity of ~ we conclude that the determinant, g = Det[~l,is non-zero. From the transformation rule (6.23) we furthermore get
(6.24)

It follows trivially that the sign of g is independent of the coordinate system we work in. Clearly the determinant depends continously upon the coordinates. Now let us assume that our manifold M is connected (i.e. any two pOints in M may be joined with a snooth curve within M.) If g was positive at a pOint P and negative at a point Q, then g would have to be zero somewhere between P and Q. Since this is excluded we obtain the following lemma:

Lemma
g has a constant sign throughout M

Being symmetric we know that ~ has n real eigenvalues AI"",An. We also know that ~ is regular, therefore none of the eigenvalues can be zero. (Remember that Det[G 1 = A1A ). It can be shown that the numn ber of positive eigenvalues is independent of the coordinate system.

280

II

(The actual value of AI, .. ,A depends strongly on the coordinate n system in question!) How let us assume again that our manifold is connected. Then we can-extend the previous argument to show that the number of positive eigenvalues is constant throughout our manifold. We will especially be interested in two cases:

Definition 9 If a~l the eigenvalues are strictly positive we say that the metric is an a) Euclidian metric. In that case the metric is positive definite. A maniford with an euclidian metric is called a Riemannian manifold. (If the metric is not Euclidian we speak of a pseudo-Riemannian manifold). b) If one of the eigenvalues is negative and the rest of them are positive we say that the metric is a Minkowski metric. In this case the metric is nondefinite. (Hence a manifold with a Minkowski metric is a special example of a pseudoRiemannian manifold).

Let us investigate the Minkovlski metric a little closer:

Lemma 2 Let M be a manifold with a Minkowski metric. Let T (M) be the tangentspace at the p point P. Then we can decompose T (M) in an orthogonal sum
p
T (M) = V_ P
~

U+

where V_ is a one-dimensional subspace, such that the restriction of g to V_is negative definite, and U+is a (n-1}-dimensional subspace, such that the restriction of g to U+is positive definite. Proof: Choose a coordinate system covering P and let gi" be the metric
coefficients with respect tolthis coordinate system. Then

G=

(gij)

has one negative eigenvalue AO and (n-l) positive eigenvalues AI, .. ,A _ . Let V_be the eigenspace corresponding to AO and let U+be n l the subspace spanned by the eigenspaces corresponding to AI, ... ,A _ . n 1 The restriction of g to V_is negative definite. To see this let ~ be an eigenvector with eigenvalue AO and coordinates a i . Then g(v,v) Hence
~ ~

i j i i i i a gija = a (Aoa ) = AO (a a ) < 0


~ ~

v has

a negative square.

Similarly the (n-l) eigenvectors UI, ... ,U _ with eigenvalues n 1

281

II

Al, ,A n _ l have positive squares. But they form a frame for l4and therefore the restriction of q to U+is positive definite.
Finally we must show that V_ is orthogonal to U.. It suffices to show that v is orthogonal to each of the frame vectors U1, . . . ,U _ . But n l i j i i g(V;U ) a gijb a (Akbi) Ak (aib ) k and i j g'(Uki V) b gija As AO*A k
~ ~~ ~

vanish.

(A k is positive and AD is negative) it follows that g(v;l1c) must

+ ...

This lemma motivates the following classification:

Defpiition 10 iet M be a differentiable manifold with a Minkowski metric: (1) A vector with negative square is called Time-like. (2) A vector with zero square is called a null-vector. (3) A vector with positive square is called space-like. The null-vectors forms a cone, the light-cone, which separates the time-like Vectors from the space-like vectors. (See fig. 97 )
x

Fig. 97

!f.lustrat1ve example:

The sphere II.

Let us look at a sphere with an arbitrary radius r and let us introduce spherical coordinates:

X] = [
y z

[r Sine r Sine r cose

COS~]
Sin~

21f

1f

282

II

Here (S,tp) are the the frame vectors


-+

intrinsic

coordinates on the sphere and (x,y,z) are

extrinsic coordinates. At the pOint (S 0 ,tpo) we have the canonical

eSam

e tp with the extrinsic coordinates: Costpo


-+

...

.. aXl as
s

[, co.a,
-r SinSo

ax atp

lY

as
The

as az

r CosS o Sintpo

<P

lY atp
az alp

[-'

Sin6 0 Sin., r SinSo Costpo


0

intrinsic coordinates are by definition


and
e tp
-+

The frame vectors radial vec~r

s and are obviously orthogonal to the

e(?

and they are or-

thogonal to each other. We also want to construct a metric on the sphere. But here we may use the usual inner product

<vlu>

+ V 2 U 2 + V 3 U 3 associated with R 3. Thu;, for any


=
V1U

two tangent vectors we put:


(6.25)

vp

and ~

This is called the ents:

induced metric. Let us compute the metric coeffici-

[: ~:::: ~:::]
-r SinSo

g(jl(jl= [-r SinSoSintpoir SinSoCostpo; 0]

-r SinS OSinlAJ] r Sins~costpo

283

II

Therefore we find
(6.26) 1 g;J' (So ,tpo) = r2 [0

0 2 ] Sin So

Observe that the metric becomes singular at the north pole and the south pole,

gi'l

01

J pole

oj

which again tells US that the coordinate system breaks down at the poles! Clearly this example can be generalized to an arbitrary Euclidian manifold, which then becomes a Riemannian manifold, when we equip it with the induced metric. Let us emhasize however that the induced metric is not the only :oossible JT!etric on an Euclidian manifold. ~'l'e may choose the family of inner products completely arbitrary.

Exercise 6.5.3
Problem: Consider a torus with the outer radius a and the inner radius b. Use the angles S and tp as shown on the figure to parametrize the surface. Compute the metric coefficients of the induced metric in the (S,tp)-coordinates. y-axis
FIG.

100

Exercise 5.5.4
Problem: Consider the unit sphere Sll-l in the Euclidian space ~. We introduce spherical coordinates as indicated below. Show that the metric components, are given as shown below Xl = r SinSISinS2 ... SinSn-2CosSn-1 x2 x3 r SinSISine 2 .. cosenr SinSISinS2 ... SinSn-2SinSn-1 2

xn- l r SinelCosS 2
xn = r cose l

r 2Sin 2S 1

(6.27)

g, ,(r ,SI, .. ,Sn-l)=


IJ

r 2Sin 2e1SinS2

284

II

6.6

THE MINKOWSKI SPACE


manifold M which we may identify with space R4 , but forget about the extrinno importance in our example. We shall of our space-time. So each pOint P in our This event could be the absorption of happened at a specific position at a
MATHEMATICAL MODEL;

Consider a four-dimensional the four-dimensional Euclidian sic coordinates as they are of use this manifold M as a model manifold M represents an event. a photon by a particle. It all specific time: REAL WORLD AT t=t o (PHOTOGRAPHY)

PARTICLE A ABSORBS A PHOTON

! ~

PARTICLE B MOVES UNDER INFLUENCE OF FORCES

THE PHOTOGRAPHY CORRESPONDS TO THIS SLICE

FIG.

101

A particle moving in ordinary space is represented by a world line in our model. The world line is a smooth curve comprising all the space-time pOints occupied by the particle during its existence. Next we want to introduce coordinates in our manifold. To this purpose we shall use an inertial frame of reference, i.e. we imagine an observer equipped with a standard rod for the measurement of length and a standard clock for the measurement of time. Equipped with these instruments he can characterize each point in space by means of three Cartesian coordinates (X 1 ,X 2 ,X 3 ) and he can characterize any event by its position in space (X 1 ,X 2 ,X 3 ) and the time xo.

Hence any event is characterized by a set of four coordinates (xo ,Xl ,x 2 ,x 3 ) . In our mathematical model such an Observer corresponds to a specific coordinate system, ~: R4~M,covering the Whole space-time manifold.

+(x ,x ,x ,x )

What do we mean by an inertial frame of reference? Physically an inertial frame of reference is characterized by the following property:

285

II

Definition 11 An inertial frame of reference is a frame of reference where any free particle moves with constant velocity. The coordinate system corresponding to an inertial' frame of reference will be referred to as an inertial frame and we speak about inertial coordinates.
Consider the worldline of a free particle. In inertial coordinates it is parametrized in the following way, xo+

[~:l
x
3

Le.

xI'" alxo+bl

where al is the constant velocity of the particle. In the coordinatespace this is the parametrization of a straight line, cf. fig.103

4 R

[:;J -

x2 - a2 >'of b 2
a3

[!l] [~ljl
b3

FIG.

103

NOW,

suppose we have been given two inertial frames of reference.

Then we have two coordinate systems, (<P, , R4) and (,/>2, R4), covering the space time manifold. The transition function .is referred to as a

<P2'

which exchanges the

old coordinates (xO,x',X 2 ,X 3 ) with the new coordinates (yO,y',y2,y3)

Poincare transformation. What can we say about the

'::ransi tion function? Observe that the world line of a free particle Ls represented by a straight line in both coordinate systems.

FIG

104

~
+ bU.

cx

<P1/)/./ <P2
/ I
/

'{i

<P2l

;1\

Thus the function

y=<P2' (x)

maps straight lines onto straight lines

and we conclude that it is a linear function

/J.

AUSXS

286

II

To proceed we must invoke the following assumptions:

BASIC ASSUMPTIONS OF SPECIAL RELATIVITY:

The speed of light has the same value in aU directions in aU inertial frames of reference, i.e. the speed of light is independent of the observer. II AU inertial frames of refenence are equivalent.

Let us introduce matrix notation. Put

(6.28 )

o I o o

0 0
I

~l

We can then show

Theorem

;3

A Poincare transformation connecting two inertial frames is given through the formula

Yl=A~1 +
where
/11.29)

51
n

A satisfies
A+nA
===+ AnA

Proof: We have normalized our units so that the speed of light is one, c=l. Suppose P and Q are two pOints lying on the worldline of a free particle. Then the speed of this particle is given by
v

(XI_X I )'+ {x'-x' )2 + (X 3_X 3 )' Q P Q P Q P

points P and

If we introduce the notatipn ~XI=XIQ-XIP we therefore conclude that the Q lies on the worldline of a photon if and only if
-(xQ-X~)'+(xQ-X;)'+(X~-~)'+(X~_X;)2

i.e.

But the conditiOn for P and Q to lie on a worldline of a photon must be valid in any inertial frame. Therefore we conclude (6.30)
But here

~y 1=

YI Q - Yip

'!'hus (6.30) can be rearranged as

287

II

iff Hence the two symmetric matrices nand A+nA generates the same lightcone K, where

But then they must be proportional. Therefore we conclude that (6.31) In this way we have attached a proportionality factor A to the POincare transformation
~21

which was given by

YI =

~21 (XI)

= Ax/.+

51

The inverse POincare transformation ~12 is gonsequently given by =_1 XI = ~12(YI) = A A-1bl

y, -

Thus ~12 is characterized by the matrix A-I lation of eq. (6.31) gives us

A simple algebraic manipu-

i-~ = (A-I) +n (A-I)


so the proportionality factor corresponding to

~12

is }. As the two i-

nertial frames of references are equivalent we conclude

A=I

Le.

A= 1

But from (6.31) we also get (Det[A])2= A Consequently ).=+1. We have therefore shown and The last formula easily be converted to give =-1 An- I A+ = n But n
-I

n and therefore A also satisfies A n A+ = = n

This theorem motivates the definition

Definition 12 A matrix A with the proporty (6.32) A+n A = A A+ n is called a Lorentz-matrix. The Lorentz matrices constitute

a group called the

Lorentz group, which we denote 0(3,1). A homogenous transformation


YI = AXI

where

A is

a Lorentz matrix is called a Lorentz transformation.

288

II

Exercise; 6.6.1
Problem: (1) Show that the matrix n itself is a Lorentz matrix. What is the physical interpretation of the associated Lorentz transformation? (2) A Lorentz matrix which preserves the direction of time is called orthochronous. A Lorentz matrix which preserves the orientation of the spatial axes is called proper. Show that a Lorentz matrix is proper and orthochronous if and only if it has the properties,
AOo > 0 ; Det

A=

and that the matrices with both properties constitutes a subgroup of the Lorentz group (which is denoted SO (3,1) ). (3) Show that the spatial rotation group SO(3) can be considered a subgroup of the Lorentz group.

Exercise: 6.6.2
Problem: Consider the matrix:

Ae;p[i il!l
Show that it is a proper orthochronou~ Lor~ntz matrix. Consider the corresponding Lorentz transformation,y = AX,and show that it corresponds to a transformation between two observers, where the second observer moves with the velocity v along the x-axis of the first observer. This transformation is called the standard Lorentz transfo~ation.

We will now construct a metric on our s?ace-time manifold M. Let and be two tangentvectors at the same pOint P. Let a, and 5, p their coordinates relative to an inertial frame and consider the number ~;nb,

This is actually independent of the inertial frame. To see this we introduce another inertial frame. Let the corresponding Poincar~-transformation be given by Then the Jacobi matrix reduces to
= av a D21 = (~) = 'A S

ax

According to (6.l4)the new coordinates of ~p and up are given by A~, and AE, . Hence in the new inertial frame we comp.lte the number

(Aa,) +n (AE,)

is really independent of the and this shows that the inertial frame. Thus we may define: The inner product between two -+ ... tangent vectors up and vp is given by -+-+ =+== g(up'v p ) = al nbl

a; cJ~+nA) b number a;nEI


=

289

II

where at and 5, are the components of ~p and ~p relative to an arbitrary inertial


j'rame.

In this way we have d~fined a metric! We can easily read of the metric coefficients:

(Compare with

(6.2~.)

The~

are clearly smooth functions

o~

the coor-

dinates. Furthermore

n is

scnnmetric and regular and it has the eigen-

values -1, 1, 1, 1 so it is a nice Minkowski metric. The four dimensional manifold R' equipped with the above metric is called Minkowski space. Illustrative example: Spherical symmetry in the Minkowski space.
Very often we study physical systems having some kind of symmetry, and it is then advantageous to choose coordinates which reflect this symmetry. If the physical problem in question has spherical symmetry, you would naturally introduce spherical coordinates. The transition functions from inertial coordinates (xo,x,y,z) to spherical coordinates (t,r,8,t?) are given by: x
r

rSin8Cosq>

y
8

rSin8Sinq> ; z = rCos8

Ix 2 +y2+z2

The corresponding Jacobi matrices are given by:


(1) From inertial to spherical coordinates:

o
(6.34 )
=

o
!L
-.Jf.._ z__

o
z
r

aya D21=(~)=

0 0

X~

rZ..;:;-r::::zz

r2Yr~2

- r i y;t:z2

2_z2

v - X2:ty2

x 2+y2

(2) Frcm spherical to inertial coordinates:


0 0 0

(6.35) ])12=

dX

Sin8Cosq> Sin8Sinq> CosS

rCos8Cosq> rCos8Sihq> -rSinS

-rSin8Sinq> rSinSCosq>
0

0 0

We can then find the metric coefficients in spherical coordinates. Using the transformation rule.(6.23) we easily get

290
-1

II

(6.36)

" ,,:m,']

Observe that the part of the metric which is associated with the polar angles is nothing but the Euclidian metric associated with a sphere of radius r. (See (6.26) ). Observe also that the metric breaks down on the z-axis: 8 = O,n. That is just the old coordinate singularity we found on the sphere. So strictly speaking spherical coordinates do not cover the whole space time manifold! Now lpt us also look at a tange~t vector A (i.e. a four vector). Let us for a moment uSe inertial coordinates (A ,Ax,AY,Az ). Furthermore we assume that the vector part ,(AX ,AY,Az) ,is a radial vector, i.e. on the form A(Sin8Costp, SinSSintp, CosS) , where A is the spatial length of the space vector. This is the typical form of a four vector in problems with spherical symmetry. Let uS try to find the spherical coordinates of this four vector. Here we should use the transformation rule (6.8). Thus we get
FIG. 105

0 A,

AO A SinScostp A SinSSintp A CosS

A" A 0 0

=,

0 0 0

SinSCostp

SinSSintp

Cos S Costp - Cos 8 Sin (jl r r Costp Sintp - rSin 8 rSin8

CosS Sin S r 0

So, in spherical coordinates, we get the simple expression: At = A;Ar = A; AS = Atp = 0

Exercise: 6.6.3 Problem: Introduce cylindrical coordinates Write down the transition functions. Show that the Jacobi matrices are given by

(t,p,~,z)cf.figure

106:

(6.37)

D21

[~
[!

a
Costp _ Sintp
p

0 Sintp
~
p

~J
!J
x-axis

P(p,<P,z,

(6.38)

D12

0 0 costp -pSintp Sintp pCostp 0 0


0

y-axiS
FIG.

106

and that the metric is given by

(6.39)

gall

[-'0

p2

,]

--------

291

II

6.7

THE ACTION PRINCIPLE FOR ARELATIVISTIC PARTICLE

Now, consider the world line of a particle, which need not be free. Let it have the parametrization, XCX = XCX(A)

t
a

FIG. 107

Then the speed is given by the formula ds dt


which
~lies

ds dA dt dA

VidA

l,dX') 2+

(dxl) 2+ (dx3) 2 dA dA

<

dxO

that

dA

Thus

the tangent vector vp is always timel ike. ,'Ie therefore call the

->

world line of. an ordinary particle a timeUke curve. Now let us look a little closer at the .,orld line: Let P, and P 2 be two pOints on the "iOrld line corresponding to the parameter values A, and A2' Then .le can define the arc-length of the arc P,P2 by the follmling formula
r----------::-

(6.40)

A2 / dcxdxS ArcJ.ength=L v'-gas(X(A)) d~ dA dA

Observe that the arc-length is independent of the coordinate system chosen because the integrand is coordinate independent:

Observe also that the arc-length is independent of the A = A(S) with A, = A(sll and A2 = A(S2) we get

param~trization:

If we exchange the parameter A with a new parameter s, such that

292

II

dx dA dA

dA ds ds

Hence it is a reasonable definition. We "lill try to <Jive a physical interpretation of the arc-length. Let us look at a free particle first. As the particle is free we may choose an inertial
fra~e

of reference which follows the particle, i.e. the particle is at rest in this inertial frame. Consequently we can parametrize its world line as
X
1.

Observe that the parameter A represents the time measured on a standard clock which is at rest relative to the particle! In this particular simple coordinate system you get

Thus we have shown:

The arc-length corresponding to a free particle is the time interval measured on a standard clock which is at rest relative to the particle!
Then we look at an arbitrary particle moving in space. h'hen the world line is no longer strai<Jht the above argument does not apply. Let us however assume that the rate of a standard clock is unaffected by ac-

celerations.
Then we may approximate the curved world line with a world line which is broken straight: The broken straight world line corresponds to a standard clock which is piece by piece free. Therefore the above argmnent applies and the arc-length of the broken straight world line is the time interval measured on this standard clock! Going to the limit
lO~

we conclude:

293

II

The arc-length of a worldline is equal to the time measured on a standard clock following the particle. This will be referred to as the proper time T of the particle.
Returning to the world line of a 9article mowing in an arbitrary way we may now use the proper time T of the particle as a parameter. It is connected to the old parameter A via the formula
(6.41)

Differentiating this formula ,Ie get

i.e. in a symbolic notation we have

This formula is often "sC!Uared" whereby we get


(6.42)

dT

a B =-<JaBdx dx

and dT 2 is then referred to as the square of the line element or just the

line element.
Using the proper time T as a parameter we get the tangent vector cx (6.43) UCX = dx dT \V'hich is referred to as the four Velocity. The four velocity has a very simple property. If we square it we get -+ -+ dxcxdx S dxcxdx S [d A] 2 g(up1llp) = "bSdT dT = gaSdA dA [dT =-1 Thus the four velocity is a unit Vector pointing in the direction of time.

FIG.

110

294

II

Even in special relativity there is nothing sacred about inertial frames. We may introduce any coordinate system we like! We should, however, be aware of the frame, then a) The metric coefficients no longer reduce to the trivial ones, i.e
gCtS

follo~]ing

fact: If (q"U) is not an inertial

'*' nCtS

b) The world line of a jreeparticle is no longer parametrized as a straight line, i.e.


d 2x (f[2
Ct

'*'

/
FIG.

1~ I

11 1

XCt=XCt(T

As a free particle has a non-trivial acceleration in curvilinear coordinates we say that it experiences fictitious forces. I!e call the forces fictitious because they owe their existence to the choise of the coordinate system. If we return to the inertial frame the fictitious forces disappear. Let us try to find the equations of motion of a free particle expressed in arbItrary coordinates. It would be nice if we could derive these equations of motion from a Lagrangian principle, i.e. the worldline should extremize an action of the form

=f

A2

Ct dx Ct L(x idA) dA

Al

Observe that the Lagrangian depends on all four spacetime coordinates and their derivatives. In a relativistic formulation time and space coordinates must be on the same footing. In an Euclidian space we know that a straight line extremizes the arc-length. This suggest the following theorem:

Theorem 4 The relativistic lagrangian


Ct

(6.44)

L(x 'dA)-

azct

ree particle is given by azS -getS d)'" dA

or a

azct

so the corresponding action is pl'OpomonaZ to the proper time

295

II

(6.45 )

The equations of motion can be written on the form d 2 xv v dza dzS


(6.46)

JT2

aadT

dT

where rV aa ' t he so-called Christoffe l fie ld, is given by dg dg dO = ~g~v[ ~ + _~ _ ~ 1 (6.47) B 'Q( ...
dX dX dX

Proof:
The Lagrangian principle leads to the following Euler-Langrage ec:uations

!~~

:" [

3(dT)J

;:~] j~

0,1,2,3

The evaluation of these equations using (6.44) is strai<,!ht fon-lard but tedious so we leave it as an exercise:

Worked exercise 6. 7. 1 Problem: Show that the Euler-Lagrange equation corresponding to the action (6.45) reduces to

To conclude the proof ces to

~le

must show that the ec:uations of motion redu-

in an inertial frame. But in an inertial frame we know that the metric coefficients are constant, gaB

naB' so their devivatives vanishes.

If we compare this with the expression for the Lorentz force in

inertial coordinates,

m --a;r-2 =
you see that we may identify -mrv 0.6 with the field strengths of the fictitious

d 2x

forces.
Let us look at the electromagnetic force once more. lIe knmr that in terms of inertial coordinates it may be rewritten as
(1. 29)

= g~V [~_~]
dX~

ax

296

II

Thus

it is expressed as a combination of derivatives of the r~axwell

field Aa . For this reason ~le say that the Maxwell field act as poten-

tiara for the electromagnetic forces. Returning to the curvilinear coordinates we have seen that the Christoffel field may be expressed as a combination of derivatives of the metric coefficients: ag ago ag~o La~V[ ~a + ~~ ~~l (6.47) rv -W dX a - ~

as

"-

For this reason

~le

say that the metric coefficients act as potentials for the

fictitious forces.
Finally we notice that if we parametrize the world line by the frame time, i.e. put A = XO arranged as:
(6.48)

(=t) t2

, then the relativistic action can be re-

-m

J!I-~2

dt v l , this reduces in the

tl EOr a non-relativistic particle, i.e. when lowest order approximation to the non-relativistic action tz t2 S ~ -m J (1-~~2)dt J~m~2 dt + m(t2- t l)
tl tl

(except for an irrelevant constant) Consider an arbitrary manifold Ylith a metric, Ylhich can be a which extremizes the arc-length is called a
~ie

mannian metric or a 11inkowski metric. A curve connecting two points

geodEsic. In the above ana-

lysis we have made no special use of the fact that ,'Ie Ylere Ylorking in Binkowski space. Consequently geodesics on an arbitrary manifold are characterized by the geodEsic equation
(6.49)

d?

d2x

= _ ra

~V

dx~
ds

v dx ds

where s is the arc-length and

ra

~v

are the Christoffel fields. The

Christoffel fields are defined through the formula (6.47) But unfortunately this formula is very inconvenient for computational purposes.
~Jhen

you want to compute the geodesic equations it is therefore preferable to

extract them directly fran a variational principle. Of course you could use the

Euler-Lagrange equations coming from extremizing the arc-length itself

But this leads to very complicated computations due to the square root. So in practice one uses the following useful lemma

297

II

Lemma 3 Let y be a geodEsic:


(1)

Then y extremizes the fUnctional


I =

(6.50)

A2

l.; gOf',(X(A)) dX F

d:/).

ax S

d)"

Al
Fza>theY'l1lore, when XCt=XCt(A) extremizes the funational I, then Ct ax ax S is constant along the geodEsic, so that we can idEntify gCtS(x(A)) 7i."A F the parameter A with the arc-length s.

Proof:

Consider the Lagrangian Ct dxCt L(x iax-) = l.;g~V(X(A


dx~ dx v ax- ~

It leads to the Euler-Lagran0e equations

Le.

but they are identical to the sreodesic e0uations if we can show pro perty (2). This is done in the following way: Ct dxo Let XCt=XCt(A) extremize the functional I. Observe that L(x iax-) Ct dx is homogenous in the variable ax- of degree 2. Thus it satisfies the Euler relation
2L

dxCt We must show that L(XCt(A);ax-)iS constant alonCT XCt=XCt(A). Consequently we consider the total derivative: dL dA

USing the Euler-Lagrage equations this is rearranged as dL


2 d)..

where we have taken advantage of the Euler relation length s.

But this im-

plies immediately that ~~ = O,i.e. A must be proportional to the arc-

298

II

Observe that when you have found the geodesic equations you can directly extract the Christoffel fields as the coefficients on the rigth hand side. By an abuse of notation the above functional is often denoted as

Exercise: 8. 7 2 Problem: Consider the two-sphere 8 2 with the line element ( cf. (6.26) ), ds 2 = dS 2 + 8in2Sd~
(1) Determine the geodesic equations and show that the meridian ~ is a geodesic. (2) Compute the Christoffel field.
Probl~m:

= ~o

Worked exercise: 6.7.3 In Minkowski space the line element is given by dT2 = _ dt 2 + dr 2 + r2(de 2 + 8in28d~2) when it is expressed in terms of spherical coordinates, cf. (6.39). Compute the aSBociated ChriBtoffel field.

Okay, you might say: This is very nice. We can work in arbitrary coordinates if we are willing to pay the price of fictitious forces and non-trivial metric coefficients and all that, but why should we do it? After all, we have at our disposal a nice family of coordiante systems, the inertial frames, which gives a simple description of physiscs! There is however a subtle reason for working in arbitrary coordinates which was pointed out by Einstein. Fictitious and gravitational forces have a strange property in
co~on:

If you release a test particle in a gravitational field its acceleration is independent of its nass. Two test particles with different masses will follow each other in the gravitational field. This is known as Galilei's principle and it has been tested in modern tines with an extremely high preCision!
INERTIAL FRAME OF REFERENCE ROTATING FRAME OF REFERENCE

(THE PARTICLES ARE

FIG.

112

(THE PARTICLES APPEAR TO BE ACCELERATED)

BO'l'Ii AT REST)

299

II

But the fictitious forces have the same property: The acceleration of the test particle depends only on the field strength -

rVas '

and

the field strength itself only depends on the metric coefficients. Hence if we release free particles, then their accelerations relative to an arbitrary observer, say a rotating frame of reference, .rill be independent of their masses. Esc:>ecially if
~le

release two free partic-'

les close to each other, then they will stay close to each other! Okay, you might say, I accept that they have this property in common, but have they anything else in cOl'!1I!lon? E'or instance I can transform the fictitious forces completely away by choosing a suitable coordinate system. But you cannot transform away the gravitational field! It is true that I cannot transform away the gravitational field -completely. But I can transform it away locally! Consider a gravitational field, say that of the earth. Now let us release a small box so that it is freely falling in the gravitational field. Inside the box we put an observer with a standard rod and a standard clock. Thus he can assign coordinates to all the events inside his box (See fig.113) But now the miracle happens. If he releases particles in his box, they will be freely falling too. Therefore they will move with a constant velocity relative to his coordinate system, and they will act as free particles relative to his coordinate system! Consequently we have succeeded in transOBSERVER EINSTEIN BOX

lLJ
SURFACE OF THE EARTH

FIG. 113

COORDINATE SYSTEM IN BOX

forming the gravitational field away inside the box by choosing a suitable coordinate system!
When we say that we have transformed away the gravitational field inside the box, this is not the whole truth. Actually if we were very carefull we would find that if we released two test particles as shown on figure 114 they would slowly approach each other. Consequently there is still an attractive force between the two test particles. This force is extremely weak and is known as the Tidal forl;J2. The Tidal force owes its existence to the inhomogenities of the gravitational field, and it is a second order effect, so we may loosely say, that we can

300
transform away the gravitational field to first order inside the box! So there is a good reason why we might be interested in studying fictitious forces:

II

EINSTEIN BOX

It might teach us something about gr(]1)i tationa l forces!

SURFACE OF
THE EARTH

Fig. 114

+
6.8 CevICTORS
The next concept we want to introduce on our manifold M is a covector:

Definition 13
A c ovector
~p

at the point P is a Unear map

~p: Tp(MJ ....

Here we will use a notation introduced by Dirac. Given a tangent vector it by

vp

we denote

!vp>'

The symbol

I>

is

<w I~ > +P P

called a keto Similarly a covector ~p is denoted <~pl, where the symbol <I is called a bra. Tl}en the value of
~p

R
Fig. 115

at

vp

is written as follows

(6.52)

where the symbol < I > is called a bracket. Using this bracket the linearity of
~p

may be expressed in the follo1fling way

To justify the name a vector space If


~

covector, we must shm"l that the and

covectors form

are

covectors, i.e. linear maps from the


+p

tangent space, we define their sum

w + 4p as follows

301

II

(~+ 4p) (irp) = ~p (vp ) + 4P (vp )


(you can easily verify that the product of a co-vector
~ ~p

+ 40 is a linear map) am. s:imilarly we define

and a scalar A in the following way:

(A~p)

(vp )

= A~ (Vp )

Using the bracket notation you may rewrite these rules as: a) Sum:

<~p + 4plvp>

<~plvp>
<A~p

+ <4p!V P >

b) Product with a scalar :

....p > = Iv

A<~p

I.... vp>

Consequently the bracket is bilinear in both of its arguments. So the covectors at the point P form a vector space, ,1hich mathematicians call the dual Vector space, and which they denote by

Let us investigate the structure of this vector spa ce a little closer. Pick up a covector
~p.

Introduce a

R
Fig. 116

coordinate system around P. Ne can then define

the COOI-

dinates of the
(6.53)

co vector in the following way:

Using the coordinates of


....

~p

we can express the value of


....

~p ~t
L

a tan\ole

gent vector vp in a simple way. If vp has the coordinates a , (6.54) Observe, that an expression like w.a
.... L

find

is coordinate independent, be-

cause <~plvp> is defined in a purely ~eometrical way without any reference to the coordinate system used on M. This is a general featza>e of

expressions where We sum OVer a repeated index.


Now let
~p

..

have the coordinates wi' and 4P have the coordinates

Xi. Then

~p

+ 4p gets the coordinates

<w + 4P y Ie.> .... L In a similar way


A~p

<wple.> + <y_le .. > " w. + .... L 4~ L

gets the coordinates

Thus

we see that once ,'Ie have introduced coordinates we may identify

* the dual space Tp(M) with the vector space Rn in the sense that the

302
covector manifold
'.,;e
~~

II
~

wn is represented by its coordinates (wl""'w

we see that Tp (M) is a n-dinensional vector space. 'lb each point p on our have therefore now introduced two n-dimensional vectorspaces

Especially

vp had a simple geometrical interpretation. It could be regarded as a velocity vector corresponding to a smooth curve A passing through P. Now we want to give a similar, geometrical interpretation of a covector. Consider a diffunction,f:M~R,

....

When discussing tangent vectors we noticed that a tangent vector

ferentiable

defined in a neighbourhood of
P and let A be a smooth curve

f(A(t

passing through P (see fig.ll7). Then t


~

f(A(t

is an ordina-

---=+-----+~

t-axis

Fig. 117

ry function fran

R into R merely describing the variation of f along A.

Being an ordinary function we can differentiate it at t df (A (t dt It = 0

=0

Let us investigate this number a little closer. It represents the rate of change of f in the direction of A. If we introduce coordinates (xl, ... x n ) around P we can represent f by ap ordinary Euclidian funcn tion, y = f (Xl, ... x ), and \\Ie can furthermore parametrize the curve A as xi xi(t). Then the tangent vector

~P has the intrinsic coordinates

ai

dx

(it

It

USing these parametrizations we now obtain


(6.55)

df(A (t dt

It=O

= dt f(x (t
d

It=o

-.a
()X~

()f

This formula shows that the number dt f(A(t tangent vector

vp

depends only on the

and not on the particular curve generating it. It

also shows that this number depends linearly on the tangent vector Ne therefore define a linear map on the tan<Jent space Tp(M) by
-+

vp.

vp

df(A(t dt

where A is any smooth curve generating tangent space Tp (M) is nothing but a

vp.

But a linear map from the covector .Till

covector: This

be denoted df , and we refer to this expression as the exterior deriva-

tive of f.

303 We have thus shown


(6.56)

II

<df

d I.... v p > = dt

f(

A(t)) It

af a i axi

If you compare this with (6.54) you can immediately read off the coordinates of df
(6.57)

. af ., -af , ... ,af df.( -) ax~ ax' ax n is nothing but a generalization of the gradient vectors

But then df

you know from elementary vector analysis in Euclidean spaces! Using gradient vectors, we may analyse the structure of the dual

* space Tp(M)a little closer. Let ?obe a point on M, and let us intron duce coordinates (xl, ... ,x ) around Po' Then we have constructed a
frame for the ordinary tangent space T Po (M) in the fOllowing 'lay: see fig. 118, First we introduced the coordinate lines on M,

AI(t) = cp(t,O, .. ,O); ... ;An(t) = cp(O, .. ,O,t) and then the coordinate lines generated the canonical frame vectors~

dAI

It=o

. .... e = ~ , ... , n dt It=o

Now we want to do something similar. First we observe that we can define a smooth function on M in the following ''lay: Q(xl, ,xn)~ Xl This function is referred to as (O,t)

the first coordinate function,


and for simplicity we ''Jill denote it by Xl. Thus Xl (0) = X6 .lhere 0 has the coordinates (xg, ... ,x~). In a similar Iolay we can define the xi, i=2, . ,n "lhere xi (Q) = xi . 0 other coordinate functions covectors

Being smooth functions these coordinate functions x~ vlill generate

dx i at the point Po' These

covectors have particularly

simple coordinates, e.g. dXI gets the coordinates axl axl axl dXI: (--, - - , .. ,--) = (1,0, ... ,0) n axl ax 2 ax and similarly

dx':
Therefore the

1, , 0) ; :;

n flx : '(

,,... ,

1)

covectors

dzl; ... ; dzn serve as the canonical frame vectors in


~P

the dual space T;rM) .


If an arbitrary
(6.58)

covector-

has the coordinates Wi' ."Ie noW get

~p

WI dxl+"'+Wn dxn

304

II

This formula is in particular valid for the gradient vector df Therefore


\ole

conclude

(6.59)

df

Here you see something strange! This formula looks exactly like the well-known formula for the differential: df ... df di i dx But in modern geometry the idea of infinitesimal increments have been burried becaUSe of the logical difficulties involved in their use. The symbol a is now interpreted as an exterior derivative and the miracle happens: Many of the formulas involving the exterior derivative look exactly as the old formulas inVOlving differentials. So in modern geometry we work with the same formulas, they have just been given a new interpretation.

Finally we may investigate how the coordinates of a covector ~p are changed if we exchange the coordinate system. Let the old coordinates be \~Ii and the ne'lT coordinate:, be (6.12)we then get
I~j = <~p, &Ij > = <~p, tfd. dyj>

I~

I~

dX

w, --, (ilL dV J

axi

So the coordinates of a co-vector are transformed according to the rule (6.60)

w. (2)J

Thus the coordinates transform their name: covectors!

covariantly, and this justifies

We can also easily find out how the canonical frame vectors transform. If we introduce two coord~nate systems, then the old coordinates (xl, .. ,x n ) are connected 'to the nelol coordinates (yl, .. ,yn) via the transition functions

yl yn But here yi

yl (Xl, ,x n ) yn(x1, ... ,xn)

yi(x1, . ,X n ) is nothing but the new coordinate func-

tions expressed in terms of the old coordinates! A simple application of (6.59) then gives: (6.61)

305

II

Up to this point we have been working with a manifold without a metric. At eact point of the manifold "le have learned how to construct two n-di-

* mensional vector spaces Tp(M) and Tp(M). Let us now assume that we have also been given a metric g on the manifold M. Then we can do something strange: We can fuse the two vector spaces Tp(M) identification procedure where each tangent vector \-lith an unique covector
~p

* and Tp(M) tois identified

gether so that they become indistinguishable! I.e. we may construct an

Vp

in a purely geometrical nanner.

To see how this can be done, let us choose a tangent vector Then ,'Ie construct a linear map, Tp (M)~ R ,in the following way

Vp.

But a linear map, Tp (M)

~R,

is nothing but a

covector, and therefore our

tangent vector vp has generated a


g(vp ; .)

covector ~/hich will be denoted as

Let us introduce a coordinate system around P. Let have the coori dinates a . What are the coordinates of the corresponding co-vector
g(Vpi')? They are easily found. Denoting them by a i . k . (6.62) a i = g(vp,e i ) = gjkaJoi = gjiaJ

Vp

we get

i Thus the tangent vector with coordinates a generates the covector j with coordinates a. = g .. a . 1. 1.J. The map, I:a1.~a. = g .. a J , is a vector space isomorphism between * 1. 1.J Tp(M) and Tp(M) because g.. is a regular matrix. Therefore we may use 1.J * this map to identify Tp (M) with Tp (M). This is called the canonical iden-

tification! Exercise: 8. 8. 1 Problem: Let wp be covector characterized by the coordinates w. Show that the corresp~nding tangent vector is characterized by the comp6nents wi where
( 6 6 3)

w =g

ki

wi

'

It should be emphasized that tangent vectors and covectors were born in different vector spaces with different geometrical properties. If there is a metric g on the manifold, we may, however, identify them. Hence you will often see that when there is a metric, one simply forgets the dual space, i.e. the covectors. But then you have only tangent vectors at your disposal. Given a tangent vector, you can now characterize it by two sets of numbers:
a)

You can

decomp~se

....

;p along the canonical

frame vectors e , 1.e.

Fig. 119

306
i .... a e.
1

II

where the numbers a b)


a
= g(vp,e ) .

are referred to as the

contp~-variant

components.

You can use the numbers

........

They are referred to as the covariant c~onents. If~. is a unit vector, then a i is simply the orthogonal projection of vp onto ~i! 1

Due to the rules


(6.64)

and

vIe often say that the metric ct:Jefficients are used to "paise and lowep indices"!

Illustrative example:

The gauge potential outside a solenoid.

Let uS consider the Minkowski space M. Here we have a metric g, so we do not hacovectors. A vector field A could for instance represent the gauge potential in electromagnetism. We have previously computed the components of the gauge potential outside a solenoid in an inertial frame (~ection 1.4).
ve to distinquish between tangentvectors and

Boa [0, _ --y---2-

The spatial part was circulating around the z-axis. This makes it natural to introduce cylindrical coordinates (t,P.~,z). After all there is nothing sacred about inertial coordinates in special relativity! In exercise 6.6.3 we have computed the Jacobi matrices for the transition functions and the metric coefficients. We are therefore ready to find the contravariant components of the potential in cylindrical coordinates. Using formula (6.8) we get
0 0 0 0 0 0 0

Ao.=B oa
121

Cos~

Sin<p
P

Sin~

P
Cos~

0 0

Sin~ Cos~

Boa 2

1
;J2

but as we know the metric coeffi~ients (6.39) we can easily find the covariant components in cylindrical coordinates. Using that A =g
0. 0.8

pf>

i.e.

A cov.
2

GAcontrav.

we then obtain At= We can also

o, A p= 0 ; A ~

Boa A 2-' z = 0

A using

cylindrical coordinates

307

II

Thus A is a space-liNe vector, and its magnitude falls off like ~. If <p is a scalar field in the Minkowski space M, we may gener~te a covector field by taking the exterior derivative d<p. This is the gradient field corresponding to <p, and in terms of coordinates it is represented by d <p. Observe, that while d~<P consists of the ordinary derivatives ~

H=l<L
~ dX~

then the contravatiant components d~<p has a more complicated structure involving the metric coefficients

d~<P
Thus

= gUV a <p =
V

g~V

l<L
axV

in spherical coordinates, for instance, you get (compare section 6.6)

Returning to the potential A we can now write down a gauge transformation in a coordinate free manner as follows

(6.65)
where X is an arbitrary scalar field. (If you write it out in coordinates, you get

which coincides with a gauge transformation in inertial coordinates. Therefore it is valid in an arbitrary coordinate system). To see how (6.65) works, let us consider the pot~ntial outside a solenoid, and let us work in cylindrical coordinates: (At;>Ap>A(?>Az) Boa = -22

(0,0,1,0)
~

As the scalar function we choose the coordinate function -Boa 2 -2i.e.

multiplied by the constant

X(t,p,~,z) = -2- ~

-Boa 2

Then the "differentiel dx gets the coordinates

X (X dt ' dP
and A+

, dip , dZ

X)

Boa -2- (0,0,1,0)

dx.

the coordinates Boa


2
2
2 Boa ~ (0 0-1 0) + -2(0,0,1,0) = 0 ",

This does not mean that we have managed to gauge away A completely. The coordinate function ~ is not defined throughout the whole of spacetime. We must exclude a halfplane bounded by the x-axis, where ~ makes a jump. Hence all we have shown, is that if we cut away a halfplane bounded by the z-axis, then we can gauge away A in the remaening part of spacetime. So we have arrived at the same conClusion as we did the first time (Section 1.4). At the same time we have lerned an important lesson: Every time you use a cyclic coordinate (i.e. a periodic coordinate) there is a singularity associated with this coordinate, and the corresponding coordinate jUnc-

tion does not dEfine a smooth scalar field on the whole of our manifold.

308

II

6.9

TENSORS

By now yOU should begin to have a feeling of how to construct the geometrical concepts associated with a manifold. We consider linear maps to be the natural geometric concept associated with the tangentspace. Metrics were constructed as bilinear maps on the tangentspace, and covectors were constructed as linear r'!e can now generalize the concepts of a metric and a covector. Consider a map Fp ( ' ; " ' ; ' ) with k arguments. The arguments are all tangent vectors
ma~s

on the tangentspace.

vJl~

.. 'Vp(k),

and Fp

is supposed to
.... (1) .... (k)

) map such a k-tuple (v p ; . ;v p -+ (1) .... (k) Lnto a real number Fp(vp"";v p ). The set of all k-tuples (V~l); . ;vp (k) ) is denoted
,Tp(M)X. "xTp(M),

Fig. 120

v
and thus Fp

k factors

is defined as a map

Fp : Tp(M)x .. xTp(M)~ R To be of any interest .re will, of course, assume that Fp is linear in each of its arguments
Fp(v~
-+

(11

, ;AU p + ~wp; .. ;vp

....

....

....

(k)

)
( ....

=
(0. . .... (k) ) vp , ... ...... ,wp, ... ,v p

AF p ( .... ....... .... (k) ) + v (.1). , ,up, .. ,v p

~'p

I:

l'1e express this property of Fp by saying that it is a multilinear map. This is the kind of map we will be interested in and we ,o[ill give it a special name:

De fini tion 1 4 A cotensor Fp is a multilinear map

Fp : TiM)x xTp(M) ~ R

If Fp has k arguments we say that Fp is a co tensor of rank aU c<i;ensors of rank k will be denoted T(O.k) (M).

k. and the set of

Clearly a

co vector

wp

is nothing but a

cotensor of rank 1 and a cotensor of rank 2.

metric gp is a symmetric, non-degenerate

309

II

The next thing we must investigate is the possibility of introducing "coordinates" to facilitate computations Il7ith show' you that these cotensors (and to cotensors actually correspond to the tensor con-

cept you are used to in ordinary physics.) For simplicity we will illustrate the machinery with co-tensors of rank 3: Let us start by introducing coordinates (x1" T?(M) we will refer to the numbers
(6.66)

n .. ,x ) around the

point P in M. If Sp is a co-tensor associated with the tangent space

Sijk

= Sp(ei:ej:ek )
n

as the components of Sp with respect to the coordinates (x1, .. "x ) Let


i
~ ~ ~ up,vp,w

a ,b,c

respectively. We can then easily express the value of


~

b e tree h ar b' ~trary tangent vectors with the coordinates

....

Sp(~p;Vp;;p) in terms of the components of Sp and the coordinates of


~

up,v p and wp
~ ~ ~ _ i~ . j~ . k~ Sp(upiVpiWp) - Sp(a ei,b ej,c e k )

_ i j k ~.~.~ _ i j k - abc Sp(ei,ej,e k ) - Sijka b c

Observe, that allthough the specific components S'J'k and coordinates i j k a ,b,c depend strongly on the coordinate system, the number i j k Sijka b c is independent of the coordinate system!This is an example of the iIrp::>rtant rule: Whenever we contract all upper indices and lower indices, is the resut-

ting number an invariant, i.e. it is independent of the coordinate system employed.


Let us check how the components of Sp transform under an exchange from x-coordinate" "bc (<r~o

y-coordinates,

Sp(e ;~h;e ) \2,a 12,... l21==

=
dx i

~ dX i ~ dx j , ~ Clx k Sp(e,--; (1)1, Clya l~b dyb' (~ClyC


Clx j dX k k i j S dx dx Clx U1ijkdya dyb Clye

Sp(e'ie,;e~)--:-E --UI~ CllJ Ill"" Clya dy dyc

where we have used (6.12) and the multilinearity of Sp' Thus you see that the components transform as the name
~-tensor)

covariant

co~ponents(which

justifies

(6.67)

(~abc

310 l'le have various possibilities for


mani~ulating

II

tensors:

1. the sum of two co-tensors: If Sp and Tp are cotensors of the same rank k, then their sum, Sp + Tp ,is defined as the
follo~1ing
-+

cotensor of the same rank k


-t

-+ -+ ~ (Sp + T p) (UI;"';u k ) = Sp(ill;"';u k ) + Tp(u1;"';U k )

and the sum is characterized through the components


S.

1.1

1.k

+ T ..
1.1

1.k

2. Product of a cotensor and a scalar: If Sp is a cotensor of rank k, and A is a scalar, then we can define a new cotensor of rank k, ASp' through the formula (ASp)

(ri I;' .. ;rik )

= ASp (ri 1 i .. ;ri )

The product of a cotensor and a scalar is characterized by the CCIIIpOnents AS ... , 1.1 1.k 3. Tensor product: If Sp and Tp are cotensors of rank sand t, then their tensor procotensor of rank s + t
)

duct,SpSTp,is defined as the following (Spit T,,)(ri ; ... ;ri


1 S

iV 1 i ... ;Vt

The tensor product has the components

4. Contraction:

If Tp is a cotensor of rank k, and are tangent vectors, III l;cl then their contraction is defined as the follo~ling cotensor of rank (k-t)

v , ... ,v,

T" (v(1) ;

-+

iV t';;";;')

-+

\9.

and the contraction is characterized through the components Til .. i~, i H 1 . . . . . i all) .. ~ k It is important to observe that the
i
l

f
cotensor space Tp (O,k) (M)

actually carries a linear structure, i.e. it is a vectorspace. Once we have the tensor product at our disposal, we can discuss how to build up cotensors of arbitrary rank systematically using the covectors as building bricks

311

II

(0,1) l'Je have previously studied the space of covectors T*(M)=T p p and especially we found the canonical frame vectors dx 1 ; idx n These are the co~ctors which we shall use as the basic units!

(M)

Consider first an arbitrary covector Wp. If it has the components wi we can decompose it in the following way
(6.58)
W"

w.dx ~

Next we consider a cotensor Fp of rank 2! Let it have the components Fkt . Then we can construct another cotensor of rank 2 Fijdx fldx
~'Jhat

are the components of this


F ij d x
i
~

cotensor?

dj-o ... ) x (ek;e t

has the components Fkt too, and we conclude that are identical cotensors, i.e.
(6.68)

This is an important result: Any

cotensor of rank 2 can be expres-

sed as a linear combination of the cotensors: dx 1 fldx 1 ; dx 2 fldx 2 ; . ; dx n 8dx n j Therefore the co tensors dxifldx generate the whole tensor spa ce Tp (0,2) (M), so you see that they act as canonical frame tensors. Now you can guess the rest of the story! Let Tp be a cotensor of rank 3 with the components T ijk , we can then decompose it in the following way
(6.69)
I

etc!

Consider the metric g. It is a cotensor of rank 2. If we choose a coordinate Gystem, the~all cotensors of rank 2 can be expanded in terms of the basic cot ensors dxafldx Especially we may expand the metric tensor a S (A ) g = gaS dx fldx This formula is interesting, because formally it looks very much like the square of the line element
(B)

But (A) has an exact meaning: It states that two well-defined cotensors are identical, while (B) only has a symbolic meaning. It is a mnemotecnic rule of how to

312
write down the arC length of a curve

II

dA
(or even worse: dt is an infinitesimal arc length, and dt 2 is the square of an infinitesimal quantity, whatever that is). Hence you see again that the formulas of modern differential geometry very often resemble formulas from old days involving infinitesimal quantities and other mysterious things. Using the notation g = gusdXUaaxS we can write the metric on a 2-sphere as

(6.26)
and the metric on the Minkowski space expressed in spherical coordinates as

(6.36)

= _ dtadt

+ dr8dr + r" ( deade + Sin 2 edq:sd q

Let us look at a point P on a manifold Mn. To this point P we have attached the tangents pace Tp(M), and the tanqent space has been used to generate all the attached the cotensor spaces. But to the point P we have also

* cotanqent space Tp(M), and that is a II-dimensional veccotensors in the fol-

torspace with a structure very similar to that of the tangent space. T'le can therefore generalize the 1iscussion of lowing way:

De fini tion 1 5 A mixed tensor Fp of type (k,~) is a multilinear map

* * Fp : Tp(M)x )(Tp(M) x TiMJx . ){Tp(M)'" R


!.'~

k-factors

i-factors

A mixed tensor of type (k,l) has rank k+l, and the space of all mixed tensors of type (k,l) is denoted T~k,2) (M).

A mixed tensor of type (k,O) is siI!\ply called a tensor of rank k Let us focus the attention on the tansors of rank 1. A tensor Ap of rank 1 is characterized by its components Ai Ap( ax i ), and these components transform contra variantly. Thus there seeI!\S to be an intimate connection between tensors of rank 1 and tangent vectors. In fact we can identify theI!\: Let Vp be a tangentvector. Then the map
~p

<w

"'p

Ivp >

313
generates a unique tensor of rank 1, which we also denote vp'

II

* is a linear map Tp(M) ~ R, since the bracket is bilinear. Therefore vp

...

Exercise 6.9.1

Problem: Conslder a tangent vector vp ....Show that the components of vp as a tangent vector and the components of vp as a tensor of rank 1 are iaentical.

...

...

Ylhen we have tensors of an arbitrary t:'pe at our disposal w" can ge-

neralize the contractions. Let, for instance, Tp be a tensor of type

(1,2). It is then characterized by its components

n with respect to some coordinates (x1, ... ,x ). Here the index a transforms contravariantly and the indices band c transform covariantly. But then we may contract the indices a and b obtalltingthe quantity:

As we sum over the index a, this quantity is characterized by only one index

The important point is now that the quantity Sc transforms covariantly! This follows from the computation j a S Ta av Ti ax axk (Toc (21 ac axi C!l jk aya a~/

a j where we have used that (~) and (~) are reciprocal matrices. But axl. aya if Sc transforms covariantly, we may regard it as the components of a
cotencor Sp' We say that Sp is generated from Tp by contraction in the first two variables. This is obviously a general rule: Whenever you

contract an upper index with a Zower index, the resuZting quantity


transforms as the components of a tensor. (where the degree of course is lowered by two!). The only trouble with contractions is that they are almost impossible to write in a coordinate free manner! As contractions play an important role in the applications, we will therefore write down many equations involving tensors using component notation.

314

II

In the following table l'/e have summarized the most important properties of mixed tensors:

A mixed tensor Tp of type termined by its components


Components

(k,~)

is completely de'

l .. i

JI'''~

. = T (dx
P

; ... ; dX1k;~. ; ... ;~. )


JI
J~

with respect to a coordinate system. The indices i l , ... ,i transform contr8:variantly , and the indices k jl, .. ,j~ transform covariantly.

Linear structure

If Sp andTp are mixed tensors of the same type and A 1S a real scalar, then you can form the mixed tensors Sp+Tp and ATp of the same type (k,~). Furthermore these mixed tensors are characterized by the components
(k,~),

(6.71)

Tensor product
(6.72)

If Sp and Tp are mixed tensors Of type (kl'~I) and (k 2 ,P-2), you can form the mixed tensor S1' ~ Tp of type (k l + k2 '~l +P-2). The tensor product is characterized by the components
S

il .. i k

il ... i
1. T
JI"Jt 1

jl"j~2

Contraction

If Sp is a mixed tensor of type (k,~), you can form a mixed tensor Tp of type (k-l ,~-1) by contracting a contravariant and a covariant index. The contraction is characterized by the components

(6.73)

In what follows we are going to deal a lot with tensor fieZds.

To con-

struct a tensor field T of rank 3 we attach to each point P in our man nifold a tensor Tp of rank 3! Let us introduce coordinates (xl, ... ,x ) n on M. Then the tensor at the point P(xl, .. ,x ) is characterized by its components T abc (x 1 , ... ,x n )

315 \7e say that the tensor field T is a smooth field if the components

II

Tabc(X', ... ,xn) are smooth functions of the coordinates. If nothing else is stated we will always assume the tensor fields to be smooth.

Illustrative example:

The unit-tensor field.

Let M be an arbitrary manifold. Then we construct a mixed tensor of type (1,1) in the following way. At each point P we consider the bilinear map

The corresponding components are


i .... To( dx ie.) J '" < ax i

" l e.> J

i.e. the Kronecker-delta! It is a remarkable fact that the components of this tensor have the same values inall coordinate systems. As the components are constant throughout the manifold, they obviously depend smoothly upon the underlying coordinates! We therefore conclude:

The Kronecker-delta oi. are the components of a smooth tensor field on M, called the unit tensor field of type (1,1).
J

Exercise 6.9.2
Problem: Show that the Christoffel fields
r
V

as

..,g

l.

v].! [a/>:].l(:t ax

--S+-a --ax ax].!

a/>:S].!

g a aS ]

are not the components of a mixed tensor field. (Hint: Show that they do not transform homogenously under a coordinate transformation. )

Exercise 6.9.3
n Problem: Consider a point P on our manifold. To each coordinate system (x', ... ,x ) around P we attach a quantity TlJk with two upper indices and one lower index. Apriori the upper indices need not transform contravariantly, and the lower index need not transform covariantly. Show the following:

If the quantity U~ given.by


Q, " RU k - T1-J S

ij

are the components of a mixed tensor of type (1,1) .Ulhenever S1.ij are the eomponents of a mixed tensor of type (l j 2), then p1-J k are the components of a mixed tensor of type (2,1)

316

II

The method outlined in exercise 6.9.3 is very useful when you want to show that a given quantity transforms like a tensor. Clearly it can be generalized to mixed tensors of arbitrary type. Suppose now that we have attached a

metria g to our manifold M. Then

we have previously shown (Section 6.8) how to identify tangent vectors and covectors using the bijective linear map

I:T;(M) - Tp(M)

generated by the metric. If we write it out in components it is given by

Exercise 6.9.4
Problem: Show that I(

axk )

is characterized by the components gki

11/e can now in a similar way identify all tensor spaces of the same rank. For simplicity we sketch the idea using tensors of rank 2. Let T be a cotensor of rank 2 characterized by the covariant components T ij . Then we identify T with the following tensor of rank 2: II (T)

(~;4)d~f.T(I(~) ;I<';~

The tensor I 1 (T) is characterized by a set of contra variant components which we denote T ij They are given by

Tk~ = II

(1)

(<h/;ax~)=T (I(dx k ); I(ax~) )=Tijgkig~j.

In this way we have clearly constructed a bijective linear map II: Tp (0,2) (M)
_

Ip (2,0)

(M)

tie can also identify T with the following mixed tensor of type (1,1)

Here I 2 (T)is characterized by the components

In this way we

hav~

generated a bijective linear map


....

12: T(O,2) (N)

T(l,l)

(M)

3"1:7

II

On a manifold M with a metric g we have therefore a aa:noniaaZ identifiaa-

tion of the tensor spaces 0f rank 2:


Tp (0,2) (M)

Tp (1,1) (M)

g \

(2, 0) (M)

Obviously these results generalize immediately to all tensors of higher ranks. If there is a metric g on our manifold M is it customary not to distinguish bet,'leen the various types of tensors of the same rank. one simply speaks of tensors and a given tensor T is then represented by various types of components. For instance a tensor T of rank 2 is represented by the following four types of components T ij ,T ij ,T i j ,T ij

and these components are related to each other through the formulas: Ti. J
(6.75)

= =

ik

kj

T.

= =

kj

Tik T k

ij

= =

:iJc

~j

Tk~

Tij

k .. gik T j T 1J

gkjT i

Tij

gikg~jT

k~

This is known as the art of raising and lowering indices!

Illustrative example:

The metria aotensor.


cotensor itself. Thus we want to components of g.

Let us apply this to the metric Raising one of the indices g j


i
\"le

find the mixed components of g and the contravariant find


ik i

=g

gkj

=a .

Consequently the mixed ca:nponents are nothing but the Kronecker-delta! As we have already used the notation gij to denote the reciprocal matrix we will for a moment use gij to denote the contravariant nents of g. Raising both of the indices we find: gij compo-

= gikg2jgk~ =

gikojk

gij contravariant components of

So we may relax: gij actually denotes the components


(6.76)

g too. Hence the metric g is characterized by each of the following

and

where gij and gij are reciprocal matrices and Oij is the unit matrix.

318

II

6.10

TENSOR FIELDS IN PHYSICS

To see how a physical quantity can be represented by a tensor, we will consider the electromagnetic field. It is characterized by the field strengths The components matrix
0

E and B
~,

measured in an inertial frame of reference.

... ,Bz are collected into a skew symmetric square


-~ 0
-Bz

-~
Bz

-E

Fas

~
E

-By

~
z

By

-Ex

Ex
0

If we exchange the inertial frame of reference

5 1 ,characterized by iner-

tial coordinates xa,With the inertial frame of reference 5 z ,characterized by inertial coordinates yS, then it is an experimental fact that the electromagnetic field strengths transform according to the rule (6.77) wherelA

VJa S

is the reciprocal Lorentz matrix, i.e.

This allows us to define a we define (6.78 )

cotensorfield F of rank 2 on Hinkowski spa-

ce. Let P be a space-time point, and

ITp,vp two tangent vectors. Then

where F S,UU,v S are the comP9nents relative to an arbitrary inertial a . frame. That the number on the rigth hand side is independent of the inertial frame follows immediately from the transformation rule (6.77). Thus in our mathematical model the electromagnetic field is represented by a skew symmetric cotensor field F of rank 2. Of course, we do not have to restrict ourselves to inertial frames. As a specific example let us consider a pure monopole field. Let us assume that the monopole is at rest at the origin of the inertial frame of reference

SI. We want to find the components of the monopole


(t,r,e,~).

field in terms of spherical coordinates

In the inertial fra-

me the electric field strength E vanishes and

319

II

....

....2......
4rr

....

r r3

where g is the strength of the monopole. Therefore

0
F
(1)06

0 0

0
~

=--.!L
4rrr2

0 0 0

x..
r

z r y r

x
r
0

x
r

Using the transformation rule (6.67) we find

i.e.

The Jacobi matrices were computed in section 6.6.Inserting (6.35) we get

0
(6.79)

0 0

F
121

J.4rr

0 0 0

0 0

0 0

Sine
0

0 -Sine

Thus only the F e(? -term is non-vanishing in this particularly simple coordinate system. The corresponding contravariant components become

i.e.

T'Te have already determined the metric coefficients in spherical coor-

dinates, see (6.36):

[~

0 0

0 0 0

-1 J..e.G
. =-1

0
0

0
0
2

1
0

r2
0

1
0
0

0 0

1
'1:

r 2 Sin 2 e

1 r 2 Sin 2 a

and from this we obtain the contravariant

components of

F:

320

II

0 0
0

0 0

0 0
~

0 ( 6.80)

paS
(2)

--.:l..
411

0 0

0
~

r"Sin8
0

r"Sin8

Observe, that we cannot find the "strength" of the field just by observing the covariant (or the contra variant) components of F. To do that you must form an invariant quantity. If you have a tensor A) The trace of T T of rank 2 you can form the following two invariants: a a

B) The "square" of T: ~TasTaS But the electromagnetic field tensor F is skew symmetric, i.e.

F(~p,rtp)=-F(Up'~p),so the trace vanishes automatically. Concerning the


square, you get:

Therefore the "strength" of the field depends only on the radial coordinate r, and it varies as !2!
r

Illustrative example: The four-momentW7l and the energy-momentum tensor'.


As another example of a tensor in Minkowski space we consider the energy-momentum tensor T . This is a s~etric tensor of rank 2. In an inertial frame the contra variant components T(l have the wellknown interpretation:

( 1.37 )

ae T

;:!e~si:tz _I:,. ___c;:r.::eEt_

Energy:

Energy]

~ Mome~tum:
[ densl.ty ,

Momentum currents

For a sourceless electromagnetic field F the energy-momentum tensor can only depend on F i.e T = T(F).In an inertial frame we have found the following relation
(1.41 )

As naB are the contra variant components of the metric g we conclude that the general relation is

II

(Since the left and the right hand sides are components of tensors which coincide in an inertial frame they must be identical).

Tp(M) and the cotangent space ~(M). You may find it strange to represent a phys1cal quantity by a multilinear map, so let us look at some examples:

By definition the tensors are constructed as multilinear maps on the tangent space

A) Let us lssume that we are observing a particle moving in spacetime with a fourmomentum P. The four--momentum was introduced as a tangent vector along the worldline
pa

= m dx dT

but we can identifY it with a co vector P i.e. a terize it by its covariant components Pa

cotensor of rank 1, and charac-

R p

Fig. 121

As observers we are ourselves moving through space-time along a certain world line ~ith a certain fo~r-velocity As P is a linear map, Tp(M)~ R, it will map into a real number <plu> . What is the physical interpretation of this number? Let us choose an inertial frame where we are momentarily at rest With respect to this coordinate system, P is described by the components Pa where - Po is

U.

the energy which the observers will measure, because the observers are at rest relative to this inertial frame. On the other hand, the four-velocity will have the coordinates ~ = (1,0,0,0) with respect to this inertial frame. We therefore get:

<P,D> = P ~ = Po

So we conclude that-<PIU> is the energy of a particle with four-momentum P measured by an observer with fOunJelocity U.

B) Consider the energy-momentum tensor T of a certain field configuration. Let us assume that we are observing this field configuration, and that we are moving with a certain four-velocity As T is a bilinear map,

U.

Tp(M) x Tp(M)

it will map the pair of tangent vectors (U,U) into the real number T(U,U). To find the physical interpretation of this number, we introduce again an inertial frame, where we ar.e momentarily at rest. With respect to this frame T is characterized by the covariant components ~B where Too is the energy density of the field which we will measure! Our four-velocity, on the other hand, is characterized by the components un (1,0,0,0). Therefore we find,

T(U,U) = TaSU~S = Too


and we conclude that T(U,U) 1:S the energy density of a field with energy-momentwn T measu:t'ed by an observer with four-velocity U.

322

SOLUTIONS OF WORKED EXERCISES,


No.

6.5.1
If) I!1J

Let gij and gij be the components with respect to two coordinate systems Then, by definition: lj,fuk =

(~2'

U2 )

0\ (and If)i 1iik 0\)


j
=

But we know that gij transforms covariantly! Consequently dxil. axm (~k = ayJ ayK 1~9.m Substituting this we get m ij axil. ax oi l~) ayj a/ (?~m = k Multiplying with the reciprocal matrix -I- we get axP
i .. e.

ak

Multiplying with the reciprocal matrix .,.Pq we further get


('i')
Le.

ij ax q ayJ

-, = -

ayi

pq

(2)

al III

Multiplying with the reciprocal matrix L we finally obtain q ax

ak
g

i.e.

ik

(21

So finally we get the formula we wanted! After a lot of index gymnastique you. see with your own eyes that gij transforms contravariantly!

can

No.

6.7.1
Taking the Lagrangian
L(xj-) =-mV-g

adx dA

~----,-

'as dA dA

dxa dx S

, aL = -X d ~a aL as a starting point the Euler-Lagrange equatlon axa d a(a:xagaS dx a dxll

r J

reduces to

~dX~

2V

323

II

:J~l
~t ---;,:s ax ax
(V
That is obviously a mess, but you should not be surprised of that because even a straight line may be characterized by a complicated parametrization: xa agalldxdx+ S a gall

2 aJ dx

= aaSin(etgA)

+ ba

Hence even a simple curve like a straight line may solve a complicated differential equation, if we choose a crazy parametrization! Let us take advanta~e of the fact that we can choose our parameter A as we like it, so let us use the proper time T as the parameter! Hence we put A = in the equation of motion and use the simple relations:

V-=1,

.'dA !..- V-= a

Then the equation of motion is simplified considerably!

d x -all d T2

(Here we have split the term:

Performing the substitutions a+a,a~ in the last term, this is rearranged as

I f we multiply this equation with gllV we get:

d x oV a 2 d.

1 llv ya [a + - g + 2 axS

ag

Sll

_ agas

ax

a ] dx di = 0 dT d. axll

2 V d x 1 llV =>--+-g

6.. 2

and we are through!

324

II

No. 6.7.3 If we extremize


s

we get the Euler-Lagrange equations

which we may rearrange as

d? = 0

d t

~ = r (~~/
(6.81)

+ r Sin2e

(~~) 2

dr

d 2e

2 dr de dIP 2 = - ;:: de dT + sin e Cos e (de)


= _

g dT
d 2x a

~ dr dIP _ 2 Cot e de <ltD


r dT dT dT dT

These equations then comprise the equations of motion for a free particle. But we can also Use them to read off the Christoffel fields. Using the formula

d?" '" - r

a ].lv

dT dT

dx].l dx

you immediately get the following nonvanishing components of the Christoffel fields: rr - r Sin 2 e - r

r~e

(6.82)

e re rU) r
rq>

er rIP
q>r

a
r

(!XI> 1 r\P = rIP r Sq> tP9

(/XI)

- sin

e Cos
Cot e

325

II

ehapter 7

DIFFERENTIAL FORMS AND THE EXTERIOR CALCULUS


7.1

INTRODUCTION

We want to extend the differential calculus, to include tensors. Consider an arbitrary Euclidean manifold M and a cotensor field T of rank 2 on this manifold. If we introduce coordinates (xl, .. ,xn ), we can characterize T by its co-variant components TetS(x I , ... ,x n) Here you might try to differentiate the components of T ,'lith respect to one of the coordinates x~, i.e. you form the components n a~ TetS(XI, ... ,X ). The question is: Is this new quantity a co-tensor field of rank 3, i.e. do the components a T Q(xl, .. ,x n ) coincide with ~ a" the components of some co-tensor of rank 3? To investigate this we must try to show that the quantity a~TaS transforms we introduce new coordinates (yl, .. ,yn). covariantly. Therefore

Wi th respect to the old coordinates (x I, ... ,xn ) the quantity has the components
--T

din

aX~(11 et

B(x

I ,X

and with respect to the new coordinates (yl, .. ,yn), the quantity has the components
--T

din

aY~(21

etS(y, .. ,y ).

As Tets(y , .. ,y ) (21
I

(II Y

I n ax Y T .s(X , ,x ) - ayet

326

II

we get

The first term is exactly the term we are after, but the second term spoils everythin0. So the quantity with components
a~Ta8

is certainly

not a

cotensor:

We can even understand intuitively what went wrong! We are looking at a co-tensor field T and trying to form something like the partial derivative of it. Intuitively you would then try to make an expression like
1

"
tensors with

[Tp(x l

,_

.x~+~,
c"

'x n ) -

T P(x l

I ,

.x~, ,

'xn) 1

But this has no meaning at all because we are trying to subtract two

different base points! But they lie in different tensor spa-

ces and have absolutely nothing to do with each other. Of course, you could still try to form the component expression
1

"

[ T a8 ( xI ; .. ;x ~ +

E; .. ;Xn) -

n) Ta8 (Xl. .~. , . ,x , .. 'x ,

This is legitimate, but observe that this expression has no geometrical meaning. You cannot compare two tensors with different base point P and Q just

by compairing their components.


Original coordinate system Changed coordinate system

Fig. l23a

Fig. l23b

Introducing a coordinate system which covers P and Q we may characterize the tensors Tp and TO by their conponents

327 and

II

You might hope that i t I.ould make sence to say that the b/o tensors Tp and TQ are almost equal if their components
T~S(P)

and TaS(Q) are

almost equal. But watch out! Let us change the coordinate system in such a way that the coordinates in the neighbourhood of P are unaffected while the coordinates in the ponents T'aS(P)
nei~hbourhood

of Q are changed dras-

tically (see figure 123). Then Tp is still characterized by the com-

T~B(P)

while TQ will be characterized by the new components

T'NQ(Q) ~"
Thus,

=T

Yo

(0) ~ ~
ay~ ayS

although we have fixed_the components Tp, we can change the cotensors

components of TQ into almost everything! This shows clearly that it makes no sense to compare with different base points, just by comparing their components. In these notes we will not try to attack the general problem of constructing derivatives of co-tensor fields but will only show that it is possible to overcome some of the difficulties if we are willing to vlOrk only with a special kind of cotensor.

7,2

K-FORMS - THE WEDGE PRODUCT


Now let F be a cotensor of arbitrary rank. He say that F is skew-

symmetric if it changes sign whenever we exchange two of its arguments:

If we characterize F by its components,

we see that F is skewsymmetric, whenever the components F aJ a k are skewsymmetric in the indices aJ, ... ,a ! The skewsymmetric cot enk sors are so important that mathematicians have given them a special name:

Definition 1 A k-form is a skewsymmetric ootensor of rank k.

328

II

Consider a specific po.int P o.n Mn. We have previo.usly attached a who.le family o.f cotenso.r spaces Tp(o.,k) (M) to. this po.int. No.w we want to. extract the skewsymmetric co.tenso.rs. The set o.f k-fo.rms at k the po.int P will be deno.ted Ap(M). Let us start with so.me elementary remarks: If F and G are skewsymmetric co.tenso.rs o.f rank k, then so. are

F + G and XF. Therefo.re the set o.f k-fo.rms

A~(M)
fo.rms a vectorspace. If k>n then A~ (M) degenerates and will o.nly co.ntain the zero.-fo.rm:

A~(M)

{O}, k>n co.tenso.r o.f rank k. Then F is

To. see this, let F be a skewsymmetric characterized by its co.mpo.nents


F

aj a k '

but as k>n two. o.f the indices must always co.inside (since an index a can i o.nly take the values l, . ,n), whence F aj a vanishes! So. in the k case o.f k-fo.rms we do. not have an infinite family o.f tenso.r spaces! We will use the convention that co.tensors of rank 1 are co.unted as l-fo.rms. Of ~o.urse, it has no. meaning to. say that a co.vecto.r is skewsymmetric, but it is useful to. include them amo.ng the fo.rms. In a similar way it is useful to. treat scalars as O-fo.rms. Co.nsequently the who.le family o.f fo.rms lo.o.ks as fo.llo.ws:
A~ (M) =R;A~ (M) =Tp (M) lJ\.~ (M) l lAp (M) lAp

n+l

(M) ={O} lAp

n+2

(M) ={O};

Wo.rking with o.rdinary co.tenso.rs we have previo.usly intro.duced the tensor pro.duct: If F and G are arbitrary co.tensors then the tenso.r pro.duct F II G def ined by

is a tria

co.tenso.r o.f rank k+m. But if we restrict o.urselves to. skewsymmeco.tenso.rs this co.mpo.sitio.n is no. lo.nger relevant because F ~ G
1.

is no.t necessarily skewsymmetric.(If yo.u interchange V. and say no.thing about what happens to. F(Vji ;vk)G(Uji .. therefo.re try to. mOdify this compo.sitio.n:

;rtm.

rt. J

yo.u can

We will

If Waj a is a quantity with indices aj, ,a k then we can co.nk struct a skewsymmetric quantity in the fo.llo.wing way (7.1)
w[ aj a 1 k

1 = k!

!(_l)TI w TI TI

(all TI (a ) k

329 where we sum over all permutations is the sign of the permutation, i.e.
(-1)
TI TI

II

of the indices al, ... ,ak,and

(-If

(-1) TI

= =

+1 -1

if if

11 11

is an even permutation is an odd permutation

For instance we get W[ab] and W[abc]


1 = 2![Wab

- wba]

1 3! [w abc + wbca + wcab - wacb - wcba - wbac ]

Observe that w[a

The quantity w[al .... ~] is called the skewsymmetrization of wal ak . ] is completely skewsymmetric, and that if ~ wal .... ak is born skewsymmetric, then

1 (This is, of course, the reason why we have included the factor k!') If we introduce the abbreviation

(7.2)

sgn[bl .. b k al ... ak

]= { -1

+l if (b l ... bk) is an even permutation of (al"

.ak)

if (bl ... bk) is an odd permutation of(al ... ak)

o otherwise

we can write down the skewsymmetrization as an explicit summation W = 1 sgn[bl ... b k ] W [al" .ak ] k! al'" <lk b l .. b k With these preliminaries we may return to the skewsymmetric sors: If F is a k-form characterized by components F ... coten-

al a k m-form with components Gb " ' b ' then the tensor product is charateril m zed by the components:

, and G a

Next we form the skewsymmetrization (k+m)! F G k!m! [al .. a bl .. b ] k m where the factor
The
stat~ct~cal
o

(kk;m~! has been included to make life easier . m.


(k+m!)
k~m!

factor

removes double

count~ng:

330

II

Observe that sgn[Cj",Ckbdj"'bdm] and F are both skewsymmetric in (Cj",ck)' aj ... ak j .. m Cj .. ,ck When we perform a permutation of (Cj",ck) we pick up two factors (_1)n. All permutations of (Cj ... Ck ) thus give the same contribution tothe sum. We need therefore only to consider a single representative, for instance the ordered set (Cj ... c ) k c~araterized by the property cj< ... <c . There are k! permutations of (Cj .. c ) gik k ving the same contribution, and similarly there are m! permutations of (dj ... d ) giving the same contribution. Using this, we obtain the formula m
= (k+m)! F G k!m! [aj . ak bj ... bml

Cj<",<ck

dj< . <dm

sgn[Cj"'Ckdj"'~]F
aj ... ak bj ... bm

G Cj .. ck dj ... ~

where we have obviously avoided doublecounting.

Observe, that

Each of the terms Frr(aj) .. G"'rr(b ) transform

covariantly. Therefore

m
(k+m)! F G k!m! [aj ... a k bj ... bml coins ide with the components of a skewsymmetric This motivates the following definition: cotensor of rank k+m.

Definition 2 Let F be a k-form with components F and Gam-form with components G b 1 .b m a r.ak then the wedge product FAG is the (k+m)-form with components:
(k+m) !

kIm!

Let us give a Simple example: If A and Bare covectors,


i.e~

I-forms, then their wedge produc AAB

has the components

From (7.3)

this you see that AAB


=

[A I B - B I Al

Exercise 7.2.1
Problem: (a) Consider a 2-form F characterized by the components Fab and a one-form B characterized by the components Bc' Show that FAB are characterized by the components

331
(7.3) FabBc + FbcBa + FcaBb (b) Let A,B,C be 1-forms. Show that
(7.4)

II

The wedge product A has some

s~ple

algebraic properties,

Theorem 1 The wedge product is associative and distributive:


(7.5)
(7.6)

(FAG)AH

FA(GAH)
=

FA(AG + llH)

AF A G + llFAH

(7.7)

If F is a k-form and G is a m-form, then FAG = (-lJ<rrG;.F Observe expeciaZZy that coVectors anticommute!

Proof: Associativity follows from the simple observation that (FAG)AH has the components, (k+~+m) ! (kH) !m!

Distributivity also follows directly when you compare the components of both sides of the equation. We omit the proof which simply consist of keeping track of a lot of indices! Finally we consider the wedge products FAG and GAF: (FAG) (GAF) Consequently (FAG) al. .akb l ... b m
=

(k+m) ! b b = -k-'-'- [F G b + .. al ... a k I'" m .m. al ... a k b I m b l ... bmal .. a


=

(k+m)! k!m!

[G

F b I ' " b m al a k,+ ...

(GAF)

b l .. bmal" .ak

Observe that it will require k'm transpositions to obtain al ... akbl ... bm from bl . bm~I ... ak (It will require k transpositions to move b through al .. a , it will then require another k transposim k tion to move b _ through al . a ect.) Using that GAF is skewsyrnmem l k tric we get (GAF) b l . .bmal" .ak (-l)km(GAF) al" .akb l .. b m

332

II

Combining these two formulas we finally obtain

(-1) kIn (GAF)


showing that

aj ... akb j ... bm

FAG

(-l)kInGAF

The most surprising of these rules is no doubt the special "commutation" rule (7.7). There is an easy way to remember this rule. If F is a form of even degree, we call it an even form and similarly a form G of odd degree is called an odd form. The rule (7.7) can now be summarized in the following form:
EVen forms aZways commute, odd forms anti-commute. So if you consider an expression like

and you want to interchange the order of the forms, then you only have to worry about the odd forms: Every time you interchange two odd forms it costs a sign. Now let us look at a 2-form F. Introducing a coordinate system we may decompose F along the basic cotensors dx i 0 dx j , i.e. F = Fijdx
i

.dx

where Fij are the components of F with respect to this coordinate system cf. (6.68) . But F ij is skewsymmetric iIi i and j. Consequently we can rewrite Fij as Fij = ~(Fij-Fji). Thus we can rearrange the decomposition of F in the following way F

\(Fijdx

0 dx

j - Fjidxi 0 dx )

Interchanging the dummy indices i and in the last term we further get i j j i j F = \(Fijdx 0 dx - Fijdx 0 dx ) = \FijdxiAdx This may obviously be generalized to arbitrary k-forms:

Theorem 2 If F has the components P.


(7.8)
~j. "~k

. we can expand F in the followirJ{f way

k! ~j '~k

P.

. d::;i IA A d::;ik

Exercise 7. 2. 2
Problem: Let us introduce coordinates on a manifold The coordinate system generates canonical frames (tl ... ,tnl and (dx1 . ,dxll)

MP.

(a) Show that

a, (dx

333
~
A ..

II

N1x)b
1'"

b
k

al"'~l = Sgn b, . b
k

is an even permutation of (l .. k) 1S an odd permutation of (, ... k)

(b) Let (Ul . Un) be a n-tuple of ~angentve4to~s,.where Uj is characterized by the contravariant components A1 j , i.e. u=eiA1 . Show that

...

...

ExerlYise 7.2.3
Problem: Let T be a form of maximal rank n. Observe that T is characterized by the single component T j n since all other components can be obtained from this using a permutation of the indices. Show that T can be decomposed as
(7.10)

In the rest of this section we will try to develope a gearet.rical interpretation of differential forms, to help you to understand more intuitively the concept of a k-form. We start by considering the ordinary 3-dimensional Euclidian space R3. Let us investigate the k-forms associated with the tangentspace at

the origin: Canonical frame: O-forms I-forms 2-forms 3-forms 1 dx ; dy dz ; dXAdy; dYAdz; dZAdx dXAdYAdz

Consider the basic two-form dXAdy. Remembering that dXAdy is defined as a linear map, Tp(M)xTp(M)~R,we want to compute its value on a pair of tangentvectors dXAdy(~,v). Observe that dXAdy has the components (cf. exercise 7.2.2). Consequently we get

But this is easy to interprete. We use that ~,v span a parallellogram, which we project onto the x,y plane (See figure 124). But then we conclude: dzAdy(7., is the area of this projection. For this reason we say that dXAdy defines a unit of area in the x-y plane. Observe also that dXAdy generates an orientation in the x-y plane. The projected area

v)

dXAdy(~,v) is positive if and only if the projection of (U,v) defines

334

II

~::-+---+-----,. y

x
Fig. 124
Exercise 7.2.4
Problem: Consider the basic one-form dx. Show that dx defines a unit of length along the x-axis in the sence that dx(~) is the length of the projection of ~ onto the x-axis.

Exercise 7.2.5
Problem: (a) Show that dxAdYlldz defines a volume-form in R3 in the senc:e that dxA~lIdz(Ul;U2;U3) is the volume of the parallelepiped spanned by Ul,U2 and U3
(b) Show that

dxAdYlldz defines an orientation in R3 in the sense that (Ul,U2,Ull is positively oriented if and only if dxAdYlldz(UI,U2;Ul) is
(Hint: Decompose the triple (Ul,U2,U3) as

positive.

U. = ~.Ai.
J
1

and show that i Det(A . )


J

Compare with the discussion in section 1.1) .

Obvio~sly

the above analysis depends on very special properties of

the Euclidian space R3. It id however possible to convert it into a purely geometrical form suitable for generalizations to arbitrary Euclidian manifolds. Consider once more the basic two-form dXlldy. The coordinate-functions x and y defines two stratifications of the Euclidian space:
. ,X

... ,y

-2, x -2, Y

-1, x
-1, Y

0, x 0, Y

1, x 1, Y

2, ... 2, .

Together these two stratifications form what is generally known as a "honey-comb structure" (see figure 125) .

335

II

Fig. 125

If we return to a parallellograrn spanned by two tangentvectors ~

and ~, then it is clear that the projected area is simply equal to the number of tubes intersected by the parallellogram. We therefore conclude:
d::;Ady (~;

v)

is the nwnber of x-y-tubes, which is intersected by the paraUeUogram

spanned by the ~ and

v.

Exercise

7.2.6

Problem: (al Consider the basic one-form dX. The coordinate function x defines a stratification: ... ,x '" -1, x = 0, x = 1,.. .Show that dX(;;) is equal ~o the number of hyperplanes x = k that are intersected by the vector
u

(bl Consider the basic three-form dXAdYAdz. The and Z generate three strali!i~ations, which tu:e. Show that dXAdYAdz~u;v;w) is equal-+t~ ta~nQd in the parallelep~ped spanned by u,v

coordinate fUnctions, x,y together form a cell-structhe Q:umber of cells conand w. _

We can nOw transfer the above machinery to an arbitrary manifold. Consider a two-dimensional manifold M with local coordinates (X I ,X 2 ) . Let Po be a point on M and consider the basic two-form dX l Adx 2 at Po' The two coordinate functions Xl and X2 generate two stratifications on M: xl= ,-l,O,l, . and x 2 = . ,-1,O,1, These two stratifications divide the surface into a great number of cells (see figure 12b).

x2
(b l ,b2 \

1\
",.

(a l

Fig. 126

336

II

Consider two tangentvectors

u and v at

Po. This time

u and v span

a parallellogram which lies outside the manifold M. We must therefore use a trick. In coordinate space (a l ,a 2 ) and (b l ,b 2 ) span an ordinary parallellogram. The number

dx l "dx 2 (i!;v)
is therefore equal to the number of cells contained in this parallellogram. But since this parallellogram is a subset of the coordinate domain U, we may transfer it to M. The image will be referred to as the "parallellogram" swept out by

u and v.

Therefore we conclude in

the usual way:


dXIAdX2(~;V) denotes the number of cells contained in the paraZZeZlogram swept

out by

u and v.

7.3

THE EXTERIOR DERIVATIVE


That was a long algebraic digression. Let us return to our main pro-

blem. We want to construct a differential operator which converts a cotensor field of degree k into a something like T al ... a k (xl; ... ;xn) .... a T (XI; ... ;X k ) 11 al .. a k cotensor fields. They cotensor field of degree k+l, i.e.

We now restrict ournelves to skewsymmetric

playa key role in differential geometry and are called differential

fonns.

If it is clear from the context that we are working with a skewcotensor field, then a differential form of rank k will be

symmetric

referred to simply as a k-form. The set of all smoth differential forms of rank k will be denoted

A~(M). Observe that Ak(M) is an infinite-di-

mensional vectorspace for D<k<n. Now, let us consider a differential form of degree zero, i.e. a scalar field
~.

Then we have previously introduced the differential ope~

rator d which converts the scalar field degree one, i.e. a covector field
d~.

into a differential form of

If we introduce coordinates then

~ is represented by an ordinary Euclidean function ~(XI; .. iXn) and d~


is characterized by the components all~(xl; ... ;xn). It is this differential operator d we want to extend! Consider a differential form of degree one, i.e. a co-vector field A characterized by the components

337

II

A (Xl, ... ,xn) o. Differentiation this we get the quantity

__ d_ A (xl, ... ,xn) _ d~Ao.(X I , ... ,xn) dX~ o.


but this is of little interest because it is not skewsYmmetric in and o.. Therefore we antisyrnmetrize it and get
~

This is skewsyrnmetric and if we can show that it transforms covariantly,then we have succeeded. Introducing another coordinate system we get the new components
23 A (2)[ Pl2jCt 1
:)

a~"

'1

(2l a

:\

\vt.

dY

o.

(V! / ~ ,

. ,,,p)

But A (y) (2)o.

ax S = AS(X)-. Inserting this we get (1) ayo. axS ax S a [A (x)-l_a_[A (x) 1 ayo. ayo. (1) s ayP Il)S ayP aA ax Y axS a&s ax Y axS a"x S ....!ll6. ----- + A - lllS ay~ ayo. ax Y ayll ayo. ax Y ay a ayP

2Ig)[\l{~o.l

lllSayaay~

a"x S A -----

Using the fact that partial derivatives commute, we observe that the spoiling terms cancel each other! Finally we get ax Y axS ax Y axS =aA ---aA-u)YU)Sax p ayo. lllYl1tl ayo. ayP Here we interchange the dummy indices Sand Y in the last term and get 2a
l2j[ \l{zjlll

= [a A - a A 1 ~ ~ mYtnS tuS(l)Y axP ayo.

Y S 2a A ~~ (u[Y(llSl ayP ayo.

So everything works! The expression 2a[~ASl coins ides with the components of a differential form of degree 2 which we will denote dA. This may immediately be generalized:

Definition J If F is a differential form of degree k characterized by the components F-40 ik(:x;J. then dF is the differential form of degree k+l characterized by the components (k+l)! k!

F
l

[p i

ikl

The differential operator d: i<-(M)-+N<+l (M) is called the e:cterior derivative.

338

II
a~.

You may think of d as a kind of 1-form characterized by the components fact a~ transforms covariantly according to the chain rule:
(2)~

In

_a_=_a_ ~=a ~ ay~ ax\} ayW (l\\}ayll

\}

The construction of the exterior derivative is then formally equivalent to the wedge-product : dAF. This also explains the statistical factor. (Compare with definition 2).

Exercise 7.3.1
Problem: Let F be a 2-form characterized by the components FaS' Show that dF is characterized by the components
(T .11)

The ",o,r-.;; '5<"'" .:, .. :" .. h,-,-V8 has several important properties. It is a linear operator, i.e. (7.12) as you can ing one:

d(F+G) = dF + dG
~sily

check. But the most important property is the follow-

Theorem 3
(7.13)

(Poincares lemma)
dd = 0

i.e. it automatically vanishes if you apply it twice. Proof: Consider first a scalar field ~ represented by the Euclidian function ~ (Xl, ,x n ). We then get (d~)ll = a~~ , which :iInplies that
(d 2~)
~v

=a \} a

~- a

a\} ~ = 0

This shows the mechanism in the cancellations. We then consider an arbetrary differential form F of rank k. Let F have the components F. . We then get for dF
11" .1

and similarly for d 2 F


(k+2) :
(k+l)

339

II

But this sum vanishes, because if we write out all the terms they occur in pairs
d d of. a " J
I

. Jk

and

As Sajl ... jk is generated from aSjl ... jk by applying one transposition they will have opposite signs, i.e. we may collect the two terms into

and this vanishes automatically.

Before we derive more rules we introduce some more notations.


Exercise 7.3.2
Problem: Let F be a k-form characterized by the components F ... an and Gam-form al characterized by the components Gbl .. b ' Show that we can decompose FAG as m 1 a ~ b b b dx I A... A da Adx I A... Adx m k' , F G b .m. al"'~ I .. m (Hint: Use the distributivity of A)

(7.14)

Exercise 7.3. J
which is not necessarily skew-symmetric. Show Problem: Consider a quantity w al'''~ that a ~ a ~ w[ _ 4x I A... Adx =w dx I A... Adx (7.15 )
al"'''j( al ... ~

If F is a differential form of degree k we can decompose it along the basic k-forms (7.8). If we form the exterior derivative dF, we know that it is characterized by the components

Therefore we can decompose dF in the following way


i

(k+l)!

(k+l) ! . j dxll"dx 1" ... Adx ~ d[ 11 F. l.1 l.k

1
-k.'

i i d[ F. . jdxllAdx I A Adx \l l.1 l.k

But according to exercise 7.3.3 we are allowed to forget the skewsymmetrization! In this way we obtain the formula

340

II

(7.16)

dF

We are now in a position to prove a generalizatbon of the familiar rule for differentiating a product (Leibniz' rule):

~ (f.g) dx

df .g + f. ~ dx dx

Theorem 4 (Leibniz' rule for differential forms) If F is a k-form and G is an J,-form, then
(7.17) d(FAGj = dFAG + (-l)k FAdG

Proof:
If you canbine (7.14') and (7.16) you .i.nrnediately obtain 1 II i ik j j J, d(FAG)-k' n, 3 (F. . G. .)dx Adx I A... Adx 'idx I A...Adx N II 11 . . . 1k J I .. J J,

According to Leibniz' rule for ordinary functions this can be rearranged as

+ -k",F.
N.

11 .1

II i ik j I J, . (3 G. .)dx Adx l A... Adx Adx A.. Adx II J I . J J,


k

In the last term we ~d like to nove dx ll through dx I A. . Ad$. anticamn.rte, and therefore we IIUlSt pay with a factor (_l)k:

But covectors

We m~ think of d as an odd form,(cf. the discUssion after def. 3). In Leibniz' rule we have two terms and

In the last term we have interchanged d and F. This costs a sign if F is an odd form.

341

II

Observe that if ~ is a differential form of degree 0 i.e. a scalar field, then ~AG is just a fancy way of writing ~(xl, ... ,xn)G. In that case Leibniz' rule degenerates to (7.18) We can now recapture (7.16) in the following way. Consider a k-form
1 i ik

-k! F. i (x)dx lA ... Adx l.l ... k

If we keep the coordinate system fixed, we may treat F. . (x) as a l.l .l.k scalar field. Applying the exterior derivative we therefore obtain
l. i,~'.
,

:d',

q.-.

i i (x)]Adx lA ... Adx k

1 i i k + k! F. . (x)Ad[dx lA ... Adx ] l.l .l.k

Using Leibniz' rule once more we see that d[dx lA ... Adx k ] automatically vanishes since it only involves double derivatives (theorem 3). We are therefore left with the first term. If we use that

we finally recapture (7.16)~ We conclude this section with a definition of two important types of differential forms:

Definition 4 (aJ A differential form F is alosed if its exterior derivative vanishes, i.e. dF

=0

(bJ A differential form G is exact if it is the exterior derivative of another differential form F, i.e.

G = dF Observe that an exact form G is automatically closed, since G


:implies that

dF

dG .. d F

but the converse need not be true!


See

(Examples will be given later on.

,e.g., 5eC"tWn 7.8 TNhcre the case of the IlDnopole field is diocussc1.)

Exercise 7.3.4
Problem: Let F be a closed form and G an exact form. Show that FAG is exact.

342
A

II

closely related type of differential form is given by;

Definition 5 A k-form W is (JaZZed sirrrpZe if there exist k smooth scaZarfieZciB <PI '~k so that W can be decomposed as

(7.19)

w = d<PIA . .. Ad<Pk

The basic forms generated by a coordinate system and the simple forms are intimately connected. If <Pl, ... ,<P are sufficiently well bek haved then

yk <Pk (Xl, ... ,xn ) k+l k+l Y X


y
n
X

is an admissible exchange of coordinates and in the new coordinates, (yl, .. ,yn), we can decompose Was W = dylA ... Adyk So W is a basic form generated by the new coordinates. Consider a simple form d<P1A Ad<Pk. It can be interpreted geometrically in the same way as the basic forms. It generates a stratification, q,l= . ,-l, 0, l' . ' ... '<P k = ... -1,0,1, .. ,and d<PIA ... Ad<Pk(Ul ; .. ;iik ) is equal to the number of "tubes" intersected by the "parellelepiped" spanned by (~l' ... '~n). Simple forms have expecially nice properties:
Lermra 1

A simpZe form is cZosed and exact.

Proof:
As d =
2

0 it follows trivially that dtPIA ... AdtPk is closed. dcPIA ... Ad<Pk = d[<Pldq,2 A Ad<Pkl

The exactness follows from the fOlllUlla

Worked exercise 7.3.5


Problem: Let flo .

,fit: R.... R be smooth real functions. Show that

is a simple form.

343

II

7.4

THE VOLUME FORM


n Suppose the manifold M is also equipped with a metric ~. Then we

can speak of areas, volumes etc. in the tangent-space Tp(M). Let us make this more precise. Consider an n-tuple (Ul"'.'~n) which generates a "parallelepiped". We want to find out how to compute the n-dimensional volume of this "parallelepiped". If we introduce coordinates n (xl, ... ,x ) it would be tempting to put Vol [Uli jU n ] = dx 1I lIdx
~

-+?

[uli .)'un ]

-+

:-1'

by analogy to the lower dimensional Euclidian cases. (See exercise 7.2.5) But here you $hou:d be careful since the canonical frame generated by (xl, .. ;xn ) needs not be orthonormal. The tangent space Tp(M) is simply isomorphic to the standard Euclidian
~pace

R . Recall the standard definitions of a volume in the n-

dimensional Euclidian space. Let (e:, .. ,e~) be an orthonormal frame and consider an arbitrary n-tuple of vectors (~l""'~n)' They span a "parallelepiped" with the volume

(7.21)

V01(VI; ... ivn)

........

Det[B]

i where the matrix elements B , are the coordinates of ~, J J

~,

= e~Bi,
].

-+ -+ ""'0 -+0) Furthermore (VI""'V n ) generates the same orientat].on as (el, ,en

if and only if Det[B] is positive. Now let us introduce coordinates around the point P. Then the canonical frame (el, ... ,e ) generated from the coordinates needs not to n be an orthonormal frame, but we can decompose it on the orthonormal frame (e~, .. ,e~) : (7.22)
.... ....0

ej

eiE j

Let furthermore (u""',;n) be an n-tuple of tangent vectors, where

~j is characterized by the
cise 7.2.2 we know that
(7.9)

contravariant components Ai,. From exer)

DetCA)

We now get uj
-+ -+

i _""'0 k i eiA j - ekE i A j

From (7.21) it then follows that


(7.23)

It remains to interprete Det[E]. Consider the coordinate exchange

344

II

Observe that the orthonormal frame (elo, ... ,e~) me generated from the (yl, ... ,yn)-coordinates. ly from (7.22) and

is the cononical fra(This follows directe-

(6.10. The metric coefficients gij corresponding

to the (yl, .. ,yn)-coordinates therefore reduce to the Kronecker-delte 0ij' The transformation formula (6.21) for the metric coefficients may be rearranged as

i.e.

g = DetLGl

[Detd~)12

i.e.

Inserting this into (7.23) we have thus determined the volume form in the tangent space Tp(M):

(7.24)
later on we shall recover this formula from another point of view. The volumeform also
contro~the

orientation of the tangentspace. This

follows immediately from (7.9) which shows US that (~I, ,Un) generates the same orientation as the canonical frame (~l, ... ,en) if and only if dx 1 A Adx n (ir 1 i ; irn ) is positive. In the preceding discussion we have been focusing upon a single tangent space. We will now try to extend the discussion of the local aspects of the volume form to its global aspects. Our final aim is to construct a global differential form for this we must first discuss Let M be a manifold and Po
~he

which at every point

re-

duces to the local volume form given by

(7.24), but as a preparation

orientability of manifolds.

a fixed point in M. Consider the set of

all coordinate systems surrounding Po. Each coordinate system generates an orientation on the tangent space Tpo(M). Now consider two sets of cOordinates: From
(x 1 , , xn)

and (y 1 , , yn)

(6.12) we know that

(2jJ

.... e.

Therefore we conclude that (x1, .. ,x n ) and (yl, ... ,yn) generate the same orientation on the tangent space Tpo (M) provided the Jacobiant

345

II

Det

[dX~l
dyJ

is positive.

Okay! Let us choose a positive orientation in our tangent space


Tpo(M). This is done by picking up a specific coordinate system
(<I\"U o ), which we declare to gene race the pos'itive orientation. If
(~,U)

is an arbitrary coordinate system, then it generates a positive

orientation provided <l Det[d x S1


dy

is positive and it generates a negative orientation provided its Jaco-

biant is negative! In what follows we will always assume that we have chosen a positive
orientatio~

in our tangent space Tpo(M).

This was the local aspect,

(i.e. we have been focusing at what hap-

pens in a neighbourhood of a point Po). Now we turn to the global aspect of orientation. First we consider a simple manifold, i.e. a manifold which can be covered by a single coordinate system
~eclare (~o,Uo).

It

generates an orientation in each of the tangent spaces on M, and we these orientations to be the positive orientations!
(~I,UI)

Then we consider an arbitrary manifold. Suppose we have two coordinate systems and
(~2,U2)

which overlap. If we are going to let


nI2=~1 (UI)n~2(U2)

them generate the same orientation in the tangent spaces we must be careful in the overlapping region that Det[dx<l/dySl .We must demand that they generate the same orientation in the overlapping region n 12 , i.e.
is positive throughout the overlapping region. This

motivates the following definition:

Definition 6 A manifold M is said to be orientable if we aan ahoose an atla~ (~i>~i)iI.suah that any two aoordinate systems in our atlas generate the same o~entat~on> ~.e.
det(~)
dys

<l

is positive throughout all overlapping regions.

When a manifold is orientable we can choose an atlas with the above property. This atlas will then generate a specific orientation in each of the tangent spaces on M, which we arbitrarily declare to be positive! Let us end this digression on orientability with an example of a manifold which is

not

orientable: The MObius strip (see fig. 127 ).

346

II

Fig. 127a

CONSTRUCTION OF THE MOBIUSSTRIP

Fig. 127b

NON-ORIENTABILITY OF THE MOBIUS STRIP

Exercise 7.4.1
Problem: Let MTI be a differentiable manifold. Let us assume the existence of an nform 0 which is nowhere vanishing. If we introduce coordinates (xl, ... ,xn) we can decompose 0 as 0 = n1 ... n(x)dx1A ... Adxn . That 0 is nowhere vanishing implies that I"lr ... n(x) is nowhere zero. (a) Let (xl, ... ,~) denote an arbitrary set of coordinates. Show that the sign of O(~l; ... ;~n) is constant throughout the range of the coordinate system. (b) Show that 0 generates an orientation of M by defining a coordinate system (xl, ... ,xn) to be positively oriented if and only if O(el; ... ;en ) is positive. (c) Show that an n-tuple (~l, ... '~n) is positively oriented with respect to the' orientation generated by Q if and only if O(~l;' ... ;~n) is positive.

Exercise 7.4.2
n be an n-dimensional manifold in Rn+1. i.e. M is a smooth hY.PersurProblem: Let M face. Let us assume the existence of a smooth normal vectorfield rt(x) to M. i.e. to each point x in M we associate a normal vector ti(x) to the tangentspace. Show that M is orientable. n 1 (Hint: Let (x 1 .. ,x + ) denote the extrinsic coordinates in ~+1. Let (~l .. ~n) be an arbitrary n-tuple tangent to M. Define a smooth n-form o on M in the following way ... 1 n+1 ...... ... o(... Vl; .. ;Vn ) = dx A.. Adx (n;vl; .. ;V ) n and apply exercise 7.4.1) Consider the sphere and the Moebiusstrip in R3 (see fig. 128.129)! You can see with your own eyes that the sphere allows a normal field. while the MOebiusstrip does not!

347
The sphere The Moebiusstrip

II

Fig. 128 Fig. 129 As a preparation for the construction of a global volume form we are going to investigate two special quantities which are not true geometrical objects, Le. they can only be defined in connection with a coordinate system on the manifold. The first quantity is intimately connected with the metriC g on our manifold. To be specific let us assume that g is a Minkowski metric. Consider a point Po in M and a coordinate system covering Po. We can then characterize g in terms of its components
ga~

1 gaS(xo, ,xn) o

and we may compute the determinant g = Det[ga~] This determinant is necessarily negative, because g is a Minkowski

metric. It is this determinant we are going to examine. Obviously g(xo) is not a scalar, i.e. it does not transform according to the rule q(yo)
(~)

trl

q(xo) when we exchange the coordinate system. To find the


ga~

transformation rule we use that (6.21)

transforms

covariantly:

=g

ax Y axe

(l)ye aya ay~

We can rearrange this as a matrix equation: (6.23) If we evaluate the determinant, we find

Le. (7.25)

g (Yo)
(21

348

II

This transformation property makes it natural to look at V-g(x), and we conclude:


Lemma 2
On

a manifold with a Minkowski metric y.:g7XJ transforms according to the pule


v-g(yJ
(21

(7.26 J

= I Det(-7,J I y.:g7XJ ay (1)


Fg by

a ().

(On

a manifold with an Euclidian metric you just replace

Wi) .

Keeping this in mind we now introduce the next quantity, the LeviCivita symbol: +l ~f (al, ... ,a n ) ~s an even permuta~ion of (l, ... ,n)
(7.27)

-1 {

~f

(al, ... ,ad

~s

an odd

permutat~on

of (l, .. ,n)

otherwise

Again we select a point Po and an arbitrary coordinate system to this pOint. Then we attach the Levi-Civita-symbol to this point Po. This does not corresponds to the components of a definition it transforms according to the rule:
(7.28)
S

cotensor, since by

al .. a n

(Yo)

al .. a n

(xo)

However we may use these two quantities,~ and sal ... an,to construct a differential form. Let us choose specific coordinates (xl, .. ,x n ), and consider the n-form T characterized by the components

with respect to this particular coordinate system. Let us try to compute the components of T with respect to some other coordinates (yl, ... ,yn). oAs T get
(*)

.an
1

transforms

covariantly, we

~b
1

(Yo)

ilIa a
1

(xo)---". - b Dl

ax ay

ax n
n

ay

ax 1 a ... a - bn I ay 1

ax n -bay n

Here we should try to study

349

II

a
-bay n

ax n

a little closer. If (b 1 ,

,b n )

(1, ... ,n)

then

by the very definition of a determinant! In the general case we conclude:

a ax 1 sal . a --b-'" nay 1

--say n

a ax n

Det

Deti~lif (bl, . ,b n ) is an even permutation of (l, ... ,n)


ayS

CI.

-Deti~lif (bl, ... ,b n ) is an odd permutation of (l, ... ,n)


aye

CI.

otherwise

Consequently we obtain the following formula a ax


I

(7.29)

-b-

ay

--say n

a ax n

CI.

b1

Det (ax 0) b n ay"

Inserting this in (*) we get


CI.

Th
(2),",1'"

b (Yo) = I-g(xo)
n
(!)

l ...

bn

Det(ax a ) ay"

But using (7.26) we finally obtain Det()


l1;b l

axCI.

b
n

(yo)

IDet(~)1 ay

I-g (Yo)
(21

E:b

1'"

Sgn (Det[ axCl.]) J-q(yo) aye (2')

bl . b n

350

II
~bl

Consequently we see that up to a sign form as lTal ,I)

... bn has exactly the same

a n! This motivates the introduction of the Levi-Civita

form:

Definition 7 Let M be an orientable manifold of dimension n. Then the Levi-Civita form associated with the chosen orientation is the n-foTIn characterized by the components V-g(x)

[c]

a1 .. a n

={
-V-g(x)

E al .. a

with respect to a positively oriented coordinate system

n
E al .. a

with respect to a negatively oriented coordinate system


n

(If the metric is Euclidian we replace V-g(x) by Vg(x) ).

Observe that the Levi-Civita form is only defined after we have fixed the orientation of our manifold. If we change this orientation then the Levi-Civita tensor is replaced by the opposite tensor, i.e.

(:

'=

-t

For this reason physicists refer to it as a

pseudo-tensor.

Exeroise

7.4.3
We define

Problem: Consider the Levi-Civita symbol E al .. ~

( 1) Show that

(7.30)

a1an E E al ... a n

n!

(2) Consider the cases n=2 and n=3. Deduce the rules

(3) Show that the contra-variant components of the Levi-Civita form

are given by

(positively oriented coordinate system)


1
aj a
n

e:

(negatively oriented coordinate system)

351

II

(4) Let (xl, ... ,xn) be positively oriented coordinates. Show that the Levi-Civita
form can be decomposed as

(7.34) n Let (x1, ,x ) denote positively oriented coordinates on a Riemannian manifold Mn. From (7.34) we learn that the Levi-Civita form can be decomposed as

= Vg dx11l lIdx

If we compare this with (7.24) we see that is nothing but the volwne

form.

Observe that is globally well defined, so that we,on an orien-

table manifold,can piece together all the volume forms on the different tangent spaces to a globally well defined n-form. On a non-orientable manifold this is no longer possible. Observe also that not only generates the volume, but also the

orientation of the tangent spaces. If (~l"" '~n) is an arbitrary n-tupIe then (~l;"';~ ) is the volume of the parallelepiped spanned by
(~l""'~n)' and it is positive if and only if (~l, ... ,Vn) is positively oriented.
The Levi-Civita form can also be used to generalize the cross pron be a Riemannian maduct from the ordinary Euclidian space R3. Let M nifold, and consider a set of (n-l) tangent vectors, Ul,""U _ . n l If we ~ontractthis set with the Lev1-Civ1ta form we obtain a I-form (7.35) Since we work on a manifold with a metric, this I-form is equivalent to a tangent vector
-+ -+

fi, and it is this tangent vector fi which genera-

lizes the usual cross product (Compare the discussion in section 1.1):

Lemma :> The vector ; is characterized by the foZlowing three properties: -+ ... -+ (1) n is orthogonal to each of the tangent vectors u 1 , .. ,un - 1 .
(2) The length ~f
(3)

is-+equal to the "area" of the "paralleLZogram" spanned by u , ... ,u _ , n 1-+ 1-+ -+ The n-tuple (n,u 1 , ... ,u _ 1 ) is positively oriented. n

Proof: ... First we deduce a useful formula. Let v be an arbitrary tangent vector. Then bl b r: un-Iva -+ -+ ...) (7.36) !I(n;~) = nava .gE b bU... = (ViUl;;un_l a 1'" n-l

352

II

The rest now follows easily:

= (~.;~ i . ;~ ) which vanishes automatically since t (1) g(;i~.l 1. 1. 1 n-l is skew symmetric. (2) Consider the normalized vector ti/litiU . It is a unit vector orthogonal to ~ , .. ,~ . Therefore the volume of the parallelepiped 1 -+ n:'l -+ -+ spanned by (n IUnll),u , ,u reduces to the area of the paral4.1;- 1 lellogram spanned by u1, .. ,U _ ' Consequently we get that n 1 -+- .. iUn_l] -+-+- ... ;u -+- - ) = g(n; -+- -+Area[uli .:.( -+nl 11-+-11 n ;u1; nl II-+-I nl) n 1

r +. IInll

g(ni n )

-+ -+

IInll

-+

(3) We immediately get that


(ti;~ i ;~
1

n-l

g(ti,ti)

> 0

Lemma 3 motivates that we define n


-+

-+

to be the cross product of

Ul, .. ,U _ n l
-+

-+

and we write it in the usual manner


-+
-+

n = Ul x ,.xu _ n l

Exercise

7.4.4

Problem: Consider the Euclidian space ~+1. (a) Let (~'~l'" .,~ ) be an arbitrary (n+1)-tuple. Show that the volume of the parallelepiped gpanned by (~';l""'; ) is given by the familiar formula
n

(bl Let Nfl be an orientable n-diemensional manifold in Rn+1. At each point in M we select a positive oriented orthonormal frame (til""'~ ) in the corresponding tangent space. Show that the unit normal vector ield
-+

n=u1x ...x U n

-+

-+

is a globally smooth normal vectorfield on M (Compare with exercise 7.4.2)

7.5

THE DUAL MAP


As another important application of the Levi-Civita form we will

use it to construct the dual map *, which allows us to identify forms of different rank. This gives a greater flexibility in the manipulations of various physical quantities.

353

II

Let F be a k-form with the components considering the

We start by components

Then we contract this tensor with the Levi-Civita form and obtain !,/g(x) e: b b Fbl . bk [PositivelY orientated] k. alan-k 1'" k coordinate system!
1 is included to make life easier! where k! As e:al"'~ is skewsymmetric we have thus produced a form of deThis form is called the dual form and it is denoted *F gree n-k * or sometimes F

Definition 8 Let
F

*)

be a k-form on an orientable manifold


*F

n M

with a metric

g.

Then the dual form nents

is the (n-k)-form characterized by the compo-

~!/g(X) e:al"'~_k
(7.38)(*F)

bl .. b k Fbl . bk

(~~~~i
metric)

al'"

n-k

~!/-g(x) e:

alan _ k b1b k F

bl' .. bk (Minkowski metric)

with respect to a positively oriented coordinate system. Exercise 7.5.1


Problem: (a)

Let ~ be a scalarfield on a Riemannian manifold and let be positive oriented coordinates. Show that

(xl, ... xn)

(7.39)
(b ) Show that

*~ = ~;gdxlA ... Adxn


E

= *1

(Euclidean metric) (Minkowski metric)

Exercise 7.5.2
Problem: Show that the contra variant components of *F are obtained by contracting the components of F with the contra variant components of 1:. i.e. _1_ e: a l , ~-k b l , bk F (Euclidean metric)

(7.41)

(*F)

ala -k
n

k!/g

b l .. ,bk

- 1

e:

a 1 , .an - k b l ,bk

k!/-g

b " .b k l

(Minkowski metri c )

*)

This definition differs slightly from the one generally adopted by mathematicians. See for instance Goldberg [1962] or de Rham [1955].

354

II

In this way we have constructed a map from obviously linear (7.42)

hk(M)

to

An-k(M)

It is called the dual map or the Hodge-duality. By construction it is

*(F+G) = *F + *G

*(AF) = A(*F)

But a more interesting property is the following one:


Theorem 5
(7.43)

**T **T

(_I)k(n-k)T

(Euclidean metric) (Minkowski metric)

= -(-l)k (n-k)

So up to a sign

is it's own inverse. Especially the dual map,

* :

Ak(M) ~ An-k(M), is an isomorphism.

Proof: For simplicity we only consider the case of a Minkowski metric. Let F be an arbitrary k-form. Then *F has the components

1. ,I-g
k!

E:

a l '~-k b l .bk

Fbl' .. b k

We want to compute the components of **F The first thing we must do is raising the indices of *F, i.e. determine the contravariant components of *F. According to exercise 7.5.2 they are given by the formula

Then we shall contract the Levi-Civita form with factor! )

*F (remembering the statistical


(*F)al .an- k

(6)

(**F)

c l .c k

I =; : g E: c (n-k)! 1
E:

l .ck a l .an- k
E:

a l .an - k b l .bk

k!(n-k)!

cl,,c k al"a n _ k

bl".b k

The rest of the proof is a combinatorial argument! First we observe that

But

is skew-symmetric and therefore

Sgn [bcl ... b = k! F ck ] F 1,, k bl" .bk c l " .c k and if we insert that into (0) we are through.

From theorem 3 and 5 you learn the following important rule whenever you deal with the operators in the exterior algebra:

355

II

"Be wise - appZy them twice"


We shall evaluate the sign associated with the double map often that it pays to summarize them in the following table
**

so

(7.44)
**F

Euclidean metric k even


F F

(7.45)
**F

Minkowski metric k even


-F -F

k odd
-F F

k odd
F

n even n odd

n even n odd

-F

Observe that the dual form,

*F ,involves the Levi-Civita form. Thus

it depends on the orientation of our manifold. If we exchange the orientation with the opposite one, then the dual form is replaced by the opposite form
*F ~ '= For this reason physicists also call the dual form

F -F

*F

a pseudo-tensor.

Introduction: Let MP be an orientable Riemannian manifold. A differential form F is called seZf dual i f it satisfies *F = F and anti self dual if Problem: Show that self dual and anti -self dual forms can only exist in Riemannian spaces of dimension 4, 8, 12, 16, .,.

&eT'Cise 7.5.3

Problem:

Let MP be an orientable Riemannian manifol~ and A a+l-form+on M. Then A is equivalent to a tangent vector a. Let (ul""'~-l) be an arbitrary in-l)-tuPle. Show that the volume of the parallelepiped spanned by (iIl , ... '~n-l,a) is given by -+ -+ -+ -+ -+ -+ -+ -+ Vol[ul;;un_l;al = *A(ul;;un _ l ) = ai u l x .. x un - l ) This gives a geometric characterization of the dual form *A'

Exercise 7.5.4

Exercise 7.5.5
Problem: a)

*(dx
b)

Show that al
A Adx

ak

_ Itg
) -

(n-k)!

E..

ll" .In-k

h .. jk

jlal jkak i l i n- k ... g dx A . . Adx

Consider spherical coordinates in Euclidean space dual map is given by


*dr

R3.

Show that the

= r 2 SinedeA<ip

*de = Sine<ipAdr

*<ip = -:Slne drAde

356

II

The scalar product between two I-forms is given by and am-form

gijA.B.
1.

Clearly this is a special case of a contraction between a k-form

where
S

jl jm (k-m) It

The result obviously is a differential form of degree would be nice to be able to express such

contractions in a coordinate

free manner, and as we shall see this is possible using the dual map.
Thus the following pattern emerges: When you have a metric on an orientable mani-

fold, you can use this metric to construct a dual map has been "coded" into the dual map.

but once

you have constructed the dual map, you can forget about the metric: It The first thing we will construct is the scalar product between 1forms. Let therefore lar * (*AABl This scalar is a coordinate invariant formed out of and therefore it must be proportional to Ai Bi , other coordinate invariants you can make out of Ai and Ai' Bi ' Bi and (since there are no gij' A *AAB and B be two I-forms. Then *A is an (n-l)-form, and is an n-form. But this can be dualized to give a sca-

. !) Okay, that was a rather abstract argument. Let us work 1 l.n out the coordinate expression for *(*AABl in, e.g., the Euclidean case C

;g

El.'

*(*AABl

i.e.

*C

*AAB

Here the right hand side is characterized by the component (*AAB) 12 .. n


1 (n-ll![(*Al 12 ... (n-ll(Bl n + ... - . 1

(*Al12 ... n_l(B)n+(-lln-l(*Al2 ... n(Bll + ..

whereas the left hand side is 'characterized by the component (*Cl1. .. n From this we conclude *(*AABl is not manifestly Observe that A

= ;g

E1. .. n C

= ;g

*(*AAB) which justifies our claim. The expression symmetric in A and A Ai B, but that is the dualization of
has been raised,

AiB. 1.

neither!

is reflected in the fact that the index of ~ Ai We have thus shown

357 Theorem 6 Lt
( 7. 47) ( 7.48)

II

A and

B be l-forms, then * (*AI\B) * (*BI\Al * (*AI\B) * (*BI\Al

AiB.

l.

(Euclidean metria) Il (Minkowski metria)

-AIlB

This theorem has an important consequence. It tells you how to reconstruct the metria from the dual map! Sometimes it is preferable to omit the last dualization, which in both cases produces the formula

(7.49)

*AI\B = *BI\A = AiB. l.

The above formula suggests a generalization of scalar products to arbitrary differential forms. Before we proceed we need a technical result: Worked exeraise 7. 5. 6 Problem: Let T,U be k-forms. Show that
(7.50)
1 il *TI\U=,T k. .. i k U..

l.l ... l.k

From exercise

7.5.6

we see that

*TI\U

is symmetric in

T and

and that it represents the coordinate invariant

The statistical factor is very reasonable, since it removes "double counting" . Let us introduce the following abbreviation for the scalar-product of two k-forms: (7.51) Then we have succeeded in generalizing theorem 6 to arbitrary k-forms: Theorem 7 Let
(7.52 )
(7. 53)

T,U

be k-forms, then

(T I u)

-- T

i l " .ik

k!

il .. i k

(Eualidean metria)

1 Ill" ' T k.

. Ilk

U (Minkowski metria) )Jl,,)Jk

Exeraise 7.5.7 Problem: Show that


a)

(dxalm(

= gab

(dxa,wPldxcNlxd ) = gacJX1 - gad,fc Remark: It can be shown in general that [] al ~ bl ~ 1 iljl ~jk !-".~ (7.54) (dx 1\ ... Nlx Idx 1\."l\dx ) = k!'1 ".g Sgn l.l'''~
b)

358

E:z:eT'Cise 7.5.8
Problem: Let T rank. Show that be a n-form on

p,f.

i.e.

is a differential form of maximal

Consider a single point

on our manifold. To this point we have at-

tached the finite dimensional vector space k-forms. By construction the map

A~(M)

consisting of all

(T,U)

(TI u)
A~(M)
k

is a symmetric bilinear map on the vector space

Thus there

is a good chance that it actually defines a metric on following holds:

Ap(M) . To check this we--Ulust show that it is a non-degenerate map. Actually the

Theorem 8 (aJ Let

M be a Riemannian orientabZe manifold. Then the scalar

product (T.UJ (TluJ defines an Euclidean metric. (bJ Let M be a pseudo-Riemannian orientable manifoZd. Then the scalar product
Proof: (a) cients gij (Po) Then the scalar product We can choose a coordinate system at Po with metric coeffi-

(T.UJ

(TIU) defines an indefinite metric.

ij .

(Tlu) reduces to (TI u) = 1:


i < ... <i l k

which is obviously positive definite. (b) Let T We must show that the inner product
(I)

is nondegenerate.

be a k-form that is perpendicular to all other k-forms, i.e.

o = (TIU)
Choose

for all

u.

U with the component

U12 k = I

and such that all other

components I2 k T = 0

Ui ... l where il< .. <i is zero. Then we deduce that k l k In a similar way we can show that all the other contraT

variant components of

vanish. But then

Observe that the dual map tively onto the vector space fore no great surprise:

* maps the vector space A~ (M) bijecAn-k(M) The following the8rem is therePo

359

II

Theorem 9 (a) Let M be a Riemannian orientable manifold. The dual map an isometry, i.e.
(7.55) (*T I *U) (Tlu)

is

for all k-forms (b)

and

U.

Let M be an orientable manifold with a Minkowski metric. Then the dual map * is an anti-isometry, i.e.
(*TI*u)
T

(7.56)

-(Tlu)

for all k-forms


Proof:

and

(The Riemannian case) (*T I *Ul *(**TA*Ul

(-l)k(n-kl*(TA*Ul

*(*UATl = (UITl = (Tlu) Observe that we have two signs involved. One coming from the double dualization, and the other coming from the exchange of T and *U

So much for the scalar product. Now we proceed to investigate the more general contractions: Let
k~m.

be a k-form and

U an m-form where

Then we can extend the above machinery and we immediately see

that becomes a out of (k-m)-form. and

But the only general coordinate invariant (k-m)-form you can make U jl " ' jm is the contraction k _ m jl .. jm T'U jljm U In the

=r T . .
1

1 m.

Thus i.e.

*(*TAUl T'U

must be proportional to this contraction.

following we will use the abbreviation

for the contraction

is the (k-m)-form characterized by the components and the differential form *U = U U

(7.57).

Observe that dualization itself is a contraction between the Levi-Civita form

you want to dualize, i.e.

Ea:ereise 7.5.9
Problem: Show that (Euclidean metric) (MinkowSky metric)

Wor'ked ea:ereise 7.5.10(For combinatorial fans!)


PrOblem: Show that (Euclidean metric) (Minkowsky metric)

(7.58)

360

II

7.6 THE CO-DIFFERENTIAL AND THE LAPLACIAN


Once we have the dual map at our disposal, we can construct another "natural" differential operator:

Definition 9 For an arbitrary k-form T the &T is the (k-1)-form defined by


~T ('1.5~)

co-differentia~

of

T,

denoted

(_l)k(n-k+l)*d*T

(EucZidean metric) (Minkowski metric)

~T =-(_l)k(n-k+l)*d*T

In this way we have constructed a differential operator &: Ak(M) ~ Ak-l(M) It is obviously linear, since it is composed of the linear The sign of Ii is a convention which makes forthmaps * and d coming formulas easier to work with. Observe that Ii has an important property in common with the exterior derivative d

Theorem 10
(7.60)

o
=

Proof:
Ii~

(*d*) (*d*)

*dd*

where we have used theorem S.

So once again you see that it pays to apply the operators twice! Note that there is no formula for Ii analogous to the Leibniz rule. This is due to the fact that *(TAS) cannot be reexpressed in a simple way (See however exercise 7.6.1 for a possible generalization).

Eft{trcise '1. 6. 1 Problem: Prove the following generalization of Leibniz rule: Ii(T'U) = (_l)n-m liT'U +(_l)n-k T'du (7.61 ) (Yes, it is an ordinary exterior derivative in the last term!)
Let T be a differential form. We have previously Seen that we can symbolically consider d as a I-form and the exterior derivative of T as a kind of wedge product: "ctr = dAT". In the same spirit we can consider the co-differential of T as a kind of contraction ~ = (-l)kd-T This follows from (7.58-59).
E:urei3IJ ? 6 2

Problem:

(7.62) (7.63)

Prove the following identities *~T = (_l)n-k+l d*T *ctr

= (_l)n-k &*T

361

II

Observe that the above formulas hold both for the case of an Euclidean metric and a Minkowski metric!

If you think of think of out what


(a)
~

as generalizing the gradient, then you may

as generalizing the divergenae. To see this we will work is in a few cases. For simplicity we work with an Euclia
B

dean metric:
Let cp be a a-form, then 5cp = gree of a form. (b) Let A be a I-form. If We put B = (-l)n*d*A i.e. has the If A components A. then 1 ponents Thus is obvious because SA , then B *B = (_l)n d*A S lowers the de-

is the scalar given by

*A

is an (n-l)-form characterized by the com-

(*A)12 ... (n-l) = .Tg E: 12 ... (n-l)i Ai = .Tg An (where the other components are obtained by cyclic permutation). Taking the exterior derivative, we get an n-form characterized by the component (d*A) 12 ... n
= _ 1 _ [d (*A)

(n-l)!

2 .. n

....

. ...

dl (*A)2 ... n + (-1)n-l d2 (*A)3 . nl + d3 (*A)4 ... n12 + (-l)n-lal(lgAl) + (-1)n-l a2 (lgA2) + (_l)n-la.(/i Ai)
1

Consequently

(-l)nd*A is characterized by the component [(-l)nd*AJ = _d.(lgAi) 12 ... n 1 On the other hand *B is characterized by the component [*BJ 12 ... n = .TgB

We thus get:

5A

= _..l..

d.(/~i)

Ii 1 Observe the sign which is conventional: 5 really is minus the divergence! (c) Let F be a two-form. If we put G SF, then G is a one-form: G Here *F

= *d*F

i.e.

*G = (-l)n-ld*F

is an (n-2)-form characterized by the components: 1 r F ij 1 '[F(n-l)n Fn(n-l)J (*F)12 .. n-2 = 2! vg E 12 ... (n-2)ij = 2! vg = Ii F(n-l)n

In the rest of the discussion we will put n=4. This makes it easy to see what is going on, and the calculations can easily be generalized. d*F is then a 3-form characterized by the components: (d*F)123

=~! [d l (*F)23 - dl (*F 32 )

+ ... -

... J

dl(*F)23 + d2(*F)31 + d3 (*F)12

dl(/i~4)

d (/gr24) + d (/iF34 ) 3 2

362

II

But here we can artificially introduce t~~ term d4(;gF44) since it vanishes automati cally due to the skew symmetry of F2J Thus we have obtained (d*F)123 = di(fgFl ) and this formula is immediately generalized to If (d*F)12 ... (n-l) = (-l)ndi(;gFin) G is a one-form characterized by components Gl ... n .
(*G)

l2 ... (n-l)

,;g E l2 ... (n-l)1

Consequently we have

Let us collect the results obtained in the following scheme:

( 7.64)

Euclidean metric Il<j> IlA = (5F)j


1

Minkowski metric 5<j>


.

o-rorm: <j> l-form: A 2-form: F

=
d

0
i (Ig Al.) Fij) SA

;q

= - __1__ d

r-a
1

(i=g A~)
F~v)

= - .l:.. d. (Ig _Lci l.

(IlF)v =

r-a

d (i=g
~

A general coordinate invariant expression for the components of

liT

is hard to derive. The details are left as an exercise for those

who are especially interested in combinatorics:

Worked exercise 7.6.3


Problem: given by Let
T

be a k-form. Show that the contravariant components of

IlT

are

(7. 65 )

Once we control the co-differential this is particularly easy. Consider a


~

II

we can now generalize the <j>

Laplacian to arbitrary Riemannian manifolds. For a scalar field Cartesian coordinates. Then the Laplacian is defined as:
L\.<j>

Euclidean space with the usual

= ;7; v

(V<j

<.. ~ i=l d (xi) Z

dZ,j,

In the language of differential forms this is immediately generalized to:


(7.66)

-Ild<j>

363

II

Consequently (7.66) is the covariant generalization of the expression

aiai~, which is valid only for Cartesian coordinates.


This suggests that we could generalize the Laplacian for arbitrary differential forms in the following way:
?

-/; ;, Sd
but unfortunately this is not correct. To see this we consider a one-form , A. characterized by the components
I

Then

SdA

is a I-form

a. [/g(aiAj_ajAi )]
l.

/g

~ a.[/gaiA j ] + ~ a.[/gajA i ]

;g

l.

/g

l.

It is the last term that spoils the game! On the other hand we may consider the differential operator know that SA is the Q-form - ~ a. (/gAi) /g l. and therefore (dSA)j

dS

When i t operates on

A,

we

= -aj(~a.

/gl.

(/gAi

In a Cartesian coordinate system reduce to:

/g

I,

and the two expressions

-a.aiAj + aja.A i
l. l. l.

aja.A i

Adding the two expressions we therefore get that characterized by the aontra-variant components -a.aiA j l. Observe that

(Sd + dS)A

is

dS

automatically vanishes for scalar fields! The correct

generalization is therefore apparently given by symmetric expression: Definition 10 *)


(7.67)

-A

Sd + dS

Worked exeraise 7.6.4 (For combinatorical fans) Problem: Show that in a Euclidean Space with Cartesian coordinates, the coordinate expression for the Laplace operator reduces to
(AT)'

ll l.k

= (a. aj iT.l.ll.k . J

*) This definition differs in sign from the one generalZy adopted by


mathematiaians. See for instanae Goldberg [1962] or de Rham [1955].

364

II

Exercise 7. 6. 5
Problem: Show that the Laplacian commutes with the dual map, the exterior derivative and the co-differential, i.e.

t.*

= *t.

dt.=M

If we proceed to consider a pseudo-Riemannian manifold with a Minkowski metric, then the above considerations are still valid, except that

-(&d+d5)

here represemts the d'Alembertian

o.

In an inertial

frame it is given by the expression


o

= oj.lo

j.l

= atT

02

and its co-variant generalization to arbitrary differential forms is (7.68) -0 = dS + Sd Okay, let us return to manifolds with arbitrary metrics. In analogy with the exterior derivative we can now introduce two important types of differential forms

Definition 11 (aJ A differentiaZ form F is co-closed if its co-differential vanishes, i.e. SF = O. (bJ A differential form G is co-exact if it is the co-differential of another differential form F, i.e. G 5F
Observe that a co-exact form is automatically co-closed, since G implies that SG SF

S2F

Exercise 7.6.6
Problem: (a) (b) Show that Show that

F is co-closed if and only if *F is closed. G is co-exact if and only if *G is exact.

In the rest of this section we consider only Riemannian manifolds. The generalization of the Laplacian to a Riemannian manifold suggests the following definition:

Definition 12 A differential form Exercise 7.6.7


Problem: (a) (b)

is called harmonic if

t.T

O.

Show that the constant function 1 and the Levi-Civita form are harmonic forms. Let T be a harmonic form. Show that *T, ctr and liT are harmonic forms too.

365

II

A very important class of harmonic forms are those where both the exterior derivative and the co-differential vanishes:

Definition 13 A differentia~ form T is called primitively harmonic if it is both closed and co-closed, i.e. dT = ~T = 0
A primitively harmonic form is obviously harmonic, but the converse need not be true. This is wellknown already for scalar functions. Consider

R3 ,{O}
d~

-+

and put

~(x)

I =r
~

Then i t is not constant and is not primitively harmonic. On that


M

therefore

Consequently

the other hand, i t is harmonic as you can trivially verify. In the above example it is important to notice is not compact. For a compact Riemannian manifold we shall show later on that all harmonic forms are automatically primitively hanronic (see chapter 8, theorem 8).

Exercise ?

(j.

8
d~.

Problem: Let A be an exact I-form, A = nic i f and only if ~ is harmonic.

Show that

A is primitively harmo-

Worked exercise 7. (j. 9 (n Complex calculus on manifolds") Notation: Let Me RZ be a two-dimensional manifold, i.e. an open subset of RZ.
We identity R2 with the complex plane (x,y)
+>-

C through the usual identification

z = x+iy

We introduce complex-valued differential forms by admitting the components to be complex-valued functions, i.e. a complex-valued differential form is a skew-symmetric multilinear map

F : Tp(M) x x Tp(M)
Consider the complex-valued I-form

c .

w = [f1(x,y)+ig1(x,y)ldx + [f2(x,y)+ig z (x,y)ldY


If we introduce the basic complex differentials

dz = dx+idY,
then we can expand where hl(Z) If and
h~)

d;:

= dx-idY

as

= ~[fl+g2l
db

~[gl-fZl

and hz(z)

= ~[fl-gzl

~[gl+fzl
dh

ay

is a complex valued function we can introduce the partial derivatives through the expansion:

dZ

= ~z z a

~z ;: i.e. ,d = 1(...... -i......) a aZ 2 aX ay

and ...... = (...... + i......)

az

ax

ay

We take it as well-known that an analytic (holomorphic) function in M is characterized by the following two equivalent characterizations (1) h can be expanded as a power-series around each point Zo in M. i.e. h(z) = ~ au(z-zo)n for z sufficiently close to zO. (2) ~~ "'0 n(Cauchy-Riemann's equations)

366
Problem: Consider the l-form w

II

= h(z)dz:

(a) Show that the real part is the dual of the imaginary part. (b) Show that h(z) is analytic (holomorphic) if and only if w is closed. (c) Let us assume that h(z) is analytic (holomorphic). Show that the real

part and the imaginary part of h(z) are harmonic functions. h(z) be holomorphic. Show that ~ is holomorphic too. (e) Show that h( z) is holomorphic if and olhy i f dh is anti -self dual in the sense that
(d) Let

*dh = -idh (Compare with exercise 7.5.3).

7.7

EXTERIOR CALCULUS IN 3 AND 4 DIMENSIONS


We start by looking at the ordinary Euclidean space

R3.

Here we have the stan-

clard veator anaZysis: The only admissible coordinate systems a~e the Cartesian coordinate systems. A vector a is characterized by its Cartesian components (ax,ay,a z ) We do not distinguish between co-variant and contra-variant components as the metric coefficients are given by
[g. ,] =

lJ

[:

because we restrict ourselves to Cartesian coordinates. There are two kinds of products: the scaZar produat ... a b = a b + a b + a b

::]
z z

... e

~
... e a
p

~-----.. y
Fig. 130

...
~

x x

y y

and the arosB pl'Oduat


x

with coordinates

with a lot of well-known

algebrai~

[aybz - azby ' az bx - ax bz , ax by - ay bx ] properties. Then there is a differential operator

V = (1-, 1- , 1-) ox ' oy , oz


Using this differential operator we can attack a scalar field tor field the
~,

producing a vec-

!t ) ' v.. W1.th coord' mates (!t. ox' 2.1. oy 'oz and we can attack a vector field !, producing a scalar field :I: ... 3a 3a oa the divorgml!!e: V a = oxx + 3Y" + az'" or a vector field
gr'cu:J;L.en :
._.1.'

:1:..,

the aurL:

Q x i with coordinates [3a 3a 3a 3a 3a _~]. oy'" - ozY 3x z oy~ This is the scheme we are going to generalize. In the e~erior algebra we can use any coordinate system: Cartesian coordinates, spherical coordinates, cylindrical

azX -

a,cv

367

II

coordinates, etc. For simplicity we shall, however, restrict ourselves to positively orientated coordinate systems. This fixes the components of the Levi-Civita form e to be IS E. 'k' We distinguish between covariant and contravariant components and we use thel~etric coefficients to lower and raise the indices. In exterior algebra we play with forms. As n=3 we have O-forms, I-forms, 2forms, and 3-forms. The scalars are represented by O-forms, but via the dual map we can also represent them by 3-forms! The vector fields are represented by I-forms, b~t via the dual map we can also represent these by 2-forms. Thus an arbitrary form in R is either associated with a scalar or a vector. To investigate the dual map we use formula (7.38). The results are comprised in the following scheme:

(7.69)

The dual map in F

R3 *F

Covariant components of O-form:


<P

Covariant components of 3-form: *$ : <"$ ) 123 = vg$ 2-form: E3 *E :

I-form: E : (E I , E2, E3) 2-form:

r. [-:' E2

-E 1

EI -"J 0

I-form: B12
0

B:

[-:"

-'" BZ3 J
0

*B : /g[B 2 3 , B3I, BIZ]

B31

-B23

3-form:
G : (Gh23= G

O-form: 1 G *G : 7g

Most of the scheme is selfevident but you should be carefu1d when you dualize a three-form. A three-form is completely characterized by just one component, which we choose to be the 123- component. Applying formula (7.38) we get: lJ 1 . 'k Now we use that the contravariant components of ari. given by 72 ElJ (See exercise 7.6.1). This is where the mysterious factor 72 comes fr~. Raising the indices of c and lowering those of G produce the equ'va1ent formula But here all permutations of reduces to *G = 3T Tg E Gijk (ijk) give the same contribution so that the formula
1 1 1

*G _ JT 1
1

vg

rE.

'k G

ijk

ijk

*G = 7g E 23G123 = 7g G You should also observe that in Euclidian spaces of odd dimension we always have **F = F for a form of arbitrary rank. (Compare this with the scheme (7.44)). Armed with the dual map we can nOw investigate the vector analysis. In the following A, B and C will denote 1-forms. Let us first consider the wedge product: A A B is a 2-form with the components (A;Bi - ~Bi)' This obviously generalizes the us~a1 cross product. Although it is a '-form, A A B actually represents a vector whlch we ean find by dualizing:

368

II

If we use Cartesian coordinates, then /g = l,~ ani the expression above reduces to the components of the usual cross product a x b. ~onsequently * (AAB) is the strict generalization of the usual cross product ~ x b. Due to the fact that 1forms are odd forms, i.e. anticommute, we immediately recover the characteristic properties of the cross product:
~

axb=-bXa

~x!=O

Next we consider the scalar product between two I-forms:

This obviously generalizes the usual scalar product a b. More generally we can apply contractions between forms. Let. F be a 2-form, then F' A is the I-form characterized by the components F.. AJ. Here F represents a vector: F = *B, i.e.
F

ij -

r.: g

lJ

ijk

Bk

and therefore

F A is actually characterized by the components

Ii E. 'k B A 1J
which are the

k j

Ii E"kAJB 1.J
~

. k

covariant components of

aX b.

E:r:!'cise 7. 7. 1 Problem: (a) Let F and that the scalar product

G be 2-forms representing the vectors

and

b.

Show

represents the ordinary scalar product: a ' b. (b) Let F 4e a 2-form and consider the wedgeproduct F A B. Let F represent the ~ect~r a and show that the 3-form F A B represents the ordinary scalar product a ' b. Then we consider the trinle product , represents a scalar. We alre~dy know that x b. To find the meaning of (A A B) A the l23-component:

CA A B) A C' which A A B represents C we use exercise

is a 3-form, i.e. it the cross-product 7.2.1 from which we get

But A A B is the dual of the cross product, i.e. We thus find

(A

A B)12=

;g(a

b)3,

etc.

[CA

B)

A CJl23 =

/g[(a x b)3 C3

(a x b)lCl + (~ x b)2C2J

Dualizing this we finally get the scalar


*[(A
A

B)

C J=

(a

b)iC.

1.

and you see that [(A A B) A CJ generalizes the triple product x b). ~ But A A B is a 2-form, and therefore it commutes with C. We thus get

(a

If we then dualize, this corresponds to the rule

... (a x

369
... b)

(~ x~)

b = (b
C.

~)

... a

II

These rules are often stated in the following way: In a

trip~e

product it is allowed

to interahange dot and arOS$. Suppose now we have three I-forms: triple product:
(A

A, Band
A

Then you can form another

B) C

This is a~th~r one-form! A A B is the dual of the cross product a x b, but then [*(a x b) J C is still another cross product, according to the preceding discussion! Thus (A A B)C generalizes

It is wellknown that the triple product

... ... (a x b)
j

... x c

(~x

b)

satistifies the rule

In the exterior algebra this follows straight forward when you work out the nents off (A A B)'C '.\.B. - A.B.)C
1.

compo~

From this you iDDl1ediately read off the formula

(A " B) C

A( BJ C ) - (

AI C )B

If you compare this with the proofs in the conventional vector analysis, you will learn to appreciate the exterior algebra! Let us collect the preceding results in the following scheme

(7.7 0 )

Scalar product and wedge product Forms: Components: A.B


1-

.. ""
<a
~
0

Scalar product Contraction Wedge product Wedge product

(A

IB )

<.!>

FA A AB F AA

j F .. A
1.J

--- ---------------- ------------- -----------------------------------Conventional vector


analysis Exterior algebra (A I B)

. ..
0::
~

0::

A.B. - A.B.
l.

E-<

l.

Fij~ + FjkAi + FkiAj

... ... a b ... ... a x b


0

~
E-<
U
H H

(1

b) ..

... c

(1 " b)
(txb)"t
=

x c

...
t(bt)

AAB (A A Bl A C

(A A Bl

.C

."

(~xb).c = (txt)'b = (bxt).! (~.~) b -

A"B A C=CAAAB=BACAA (A A B)' C = ACB I C)

- <AI

C>B

370

II

We now oroceed by investigating the differential calculus. In the exterior algebra we h~ve the exterior derivative d. It converts a scalar field ~ into a one-form with the components

d,

a~

(axl> Thus

a?' 3x3

a,

a,)

dA with components

d~

generalizes the gradient ~~

It converts a I-form A into the 2-form

But- this 2-form obviously represents the curl

V x t.

Finally d converts a 2-form into a 3-form. If the 2-form represents a vector field, i.e. we consider the 2-form *A. then the 3-form d*A represents a scalar. It is this scalar we want to examine. As d*A is characterized by the components ai(*A)jk + aj(*A>ki + ak(*A)ij the 123-component is given by

Dualizing this we find that

*A

represents the scalar:

*(d*A)

=~ rg

a.(lgAi)
l.

If you remenber that I:i = 1 in ~art~sian coordinates, you immediately see that d*A generalizes the divergence g. a. Notice however that we may express the divergence more directly by means of the co-differential 5. If A is a I-form we know that

-&A

= ....,...

1 rg

a. (/gAl.) l.

(7.59) we have

(Compare (7.64. This is in agreement with the result above since according to -5 .. *d*.

Exercnse 7.7.2 Problem: (a) Let F be a 2-f~rm r!pre~enting the vector a. Show that 5F rerepresents the curl V x a. (b) Let G be a 3-form representing the scalar field ,. Show that 5G represents the gradient
~

V,.

We know that the exterior derivative

d has some simple properties

rJ.2 =
d(F" G) = d F"G+
(-1)

deg F

F"dG

Almost all of the identities involving grad, div and curl are special cases of these two rules. see the scheme two pages from thisl

371
(7.71)

II

The exterior derivative in R3. l-form:

O-form:

4>
l-form:
al ... p
OJ bD

d4> : (a14>, d24>, 3 3 4

2-form:
0

"1 ~ -"zA1
0

xxx "Z A3- d3AZ


0

rl

A: (AI, A2, A3 ) 2-form: *A:

... 0 ..... ... OJ


~
+'

al

Ii

~,

xxx ",A,-3 A, 3-form:

dA,:

xxx

A3
0

-A'j
AI
0

Z _A1 A

i (d*A)IZ3 '" "i (y'gA ) 4-form (Zero !)

3-form:

- - --------------------Conventional Vector analysis Gradient:

G : (GhZ3= G

-------- ------------- -------------------------------Exterior algebra Covariant formulas

dG

= 0

... al
0: 0 A

i>o

VtP Divergence: 'i/ a


~ ~

dtP

a itP
1 ai(lgAi ) n

..... +' ..... "

- SA
dA

Curl:

vx;:
Laplacian: lltP

.rg E:ijkaJ A
1 .

- &H

~d i Vial.,)

The preceding discussion of the dual map, the scalar product, the wedge product, and the exterior derivative should convince you that exterior algebra is capable of doing almost anything you can do in conventional vector analysis. But exterior algebra is not just another way of saying the same thing. It is a much more powerful machinery for at least two reasons I 1. 2. It works in any coordinate system, i.e. it is a covariant formalism. It works in any number of dimensions.

372

II

(A)
l.

d2 = 0
If 'We apply

(A)

to a scalar field
d(d.p) = 0

.p.

'We get

(*)
0 .... '" ....
()

<l

As d.p represents the gradient of .p, the second curl . Thus (*) general ize s the rule
~

represen ts the

....

V x (V.p)

..:

'"

0.

.
we get

2.

If we apply (0)

(A)

to a vector field A
d(d A) : 0

As dA represents the curl of A, the second d represents the di vergence. Thus (0) generalizes the rule x!) = 0 gF FA dG dF A G + (_l)de to scalarfields we get

V (V

(B)

d (F A G) =
If 'We apply

l.

(B)

d(,l/I) = (d,)l/I + <p(dl/l)

which generalizes

~('l/I)
2.
If 'We apply

(V.p)l/I + <P(Vl/I)
(B)

to a scalarfield
=

.p

and a l-form A 'We get

ilJ,.pA) Which generalizes


~

(d.p)AA+<pdA

V x (.pa)

(V.p) x a + <l>V x -;
(B)
=

.;

.... '" .... ....


()

<l 0

3.

If we apply

to a scalarfield
d ,A (*A) + .pd *A

,
,

and a 2-orm *A we get

ilJ"

* A)

'"
4.

which generalizes
~ (.p!) = V.p
If 'We apply
(B)
=
~ ~

a + ojlV

to l-forms we get dAAB-AAdB

d(AAB)

,
x

which generalizes

V (!

ib

(V

x 1) E"

- !. (V

E")

373

II

The only drawback in comparison with the conventional vector analysis arises when you discuss the equations of motions for particles. In conventional vector analysis Newton's equation of motion for a particle moving in a potential u(~) is simply given by m= -Vu dt 2 This equation of motion can be solved very elegantly in various cases using the vector calculus. It corresponds however to the coordinate expression
(7.72)

d 2 :t

(7.73)

dt2

which is valid onLy in cartesian coordinates. To geometrize the formula (7.72)we must first extend the above coordinate expression to a aovaria:n.t formula!
Exel'aise 7.7.3
Problem: Show that the covariant generalisation of (7.73) is given by
2 k m d x =

-tnt:<

dx

dx

li

dt 2

ij dt

dt

aU axi

3 where r ~. are the chris toffel fields in the Euc lidian space R They sh6dld not be confused with the Christoffel fields in Minkowski space! (Hint: Use that the equation of motion (7.73) extremizes the action t2 did j
S =

![

t1

!mg .. (x) .....!.-......!.-. ~J

dt

dt

U (x)}! t

and copy the discussion in Section 6.7 ).

field

But a covariant formula explicitly containing the Christoffelfalls beyound the scope of the exterior algebra.

Worked exeraise 7.7.4


Problem: Compute the Laplacian in spherical coordinates.

In the case of the Minkowski space the dimension is n = 4. Thus we have O-forms, I-forms, 2-forms, 3-forms, and 4-forms. Scalars are represented by O-forms and 4-forms. Vectors are represented by I-forms and 3-forms. Finally we have 2-forms at our disposal. They represent a new kind of concept, which did not exist in conventional vector analysis, but which plays an extremely important role in 4-dimensional geometry.

374

II

The effect of the dual map is comprised in the following scheme:

(7.74)
a-form:
~

The dual map in Minkowski space: 4-form: **$ = - $ *$ : (*$) 0123 = 3-orm: **A =A *A : (*A) 012 = 2-form: **F a *F : r-g -F -F -F 1-form: *6
:
~

r-g

1-form: A: (Ao, Ab A2, A3) 2-form: a F : -FOl -F02 -F03 3-orm: 6 : (6)asv 4-form: H : (H)0123= H FOI a -F12 F31 F02 F03 FI2- F31
0

.;::g A3
F 23
F F

etc. 12 F 02 -F 01 F a

-F

31

23 31 12

a -F F 03 02

03

F23

-F23 a

-F

01

**6 6 r-g[G1 23 , -C 23O , G301, _GO I2 ] **H


=

O-form: *H :

-H

-~

The only thing your should be careful about is when you dualize a fourform. The discussion of this is left to the reader (Compare the discussion following the scheme (7.69. The rest of the scheme should be self-evident. The minus signs associated with the double dualization are characteristic for Mi,nkowski spaces and you should be careful about them! Then we will investigatejthe differential calculus. First we look at a scalar field $ . It is converted into a vector field,

which we may naturally call the gradient. Then we look at a vector field. We can represent it by a I-form A. Then it is converted into a 2-form ,

375

II

which we may naturally call the curl. But we may also represent it by a 3-form *A. Then it is converted into a 4-form, i.e. essentially a scalar field. Dualizing this 4-form we get
*d*A
= -

fiA

d*A essentially represents the divergence. Finally we can look at a 2-form F which we can also represent by its dual form *F. The exterior derivative of F is a 3-form, which we may call the curl,

Thus

The exterior derivative of *F is similarly related to the divergence of F:


*d*F = - 5F (7.75)
O-form: The exterior derivative in Minkowski space: 1-form:

d.p

(a o"

a1"

a2<1> a3')

1-form:

2-orm:

A: (A o AI' A2 A3)
2-form:

dA: (aa AA - e A Aa)


3-form:

F *F

: (Fall) :

!Fge: aByO

F Yo

dF : (aa F lly + all Fya (d*F) 012 = a a (;.:g fl~


4-form:

+ ay Fall)

etc.

3-form:

(*N 012
Concept: Gradient Curl

Fg A3

a (d*fo)0123 = - ea(r-g A ) Exterior algebra Covariant formula aa'

d,

dA if

ea.

AS -as Aa

aa

Ily

Divergence

- &A

kg

+ all F ya + ay Fall

aa (;.:g A")

D' Alembert ian

- &F - &d,

a -/:; a (;.:g F -g a 1 aa( M -r-; -ga a. <P~

'1

37b

II

7.8

ELECTROMAGNETISM AND THE EXTERIOR CALCULUS

In this section we will first exanplify the manipulations of the exterior calculus by studying electromagnetism in the Euclidean space R3. The electric field strength will be represented by the I-form

where as the magnetic field strength will be represented by the 2form

There are several reasons for this, but let us just mention one of them. In Minkowski space the field strengths are represented by the 2-form:

~l -~l
E2 -B3 E3

-::

~::l
Bl 0

B2 -Bl

and here you clearly see that the electric field strength becomes a I-form while the magnetic field strength becomes a 2-form when we decompose F into space - and time - components;

_:;:_~:r--t---_~~~~_~
space I-form space 2-form

1 I+'E E B
dB

If you remenber that B is a 2-form it 1s not difficult to translate Maxwell's equations: j conventional vector analysis Exterior calculus (7.76) (7.77) (7.78) (7.79)

V B =
>0
~

0 0

C<J

0 .-i
.jJ .,.j

s::

ClB at +
aE _ 2 c at

V V

E = (\"
-p
Eo

~+dE = at

.E=
x B
+

-1 Eo

-t

- &E

I = ;p

-c

SB

-1 J
Eo

377

II

You can now reshuffle the Maxwell equations in the usual way as shown in the following exercise:
Worked exeraise 7. 8. 1
Probe 1m: Use the exterior algebra to re-examine the introduction of gauge potentials the equation of continuity and the equations of motion for the gauge potentials.

Illustrative example:

The magnetia fieLd outside a wire.

Suppose we have a uniform current j in a wire which we identify with the z-axis. It is well known that the current produces a statiC magnetic field ~ circulating around the wire. According to Biot and Savart's law we get:
~

----L21TEoC2

(---=Y-; _x_; 0) x '+y 2 x 2+y2

At this level We do not have to distinguis between covariant and contravariant components. In the exterior algebra the magnetic field is described by the 2-form:
~

-x
.c+y2

-L
21T .,c2

-L.c+y2

x
X~2

...:..Y..-

x"+y2

or, if it is preferable, by its dual form *B : ~ [- ~


21TEoC2 x 2 +y'

x
x ; 0] X 2 +y2

Fig. 131

which reproduces the original vector field directly. It is convenient to decompose Band *B on basic forms
B=

---i2
21TE o C

[--=.Y..- dy X2+y2

1\

dz +

*B =

. [....=:."L~ X'+y2

dx

x
X2+yZ

dY]

Then we introduce a new coordinate system, cylindrical coordinates, which has the symmetry of the problem. First we notice that

378 Cartesian coordinates and cylindrical coordinates are related by

II

x
y

PCOSql pSin<p

dX
i.e.

=z
. 1 --.J.- _ 21TE oC 2 p

tlY dZ

COSql dp Sinql dp + dz

pSinql dql pCOSql dql

Inserting these formulas we get


B

dZAdp

and similarly *B 2 11~ oe2 <ip

Using cylindrical coordinates, we havecons~Uy succeeded in writing Band *B as simpLe forms. To find a vector potential which generates the magnetic field, i.e. B = dA, we simply rearrange the expression for B (compare exercise 7.3.5): dz A d(lnp) From this we read off A
j 21TEoC2 Zdlnp

= ~2' (zdlnp) 1I E oC
=
~z

21TEo~p dp

Ouside the wire we have a static magnetic field. Thus we conclude that outside the wire the Maxwell equations reduce to: dB Consequently B

= 0 and

&B

= 0

is a primitively harmonic form.

Worked exercise 7.8.2 Problem: Show that B is almost co-exact, i.e. determine a three-form
B
=

S so that

Ss

Show that S is necessarily singular along a Dirac sheet, i.e. a halplane bounded by the z-axis. Show that S can be chosen as a harmonic form.

Finally we can investigate the geometrical structure of Band *B' Here B is a simple form generated by the functions Z and lnp (for simplicity we forget ~ for a moment!) These functions produce the 211E oC honey comb structure shown on fig.13~Similarly *B is almost a simple form generated by the function ql, which, however, makes a jump somewhere
between 0 and 21T (cf. fig. D2b). '!his exaIti'le ooncludes our discussion of electro-

magnetism and the three-dimensional geometry.

379

II

ql=-

11

-------Fig. l32a Z-axis

Fig.132b

In the rest of this section we return to our main interest: electromagnetic fields in Minkowski space. The field strengths are represented by a skew symmetric cotensor field of rank 2

Therefore F is nothing but a differential form of degree 2. We may decompose it along the basic 2-forms associated with an inertial frame:
(7.80) F

+ B

z d:x

1\

dy + Bx dy

1\

dz + By dz

1\

dx

The gauge potentiaL is represented by a covector field A. Therefore A is nothing but a differential form of degree 1. We can decompose it along the basic l-forms associated with an inertial frame:
(7.81)

NoW we want to compute the exterior derivatives dA and dF We will use an al'bit.rary coordinate system. Then dA is characterized by the components d~AS - dSA~. previously we have introduced the gauge potentUll A~ through the relation
(1.30)

valid in an inertial frame. But this can be geometrized as


(7.81)

Now the components of F and dA must coincide in any coordinate system. Thus we learn that (1.30) not only holds for inertial coordinates but is valid in an arbitrary coordinate system:

380

II
F~~

Furthermore
d~F~y

dF

is characterized by the components In an inertial frame we know that solves

+ aaFy~ + ayF~~.

the

~iaxwell-equation

(1.27)

This can be geometrized as


(7.83)

So

is a closed form. Again we see that (1.27) not only holds for

inertial coordinates but is valid for arbitrary coordinates. The fact that the ansatz (7.82) solves the I-1axwell equation (7.83) now becomes a trivial consequence of the rule To equation:
(1.28)

d2

o.

finish the discussion of the electromagnetic field we should

also try to give a geometrical interpretation of the second Haxwell

As it involves the divergence of the aontravariant components

F,

we are suggested to look at

SF.

In an arbitrary coordinate system this is a I-form, characterized by

In inertial coordinates variant components of simply states that (7,84)

;=g
SF
and

1.
J

Therefore we see that the contracoincide in an inertial frame. But

then they are identical. Consequently the second Haxwell equation

SF = J SF
and

Observe, that as we know the components of

in an arbitraform:

ry coordinate system we can write (7.84) in a


(7.85)

aoval'iant

If we take the co-differential of both sides of (7.84) we immediately get (7.86) But

o = SJ
is the divergence of

-&J

so this equation is equivalent to

(7.87)

381

II

which reduces to the usual equation of continuity in an inertial frame. We have therefore succeeded in comprising the structure of the electromagnetic field into the following elegant form:
(7.88)

Theory of electromagnetism Geometrical form Covariant form da. F f3y + di3 F ya. + dy Fa.f3 1 a (,t=gFa.i3) = Jr:I.
0

r-Iaxwell's l. eq. Maxwell's 2. eq. continuity eq.


Eq. of ga.uge r-otential

dF

=0 ,sF = J &J = 0
F

v=g a

1 /=a a. T-a -g da. ( -g ::r ) = Fa.i3 = d a.Ai3 - df3 Aa. A ... A + da. X a. a.

= dA

Gauge transformation

A ... A + dX

Illustrative example: The monopoLe fieZd. Observe first that we can now work in arbitrary coordinates. A relation like Fa.i3 = da.Af3 - di3Aa. is also valid in spherical coordinates or even in a rotating coordinate system if you prefer! We have previously computed the field strengths of a monopole expressed in spherical coordinates (Section 6.10), cf. cq. (6.79). Therefore we can decompose the monopole field as:
(7.89)
F

:fir SinB de

1\

d)

This is a simple form which may be rewritten as


F =-..5l. d(Cos6) 41T
1\

cip

AS it is a static field we can suppress the time coordinate for a moment, i.e. we look at space at a specific time to. Apart from the normalization factor, F Fig. 133 1s generated by the functions CosB and ). These two functions generate the honeycomb structure shown on fig. 133 , (i.e., CosS and
Tubes

= 0,

, 2,

0, , 2, )

pointing in the radial direction

382

II

The fundamental tubes are defines by cose = +1, 0, -1 and = 0, 1, 2, 3, and the number of fundamental tubes emanating from the origin is 2x2~. Thus the number of fundamental tubes which cross a sphere around the origin is 4~. Observe also that the Etubes all intersect the unit sphere in the same area. To see this we
~

use the wellknown fact that a region n = {(e,~) lel~e~e2; ~1~~~~2} covers the area: 9 2 (j)2 -COSe2 ~2 r r Sine ded~ = r r d(-Cose)~ -cosel ~l e 1 ~l Therefore an arbitrary E-tube covers the area U=-kE ~=mE+E r dud~ = [-kE + kE + E][mE + E - mE] r U=ke:-E ~=m ThiS, of course, again reflects the fact that the monopole field is a spherically symmetric field! Now observe that because we have written the monopole field as the simple form
F

Sine de

1\

d~

we can immediately find a gauge potential

generating the monopole field

Thus we can use the gauge potential


(7.90)

A = -.3.. Co s e d~
4~

Now we should be very careful because we have previously shown that the monopole field cannot be generated by a gauge r-otential throughout the whole space time manifold. (~ection 1.3l. Therefore we conclude that A = ~ cose cAp cannot define a smooth vector field throughout our manifold. The component ~ CosS is certainly a nice smooth function, so this must be related to the fact that the coordinate system itself is singular! We have several times seen that this coordinate system in fact breaks down at the z-axis, and therefore the gauge r-otential (7.90) 1s not defined on the z-axis. To investigate this a little closer we return to the Cartesian coordinates. Using that

383

II

cose we easily find

(jl= Arctan Y
x

so that the gauge r-otential has the Cartesian components


..5l..[0 - 41f '

z~ r (x +yz)'

r (X2+yZ)

zx

0]

But these components become highly singular when we approach the zaxis! If we approach the pOint (O,O,zo) where Zo > 0 we can safely put: ~ '" 1. Thus A varies like r a.
Aa.

"*

[0, - xl;y2

, XZ:y2 , 0]

and we see that

JAx J , JAy J ... .,


when we approach the z-axis! Consequently A is highly singular on the z-axis in accordance with our general result: There aan exist no gLobaL gauge potential- generating the monopo Le fie Ld. However, if we are willing to accept a string, where the gauge potential becomes singular, then we have just found suoh a gauge potentiaL This string is, of couse, the famous "Dirac string"! (Observe, that in the preceding discussion we have "excluded" the origin from the spacetime manifold, because at the origin the monopole field itself is singular). You should also observe the following: The spherical coordinates not only breaks down at the z-axis, but the cyclic coordinate (jl also has a singularity at a halfplane extending from the z-axis. This is connected with the fact that (jl has to make a jump somewhere between o and 21f. Thus d(jl is not defined on the halfplane where (jl makes a jump. But if you return to Cartesian coordinates you find d(jl = - ~2 dx + x 2 Therefore d(jl we conclude:
1.

y2 dy

is perfectly well-behaved (except on the z-axis) and

itself is a coordinate function, which makes a jump at a halfplane, and the-refore it aannot be extended to a smooth function on the whole space surrounding the z-axis (Compare , section 1. 4)

384 2.

II

d~ does not make a jump at a halfplane and consequently it can be extended to a smooth I-form on the whole space surrounding the z-axis.

In what follows we will let d~ denote the smooth I-form defined on the whole space surrounding the z-axis. The singular gauge p:ltential A =:5l.. cose d~ is obviously not the on41T ly gauge p:ltential generating F. To get another example we observe that d(~) = O. Therefore we consider AI = A + ..5l.. d~ = -..5l.. cose d~ + ..5l.. ~
41T

41T

41T

Although it formally looks like a gauge transformation, it is a singuLar gauge transformation because ~ is a singuLar coordinate function. Nevertheless A' generates the same field strengths. Observe, that:
(7.9ll
A'
= -..5l..

41T

(Cose - 1) d~

and that the coefficient cose - 1 vanishes on the positive z-axis. This suggests that A' is onLy singular on the negative z-axis. To confirm this we return to Cartesian coordinates:
A'

=:5l.. (~ - 1)
41T r

(- xZ;y2 dx + X2~y2 dy) where


X 2 +y2

If we approach the pOint proximation

(O,O,zo)

Zo > 0

we can use the ap-

+
r

2z

to get the asymptotic behavior of


Z

~ - 1:
X2

r - 1 ,,-

2zo

+y2

It fOllows that

A'

varies like
AI

,,-*

(~ dx - 2 zX

dy)

and the singularity on the positive z-axis has disappeared. Thus we are left With an semiinfinite Dirac string: The negative z-axis. This is the best result you can obtain! (If you take any closed surface surrounding the monopole, then it must contain at least one singularity of the gauge field. Otherwise you would contradict theorem 1 of section 1.3).

SOLUTI ONS OF "lORKED EXERC ISES: -

385

II

No. 7.5.6 Since *TAU is an n-form we only need to find the (12 ... n)-component. This is given by the expression

(7.92)

n! (*T) _ (U) (n-k)!k! [12". (n k) (n-k+l)."nl

Now observe that alan_kblbk

a l .an_kc l .. c k (C " .c ) I k Clearly the following

vanishes unless (b l " .b k ) is a permutation of relation holds (cf. exercise 7.4.3): I (n-k)! alan_kblbk

where we have used that there are (n-k)! permutations of (al ... a n _k ) , all contributing with the same value. Furthermore We can USe that TC1 .. Ck 1S skewsymmetric so that we get:

Inserting these

reduces to

We have thus shown

No. 7.5.10 For simplicity we discuss only the Riemannian case. We want to compare *(T'U) and *TAU For definiteness we only compare their l ... (n-k+m) component, but the argument can easily be generalized to arbitrary components. The contraction T'U is characterized by the components
-T

fl .. f

m!

e l .ek_mfl .fm e l " .ek_mfl " .fm

The dual form is then characterized by the components


~ ). (k

1
---,

-m.

m.

cIcn_kgl~elek_m

f 1 f m

If we specialize to the

(cl ... cn-kgl"'~) = (l .. n-k+m) l ... (n+k+m) m!

component we finally get fl ... f m

*(T'U)
*TAU

= ~ T(n-k+m+l) ... nflfm U

is characterized by the components

306

II

We must then show that (7.94) reduces to (7.93) except for a sign. This is done by brute force! We attack (7.94): Observe ~irst that (al ... ao-k) is an ordered subset o~ (1 ... n-k+m) and that (el .. ek-mfl ... fm) is an ordered subset which is complementary to (al ... ao) i.e. they are mutually disjoint, and together they exhaust (l . n). Especially (el . ek-mfl ... f m) must contain the numbers (n-k+m+l ... n) . Now fix (bl ... bm) for a moment. Then (al . ~-k) is determined up to a permutation. All these permutations have equal contributions when we perform the summation, and we can therefore pick out a single representative shich we denote (al(b), ... ,ao-k(b)), since it depends on {bl, .. ,b m} . Now that (al(b), ... ,an-k(b)) has been fixed we observe that (el . ek-mfl ... fm) is determined up to a permutation. Again all these permutations give the same contribution, and we need only pick up a single representative. Since (el ek-mfl ... fm) contains (n-k+m+l, ... ,n) we can simply put e l

= n-k+m+l

, .. , e k m

=n

Furthermore we pick up a special representative for f , ... ,~ which we denote (fl(b), ... ,fm(b)) , since it too depends on {bl, ... ,tm}. wrth these preparations ('r .94) now reduces to

~ Sgn
m.

;,-

[al (b) . a n- k (b )b l bm] 1. n-k n-k+1. .. n-k+m

al(b) ... an_k(b)(n-k+m+l) ... n~l(b) .. fm(b

(n-k+m+l) .. n~l(b)~m(b) U x'T b l .b m where we only sum over b. Observe that (a (b) . ~_k(b)b . b m) is an ordered subset of (l, .. ,n-k+m) and similarly (al(t) ... ao-k(b)fl(t) fm(b)) is an ordered subset of (l, . ,n-k+m) . But then (fl(b), . ,fm(b)) is.a permutation of (bl .. b m) and we can safely put fl(b) = bl' . '~m(b) = b m Thls means that (7.94) can be further reduced to

i&
I

Sgn

[al(b) an_k(b)bl ... bm] 1. . n-k. n-k+1. . n-k+m

m.

387

II

Let us focus on the statistical factor: (No summation!) al(b) ... an_k(b)blbm} Sgn [ l ... n-kn-k+l ... n-k+m Eal(b) ... an_k(b)(n-k+m+l) ... n bl ... b

This statistical factor is easily evaluated if we rearrange the indices of the Levici vi ta symbol al(b) ... an_k(b)(n-k+m+l) ... nblbm
+

al(b) ... an_k(b)blbm(n-k+m+l) ... n

This costs a factor (_l)m(k-m), but apart from this sign the combinatorial factor is now recognized as a complicated way of writing the number I ! We have thus finally reduced (7.94) to the form C_I)m(k-m);g T(n-k+m+l) .. nblbm U m! b ... b l m If you compare this with (7.93) you deduce the following formula (7.95) *TAU = (_I)m(k-m)*(T'U) which is in fact valid for both Euclidean metrics and Minkowski metrics. The desired formula now follows by performing a dualization on both sides .

No. 7.6.3

Let us work it out for an Euclidean metric. Putting S &T = (_l)k(n-k+I)*d*T

we obtain (Observe that this actually holds for Minkowski metrics too). Here (n-k+l)-form characterized by the components (d*T) . (n-k+l)! 3[(*T) . 1 Jlll n _k (n-k)! rJ llfn-k
= __ 1_ Sgnlabl' .. bn-kj 3 (*T)

d*T

is an

(n-k)!

.. . Jll .In-k

bl ... b_k n

= (n-k)!k!

[ab l ... b k) E 3 (lgrclc k ) Sgn ji l ... i~=k b l ... bn_kc l ... c k a (l, ... ,n-k+l) - component, but the results

For simplicity we investigate only the are easily generalized

(d*T) 1. .. (n-k+l)

I Sgn[abl ... bn-k]E 3 (/gT (n-k)!k! 12 ... (n-k+l) b l .bn_kcl.c k a

cl

ck

For fixed a the ordered set (bl ... b n- k ) is determined up to a permutation, since a and (bl ... b -k) are complementary subsets of (l, ... ,n-k+l) . All these permutations give th~ same contribution so we need only specify a single representative (blb n _k ) = (bl(a), ... ,bn_k(a)) Once (bl ... b n- k ) is fixed the ordered set (cl ... c k ) is also determined up to a permutation. All these permutations give the same contribution too, and we need only specify a single representative. Since bl .. b n- k are all different from a. we conclude that a coincides with one of the numbers cl .. ck We can therefore always achieve that

388
The rest then can be chosen as c2 = n-k+2 , ... , c k = n With these preparations (7.97) now reduces to
(d*T)l. . . (n-k+l)

II

ra bl(a) ... b (a)] Sgnl n-k g a (/gF a ( n-k+2 ) .. n) 12... (n-k+l) b l (a) ... b n_k (a)a(n-k+2) ... n a Using that g we
~inally

~l (a) ... b _ (a~a(n-k+2) ... n

= (_l)n-k g

n k

ab l (a) ... b - (a) (n-k+2) ... n n k

obtain (d*T)l ... (n-k+l) *s is a (n-k+l)-form characterized by the (1, ... ,n-k+l)-component _ ~ jljk-l (*S)l. .. (n-k+l) - (k-l)! gl. .. (n-k+l)jljk-IS

(7.98)
Similarly

(7.99)

;g S(n-k+2) ... n
If we compare (7.98) and (7.99) and use the relation (7.96) we finally obtain the ~ollowing result

which we can immediately generalize to the result (7.65)


No.

7.6.4

When we use Cartesian coordinates in an Euclidean space we do not have to distinguish between co-variant and contra-variant components. Furthermore we can put ;g = 1 If T is a k-form the derivative dr is then the (k+l)-~orm characterized by the components

~, Sgn[~J ll ~l""'~] . lk
Thus -&dT is the
k-~orm

acT

dl"'~
.~] a
T c
dl"'~

characterized by the components


=

Hdr)

al"'~

.l..
k!

a
b

Sgn[C d l
j

bal"'~

For a ~ixed c the ordered set (d ... ~) is determined up to a permutation. As all permutations give the same contribution we need only pick up a single representative
~ = dl(c), .. ,~ = ~(c), and the above expression reduces to

where we only sum over

and

c.

Clearly this sum decomposes in the

~ollowing

way

389
It is the last term we must now get rid of. Since king one transposition we get cfb we can fix dl(c) = b.

II

Ma-

Inserting this we have shown:

On the other

For fixed c the ordered set (d2 ... ~) is determined up to a permutation. All permutations give the same contributlon, and we need only pick up a single representative d 2 = d2(c), ... ,~ = ~(C) The expression above then reduces to

Adding the two results obtained we finally get:


(AT)

a l .. ~

= (-d&T)

a l ... ~

+ (-&<iT)

a l ... ~

3 3 T

b b a l ... ~

No.

7.6.9
w

(a)

= [f+ig][dx+iay] = [fdx-gay]
*ay

+ i[gdx+fayl we immediately get

If we use that
(b)

= dx.

*dx

= -ay

*[gdx+fdy] ~ [-gay+fdx] 3h dh dJM ~ dbAdz = a; dZMZ + dZ- <lzAdz Here the first term vanishes automatically because
. lS

dz

is a l-form, so that

Consequently (c) If h(z)

dJM = ~~ dzAdz closed l. f and only ~f ~

dh = 0 3Z-

is holomorphic we get

db
which implies that

= dh
3z

dz

'

But Thus
(d)

h Let

is a harmonic function and so are its real and imaginary parts. h(z) be a holomorphic fUnction. Then db

= 3z dh

dz

and that is a closed

390
l-form. According to (b) this means that use that h is a harmonic function:
0" a h
2

II

ah az

is a holomorphic function. We can also

azaz

a (ah) = a-z az ah az

but this is exactly the Cauchy-Riemann equation for the function (e) If we once more use that *dz Consequently the decomposition db = ah dz + a~ clZ az az
*dx = -dy

and

*dy = dx

we immediately get

-idz

is a decomposition of db in an anti-self dual and a self-dual part. It follows that db is anti-self dual if and only if = 0 0

NQ.1.,1.4

We want to compute the Laplacian in spherical coordinates. We have previously computed the metric coefficients, cf. (6.36)

We also need the contravariant


i.e.
1

components of

df,

aif

a
(\

[aif] -

a a

rz
a

r2Sin 2a

["

a f r
[

" ref

rz I
r2s~n2a

"a f a<pf

"qf

We thus get

(1.100)

No. 7.8.1 The first two Maxwell equations can be solved using gauge potentials. The equation, dB = 0, is solved by the ansatz
B = dA ,
the second Maxwell equation

where A is a vector field! Substituting this into (1.T1)you get

391

II

o ,
and this equation can be solved by the ansatz

where

is a scalar field. Thus we have rediscovered the gauge potentials $,A:

= dA

The last two Maxwell equations(7.78) and(7.79) can be used to find the equation of continuity. Rearranging them we get

and

- Eo" at

1 ap

from which we deduce that

~ at

SJ

i.e.

which is the proper generalization of the equation of continuity. Finally we deduce the equations of motion for the gauge potentials. Substituting eq. (7.101) into the two remaining Maxwell equations we get

i.e.

1 a2 ~ - SA) <cz acz + sd + dS)A + d<-!z. c at which can be further rearranged as

{ {

&(1A + d,)
at

....!..
EO

alA -at<at + d,) + c 2SdA 1 a2 <~

= +-J = -

1 EO

acz +

Sd), - ..2....<A ~ - SA) at c at

1 p EO 1
= EOC2 J

If

<"A)

satisfies the Lorenz condition

we thus recover the waVe equations

392

II

!'!.n. 7.8.2

First we observe that

(*)

B = &S

iff

*B

- d*S

Therefore B is co-exact in a space region same region. But

if and only if

*B

is exact in the

*B

= 21TEjoc2 d(jl

Apart from an irrelevant constant this is our old friend d(jl (See section 1.4) which isonly almost exact. Consequently B is only almost co-exact. From (*,**) we also see that

= -

21TEOCL

*(jl
In cylindrical coordinates we know

It remains to determine the Levi-Civita form [. that ,.g = p (Compare exercise 6.6.3). Thus

= pdPAcIQ A dz

so we nave finally found

= -

21TEjocL

p(jl

dp A

<ill

A dz

and at the same time we have managed to write S as a simple form. In particular S is closed and it follows that S is a harmonic 3-form

- dSS

->tS = -dB

3:

II

ehapter 8

INTEGRAL CALCULUS ON MANIFOLDS


8.1 INTRODUCTION
The next problem we must attack is how we can extend the integral calculus to differential forms. Before we proceed to give precise definitions we would like to give you an intuitive feeling of the integral concept, we are going to construct. In the familiar theory of Riemann-integrals we consider a function of say two variables defined on a "nice" subset Then we dvide this region cells Ai with area
2

f(x,y)

of into

R2

z = f(x,y)

and in each

cell, Ai' we pick up a point We can now form the Rie(xi'Yi) mann-sum:
L f(xi'Yi)&

2
U

If

is "well behaved" and

is

a sufficiently "nice" region, then this sum will converge to a limit as

and this limit will be indepen-

x-axis

Fig. 134

dent of the subdivision in cells. It is this limit we define to be the Riemann-integral:

lu
sion 2 embedded in
in

f(x,y)dxdy

Next we consider a two-dimensional smooth surface dinary Euclidean space R3 R3 As R3

in the orF

i.e. a differentiable manifold of dimen-

Similarly we consider a two-form

defined

is a two-dimensional surface, we may introduce coorwhich we assume cover

dinates surface

(A1'A2)

n,

i.e. we parametrize the

n.

(See figure 135) . U into cells &2 In each of these cells we pick up a point

Then we make a subdivision of the coordinate domain of area

(A 1 ,A 2 ).
(i)
(i)

394

II

At the corresponding point P;

~(A 1 ;A 2 )
(i)
-+

in
-+

n
2

we

1.2

(i)

have two basic tangent vectors F


e
(i)

and

As

(i)

is a two-form, it maps into a real numSo we can


(i)

(e ,e) (i) 1 (d
ber
(i)
-+

F(e 1,e.2 )

form the sum


L F(e 1,e 2 )& i (i) (i)
-+

x
n
0

Fig. 135

If

is "well-behaved" and
&
-+

is a "nice" surface, then this will and this limit will be independent of

converge to a limit as gral of F:

the subdivision in cells. It is this limit we define to be the inte-

(8.1)

J n

def -+ -+ 2 "" lim L F(e 1 ,e 2) & &-+0 i (il (i)

At this point you might object that


(;\1,1. 2 )

fn

cannot be a geometric

object, because we have been extensively using the special coordinates

in our definitions. However, it will be shown that fn F does not depend on which coordinate system we choose, provided they define
the same orientation on

n.

Of course the above formula is of little

value in practical calculations. Therefore we should rearrange it! 1 2 To each point (1. ,1. ) corresponds a set of canonical framevectors

(e 1 ,e2 )

and therefore

generates the ordinary function

U ~ R

given by

Using this function we can ry Riemann-integral:


(8.2)

conv~rt

the integral of

into an ordina-

Jn

F""l1mLF(e1,e2)&2
& .... 0

(i)

(i)

We can also express the integral of of

F
Let

in terms of the components

F.

Here you should observe that R3

is defined as a two-form on

the whole of the Euclidian space Then

(i ,i 2 ,i 3 ) 1

-+

-+

-+

denote the R3 .

canonical framevectors corresponding to a coordinate system on

is characterized by the components

395

II

The canonical framevectors in the following way:

(e1 ,e 2 )
and

on

are related to

(1 1 ,1 2 ,1 3 )

Using this we can rearrange (8.2) in the following way:


Fab(A , A )

;;:'f;;?
0 "

'x a 'x

i.e.
(8.3)

We shall reconstruct these formulas from another point of view in a moment.

8.2 SUBMANIFOLDS - REGULAR DOMAINS


In the general theory we have a n-dimensional differentiable manifold

M.

We want to characterize what we understand by a k-dimen-

sional surface R(n,k)


R~n,k)

in

To do this we must fix some notations: x x

n {x E R n {x E R

= =

1 2 (x ,x , .. , xk , 0, , 0) }

2 (x 1 ,x , ... ,xk,o, ... ,0)

~ a}

(See figure

l36.~

n=3,k=2.

Fig. 136 We can then define the concept of a submanifold:

Definition 1 point A subset P in

n n

=M

is said to be a k-dimensional submanifold if each has the following property:


neighbourh()~d

There el1;i"sts an open

of

and a coordinate

396

II

system
V

($,U)

such that

maps

U n R(n,k)

bijectivelyonto

n n.
137a-b below.)

(See figure

n=3,k=2

Fig. 137a

n=2,k=1. X2

Fig. 137b A coordinate system ted to

($,U)

with the above property is said to be adap-

n.

According to the definition the adapted coordinate systems

actually cover that we call

n,
n

hence they constitute an atlas. This justifies

a submanifold. As

itself is a manifold, we can in-

troduce arbitrary coordinate systems on

n,

as long as they are com-

patible with the adapted coordinate systems. Clearly the k-dimensional submanifolds is the eXact counterpart of the loose concept: a k-dimensional smooth surface in

M.

As a specific example you may' 2 consider M 8 Then the 1 equator 8 , is a one-dimensional submanifold. As adapted coordinate systems we may use the standard coordinates on the sphere. (Compare the discussion in section 6.2. 138 too.) See figure Fig. 138 xl

Next we should try to characterize the domains of integration. From the standard theory of integration you know that these domains

397

II

integral. If

should be chosen carefully, otherwise we will get into trouble with the b . f: R ~ R is a continuous function, then fa f(x)dx is

well defined, while

f+OO

-00

f(x)dx [a;bl

is somewhat dubious. In standard inte. Then you will have no trouble beR and continuous functions on com-

gration theory you would therefore start by restricting yourself to closed, bounded intervals cause these are compact subsets of

pact intervals have many nice properties; especially they are bounded and uniformly continuous. Furthermore, you know that there is an intimate connection between integration and differentiation (the fundamental theorem of calculus) : b df dx = f(b) dx - f(a)

This relates the integral of a and

df to the values of f at the pOints dx b, which constitute the boundary of the interval [a;bl . (If MS RN then a subset of M N R .) Further-

Back to the general theory! To avoid troubles with convergence we will only consider compact domains. is compact if and only if it is closed and bounded in

more we should consider domains where the boundary itself is a nice "smooth surface". Otherwise we will not be able to generalize the fundamental theorem of calculus. This motivates the following definition: (See figure 139 too.) Definition 2 A compact subset lowing two properties:
Q

of a n-dimensional manifold

M is called a

k-dimensional regular domain, i f each point P in Q has one of the fol-

1.

There exist an open neighbourhood Vof P in M and a coordinate system

($,U) on M suah that:


bijectiv~ly V

$: U n R(n,k)
2.

n n .

There exist an open neighbourhood Vof P in M and a aoordinate system


(~>U)

on M such that:

~: U n R~n,k)

bijective~y

V n nand

x 1 (P)

=0
n
n
Clearly it is a

The points with property 1 are called interior points of The set of interior points of k-dimensional submanifold. The points with property 2 are called boundary points of

is denoted

int

398

II

P satisfies property 1.
Q satisfies property 2.

(Observe that $-l(Q) is lying 2 on the X -axis, i.e. x 1 (Q) = 0)

Fig. 139

The set of boundary points is denoted ry point and use


($,U)

an

and it clearly is a

(k-1)-di-

mensional submanifold called the boundary of

If

is a bounda-

is a coordinatesystem with property 2, then we may

as local coordinates on disjoined components:

an.
int

Thus we have decomposed! a k-dimensional regular domain into two

and

an.

The boundary

an

has an ex-

tremely important property! It is itself a compact subset:

Theorem 1

an
Proof:

is a compact

{k-1) -dimensional submanifold.

The proof of this is somewhat technical and you may skip it. As n is compact it suffices to show that an is closed in M. Let (Pn)nEm be a sequences of points in an and assume that this sequence converges towards P, i. e. Pn'" P. We have finished if we can show that P E

an .

399

II

As n is compact, it is closed in M and therefore we know that PEn But then P is either an interior point or a boundary point! Thus we must rule out the possibility of P being an interior point. Assume that P is an interior point. Then there exist an open neighbourhood V of P and a coordinate system (~,U) such that
~:U () R(n,k)

bijectively ..

V () n

Fig. 140
and P n If follows that all the points in V n n are interior points. But PEn ~ P therefore there exist an integer N such that n>N implies tRat
P E V () n

n
n
~

Thus we conclude that tion! i

is an interior point if

N and that is a contradic-

This property of see that

an

has important consequences. As

an
an

is a com-

pact submanifold all the points in

an

has property 1. Therefore you has no

an

itself is a regular domain! But as all the point in the

regular domain

an

are interior points, it follows that

boundary. Consequently Theorem 2 FOr a in. But (Cartans


regu~ar

a(an)

is empty:

~emma)

domain

the boundary

an

is also a regular doma-

a>J

has vanishing boundary

a (an)

o.

We shall include a point as a zero-dimensional regular domain, although it really falls outside the scope of our definitions. We do not assign any boundary to a pOint, so that the rule

aan = 0

still holds

in this very special case!


As this point we had better look at some specific examples all using
M = R3

400

II

Examp l~ 1: If n is the closed unit ball,


n then

= {xl
intO

<xix> ~ I}, is the open unit ball, <xix> < I},

intn

= {xl

and the boundary of two-sphere, an

is the

=
ao

{xl <xix>

= I}

,
s2

i.e.

= S2.

Observe that aan

is itself a two-dimensional regular domain without boundary

0.
Fig. 141

Examp Ze 2:
If of n is the northern hemisphere

82,

n = {x then

E R31 <xix>
intO

= 1, x

3 > oJ,

is given by,

intn = {x E R31 <xix> = 1, x 3 > and the boundary of n is the unit circle, an

oJ,

= {x
an

E R31 <xix>

1, x 3

OJ,

Le. where

sl , ao=sl

sl is itself a onedimensional regular domain without boundary: aan = 0. a(an)

Fig. 142

Finally we observe that the rule rule

=
dF

reminds us of the If with , we called it a 0

d(dF) = o.

They are, in fact, very intimately connected. F had the property

a differential form

=0

cLosed form. the property

In a similar way we will call a regular domain an

o ,

a closed domain.

This is in agreement with

the familiar expressions "a closed curve" and "a closed surface". At this point you should begin to have a feeling for the regular domains which are going to be the domains of the integration! There is, however, still a problem which we will have to face, and that is

the orientability of

n.

401

II

If we return to our intuitive discussion of how to integrate a twoform on a surface, it is clear that if we interchange the coordinates Al and then we also change the sign of the integral because
~ F(e l ,e 2 )E
~i

-+i

Fig. 143

x
Al and A2 , this simply means that we

But, if we interchange

construct a new coordinate system with the opposite orientation! This clearly shows that the integral depends on the orientation. Consider now a regular domain ordinate systems. 0 in a manifold

Then int 0

is a k-dimensional submanifold which can be covered by adapted co-

provided

int n

We say that the r~gular domain 0 is orientable is an orientc.ble manifold. (Compare with the

discussion in Section 7.4). The interesting point is that whenever

is an orientable regular domain, so is int 0

aO!

In fact, the

orientation chosen on orientation on int 0 then: (xl, . , xk, .. , xn) Now let

induces in a canonical manner an

ao

Tb explain this we choose an orientation on


(~,U)

be an adapted coordinate system on

ao,
M

serves as local coordinates on

(xl, . ,xk,o, ,0) xl < (0,x 2 , . ,xk,o, .,0)

serves as local coordinates on serves as local coordinates on

int 0

ao

The local coordinates generate a positive orientation on ao if and onZy if the loaaZ coordinates (x l ,x 2 , .. ,x k ,0, ,0) generate a positive orientation on int Q. We have tried to exemplify this on
This makes i t reasonable to define:

(0,x 2 , ,x k ,0, ,0)

fig.

144a-c.

402

II

Orientation of intll induced by a coordinate system.

Fig.

l44a

11111':--1, ,--, 1I

by a coordinate
k = 2

12 has a hole! Observe that the


:s)inner boundary

and the outer boundary get opposite orientations.

Fig. l44b

I ,

III
Q(+)

k
dO

=1
P UQ

Orientation of intO induced by a coordinate system.

PH
A point is negatively oriented when we move away, positively oriented when we move towards the point.

Fig. 144c

Observe, that in the case where

dO

reduces to a finite set of

points, we will have to construct a special convention, as you cannot introduce coordinate systems on a single point!

403

II

8,3 THE INTEGRAL OF DIFFERENTIAL FORMS


With these preparations we are ready to define integrals of forms. , n and So let n be an orientable k-dimensional regular domain in M n let F be a differential form of degree k defined on M We assume for simplicity that denote the coordinates of
(xl ,xk xn).

int int

n n

is a simpZe by

manifold which can be M by

covered by a single coordinate system (not necessarily adapted!). We (Al, ,Ak) and those of

Fig. 145

The coordinates tated.

(Al . ,Ak)

To each point
~

are assumed to be positively orienk (Al, ,A ) in U corresponds a set of


~

canonical frame vectors el,,e k As F has degree k, it maps (~l""'~k)


. cons~der

into a single real


-+
~

number: as
(8.5)

F(el, ,~k).

ordinary function on

Thus we may F(el, e ) as an k U and therefore we can define the integral

Jn F
~l'

def.

f
u

F(el, ,ek)dA dA

-+

We may rearrange this formula by introduci~g the canonical frame ~s


-t

. , ~n

-t

on M.

404
(8.6)

II

Inserting this we get

Therefore we finally obtain:


Definition 3 Let metrized

=Mn
by

be an orientable k-dimensional regular domain paracoordinates (A1 , ,)..k). k Let


F

be a differential
F

form of degree

characterized by the components

al ak

with

respect to coordinates integral of

(xl, .... xn) on

Then we define the

F
~

through the formula

\J
(8.7

''i n J,

fu
Ju

-+ F(e l ,, e k ) dAl ... dAk

a r .. a

ax

al

ax

al

dAl ... d).k

The formula (8.5) clearly shows that the integral is independent of the coordinate system (xl, ,x n ) on int

but we still have to

show that it is independent of the coordinate system ().l, ,).k) on

n . Therefore we consider two sets of coordinates (AI, ,).k) and (~l, ... ,~k) which both cover int n and which specify the same
int

orientation on positive.

n ,

i.e. the Jacobiant

Det

(a~~)
d).J

is always

First we observe that i dA 1 i dA k k

F(el;:e k )
(2) (2)

-+

-+

= F(~i

(1)1 aliI

, .... , e

-+

i (l)k aJ.1

-+ = F (e

i (1)1

; ; e. )

(If

~ ... dA
aliI

i k

all

But F is a k-form and Therefore we conclude


F (~.
~l

i l , ,i k
.

can only take the values

1, ,k.

; ; ;;.

~k

E.

(1) F(tl ;;
(2)

(1)

1 lk

F (t ; ; ~k) l

(1)

(1)

Inserting this we get

~k)
(2)

F(E!l;:
(1)

405

II

Using the transformation rule for Riemann integrals we then obtain k e k ) Det (Olli) ---j I dA l dA OA

(2)

But

Det(dll~) J
OA

is assumed to be positive and therefore we can forget We then finally obtain k

about the absolute value.


(8.8)

f F(el ;; ek )
U2 (2) (2)

d]Jl . dll

f F(el ;; ek )
Ul (1) (1)

dAl dAk

and this shows that int

IF
n

is independent of the coordinate system on

n ,

as long as it generates a positive orientation!

up to this point we have only considered the situation where int Q is a simple manifold, i.e. it can be covered by a single coordinate system. We now turn to the general situation where we have to use an atlas (cj>i,Ui\EI consisting of overlapping coordinate systems, which together cover integral of a differential form over n.

IF
n

int

n.

We want to construct the

Now there is one case where the answer is obvious: Suppose there exists a special coordinate system (cj> o,Uol so that the restriction of F to n vanishes outside cj>o(U )' Then of course, we define o the integral as

fn Iu F(~ l';~
F =

kldAl dAk

Fig. 146 F ;,t 0 Next we want to consider the general case of an arbitrary differential form F This requires, however, the introduction of a oechnical but important machinery: Recall that an open covering of a topological space subsets (OiliEr that covers M, i.e.
M

M is a family of open

iEr

U O.
1

(Open coverings play an important role in the characterization of aompaat spaces. A topological space M is compact if an arbitrary open covering of M can be exhausted to a finite covering) Then we consider a family of functions (WiliEr on M,

1/\ :

M~

We sfI that such a family is ZceaUy finit6, when at an arbitrary point P E M only a finite number of the functions Wi have a non-zero value. Observe that this property guarantees tllat the sum

406
l: 1/J.

II

iEI

is well-defined in a completely trivial sense, even if the index set I is uncountable. (At each point P the sum reduces effectively to a finite sum). Consider once more a family of functions (1/Ji)iEI on our toplological space M. We say that it is a partition of unity if it is a locally finite family of non-negative functions such that l: 1/J. = 1 iEI 1 {i.e. for any point P we have
L1/J.{P)

= 1)

Now we can filially introduce paracornpaat spaces. As you might expect,they are characterized by a suitable property of their open coverings: A topoboeioaZ spaae M is paraaompaat if we to an a:t'bi trary open covering {O i \EI Clan find a partition

of unity {1/Ji)iEI

s~ah that 1/J i vanishes outside 0i' (We say that such a partition of unity is subordinate to the open covering. Observe that the partition of unity and the open covering have the same index set I). To get used to the concepts involved, you should try to solve the following useful exercise:

Exeraise 8.3.1
Introduction: Let (<p i ) iEI be a partition of unity subordinate to the open covering (Ui)iEI and let (1/Jj)jEJ be a partition of unity subordinate to the open covering (OJ) jEJ . Problem: a) Show that (Uinoj)(i,j)EIxJ is an open covering. b) Show that (<P. o 1/J. \. ')EIxJ is a partition of unity subordinate to the
1

Jh,J

open covering

(uinoj)(i,j)EIxJ

Now what have all this to do with Euclidian manifolds? Well it turns out that they satisty the following important theorem:

Theorem 3
An Euclidian manifold M is paracompact, i.e. to every open covering we can find a corresponding partition of unity. (For a discussion of theorem 3 including a proof, see Spivak [1970]). Okay, let us use this heavy artillery: Consider an atlas int (<Pi,Ui)iEI covering

n.

Then

int

is a submanif91d and

[<Pi(Ui)]iEI

an open covering. Thus we

can find a partition of unity (1/Ji)iEI subordinate to this open covering. We use it to cut F into the small pieces: Fi = 1/J i F Here 1/J i F vanishes completely outside the coordinate system <P i (U i ). Consequently the integral is trivially well-defined. And furthermore We have F = (l: 1/J. )F . 1
1

~(1/JiF)
1

on

int Q

Thus We can finally define:

407

II

Definition 4
(8.9)

def.

This is a legal definition provided we can show that the number independent of the atlas ($i,Ui)iEI we choose to cover int

~
1.

fa

~i F

is

and also independent

of the partition of unity. This requires lengthy but trivial calculations, where We make extensively use of exercise 8.3.1. For details see Spirak [1970]. So now we can in principle calculate the integral of a differential form over an arbitrary regular domain.

Suppose that we want to integrate a differential form over a surface which is unbounded

Provided the surface is sufficiently nice this may is closed 0

still be done in the usual way. Sufficiently nice means that 0 (in the toporcgicar sense). and that arr points in

have either
and

property
If int 0

or property 2 from definition 2. section 8.2. Then we

can still divide

in two disjoint submanifolds: int 0 int O.

ao.

is a simple manifold.we can choose coordinates We can then formally write down the

(AI, .. Ak) which cover integral


(8.10)

If the integrand ful.

F(el, . ekl falls of sufficiently rapid at "infinity". F has compact support. i.e. Otherwise the integral is
(~i)iEI' This

the Riemann integral will converge and everything is nice and beautiThis is especially the case. if F vanishes outside a compact subset of divergent.
If int 0 is not simple, we must again use a partition of unity
~i

M.

can be chosen so that

F has compact support on

int 0

Therefore the integral

is trivially well-defined and the complete integral provided


L f~ F i 0 1

f F a

is well-defined

is absolutely convergent, i.e.

LIN FI i a 1.

<

CD

In that case the value of the sum is

independent of which partition of unity we use, and we can therefore define:

Jl

~ JO~i

F
=
{(x.y.z) 10~x~1, o~y~l. O<z<l} O<x<l, O<y<l,

We can also use the integral concept on non-regular compact domains. Consider e.g. the unit cube in

R3: 0

The interior of the cube consists of all points with

O<z<l. The boundary consists of the remaining points, i.e. of the six faces 6 = {(x.y.z) I x = 0, O<y<l. O<z<l} etc. 1

408 Notice that the points with property (2) in exhaust


(j(l

II

We still have the points on

the edges, which form the sceleton of the cube. The interior is a nice submanifold, but the boundary is not a submanifold. The integral over

oQ

is therefore apriori

not well defined. However,

oQ

is "smooth

piece by piece", and the edges are of a lower dimensionality than the faces. Therefore we can neglect their contribution to the integral. We can then define

Fig. 147

JaQT =
If int Q
Q

i=l

6 1:

is a regular domain, then it is composed of two parts:


aQ.

and

Let us concentrate on the interior for a moment. To each point

It is a k-dimensional submanifold.

on

int Q

we

have attached a k-dimensional tangent space T (int Q). If we introduce coordinates (AI, .. ,A k ) on int Q, they generaie a canonical frame

(~l""'~k) for the tangent space

Tp(int Q). Clearly

Tp(int Q) is a

k-dimensional subspace of the n-dimensional tangent space Tp(M). I f (xl, .. ,xn ) denote coordinates on M, then they generate a canonical frame
(1 , .. ,1 ) for the full 1 n
-t-t-

tangent space
-+-

T (M).
-+- p

The

frame (el, . ,e ) is related k -t-tto the frame (1 1 , .. ,1 ) in n the following way (8.6) Now, if
F

ei

-+-

-t-

a Hi

21L
Fig. 148

is a differential k defined on F

form of degree

M, then

is a multilinear map on each of the tangent spaces Tp(M). as a differential form defined on int Q
F F

But we may also consider simply because spaces


restriction of

defines a multilinear map on each of the tangent


~

Tp lint Q)
F

Tp(M). This new differential form is called the


to

int Q and it is convenient to denote it by the

same symbol When F

is considered to be a differential form on

M, it is

characterized by its components

409 and we may decompose it in the following way

II

(7.8)
In a similar way the restriction of ized by the new components
F

to

int Q

can be character-

F(~. ; ;~. ); 1 < i. < k 1.1 1. J


k and we can decompose the restriction of
F

as

(8.11)
since it is a k-form on a k-dimensional manifold. definition:
(8.12)

(Compare exercise

7.2.3). This is very useful when you want to integrate


+ 1 k def. ;ek)dA A AdA =

F!
1

By

J
/

/ ( e l ; J

/ ( e l ; .. ;ek)dA J

.. dA

Thus the integration is performed simply by replacing dAIA .. AdA k by the ordinary Riemannian volume element dAl dAk! Observe, that dAIA . AdA k itself can be considered as a volume element in each of the tangent spaces Tp(int Q) (Compare the discussion in section 7.4).
+

Consequently the geometrical integral,

JQF (e

l; ;e k)dA A.. AdA

may be considered as a generalization of the Riemann integral defined for ordinary continuous functions on ordinary Euclidean spaces! To use this point of view in practical calculations we must have a method to find the restriction of is of course to compute
F

to

int Q.

One way of doing i t

F(tl

, . '~k)

directly using the formula:

But in many applications it is preferable to use the decomposition

= k!

a l .ak

dx

al

A Adx

ak
What is the restriction

as a starting point. The question is then: of dx i to int Q ?

E:r:ercise 8.3.2
Problem: (a) Show that the restriction of dx i
(8.13)
dx~~...e... a
(lA (l

to

int Q is given by

dA a

410
(b)

II

Insert the result obtained in (a) in the decomposition (7.8) of F, and show once again,by rearranging the formula obtained,that the restriction of F to int Q is given by -+- d 1 (8 ) F = F( -+.11 el, ,e ) A A AdA k k

The calculations are especially easy if you use an adapted coordinate (xl, . ,xk , ... ,x n ) parametrizes M and (xl, .. ,x k ) parametrizes int Q. Then the restrictions of dxl, ,dxk , .. ,dxn system, i.e. to

k and int Q are simply given by: dx l = dx l , .. ,dxk = dx k l n O. dx + = O, ,dx So if you want to find the restriction of

1.
F to

k.' F

a l .a k

dx

al

A Adx

ak
dxk+l, ... ,dx n

int Q , you simply remove all terms containing Suppose


Q

We conclude with one more remark on how to calculate the integral in practice. int Q is a nice orientable regular domain, where Then int Q cannot be covered by a In that case we ought to use overlapping In praxis we will compute the integral
Q

is not a simple manifold.

single coordinate system.

coordinate systems, find a partition of unity etc. This is. very technical and very complicated. using our common sense! int Q Thus int Q S2 . which is its own interior, i.e. is not simple. Then we But it Consider, for instance, the case,

s2,

But let us introduce spherical coordinates anyway! exclude an arc joining the North Pole and the South Pole. constitutes a subset of lower dimensionality and therefore it does not contribute to the integral anyway. Thus we simple way:
Tl 2Tl

x
Fig. 149

comp~te

the integral in the following'

J F J J F(~8;~~)d8d~
Q

8=0

~=o

Worked exercise 8.3.3


Problem: (a) Let Q = S2 be the northern hemisphere of the unit sphere in Show that + R3.

411
(b) Let
Q = B3

II

be the closed unit ball in

R3.

Show that

Jdx
B3

dy

dz

Tl

8.4 ELEMENTARY PROPERTIES OF THE INTEGRAL


The integral has, of course, several simple algebraic properties. It is obviously linear:
(8.14)

JQ (F+G) =

J F + J G Q Q
Q aQ we have attached Suppose F dF is a is

Now we have been very careful in our discussion of regular domains. To each orientable k-dimensional regular domain an orientable (k-l)-dimensional regular domain a differential form of degree and The question is then: What is the connection between these two integrals?
Worked exercise 8.4.1
Problem: Let Q be the unit cube in two-form. Show that

differential form of degree k-l, then the exterior derivative k.

Thus we can form two integrals:

R3. and let B be an arbitrary smooth

dB

aQ

It follows from exercise 8.4.1 that the two integrals are always identical when theorem:
Theorem 4 Let
Q

is a cube.

This is a special case of the famous

(Stokes' theorem)
F

be an orientable k-dimensional regular domain and let then

be a

differential form Of degree k-1,

(8.15)

J
Q

dF

For a complete proof see Spivak [1970], but the formula can also be derived from the following naive argument. The regular domain Q can be apprOXimated by a family of cubes: (Qi)iI' cf. fig. 150

412 The boundary of the cubes consists of two kinds of faces: a) b) The face lies at the boundary of The face lies in the interior of

II

an

n. n.

v
1

In case (b) the face always coincides with a face from the neighbouring cube, but they will have opposite orientations and their contributions to the integral therefore automatically cancel. When we perform the integration, we can therefore replace the boundary of with the boundaries of the cubes:

Fi

g.:l

0 1"-./ 0
/\
\......./

11'""\

i\

fdF ~ ~

I dF = ~ I F ~ I F n. an. an
~

--

If the manifold

M is equipped with a metric, we obtain the

following important corollary: Theorem 5 (Corollary to Stokes' theorem) Let

be an orientable (n-k+l)-dimensional regular domain and let

be a smooth k-form. then (_l)n-k+l

Ian *

Proof: From (7.62) we learn that

6 F

(_l)n-k+l d

and theorem 5 is now a trivial consequence of Stokes' theorem. Finally we have the rule of integration by parts. generalized the Leibniz rule in the following way
(7.17)

We have

where

and

G are differential forms of degree

and

Let

be a

(k++l) -dimensional orientable regular domain.

Then

Ind(F A G)

In d

FAG

+ (_l)k

In F A d G .

But we may transform the left hand side using Stokes' thaorem whereby we get

aQ

FAG

d FAG + (_l)k

I FAd G
Q

This we can rearrange in the following way:

413

II

Theorem 6

(Theorem of integration by parts)


IanF A G -

(_l)k In F A dG

This, of course, is a direct generalization of the familiar rule:


b b

I
a
either because
Q

df g ()d x x = [f(x)g(x)b - If(X) Qg dx dx a dx

In many applications we will be able to throwaway the boundary term,


has no boundary (Le. an

= 0),

or because one of the differential

forms vanishes on

an!

Let us discuss this new integral concept in a few simple cases: Consider first scaZar fields. represented as a zero-form. mensional regular domain. the integral reduces to A scalar field,
~

: M

R,

can be

It can then be integrated over a O-diBut a O-dimensional regular domain consists

simply of a point (or a finite collection of points). So in this case

I
A scalar field
(7.39)

4>

4> (P) 4> This is the

i.e. it just reproduces the value of the function most trivial and most uninteresting case. 4>

can also be represented as the n-form

But then it can be integrated over n-dimensional domains including itself.

Usually such a domain n can be parametrized by the coordinates (xl, .. ,x n ) on M itself and the integral reduces to (8.18) where In

4>

4>

;,g

dx

A A dx

Iu

4>(X);,g dxl dx
n.

is the coordinate domain corresponding to l n I 4>(X)dX dX , n

This shows

that we can generalize the ordinary Riemann integral in Euclidian space

which is valid only for Cartesian coordinates, into the covariant expression

which is valid in arbitrary coordinates.

All you have to do is to

replace the Cartesian volume element with the covariant volume element ;,g dxl dx n When we integrate a scalar field 4> on a manifold M

414 with metric

II

we always use the covariant integral (8.18).

As an application we can choose

=
n

1.

The integral

JQ
7.4) (8.19) Here Q

JQIcJ dxl . dx
Q

then represents the volume of

since

= ;.g d

xlA . Ad xn

is the

volume element in the tangent space, (compare the discussion in Section

Vol [Q]

J ;.g dxl .. dxn


Q

is restricted to bean n-dimensional domain, but the formula

(8.19) can easily be generalized to cover the "volume" of arbitrary k-dimensional domains. i.e. the length of a curve, the area of a twodimensional surface etc. fold way
M
Q

To see how this can be done we consider an Q in our n-dimensional maniand in this Let p,l, Ak ) on Q.
-+-

orientable k-dimensional regular domain The metric


M

induces a metric on

becomes a k-dimensional manifold with metric.


-+-

be a parametrization of

The intrinsic coordinates (Al .. ,Ak)

generates a canonical frame (el ,e ) in each of the tangent spaces k (compare figure 148). The induced metric is characterized by the intrinsic components

g(~i;~j)
The k-dimensional volume of
)

Q
~

is therefore given by
1 k

(8.20)

Volk[Q]

JI
Q

Det[g(e.;e.)] dA . dA
J

-+--+-

This can also be reexpressed in terms of the extrinsic components of

g:

The volume formula is then rearranged as


(8.21)

/ Det [g b J Q a

ox ---j ox ] dA 1 dA k ---.
oA~ 01..

Problem:

E:x:ercise 8.4.2 Consider a curve r connecting two points P and Q Show that reproduces the usual formula (6.40) for the length of a curve.
Next we consider vector fields. Let

VOll[r]

t..

be a vector field.

We can

represent

t..

by a one-form

A.

It can then be integrated over a one-

dimensional domain, i.e. a curve

If we parametrizes this curve

with the parameter

A, (cf. figure 151), the integral reduces to

415

II

(8.22)

A2 r J A=J

< A

AI

~ > dA

JA2
"I

Ao ---d~ dA

d~i
1\

l.

Observe that in a space


->->-

Euclidian

Q
A-

dr dA

and the integral (8.22) therefore generalizes the usual line integral

12 ..
Fig .151

We shall also refer to Suppose


A

fA

as a tine-integraL

is generated by a scalar field

i.e.

d$ ,

then Stokes' theorem gives

Jr d$
The boundary therefore put ar

= Jar $
P
and

consists of the points Q

Q, where

is

negative oriented and

is positive oriented.

By convention we

Jar
(8.23)

= $ (Q)

$ (P)

Stokes' theorem consequently produces the formula

which generalizes the usual theorem of line-integrals use that

(sec. l.l}It

does not, however, use the full strength of Stokes' theorem. If we

Jr

d$

=J

A i 2 dx AI axl. dA

dA

J"2 AI

~
1\

dA

we irrmediately see that it is a simple reformulation of the fundamental


theorem of Calculus. Worked ezercise 8.4.3 Problem: We use the notation of exercise 7.6.9. Especially manifold in R2, which we identifY with C, (a) Let r be a curve in M Show that

M is a two-dimensional

Jrh(z)dz = frh(z)dz
where the integral on the right hand side is the usual complex line-integral.

416

II

(b) Consider an analytic (holomorphic) function curve connecting A and B . Show that Ih'(z)dz = h(B) - h(A)

h(z).

Let

be a smooth

(c) Consider a two-dimensional regular domain Q bounded by the closed curve r . Let h(z) be an analytic (holomorphic) function in M. Prove the following version of Cauchy's integral theorem: fh(z)dz=O

A vector field dimensional surface i.e. at each point to the tangent space

A
Q P

can also be represented by a (n-l)-form is usually called a hypersurface. on Q the normal vector ti(P) If

*A. Q is
I

Then it can be integrated over a (n-l)-dimensional domain.

A (n-l)-

an orientable hypersurface then it allows a normal vector field

is orthogonal

Tp(Q).

Let us parametrize the surface Q with the parameters (AI, ,A n - l ). They generate the canonical frame vectors -+-+(e l , ,e n _ l ). Furthermore we let ti denote the normal vector n = elx xen_l (Compare the discussion in

r
Fig. 152 we have
(-+-+-) elx xen_l

...

...

section 7.4). According to exercise 7.5.4

*A (-+-+-) ~ el,,e = A' n_l


(8.24)

The integral of *A therefore reduces to

fQ* A -- fu*
I* A
Q

-+-) ,1 d n-l A (-+el,,e n _ l dA A

fuA 7,-+,n-l n d,l A . dA


A
through
* A

Thus

simply represents the fZuz of the vector field


Q

the hypersurface

For this reason we refer to the integral of

as a fluz-integral. Having introduced flux-integrals we can now explain why theorem 4 is called Stokes' theorem. If S is a surface in R3 bounded by the curve r theorem 4 gives us fs dA = Ir A The right hand side is the line integral of A along

r.

We also have

d A = *(ilxA)
(Compare the discussion in section 7.7). The left hand side therefore represents the flux of the curl ilxA. Consequently theorem 4 generalizes the classical Stokes' theorem from the conventional vector calculus, cf. (1.16).

417 Let surface

II

be an n-dimensional regular domain bounded by the hyper If A .


-+-

an

is a vactorfield, then

&A

represents the and get

divergence of

We can integrate this divergence over

the volume integral

According to the corollary to Stokes' theorem this can be rearranged as (8.25)

(_l)n-l J

on

up to a sign we have therefore shown that the integraL of the divergence is equal to the flux through the Qoundary. This is the gene-

ralization of Gauss' theorem to an arbitrary manifold. (8.18) , (8.26) (8.24) and (7.64) we get Jn

We can also Using

work out the corresponding covariant coordinate expression.

a. (/g Ai)dnx
~

= JAin.
on

dAl .. dA n - l

where the normalvector (8.27)

...

is characterized by the covariant components

n. ~

= Ig

e:. .

~Jl'"

n-l

We can also reproduce the integral theorems from the two-dimensional vector calculus. Consider for instance a region n bounded by a smooth curve r (see figure 141). Then it is well-known that the area of n is given by the line-integral

(8.2 8) Area
(Here
to ;

-+- -+- . [n) = ~ r*roar

'f

*;

is the vector orthogonal

and with the same length. Remember

also that

~*;oAt represents the area of

the triangle spanned by ; and 1+6;). It is convenient to write out this formula in Cartesian coordinates

(8.29) Area

[n)

= ~Jxdy-ydx

Fig. 153

To deduce this formula we use that dx. A dy is the "volume form" in R2. Consequently Area [n] i.e. Area [n)
An application of Stokes' theorem now gives

Area [n)

= ~Jrxdy

- ydx

which we immediately recognize as being equivalent to (8.29).

418

II

We conclude this section with a discussion of simple forms, where we can give a naive interpretation of the integral. For simplicity we consider the threedimensional manifold M = R3. If we choose a coordinate system on M, then the coordinates (Xl, x 2 , x 3 ) generates several simple forms. for instance: dxl.

dx l

dx 2

dx l

dx 2

dx 3

Consider first the basic one-form Q. (See figure 154).

dxl.

Let

r be a smooth curve from P to

a
Fig. 154

We want to interpret the integral [a;a+ll. [a+l;a+2l, .

Ir

dx l

We divide [a,bl in unit intervals

Then the integral is roughly given by:


dxl J r "" L
< dxll

"tli

>

The coordinate

xl

generates a stratification characterized by the surfaces: xl = -1. xl O. xl = +1, ..


<

According to our analysis in section 7.2 , the number

dxll ... e i > l

can be inter-

preted as the number of surfaces intersected by the curve segment corresponding to the i'th unit interval. Therefore the total integral can be naively interpreted as the number of surfae~8 intersected by r. This can be demonstrated rigorously using Stokes' theorem. It immediately gives us

JIl

dxl

= J xl = xl(Q) or

- xl(p)

and the number xl(Q) - xl(P) cle~ly represents the number of surfaces intersected by r. Observe that if r is a ciosed curve, then each surface is intersected twice from opposite directions. Therefore

fr

dx

=0

This is also in accordance with Stokes' theorem

Next we consider the basic form dx l A dx 2 . (See figure 155).

Let

Il

be a smooth surface in

M.

419

II

A2

AI

Fig. 155

We want to interpret the integral

III
We divide U into unit squares.

.bel A

dx'

Then the integral is roughly given by:

I
The coordinates (x l ,x2 )

dxl

dx 2 "" 1: "-A ....1


i

Il

dx2 ( + . + e i l' e i2 )

generate a honeycomb structure in M According to our I 2( +i +i ) analysis in section 7.2 the number dx A dx e ; e can be interpreted as the 2 l total number of tubes intersected by the surface segment corresponding to the i'th square. Therefore the integral can be naively interpreted as the number of tubes intersected by Il. This can also be justified more rigorously by the following argument: For simplicity we assume that we can use (x l ,x2 ) as adapted coordinates on Il, i.e. Il is parametrized on the form: x3

= x 3 (x l ,x2 ).
A dx
2

Then we get

III dxl
Here the coordinate domain

Iu dxl ax2

U is an open subset of the xl -x2-Plane. Tubes in M correspond to unit cells in the x l -x2-plane. The number of tubes intersected by Il is equal to the number of unit cells contained in U. On the other hand f ax l ax2 U is equal to the area of U and this number also represents the number of unit cells contained in U. If Il is a closed surface, then

In dxl
Stokes' theorem for the basic form

dx

because each tube is intersected twice with opposite orientations. This illustrates

dx l A dx 2 ,:

420
Exercise 8.4.4 Problem: 8.3.3. Exercise 8 . .4.5 Problem: Let
Q

II
R3. Show

= S!

be the northern hemisphere of the unitsphere in S! is


Tl.

that the "number" of tubes intersected by

Compare this with exercise

2 Consider the basic three-form dxl A dx A dx3 The coordinate functions xl,x2 and x 3 generate a cell-structure in M. Let Q be a

3-dimensional regular domain in

M.

Show that

f
Q

2 l 3 dx A dx A dx

can be interpreted as the number of cells contained in

Q.

8.5

THE HILBERT PRODUCT OF TWO DIFFERENTIAL FORMS


Consider two differential forms T and U of the same degree k introduced the scalar product between

We have previously (sec. 7.5) T and U, cf.


1

(7.52-53):
i l i k

(Tlu)= k: T

U..

~l ~k

~*(*T

U)

This is the relevant scalar when you look at a specific point space
k Ap(M).

P,

i.e.

when you want to introduce a metric in the finite dimensional vector

Globally the differential forms of degree dimensional vector space

span out an infinite

Ak(M)

and we would like to associate an

inner product with this space.

To motivate it we consider real-valued smooth functions on a closed interval [a;b]. Here we have the natural inner product:
< fig>

.fbf(X) g(x)dx , a

=f
[a,b]

*f A g

In analogy with this we consider the n-form (7.50) *T


A U = -- T

i l .i k

k:

il i k

If it is integrable, we define the inner product in the following way: def. 1 i l .ik 1 n A U = --k' T U . . Ig dx dx (8.30) < T I u >

~l ~k

We shall refer to this inner product as the Hilbert product. compact, then the Hilbert product is always well-defined. If compact, the integral is not necessarily convergent.

If M is M is not

421 has compact support.


Consider a Riemannian manifold. with compact support.

II

It is, however, always well defined if one of the differential forms


Let Ak(M) denote the vector space of k-forms o The Hilbert product defines a positive definite metric in is a pre-Hilbert space. We can

this infinite dimensional vector space.

Thus Ak(M) o complete it and obtain a conventional Hilbert space

~(M)
called the Hilbert space of square-integra?le k-forms . An element of the form 1 l.l l.k T = k' T. (x) d x A A d x l.l .. l.k with measurable components

~(M) is on

is convergent. For a pseudo-Riemannian manifold things are slightly different. Here the scalar product (Tlu) is indefinite and therefore the Hilbert product is indefinite too. If we complete Ak(M) o we therefore get a Hilbert space with indefinite metric.

In the following we shall always assume that all the differential forms we are considering have a well defined Hilbert product. preserves the inner product between two differential forms: Observe first that up to a sign the dual map is a unitary operator, i.e. it

Theorem ? (a) The dua'l map * is a unitary operator on a Riemannian manifo'ld: <* T I * u > = < T I u > (b) The dual map * is an anti-unitary operator on a manifold with Minkowski metric: <* T I * u > = -< T I u > Proof:
This follows immediately from the corresponding local property (Theorem 9, section 7.5):
< *T

*u > =

(*T

*u)'

<Tlu>

Then we finally arrive at a most important relationship between the exterior derivative -form, U

and the codifferential

5.

Let

be a

(k-l)

a k-form and consider the inner product:


<

dT >

I*U
M

A dT

This can be rearranged using a "partial integration":

422 d(*UAT)

II

d*U A T + (_l)n-k *U A dT

From exercise (7.6.1) we know that

d * U
Furthermore

(_l)n-k+l * IS U

J d(*u M
either because M

T) =

JaM

*U

T = 0
*U A T Therefore we obtain: i .e . < UI d T > = <
& U IT>

is compact without boundary or because

"vanishes sufficiently fast at infinity".

(8.31)

J * U A dT = J *
d.

& U A T

Thus we see that due to our sign convention, 6 operator of

is simply the adjoint

It should be emphasized that this holds both for Let

Riemannian manifolds and manifolds with a Minkowski metric. This has important consequences for the Laplacian operator.

M be a Riemannian manifold.
Laplacian as

Using (8.31) we can rearrange the

where

&'

...
denotes the adjoint operator. Consequently -A is a

positivQ
k-forms.

B~rmitian operator.

The above argument applies to arbitrary

In the special case of scalarfields it is well-known from

elementary quantum mechanics, where the Hamiltonian, H

=-

i'J2

2mA ,is a

positive Hermitian operator reflecting the positivity of energy! We can also use (8.31) to deduce an important property of harmonic forms on a compact manifold.

Theorem 8 Let M be a compact Riemannian manifold.


A T

A k-form

is harmonic

i f and only if it is primitively harmonic, i.e.

=0

iff;d T

= &T =

Proof:
Let us first observe that <-ATIr> = <d5TiT> + <5dT1T> = <ST 1ST> + <dT IdT> If T is harmonic, i.e. AT 0, then the left-hand-side vanishes

automatically. But the right-hand-side consists of two non-negative terms: Hence they must both vanish: < & T I & T > = <d Tid T > = 0

423

II

But as the inner product is positive definite, this implies

&T

=d

=0

0
i. e. a smooth function, <p
:M~R.

Consider for instance a zero-form,


Ii

Then

q,

vanishes automatically and

d <p

vanishes if and only if it is This

constant. Consequently <p is harmonic if and only if it is constant. is a generalization of If Liouvillestheorem in complex analysis.

M is not compact the above consideration breaks down (unless

itself has compact support) . Then you should be more careful! E.g. in

potential theory we are interested in the electrostatic potential generating a static electric field

q,

The equations of motion for (cf. (1.19:


A typical

q,

reduces to the Laplace equation

A<P

= 0

in space-regions where there are no electrical sources. dimensional regular domain potential more since

problem starts with a space-region, which we represent as a three-

=R3.

We have been given an electrostatic and we want to reconstruct

q,
<p

on the boundary of

q,

in

the interior of

n.

But observe that int

is not compact!

Further-

does not necessarily vanish on the boundary we cannot

neglect the boundary term. We will not go into any details, but the starting point is Green's identities which you can have fun working out in the following exercise:

worked exercise 8.5.1


Problem: (8.32) (8.33) Let n be an n-dimensional orientable domain in an n-dimensional Riemannian manifold n. Deduce Green's identities: (1) (2) If
<

dq,1 d~>

<IA~>

= (_l)n-l

fan*[d~]

<<pI~~> - <~<pr~> = (_l)n-l Jan*[q,d~-~q,]

Use Green's identity to show the following property of harmonic functions:

q,

=0

on

on

then

=0

in

n.

We can also investigate singular solutions to the Laplace equation

in

R3.

For instance we know that

q,(i) = 1 r

is a harmonic function

(generating the Coulomb field!) But it is not primitively harmonic, as it is not constant. observe also that Hilbert product since it is only well defined on Thus it "violates" theorem 7. This is no disaster M R3 ,{O} which is not compact.

q,

and

are not square-integrable on

M. The

424
<

II

EIE >

<d~ld~> =

R3 ,{O}
point-charge!

!4 dxdydz

= 4TI

J~ ~
0

diverges at the origin, reflecting the infinite self energy of a If you try to repair the proof of theorem 8 through a regularization procedure, i.e. put

then you should observe that the "partial integration" leading from
<& d ~ ~> to <d ~ d ~>

is no longer valid because the differential

forms no longer vanish "sufficiently fast".

Worked exercise 8.5.2


Problem: Let (a)
~

be a smooth scalar field on M

Show that
<A~ I~> = - < d ~ Id P
Put

f
dM

*d ~

~(;t)

= 1. r =J

. Show by explicit computation that


~

< d ~ Id

* d ~ = 4TI

dM
Moral: Whenever you work on a non-compact manifold, you should

look for "boundary" contributions.

8.6 THE LAGRANGIAN

FOR~ALISM

AND THE EXTERIOR CALCULUS


as an important

We will now investigate the Lagrangian formaLism example of how to apply the integral formalism. In a field theory the Lagrangian density S
equat~ons

of motion are determined by the

principle of least action.

The action itself is constructed from a

L, which is a scalar field:

JQLE = JUL(x)i=g dxo dx l dx 2dx 3


L

This scalar field function of


~

is constructed from the fields and their


~

derivatives. In the simplest case of a scalar field and


a~~.

it is thus a

Using the covariant action we can now They are obtained using
Q.

construct the covariant equations Of motion:


~

the now familiar variational technique: We replace the scalar field by


~+~,

where

vanishes on the boundary of

425 This generates the new action

II

S(E)

= JU L(~+E~,
=

].l].l

~+E3 W);:g
E

4 d x

which is extremal when

= 0,

i.e.
d 4

~~IE=O

J (~~
U

W +

o(~L~)o].lW);:g ].l
d
].l

JU[(;:g
o=
As

~~)lji ~~
-

(;:g

(~L ~) lOlllji Jd 4 X
4 lji d x

Using partial integration on the last term we get

fJ

I-g

d].l

(/-g

~) ]
(;:g
0

lji

is arbitrary, this is only consistent if:

(8.34)

~~ = ?:g

d].l

(~\))
].l

If we have several fields

~a'

then each of the components have to Thus we have shown:

satisfy the appropriate Euler equation!


Theorem 9 The
~a

aovariant equations of motion for a aolleation of saalarfields

are given by

(8.35)

Although it is much more dubious,the interpret the index a

covariant equations of too, when we

motions (8.35) are actually valid for vector fields, as a spacetime index!

Exeraise B.6.1 Problem: Consider the Lagrangian density for the electromagnetic field
L

,,_!
4

F
].l\!

F].l\!

Re-derive the covariant Maxwell equations (7.85). The geometrical point of view can in fact be used to throw light on one of the more subtle points in the derivation of the Euler equations. To pass from

to

we used a "partial integration", but we neglected the boundary terms, because vanishes at the boundary", This is justified in the following exercise:

"1j.I

426 Worked exeraise 8.6.2


Problem: a) b) Show that Let
()L ~

II

are the contravariant

components of a vector field A.

II

be a scalar field vanishing on the boundary of

n . Show that

( 8.36)

Jn(~ A~)d ~

4 $ d x =

-Jn3~ ~ A~)$

4 d x

As we have seen in section 6.7 we can also write down the equations of motion for a free particle in a field. covariant form. Consider now the case of an electrically charged particle moving in an electromagnetic Combining (1.26) with (6.46) we are led to the following covariant equations of motion

(8.37 )

We would like to derive it from a Langrangian.

Now we have previously

determined the non-relativistic action for an electrically charged particle interacting with the electromagnetic field:
(2.13)

This suggests that for a system consisting of an electrically charged particle and the electromagnetic field we should use the following reZativistia action

S Sp

= Sp

+ SI + SF

with dx S dA dA

-mt/-g~s~(
Ai A2 ll q

(8.38)

SI

J All dx w:- dA A1 Jn
Fll\! Fll\!

SF

- 14

;=g

d 4x

Here the dynamical variables to be varied consist of a) b) The position of the particle: Xll(A) . The gauge potential: All(X).

427

II

We leave it as an exercise to the reader to verify that the invariance of the action (8.38) under these variations actually leads to the desired covariant equations of motion!

worked exercise B.6.3


Problem: a) Show that the interaction term can be rearranged as (8.39) b)
8

= In All Jll

r-g

4 x

(Compare this with exercise 3.10.2). Show that the relativistic action (8.38) leads to the following equations of motion:
d 2 x C!
C!

(8.37) (7.85)

m d,2

= qF

8 dT FllV)

ax~

C!

mfllV

a:r dT

axll axV

-L a (Fg
Fg v

Jll

It is also interesting to see that the Lagrangian formalism can be thrown into a purely geometrical form. This means that we can discuss the equations of motion completely without introducing a coordinate system! Of course. the first step consists in reexpressing the action in a purely geometrical form. Let us look at some specific examples:
~,where

The KZein-Gordon field:


The
S

This is a scalar field

the

action is based upon the Lagrangian density (3.49). covariant action is then given by

fnL
S

e:

= fnb(all~) (all~)

~ m2~2];:g
=

4 d x

which we rearrange as (8.39)

=-~<d~ld~>- ~m2 <~I~>

f-~ * d~Ad~ n

~m2 * ~

(Compare the discussion in section 8.5). This was the first step. Next we perform a variation
~ ~ ~

+ e:1}i

where

1}i

is a scalar field, which vanishes on the boundary of

n.

We then get S(e:) =-~<d~ + e:d1}ild~ + e:d1}i > - ~m2 <~+ e:1}iI~+e:1}i >

=-~<d ~Id ~> _~m2<q51 p + d< d ~ Id1}i > - m2<~ 11}i>] + ~2E-< d ljild1}i > _m2<ljillji>]
From this we immediately deduce

dS =-<d~ldlji> de: I e:=O

428

II

( H ere we have used that 1jJ vanishes on the boundary to throwaway the boundary term coming from the partial integration). But this is only consistent if ~ satisfies the equation (8.40) -Sd~ = m2~ , which is the geometrical form of the Klein-Gordon equation: This is a one-form A , and the action is based The Max~eZZ fieZd: upon the Lagrangian density (3.50) which leads to a covariant action given by

This is rearranged as
(8.41)

Then we perform a variation,

A
where

A+E:U
:2

u
S(g)

is a one-form, which vanishes on the boundary of


-~<dA+dUldA+gdU>

-~<dAldA>

- <dAldu> - :f<duldu>

From

this we immediately get


dS o = cr---= -<dAldu> = -<5dAlu>

I=o

( Here we have used that U vanishes the boundary of n to throw away the boundary term coming from the partial integration). But this is only consistent if
(8.42 )

Le.

-OF

which is nothing but the Maxwell equations!


Exercise 8.6.4
Problem:a)Consider the massive vector field by

Show that the action (3.51) is given

(8.43)
b)Perform the variation tion A ~ A+:U and deduce the following equations of mo-

(8.44)
c)Show that they are equivalent to

(8.45)

2 cA= m A

-6A=O

We may summarize the preceding discussion in the following scheme:

429

II

FIELD
(8.39)

Action S

Equation of motion
-6d~ = m2~

Klein-Gordon
~

= f~~(*d~Ad~)-~m2(*~A~)
=

(8.40)
(8.41)

Maxwell A Massive vector


A

-Is, ~ (*dA) AdA


n

-odA

( 8.42) (8.43) (8.44)

=J

-~(*dAAdA)-~m2(*AAA) -odA

= m2 A

8.7 INTEGRAL CALCULUS AND ELECTROMAGNETISM


As an other example of how to apply the integral calculus we will use it re-express in a geometrical form various electromagnetic quantities like the electric and magnetic flux through a surface and the electric charge contained in a 3-dimensional regular domain. We start out peacefully in 3-space to get some feeling for the new formalism. Remember that tially B E is a I-form, but B is a 2-form. Essenis the dual of the conventional magnetic field. Now let

be a 3-dimensional regular domain. From the discussion of fluxintegrals (8.24) we get (8.46) (8.47) (8.48)

fan

The magnetic flux through the closed surface The electric flux through the closed surface The electric charge contained in

an . an .

Jan *E
f;p

We can now use the integral calculus to deduce some wellknown elementary properties:
Example 1 If n
~ke

c~ntains

no singularities, then the magnetic flux

through

closed surface

an

vanishes.

This follows from an application of Stoke's theorem:

~
(due to (7.76.

J an B = JndB =

430
Example 2

II

(Gauss' theorem)

The electric flux through the closed surface electric charge contained in

an

is

n.
(7.78)we get

This follows from an application of theorem 5 (Corollary to Stokes' theorem). From the Maxwell equation
;

[the charge]

lJ *p
0

-J n*QE
S be a sur-

Example 5:

This time we consider a static situation. Let face with boundary


to

r . Then the flux of current through S is equal

oc 2 times the circulation of the magnetic field along Observe first that the Maxwell equa1 oc2 J

r.

tion (7.79) reduces to


&B

for a static configuration. The proposition then follows from an application of theorem 5 (Corollary to Stokes' theorem) :

Fig. 156
We conclude the discussion of electromagnetism putation of two important integrals: a) Consider the spherically symmetric monopole field. Let 52 be the closed surface of a sphere with radius r . Then polar coordinates (r,8,~) are adapted to the sphere, and we can choose (8,~) to parametrize it! We Can now compute the magnetic flux through the closed surface 5 First we observe that the monopole field B is gi ven by (cf. (7. 89) ) :
B

in 3-space with the explicit com-

= t,;- 5in8d8AiAp
2Tf Tf

Then we immediately get:

(8.49)

J B = tL f J 5in8d8d~ = g 52 ~O 8=0
Tf

Fig. 157

431
b) This time we consider the magnetic field around a wire with current j. Let r be a circle around the wire with radius p. Then the cylinder coordinates (p,~,z) are adapted to the circle, and we can choose ~ to parametrize the circle. We can now evaluate the line integral of the magnetic field along the closed curve r. For this purpose we must find the dual form *B which is the one-form representing the magnetic field. This has been done previously (cf. section 7.8) *B Thus we get
- 21TE: o c 2

II

-L
2'TT

dip
-t

fr *B = 21T~ c 2 fd~ oo

E7 n
0

Fig. 158

Then we proceed to consider electromagnetism in Minkowski space, i.e. 4-dimensional space-time. Here it is more complicated to express suitable quantities, so we shall adopt the following terminology: Suppose we have chosen an inertial frame S. Let (xO,X 1 ,X 2 ,X 3 ) denote the corresponding inertial coordinates. We say that the three-dimensional submanifold
F,; is a space slice if it is a subset on the form F,; = {XEMlxo=tO} F,;

One dimension

suppressed~

i.e.

points at a specific time

consists of all the spatial to . Ob-

serve that the spatial coordinates (X 1 ,X 2 ,X 3 ) are adapted to F,; The fundamental quantities describing the properties of the electromagnetic field are the field strengths A , and the current F and *F Fig. 159 ,the Maxwell field S Then we

J . Now let

be a 3-dimensional regular domain

contained in a space slice relative to the inertial frame

can form the following integrals which we want to interpret:

Ian F

Ian*F

and dxo

fn*J
to

Observe first that the restriction of grands. But


(8.
50~

vanishes. Conse-

quently the integrals involve only the space-components of the inte-

and

*F

are decomposed as and

[*]

*F

[*J

432 So the restriction of (respectively striction to intn


*JI~

II *F) to an is given by B

(respectively

*E). Similarly the dual current has the following re-

..

(*J)

123

dX ' Adx 2 Adx 3

= -JOdX ' Adx 2 Adx'

We can now generalize the results obtained in (8.46)-(8.48) to the following


Lemma 1 Let Then
(8.51)

be a J-dimensional regular domain contained in a space slice.

JanF

The magnetic flux through The electric flux through

an an n

(8.52)

Jan

*F

(8. [, J)

-In*J

The electric charge contained in

Exercise 8.7.1 Introduction: Let n be a. 3-dimensional regular domain obtained in a space slice and let F be smooth throughout n. Problem: Use the 4-dimensional integral formalism to re-examine the following well known results: a) The magnetic flux through an is zero. b) The electric flux through an is equal to the electric charge contained in n (As usual we have put EO = c = 1 ).

ILLUSTRATIVE EXAHPLE:

MAGNETIC STRINGS IN A SUPERCONDUCTOR

We have earlier been discussing some of the features of superconductivity, especially the flux quantization (see section 2.12). Recall that in the superconducting state of a metal, the electrons will generate Cooper pairs. These Cooper pairs act as bosons and we can'therefore characterize the superconduc~ing state by a macroscopic wave function the order parameter,
1~12,

called the order parameter of the superconducting metal. The square of represents the density of the Cooper pairs. We want now to study equilibrium states in a superconductor. Consider a static configuration ~(~), where ~(~) is a slowly varying spatial function. In the Ginzburg-Landau theory one assumes that the static energy density is given on the form:
(8.54)
H

433

II

Here a is a temperature dependent constant


(8.55)

aCT)

a ~c
T

(with a positiv)

where Tc is the socalled critical temperature. The constant y is just inserted to normalize H to be zero at its global minima. The equilibrium states are found by minimizing the static energy. This leads to the Ginzburg-Landau equation
(8.56)

Now consider the potential


(8.57)

It has the well-known shape shown on figure 160. Above Tc the vacuum

T>T

T<T

Irn1jJ Re1jJ configuration is given by 1jJ = 0


i.e. the metal is expected to be found in its normal state where the

Re1jJ

Fig. l60

Cooper pairs are absent. Below Tc the configuration 1jJ=O

becomes un-

stable and we get a degenerate vacuum, the superconducting vacuum, with a temperature dependent density of Cooper pairs given by
a a T - T

- B

-.~

All this is in good accordance with the experiment!


1jJ-axis

The coherence Zength:

Consider a specimen, a semiinfinite slab, bounded by the y-z-plane. For negative x we are in the normal region, where 1jJ=O , and for positive x we are in the superconducting region. So for sufficiently large x we expect the order parameter to be in its vacuum state

Nannal region

Superconduc-

ting region.

x-axis
Fig. 161

434 IljJl =

II

~ If

The coherence length, s(T), is the characteristic length it takes the order parameter to rise from its normal vacuum, ljJ=O, at the boundary to its superconducting vacuum, 11jJI=V-;, inside the slab. The problem is essentially one-dimensional so the Ginzburg-Landau equation (8.56) reduces to

with the boundary conditions ljJ(O) = 0


H:o)

= If

r=i

The solution to this problem is simply a "half-kink", cf. (4.23)


if

x>O
x<:O

ljJ(x)
if

Consequently the coherence length is given by (8.58)


S{T)

Then we consider what happens when we put on a magnetic field B. The order parameter will now couple minimally to the magnetic vectorpotential, i.e. we must exchange the exterior derivative d with the gauge covariant exterior derivative
D

d -

ifiA

(where A is the magnetic vector potential)


Exercise 8. 1'.2 Problem: Show that

(8.59)
Notice that it follows from exercise 8.7.2 that the square of the gauge covariant exterior derivative does not vanish in general. When including a magnetic field the static energy density (8.54) is modified to
(8.60) H

435
Worked exercise 8. 7. J Problem: Show that the equations of equilibrium configurations are given by

II

(8.61a) (8.61b)

D*D1jJ

(Shjll 2 + a)ljJ

*
o

i.e. i.e.

[$D1/J -

ljJD1jJ]

Let us now assume that the order parameter is in its ground state (characterized by a vanishing energy density), i.e. (8.62)
DljJ

=0

U(ljJ) = 0

USing exercise 8.7.2 we then immediately get (8.63)

i.e. either B vanishes or ljJ vanishes! In the normal state ljJ=O , so here the magnetic field can propagate freely, but in the superconducting vacuum IljJl=YIf ' and B has to vanish. Consequently the magnetic field is expelled from any region where the order parameter is in the superconducting vacuum! This, of course, is the famous Meissner effect. Observe too, that if the order parameter is in its vacuum state, then the super-current,given by (8.64) automatically vanishes.

r=a

ljJ-axis
The penetration length:

We can now introduce the second characteristic length in superconductivity. Again we consider a semi-infinite slab bounded by the y-z-plane. We now apply an exout how far into the superconducting region x ~~-ax~i~s~==~~~E=====;~~~~~ ~ A(T) Fig. 162 that the magnetic field penetrates. This ~ time we use the Ginzburg-Landau equation for the magnetic field:
..-::

ternal magnetic field B and we want to find

In this equation we can approximate IljJl with its equilibrium

value~

This does not mean that the order parameter is in its superconducting vacuum since the gauge covariant derivative DljJ need not vanish. EspeCially we can still have super conducting currents floating around. Taking the exterior derivative on both sides we now get:

436

II

- d&B

-flfl2B

~2

Here we have used that Wis proportional to exp{i~} so that ~d~ is proportional to d~ , i.e. it is a closed form. Since furthermore dB vanishes on account of Maxwell's equations we arrive at the Laplace equation: (cf. 2.77)
(8.65)

Again the problem is essentially one-dimensional and the Laplace equation therefore reduces to
2 .. d 2Bij - ( x ) =-~2 B~J (x)

dx 2

Since B is a closed form we furthermore have

ax

dB23

This leads to the following solution:


; Bl2 (x) = Bl2 (O)exp[

-ljix1

Consequently the penetration length, A(T), is given by (cf. 2.78)


(8.66)

t:Sfl A(T)=I/-'=-Ct

(Notice that the above considerations are strictly speaking only valid when tIT) AIT) ). We have now introduced two characteristic lengths, the coherence length tIT) and the penetration length AIT), both of which depends on Ct and hence on the temperature. We can now form a third parameter, their ratio
(8.67)
K =

li'!l
HT)

fl Vj3
q

which is temperature independent and is called the Ginzburg-Landau parameter. Recall that there are two different types of superconductors, type I and type II. If we apply an external magnetic field, we know that the suoerconductivity is destroyed for sufficiently strong magnetic fields. In a type I superconductor there will be a critical field strength, Bc' above which superconductivity breaks down and the magnetic field is uniformly distributed throughout the metal. In a type II superconductor there will be two critical field strengths, BCland BC2 When we pass Bc

437

II

superconductivity will not be completely destroyed but rather the magnetic field will penetrate into the metal in form of thin magnetic strings, vortices. First when we pass BC2 the superconducting regions break down completely. We want now to investigate the structure of a magnetic string in a type II superconductor. We will assume that ;(T) A(T). Thus a single string
; (T)

consists of a hard core of radius

where the density of Cooper pairs vanishes. Inside this hard core we have therefore re-established the normal vacuum. Outside the hard core the magnetic field falls off and it vanishes essentially in a typical distance also a soft core: A(T). Thus we have ; < P < A . Observe

that we have circulating supercurrents in the soft core which prevents the magnetic field from being spread out ( the Meissner effect ). Finally we reach the sUperconducting vacuum outside the soft core, where neither the magnetic field nor the order parameter contributes to the energy.

~ '< ; . ~
soft core ACT)

V
H

11jJ1

;(T)

..
Fig. 163

hard core

Remark: Consider a magnetic string with flux ~ and let us for simplicity assume that we have a constant magnetic field inside the gtring. Then

~o Bo'TTP2 where p is the radius of the string. Thus the magnetic energy stored per unit length is given by ~2 2'TTp2 = --2 (8.68) J = ~ 'B0 2'TTp2
Thus we can diminish the energy of a magnetic string with a constant flux by spreading out the string. If nothing prevents it a magnetic string will thus grow fat. In the case of a solenoid it is the mechanical wire, in which an electric current flows, which prevents the magnet~c field ~rom spreading out. In a superconductor it is the Meissner effect which prevents the magnetic field from spreading out.

Now let us try to compute the energy per unit length in a single vortex line. We shall neglect the hard core which by assumption is very thin. Thus we concentrate exclusively on the soft core. Let us denote the region outside the hard core with expression for the static energy:

n .

We start by rearranging the

Hn =
But

~<BIB>n

~<D1jJID1jJ>n

iJ~I1jJ12

~}2

1jJ =~exp[i~]

outside the hard core so we can neglect the pot en-

438

II

tial energy. Furthermore the supercurrent reduces to


(=

&B)

and the gauge covariant exterior derivative of the order parameter similarly reduces to DljJ = qiexp {iql}[ <l<P We therefore
g~t,

~Al

by direct inspection, that

;<DljJIDljJ>n = ;A2(T)<JIJ>n = ;A 2 (T)<oSBI&B>n As in the discussion of the penetration length we also have

- <loSB = A2h)B
Finally the expression for the static energy therefore reduces to

;A2(T)[<BI~&B>n +

<&BI&B>n1
*&BAoSBl =

~A2(T)(Jn*BAd&B

~A2(T)Jn<l[*BASBl

U sing Stokes' theorem this is rearranged as

(8.69) Thus all we have got to know is the magnetic field and its derivative on the boundary of the hard core! To proceed we must determine the field configuration more accurately, i.e. we must in principle solve the Ginzburg-Landau equations (8.6la-b). It turns out to be impossible to write down an explicit solution representing a magnetic string, so we shall be contend with an approximative solution. We shall concentrate on a cylindrical symmetrical string. Notice that the flux is necessarily quantized (2.80) where n is related to the jump in the phase of the order parameter when we go once around the string~ For a string with n flux quanta we therefore use the ansatz: o lim ~(p) P -.. 0 (8.70a) ljJ(r)= ~ ~(p)exp[inqll with {
lim <p (p)
p -..

(Notice that the boundary condition at p=O, which is in accordance with our general description of a magnetic string as shown on fig. actually removes the singularity of exp[inqll at the origin). The exterior derivative of the order parameter is given by

439

II

d~
d~

~{~ dp
1jJ{ind~}

ind~}

outside the soft core it is therefore approximately given by


...

Since the gauge covariant exterior derivative of the order parameter must vanish outside the soft core, the magnetic vector potential is asymptotically given by A ... n~d~
q

This suggests the following ansatz for the magnetic vector potential lim A(p) 0 with { p.... 0 (8.70b) lim A(p) 1
p .... '"

(The boundary condition at p=O removes the singularity of d~ at the origin) We proceed to determine the static energy density for this type of configuration. Notice that (dpldp) = 1 ;
(dpld~)

= 0 ;

(d~ld~)

= ;i2 ;

(dp"d~ldp"d~)

;i2

(cf. the exercises 6.6.3 and 7.5.7). ISing that (8.7la) (8.7lb)
D~

~{ ~ ~(p)dp +

in[l - A(p)ld~}

nfl A' (p)dp"d~


q

the expression (8.60) for the static energy reduces to


H

;n2~~.(~')2 _ ;~.(~')2 _ ;~n2

(1

~2A)2 ~2 + ~~[~2_ lF

(Notice that ~ is negative:) As expected this only depends upon p so that it is manifestly cylindrical symmetric. It follows that the static energy per unit length is given by
(8.72)
J[~,Al

2'TTfJ[<p,AldP

o
2Tff{~n2~~,<~')2- ~!:!p.(<p')2- ;n2~(1-A)2~2 + a2pH2_1l2}dp

o
Next we functions is to plug (8.6la-b). *dp

48

want to find the equations of motion for the unspecified <P(P) and A(P) . The safest, but most complicated method, the ansatz (8.70a-b) into the Ginzburg-Landau equations Notice first that
pd~"dz

*d~

*(dp"d~) = dz p
and

* (dz"dp) = pd~

(cf. the exercises 6.6.3

7.5.5). It follows from (8.7la-b) that

440

II

D*D1jJ

Thus we obtain the following equations of motion (8.73a) (8.73b)


p[pl
A"

)?~T)<jl~[l-Al
p

(p<jl')'

-a[<jl2-l1<jl + n 2 (l-t) 2<jl

There is, however, a quicker way to obtain these equations. We know that the cylindric symmetrical equilibrium configuration must extremize the energy functional J[<jl,A] given by (8.72). The corresponding EulerLagrange equations are given by

Thereby we recover the equations of motion (8.73a-b). One should however be very careful when using this strategy. It only works because in this particular case the ansatz (8.70a-b) is in fact the most general cylindrical symmetric ansatz. In general a variation among a restricted subset of configurations need not extremize the action against arbitrary variations. We must then show that the second order differential equation (8.73) possesses a solution with the appropriate boundary condition. This is a somewhat difficult task. uSing advanced analysis ( Sobolev space techniques) it can be shown that the energy functional J[<jl,A] , which is positive definite, has a smooth minimum configuration satisfying the appropriate boundary conditions. (For details see Jaffe and Taubes(1981) In praxis one is often contend with computer simulations. Notice that <jl (p) " 1 A(p) " 1 is a trivial solution to equation (8.73) allthough it breaks the boundary condition at the origin, As a consequence the corresponding string is singular at the origin. Neverthe the less the energy density vanishes outside the origin. It is called a vacuum texture and obviously represents an infinitely thin string carrying n flux quanta. Although we cannot solve the equation of motion explicitly we can easily determine the asymptotic behaviour. In the limit of large p the equations of motion simplify considerably. If we introduce the functions
f (p)

= l-A(p) IP

and

g(p) = /-a:{l_<jl(P)

it is easy to show that asymptotically they solve the differential

441

II and g"(p)

equations
f "( p)

...
p .... '"

),2(T)f(P)

This implies the following asymptotic behavior f (p) ...


p....'"

Cl exp [-

-rf.ri 1

and

in accordance with our previous discussion of the coherence and penetration length. Okay, at this point we return to our estimate (8.69) for the static energy. Using the equation of motion (8.73a) it follows that n2h2 1 *BASB = q2),2 (T) pA' (p) [l-A(p) J<p2 (p) dZAdqJ Consequently the energy per unit length is given by (8.74)
J ...

1fn2~{:!:AI
q p

(p) [l-A(p) ]cp2 (P)}I

p=i; (T)

To proceed further we notice that the equation of motion for A(p) does not contain n explicitly! (It does contain n implicitly through the function CP(p) .) If we impose the condition CP(p)=l we get a vacuum texture for the order parameter, but we can still retain the boundary conditions for A(p). The solution to the reduced equation of motion (8.75)

p[;i]

A'

'" - T2(T) [I-A]

with

lim
p.... O

A(p)=O

and

lim
p.... '"

A(p)=l

is then likely to reproduce the vortex outside the hard core. In this approximation A(p) is completely independent of n. It follows that J is proportional to the square of n: (8.76) AS a consequence a vorte~ string with a mUltipZe fZu~ is unstable. E.g. a single vortex string with a double flux quantum has twice the energy of two widely separated vortex strings with unit fluxes. This shows that when the magnetic field penetrates into the superconducting region it will generate a uniformly distributed array of vortices all carrying a single flux quantum.
Remark: Actually equation (8.15) can be solved explicitly using Bessel fUnctions. The expression (8.14) can then be estimated using the known asymptotic behaviour of the Bessel fUnctianin the limit of small p. In this way one obtains J n21~[K] ( 8.11)
... q2),

(Tl

where K is the GinZburg-Landau parameter (8.61). For details see de Gennes (1966)

442

II

8,8

THE NM1BU STRING AND THE NIELSEN-OLESEN VORTEX

The soliton was introduced as a smooth extended version of a relativistic point particle. Now we will show that a similar interpretation is possible for a vortex string. first we must introduce a suitable generalization of the relativistic point particle known as a relativistic string or a Nambu string. .... A point particle is characterized by its position x at a the time t. In the four-dimensional space-time it sweeps out a time-like curve, the worZd line, cf. fig.164~ The action of free particle is particular simple being proportional to the arc length of the world line: dx].Jdx v S =-m V-g].JvdA dA dA

I/

Varying the world line we then obtain the equation of motion,which inertial frame reduces to 2 u XO x md 0 dT 2
XO

xO=t

Y""d

I:B
eT
.... eo

in

an

world sheet

~:~:"""

lino

Xl

Fig. l64a In a similar way a relativistic string will be characterized by its position ~(o) at the time t, where ~=*(o) parametrizes the spatial curve representing the string. It will be convenient to assume that the string is finite and that its endpoints correspond to 0=0 and o=rr In the four-dimensional space-time the string sweeps out a time-like sheet, the worZd sheet,cf. fig~64b. The action will be chosen to be proportional to the area of the sheet. If we parametrize the sheet by x].J = X].J(O;T) where 0 is a space-like and T is a time-like parameter, it follows that the induced metric on the sheet is characterized by the components h ........ dx].Jdx v ........ dx].Jdx v 00 = g (eo;e o ) = g].Jvdo do hOT = g (eO;e T ) = g].Jvdo dT ' etc. From the induced metric we then get the area element

\,

x2

~\Js

x2

443 (8.77)

II

We then define the action to be ." a=7T,--_ _ _ _ __ (8.78) S /(h a .)2_ haa h dad. '1 a=O

2;aJ J

FLom the action we get the equation of motion by varying the world sheet:

Notice that there are no restriction at the endpoints a=O and a=rr Thus we get not only equations of motion for the string but also boundary conditions for the end points. The general covariant equations of motion are extremely complicated, so we shall restrict ourselves to an inertial frame where the metric coefficients reduce to
g].lV(X)

Tl].lV

Then the Lagrangian density depends only upon


dx].l = x,].l and da We now obtain the displaced action

S(e:)

'1

r~(~~].l

a=O
3L a ].l

We must then demand


+ -].l2X ldad. ax d.

As Y tion
(8.79)

J J {aa '1 a=O


a

'2

a=rr

dL a~

+.... a.

aL } ].l d ax].l y ad.

is arbitrary this leads on the one hand to the equations of moa aL + d aL aa d~ a. dX].l

On the other hand it leads to the edge conditions:


(8.80)

~~

;~~

= 0

la=O

la=7T
I

Let us introduce the following abbreviations for the conjugate momenta:

444

II

(B.Bl)

Then the equations of motion supplied with the boundary conditions simplifies to
(B.B2)

Observe that we also have the following identities at our disposal:


(B.B3)

P ~ TA~

p~x a ~

= p2 +
T

4n 2 a

~ 2 = p2 a

~~

= 0

which are trivial consequnces of the explicit expressions for P~ and P~. Notice especially that at the ed the edge conditions Lffiplies that P~=O a ge. As a consequence x 2 vansihes at the edge, so the end points move with the speed of Zight. We may also introduce the four-momentum of the string:
(B.B4)

p~ =

a=n

p~da

(T

fixed)

a=O

The four-momentum is conserved (as it ought to be)


a=ndP~

a=O

f-

a=7Tap~

dT

Tda

--ada da J a=O

{p~(a=O) a

P~(a=n) a

Similarly we may introduce the angular momentum of the string:


(8.B5)

J~\l =

a=n

f [x~P~
~~\l ~ 0 dT .

a=O Exel'oise 8.8.1


Problem: Show that the angular momentum is conserved, i.e.

Exeroise 8.8.2
Introduction: Problem: (a) Consider a rigid rotating otring parametrized by XO~T ; xl= A(a-!n)CoswT ; x 2= A(a-~n)SinwT ; x 3= 0 Show that it solves the equation of motion provided
1 = !AwT

(b)

(Hint: Show that this represents the edge condition) Show that the spinning string has the momentum pO = ~ pI: p2= p3= 0 Show that the spinning string has the angular momentum
J12:

(c)

A2 n 2

16a

J23

= J31

= 0

445

II

from the above exercise it follows that a spinning string behaves like a particle with the rest mass
M

= =

Arr

4a

and an intrinsic spin

A2rr2
16a
aM

Notice especially that the spin grows ZinearZy with the square of the
mass.

Now what is the relevance of such relativistic strings in high energy physics? If we plot baryons (i.e. strongly interacting particles) in a diagram with the mass square on one axis and the spin on the other axis, then the baryons with the same isospin, strangeness etc. fallon straight lines! These are the so-called Regge trajectories. This remarkable property

J(spin)

9!
8!

b./

71

Regge trajectories

6l

suggests that we consider the baryons to be somehow composed of relativistic strings. Okay, with this simple minded remark I hope to have convinced you that relativistic strings,
Nambu strings, are very interes-

3l

4!

5!
./ //
/"

21 2

~/

./

///A
6

0/0

Fig. 165

10 M2(GeV2)

ting objects. As in the case of the point particles one would now be interested in smooth extended solutions to field theories that behave like thin strings.

ILLUSTRATIVE EXAMPLE: THE NIELSEN-OLESEN VORTEX.


The first example of a string-like solution in a classical field theory was found by my respected teachers, Holger Bech Nielsen and Poul Olesen, in 1973.*) They considered a relativistic field theory based upon a complex charged scalar field coupled minimally to electromagnetism. In analogy with the
~4-model

they furthermore included a non-trivial potenti-

al energy density given by

U(~)

i(I~12 - r2)2

Models with such a potential term are generally referred to as Higgs'


modeZs and the corresponding scalar fields are known as Higgs fieZds.

*) Vortex-line models for dual strings, Nucl. phys. B61 (1973) 45

446 The model based upon the Lagrangian density (8.86) where the Higgs field
~

II

, with

D~

d~

ieA~

is coupled to an abelian gauge field, is spe-

cifically referred to as the abeZian Higgs' modeZ. The associated field equations are highly non-linear (8.87a) (8.87b) SF
D*D~

ie = :f[~D~

~D~l

i.e. i.e.

V-g

1:.....3 \I (V~FJl\l) g

(AI~12_ Jl2)~

In this model we now proceed to investigate the purely static configurations, where furthermore the electric field is absent. Such a configuration tial one-form A is represented by a spatial function ~(~) and a spa-

= A(~).d~

. Iurthermore it corresponds to a solution

of the equations of motion (8.87a-b) precisely when it extremizes the static energy, which in the present model is given by (8.88) As a consequence the theory of static equilibrium configurations in the abelian Higgs' model is completely equivalent to the GinzburgLandau theory for superconductivity, provided we make the identification:

(This follows immediately from a comparison of

(8.60) with (8.88).)

The first important observation is that the abelian Higgs' model possesses no soliton solutions in the strict sence, i.e. there are no non-trivial stable static finite-energy configurations. This follows from an argument which is typical for gauge theories: A static finite-energy, configuration must satisfy the boundary conditions (8.89) lim B
r ......

=0

lim
r+co

D~

lim I~
r+co

I =

fA

Asymptotically the Higgs field is therefore completely characterized by its phase factor

In general the phase, ~(~), need not be single valued but can make a quantized jump wed! 2nn. But in the present case such a jump is not allo-

447

II

To see this, suppose the phase makes a jump when we go once around a distant closed curve

r . This distant closed curve can be shrunk to


(Topolo-

a distant point, while the shrinking curve remains distant!

gically it is a closed curve outside a ball and such a curve can clearly be contracted without intersecting the ball.) Notice that by continuity the phase must also make the jump duce a discontinuity in the phase factor field! Since the asymptotic phase 2nn when we go once around the shrinking curve. As the curve shrinks to a point we thus pro-

exp[i~(~)l. But such a dis-

continuity in the phase factor contradicts the smoothness of the Higgs

~(~)
=

is singlevalued the phase factor

exp[i~ (~) 1

-<lL(iL
I<p (r) I

is in fact trivial, i.e. it is single valued throughout the whole space. Thus we can remove this phase factor completely by a gauge transformation. To conclude we have therefore shown the existence of a ge where
<p
gau~

is real!
lim A(~)
r+'"

In this particular gauge the boundary conditions (8.89) reduces to (8.90)

= 0

lim <p (~)


r .... '"

Consider now the deformed configuration

It satisfies the boundary condition (8.90) for any choise of E. At E=O it reduces to the vacuum configuration, while for E=l it reduces to the given configuration. Consequently we can find a one-parameter family of finite energy configurations which interpolate between the vacuum configuration and the given configuration. This shows that a sta-

tic finite-energy configuration cannot be stable since we can "press" it down to the vacuum. To obtain interesting configurations we must therefore relax the boundary conditions (8.89). Rather than looking for point-like configurations we will now look for string-like configurations. Consequently we will concentrate on configurations which are independent of z. This time we therefore impose the boundary condition that the energy per unit length along the z-axis is finite. This restriction implies the following boundary conditions: (8.91) lim B p.... '"

lim
p.... ..,

D<P

lim
p.... '"

I <p I

448

II

where p is the distance from the z-axis. As before we conclude that asymptotically the Higgs field is characterized by its phase factor:
cj> (r)'"

p-+co

If exp[ iIP (r) 1


2nn when we go once around a distant closed

This time, however, there is nothing to prevent the phase of the Higgs field to make a jump curve. A distant closed curve cannot be shrunk to a point without

intersecting the z-axis.) Notice,however, that if the phase makes a non-trivial jump then the Higg's field must necessarily vanish somewhere inside the string corresponding to a discontinuity in the phasefactor

Since the gauge covariant exterior derivative of the Higgs field vanishes outside

the string we get as usual


dIP - eA

...
p-+co

Consequently the jump of IP is related to the magnetic flux in the string:


~

lim
Po -+(0

B P<Po

lim
Po-+CO

f A
P=Po

lim
Po-+
CO

ef dIP
P=Po

21T ne

This is also in accordance with the equivalence between the abelian Higgs' model and superconductivity. The number n can consequently be identified with the number of flux quanta in the string.
Since the flux is quantized it is impossible this time to interpo-

late between the vacuum configuration and a configuration with a nontrivial flux. A string with a non-trivial flux is thus topologically
stable!

We proceed to look for configurations which minimize the static energy per unit length. As in the Ginzburg-Landau theory we shall assume that the penetration, length is considerably larger than the coherence length, i.e. in the present case we assume
(8.92)

We can then carryover the conclusions obtained in the previous section concerning magnetic strings in a type II superconductor. In the sector conSisting of configurations carrying a single flux quantum the groundstate is the cylindrical symmetrical configuration. In the abelian Higgs' model this is known as a Nielsen-Olesen stping. In a sector consisting of configurations carrying multiple flux quanta there can be no exact ground state. This is because a cylindrical symmetrical configuration is unstable since its energy grows with

449
the square of the number of flux quanta, cf. Gordon model one can however construct approximative groundstates consisting of n widely separated Nielsen-Olesen vortices.

II
(8.76). As in the sine-

Notice that a string-like excitation such as the Nielsen-Olesen vortex cannot itself represent a physical particle since it has infinite energy due to its infinite length. So if the Nielsen-Olesen vortex is going to be physically relevant we must find a way to terminate it. This can be done by including additional point particles in the model, which then sit at the endpoints of the string. As the Nielsen-Olesen vortex carries a magnetic flux these additional particles must necessarily be magnetic monopoles (or anti-monopoles). Furthermore the magnetic charge must be quantized since the flux carried away by the string is necessarily quantized. As we shall see in the next chapter that is fine: Magnetic charges are in fact quantized according to the rule
g

211 = ne .

Suppose then we introduce point-monopoles in the abelian Higgs' model. Notice first that it is impossible to introduce just a single monopole. This is because a single monopole is characterized by a long range magnetic field and that cannot exist in the abelian Higgs' model due to the Meissner effect. Let us clarify this point: You might object that the Meissner effect could just squeeze the magnetic field into a thin string extending from the monopole to infinity. But being infinitely long the string would carry an infinite energy and that is not physically acceptable. (The Coulomb field created by an ordinary charged particle is acceptable because the energy stored in the field outside a ball containing the particle is always finite. The infinite self energy of the electron comes from the immediate neighbourhood of the electron, not from infinity.) This means that we can only introduce monopoles in terms of monopole-anti-monopole pairs. A magnetic field line extending from the monopole can then be absorbed by the anti-monopole. Suppose then that we try to separate such a monopole-anti-monopole pair. This will necessarily create a thin magnetic string between the monopole and the anti-monopole. The string has a typical thickness given by the penetration length, i.e. the radius of the string, r o ' is
R<--

f'A
e

Since the string carries the magnetic flux order


B
R<

g, where g is the magne-

tic charge of the monopole, the magnetic field strength will be of the

iT?o
in the string is of the order

Thus the magnetic energy stored

450

II
=

E ~ ~B2(wr~~)
where i

~ w;z

is the distance between the monopole and the anti-monopole. It

will therefore require an infinite energy to separate the monopole-antimonopole, i.e. it is physically impossible to separate them and thereby produce free monopoles (or anti-monopoles). One says that monopoles are confined in the abelian Higgs' model. This is to be contrasted with ordinary electrodynamics. There one may also consider bound states consisting of two oppositely charged particles, say hydrogen or positronium. But in that case you can easily knock off the electron and thereby produce free electrons. To summarize: (a) In standard electrodynamics the binding energy for a bound state consisting of oppositely charged particles is given by the Coulomb potential VCr)
= -

~! 4WEo r

and it only requires a finite amount of energy to separate them. (b) In the abelian Higgs' model the binding energy for a bound state consisting of a monopole-ant i-monopole pair is given by the linear potential VCr) = l~ r
2

and it requires an infinite amount of energy to separate them, i.e. the monopoles will be permanently confined. Returning to the monopole-anti-monopole pair in the abelian Higgs' model we see that the potential generates an attractive force (which is independent of the distance for large distances). To prevent the pair from collapsing we must therefore put it into rapid rotation. In this way we produce a composite particle consisting of a spinning magnetic string with a monopole at bne end and an anti-monopole at the other. lor a sufficiently long rapidly spinning string we have thus come back to the Nambu string. Now what is the relevance of these considerations in high-energy physics? Ebr various theoretial reasons hadrons (i.e. strongly interacting particles) are generally thought to be composite particles, the constituents of which are called quarks. This so-called quark model has had a considerably succes in explaining the observed spectrum of hadrons (including some predictions of hitherto unknown hadrons) .

But there is one great puzzle concerning the quark model: A free

451 II quark has neVer been observed as the outcome of a scattering experiment in high-energy physics.

(Some famous solid-state physicists claim

to have observed free quarks in a very beautiful but delcate experiments*! but their observations have not been confirmed by other groups and they are not generally trusted.) Thus quarks seem to be confined. How can we explain this quark confinement? One possible idea looks as follows: The quarks interact through the strong interactions, and it is customary to introduce a field, the so-called colour field, which is responsible for this interaction (in the same way as the electromagnetic field is responsible for the electromagnetic interaction.) In analogy with the electromagneitc field the colour field comes in two species known as the colour electric field and the colourmagnetic field. The quarks carry colour electric charges and thus act as sources for the colour electric fields. One then adds as a basic hypothesis that in the quantized theory the coloured vacuum will act like a superconducting vacuum, but in contrast to the abelian Higgs' model we are supposed to reverse the role of electric and magnetic fields,

i.e. this time it is the colour electric field which is expelled due
to the Meissner effect. (0 course this must eventually be proven directly from basic principles if we are really going to trust the argument. This is the real hard part of the game and only little progress has been made due to great technical difficulties in the quantum theory of coloured fields.) Okay, suppose the above hypothesis concerning the coloured vacuum is correct. Then we can verbally take over the arguments concerning magnetic monopoles in the abelian Higgs' model. gion. It is however possible to collect free quarks cannot exist because the colour electric field is squeezed into a finite rethem in a quark-ant i-quark pair, and, due to a greater complexity in the structure of coloured fields, it is also possible to put three quarks together in such a way that the net colour charge is zero. (Incidentally this is where the name "colour" comes from: There are three basic colours in nature - red, green and blue - and if you "add" them you get white, i.e. the net colour vanishes:) Thus the following picture arizes: Space is divided into two regions. The dominating region filling up space almost everywhere consists of the superconducting coloured vacuum. But here and there we find small regions filled with quarks and colour electric fields. These "normal" regions, where the superconductivity breaks dowvn, are called bags, and they act as prisons for the quarks. Thus a hadron is an
*) G.S. LaRue, W.M. Fairbank and A.F. Hebard, "Evidence for the existence of fractionally charged matter", Phys. Rev. Lett. 38 (1977) 1011

452

II

extended object consisting of a bag containing pair or a couple of three quarks.


BAG WITH TIIO QUARKS

either a quark-ant i-quark

BAG WITH THREE QUARKS

(meson)

(baryon)

Fig. 166

Superconducting vacuWll

"

I
-~

Consider a quark-ant i-quark pair. If we try to separate the quark from the anti-quark we necessarily produce a thin colour electring string between them and just as in the abelian Higgs' model it can be shown completely elementary that the energy stored in the string is proportional to the length of the string. cal linear quark potentiaZ: Thus we arrive at the typi-

VCr)

kr
the bags. Such rapidly

Finally we can excitate hadrons by rotating like spinning Nambu strings!

spinning bags will be elongated and we therefore expect them to behave So I hope you now see the importance of the abelian Higgs' model. It serves as a theoretical laboratorium where we in an elementary way can test the consistence of various properties of the model, before we try to examine these properties in the more complex models which are thought to describe systems, that arp. actually found in nature.

We conclude this section with yet another argument which supports the interpretation of a Nielsen-Olesen vortex as a smooth extended Nambu string. First we rearrange the action of the Nambu string slightly. Consider an inertial frame and let us identify the parameter T with the corresponding time in the inertial frame, i.e. we put T=t The string is therefore parametrized as x = x(O",t) The arclength of the string is given by ds = The tangent vector

jdXdX dO"dO"

dO"

~~

thus becomes a unit vector. The velocity of a

point on the string is given by

453
-+

II

ax at

-+

In our case only the component perpendicular to the string is physically relevant. Evidently it is given by -+ ax ax axax v.! = at - as (asat)
-+

-+

-+-+

-+2 v.L

(ax)2 _ (axax)2 at asat

Next we observe that metric coefficients are given by 3s (3X3X) ao asat ; h


TT=

-1 +axax atat

(sing this we can rearrange the area element as follows

V-h = v' (hOT )


t2

00

TT

Consequently we finally obtain (cf. (8.93)

(6.49)):

- .1:...J )1 _ ~2 2rro 1tl

dsdt

Next we consider a thin Nielsen-Olesen vortex string moving around in space. It could be a static vortex which we have "boosted" into a uniform motion or it could be a more complicated motion specified by some appropriate initial data. The vortex will be a solution to the equations of motion derived from the Lagrangian density (8.86), but we are not going to specify the potential in this argument. All we use is that the model aloows stringlike solutions, where the fields are almost in the vaccum state outside the thin string. The deviations from the vacuum state will be exponentially small and we shall simply neglect them. Thus the Lagrangian density vanishes outside the string and the Lagrangian density therefore acts as a smeared out a-function. We can now find an aprroximative expression for the action of the string. First we introduce a rest frame for a small portion of the
~

string. The rest frame moves with the velocity yO,zO and notice that ZO

, i.e. perpendicular

to the string. The coordinates in the rest frame will be denoted to,xO, is simply the arc length for the string! In the rest frame the string reduces to a static string. The Lagrangian density thus becomes equal to minus the energy density,
L = - H

If J denotes the energy per unit length in the rest frame we have therefore shown

But then we get the following expression for the action per unit length

454
S

II dxOdyOdsdt

= J LdxOdyOdzOdtO = J LJi-vl
dt =

where we have used the transformation formula

dtjl-~~

Consequently we have deduced the following approximative expression for the total action
(8.94)

But that is precisely the action of the Nambu string! terms of the energy density per unit length, J:
C/.

furthermore

it allows us to re-express the slope of the Regge trajectories, a, in


1 271J

SPACE
DIAGRAM

~rest frame
,

'~
\

,/

-+

Vortex

x
8.9 SINGULAR FORMS

Fig. 167

We can also extend the exterior algebra to include singular differentiaZ forms. A complete discussion falls beyond the scope of these notes. First we will discuss the naive point of view generally adopted by physicists and then we wfll give a brief introduction to the framework of distributions, which is a concise formalism adopted by mathematicians. In the naive point of view a differential form of degree k

1
T = -k' T. 1
1 .. 1

.
k

(x)dx

i
A

Adx

k
T.
1

is called a singurar form if the coordinate functions the Heaviside function or the singularity of the

not smooth, i.e. they possess singularities like the discontinuity in a-function. We will assume that the usual theory of exterior calculus can be ex-

l.. k (x)

are

455 II tended in a reasonable way to include singular forms, so that we can


use rules like
d2

=0

or Stokes'theorem even if singular forms are a-function:

involved. The first thing we will generalize is the Definition 5 If M is a manifold with metric g, then the

a-function,

ap

peaked at the point

Po

is defined to be the singular scalar fie9d apo : M ~

R
n

characterized by the property that


=

fM

~ap /gdx
0
~

A Adx

for any smooth scalar field

. a ' In coorpo is represented by a singular Euclidean From the integral property of ap

We can easily work out the coordinate expression for dinates the scalar field function, which we denote we then get apo
~(x)

But then

i.e.
where

C!-(x)

_i_on (x-x )

,fg('Xj"'

an (x-x o ) is the usual Euclidean a-function. Using singular forms we can now associate an electric current

to a point particle. This is a highly singular one-form, which vanishes outside the world line of the particle. Using an inertial frame we saw

in section 1.6 that the electric current was characterized by the contravariant components
(1. 34)

The covariant coordinate expression for the

a-function suggests

that we define the components of the current to be 1 dx ll dT J 11 (x) = q - - , a (X-X(T(8.96 ) I-g(x) dT

where this expression is valid in an arbitrary coordinate system. To justify that this is the covariant expression representing a singular vector field we must show that Jll(x) x dd : transforms contravariantly. Observe that a general expression like

f~ (x-x (T
where
~

dT

is a scalar field, does not transform contravariantly! That

everything work out all right is due to very special properties of the

456

II

8-function. So let us introduce new coordinates the components

(yl, .. ,yn)

and see

what happens. In the new coordinates the current is characterized by

But __ 1_8 4 (y -y(T)) I-g (~ 0

= __ 1_0 4
I-g (XJ

(x -X(T))
0

since it represents a scalar field and therefore we get JJ.l(yo)


(2)
=

qJ

I-g(xo )
8 4 (x O -X(T)) by

8 4 (x O -X(T))

*~(X(T))d;TV

dT

Now we use that replace

is peaked at
O)

Xo . Consequently we can

~ dXV(X(T))

~ dXV(X

and then move it outside the integral

whereby we obtain Jll(y )


(2)
0

~ (x
dX V
0

v 1 _ 8 4 (x -X(T) )~d dx T )q _ _

l-g(1:)

~(x V
dX

) JV(X )
0

(1)

So everything is okay!

Worked exercise 8.9. 1


Problem: Show that the singular current i.e. 5J = 0

with components (8.96) is conserved,

In a similar way we can introduce a singular 2-form particle. The singular forms tions:
dF = 0

representing

the electromagnetic field generated by an electrically charged point


F

and

then obey the Maxwell equa-

ILLUSTRATIVE EXAMPLE:

HOW TO HANDLE SINGULARITIES.


F. It is
x

You might wonder how we can control the singularity of

infinite at the position of the particle. How can we differentiate a function which is infinite at a single point like the function Usually this is done by a regularization procedure. Consider the static Coulomb field as an example. For simplicity we work in the ordinary Euclidean space
Ei _ - L
-41TE O

I?

R3

and use Cartesian coordinates

etc. The Coulomb field is characterized by the components

'?'

xi

We regularize the field making the exchangement


r ~

f:rZ+ET

457 In this way we get the regularized field strength

II

which is now smooth throughout the complete Euclidian space. If we differentiate this,we get

Consequently the regularized electric field corresponds to a smooth charge distribution

which asymptotically vanishes like around s-3.


Q(E)

2r -5 s .

For small
pes)

it is peaked

r=O

where it varies like

Observe that the total charge


4n

foo
o

p(s)r 2 dr
E 1

is independent of desired result lim


peE)

s.

(Use

r=sx).

Thus we finally obtain the

= q03 (x)
Fig. 168

i;2m which we conclude that

Observe that the electric field panied by a magnetic field from the equation B(s).

E(s) in this example is not accomThe magnetic field can be obtained

But the curl vanishes automatically as a consequence of the spherical symmetry.

Finally we will sketch how to introduce singular forms within the framework of distributions. Consider first the case of ordinary functions on the Euclidian space n ~ R , i. e. each point Rn. Here a function f is ~haracterized as a map R x in Rn is mapped into a real number f(x) We want now to construct singular (or generalized) functions. This is done in the following way: First we choose a test space. In the theory of generalized functions this test space consists of all smooth functions with compact support, and it is denoted D(Rn ).

458
The functions in the test space are called test funations. We can now formulate the following definition Definition 6 A generalized function (i.e. a distribution) T is a Zinear functional on the test spaae, i.e. T is a linear map:

II

Now, what is the connection between ordinary functions and generalized functions? Apparently they are defined in 2 completely different ways. To understand this connection we observe that any ordinary continuous function f : RO ~ R generates in a canonical fashion a linear functional D(Rn ) ~ R which we denote by T to f avoid confusion. This linear functional is defined through the Hilbert product (8.97) Tf [$ 1 =
-+ f n f (xl

$(x)d x

-+

where $ is any test function. The integral is well-defined becaUSe $ has compact support. Thus any continuous function f generates a generalized function T , f which by abuse of notation is usually denoted f too. But the conVerse is not true. Consider e.g. the 8-function. It is defined as the linear functional
O[$l = $(0)

In analogy with (8.97) this is often rewritten in the more informal w~

o(x) $(~ldnX = $(0)

But there exist no ordinary function " Ii (x) " satisfying this integral identity, so strictly speaking it makes no sence!
Ok~, so much for the generalized functions. Consider a manifold M with metric g. We want to introduce singular k-forms. As the test space Dk(M) we use the set of all smooth k-forms with compact support. They are called test forms. We can then define:

Definition 7 A weak k-form T is a linear functional on the test space i.e. T is a linear map

Dk(M)~R

Let us fir.st check that an arbitrary smooth k-form T can be represented as a linear functional so that the weak k-forms include the smooth k-forms. Let T be a smooth k-form, then the linear functional is defined through the Hilbert-product (8.99 )

T[Ul

def

<Tlu> =

fk
1
M

i l i k

(x)

u.
J. l

'~n

. (x)

vg d nx
U has compact

where the Hilbert product is well-defined because the test form support.

Let us discuss some specific examples of weak forms. First we consider the electric current J associated with a point particle in Minkowski space. In the naive approach it was characterized by the contravariant components

459
.J.l( )

II

J'

rL=q J .r--::r='-g\X}

<5

(X-X(T)) . dT

dx].l

dT

Let

U be a test form.
=

<Jlu>

t
q

Then formally we have:

J v'-~(X)
M

<5

(X-X(T))
<5

:].1 dT

UjJ (x)

Fg"(XT d 4x

f :~ [f

U].I(X)

4(X-X(T))d 4X] dT

f :].1

Uj.l(X(T))dT

=q

fr U

Within the frame work of distributions we therefore define the singular current as the linear functional:
(8.100 )

where r is the worldline of the particle! Next we consider the singular Coulomb field in the ordinary Euclidian space R3. Its components are infinite at the origin and we must therefore define the Hilbert product through a limit procedure. Let us put M = {~I " ;t iI > d. Then we represent E by the linear functional E
(8.101)

ElUl def. = l~m ~


E+O 0

~ Ui(x)d x

xi

...

ME r

(That the limit is well-defined follows easily if you work out the integral in spherical coordinates!) Now if the weak forms are going to be of any use, we must be able to extend at least part of the exterior calculus. The differential calculus is the easiest one to extend:
(a)

The

e~terior

dBPivative:
Then

Suppose first that T is a smooth k-form. functional dTlul = <dTlu> = <TI 6u>

dT

~s

represented by the linear

where we can throw away the boundary term, because the test form U has compact support. But test forms are always smooth, so if T is a weak form, we can immediately generalize the above result and define dT to be the linear functional
(8.102)

dTlUl d~f. Tl&Ul


T is a smooth k-form.
Then

(b) The ao-differentia.Z:

Suppose first that functional 5Tlul = When

&T &T

is represented by the linear

<&Tlu> = <Tldu>
to be the linear functional

T is a weak form, we can therefore define

(8.103)

Exercise 8.9.2 Problem: Let M be a Riemannian manifold. Show that the Laplacian of a weak form T is represented by the linear functional
(8.104)

460

II

Exercise 8.9.:3 Problem: Show that we can extend the dual map to weak forms in the following way: If T is a weak k-form, then *T is the weak (n-k)-form represented by the linear functional
(8.105)

*T[ul d~f { (-I)k(n-k)T[*Ul on a Riemannian manifold


-(-I)k(n-k)T[*Ul on a manifold with Minkowski metric.

Having generalized the differential operators we can easily check some of their fundamental properties. For instance we get

d T = 0

and

&2T = 0

even if T is weak. This fOllows immediately from the definitions (8.102)and (8.103) and the corresponding properties of the test forms. [E.g. we obtain d 2T[Ul = T[&2Ul = T[Ol = 0 so that

d 2T

is the zero-functional and similarly for

&2 T .l

Worked exercise 8.9.4 Problem: (a) Let J be the singular current (8.1.00) associated with a point particle in Minkowski space. Show within the framework of distributions that it is conserved i.e. ~J = 0 (b) Le~ E be the singular Coulomb field (8.101)in ordinary Euclidean space R3 . Show within the framework of distributions that
-&E

=~ Eo

8(x)8(y)8(z)

Observe that the space of weak k-forms incorporates not only all smooth k-forms but also all orient able regular k-dimensional domains! If ~ is a k-dimensional orient able regular domain we represent it as the linear functional (8.106)

mul

= f~

Since test forms have compact support we can even allow ~ to be an unbounded noncompact domain. This is a very powerful generalization of the usual 8 -function: A O-dimensional regular domain ~ consisting of the single point Po is represented by the linear functional
~[~l

= fp o ~ = ~(Po)
qr

i.e. it generates the usual 8 -function. Using this notation we observe that the singular four-current (8. 100) associated with a point particle is essentially identical to the worldline of the point particle since
i.e.

J =

Observe that for regular domains we have introduced the boundary operator,a:~a~ This can now be interpreted as a differential operator! The boundary of ~ is represented by the linear functional

Using Stokes' theorem and (8.102)this can be rearranged as


a~[Ul

= f~du = ~[dul = &~[Ul

so that the boundary Operation coincides with the co-differential! The important property aa~ = ~ is now seen as a special case of the rule &2T = 0 Coming this far you might think that everything can be done using singular forms. But that is not true! When we try to extend the wedge product or the integral calcu-

461
lus we run into trouble: Consider the wedge product of a smooth k-form T represented by the linear functional TAS[U] d~f'<TASlu> = fM*UATAS (-l)kmf*uASAT Using (7.58) this can be rewritten as TAS[U] = f*(U'S)AT = <Tlu's> = T[U'S] and a smooth m-form

II

S . It is

This shows immediately that we can generalize the wedge product to the case where one of the factors is weak and the other smooth. If T is a weak k-form and S a smooth m-form then we define TAS by the linear functional (8.107) TAs[Ul d~f'T[U'S] but if S is singular too, then this will not work, because U'S is then no longer a test form. If we introduce a suitable topology on the testspace vk(M) and use limit procedures then one can sometimes extend the wedge product to a pair of weak forms or similarly extend the integral to m integral of a weak form over a regular domain, but it is not always possible! These deficiencies of the extended exterior calculus should not be underestimated. "Distribution" is not a magic word you can use to justify any calculation you want to perform with singular quantities. Consider e.g. the derivation of the electromagnetic energy momentum tensor in section 1.6. Here we discussed the electromagnetic field generated by a collection of point particles. The equation
(1. 40)

had a central position in the argument. Consider the right hand side. Here the electromagnetic field strength Fa is singular at the position of a particle and so is the current JY. But then theIr product is not well-defined, not even in the sense of distributions! Consequently the argument in section 1.6 is only a heuristic argument of didactic importance, not a proof in the strict sense. Weak forms were introduced by de Hham. Among other things he was motivated by the electric current associated with a point particle. He therefore used the name "currents" for weak forms in general. That is however misleading in a physical context and I have therefore adopted the name "weak form". If you want a more rigorous treatment you should consult de Rham [1955] or Gelfand and shilov [1964].

SOLUTIONS OF WORKED EXERCISES:


No. 8.3.3 (a) We use spherical coordinates adapted to
2 S+ .

S! = {(r,e,\p) Ir=l; o<e<1!. - -2' 0::~27f} y = rSineSin<p x = rSineCos<p


dx

SineCoslIdr +
SineSin~r

rCoseCos~e rCoseSin~e

rSineSin~ rSineCos~

dY

When we restrict dx zero, whereby we get


dx

and -

dY to S! we put
SineSin~ SineCos~

equal to

and

dr

equal to

cosecos~e CoseSin~e

dY

462
i.e. dxNly = Cose'Sinedelld4l

II

Then we replace the geometrical volume element lume element ded~ and finally get

delld4l with the "Riemannian" vo-

dxNl:y =

e=o ~O (b) Spherical coordinates are also adapted to B3 = {(r,e,~)lr~l


o~e'~71

s!

t71 CoseSin6'dedlP = 71 B3 :
0~~271}

Observe that dxNlyNlz is the Levi-Civita form in coordinates we can therefore rearrange it as dxNlyNlz = /'idrNlelld4l Using (6.36) we now get g = r4 Sin z e volume-element
drded~

R3 . In terms of spherical

i.e.
drAd6A~

When we replace the geometrical volume-element

with the "Riemannian"

we finally arrive at

I1
No. 8.4.1 A smooth 2-form B

I7I

I2:2sinedrded~ = ~
~O

r=O e=o

which we recognize as the volume of the unit ball.

can be decomposed as

B = B3(x,y,z)dXAdy + B1(x,y,Z)dyAdz + Bz(x,y,z)dzAdx Due to linearity it suffices to verifY Stokes' theorem term by term. We have thus reduced the problem to the consideration of a smooth 2-form on the form B = B(x,y,z)dxAdy

We then get

Cartesian coordinates are adapted to the unit cube and the two integrals can now easily be computed.
(1)

IadE = IIl~ZAdxAdy

I
I

f f x=O y=O z=O

f~Zdxdy
y
.;
/'

J [B(x,y,l)-B(x,j,O)ldxdy
'" '"
Fig. 169
I

x=o y=O The boundary all consists of six surfaces but only two of them contribute, since dx=o (or dy=O) along surfaces where x (or y) is constant. The bottom and the top of the cube are oppositely oriented: The Cartesian coordinates (x,y) are positively oriented on the top but negatively oriented on the bottom (see fig. 169). We now get (2f B = JaIlB(X,y,Z)dxNly = all

x
I

J J B(x,y,l)dxdy -

J B(x,y,O)dxdy
0

x=o y=o x=o y=O from which we see immediately that (1) and (2) are identical.

463

II

No. 8.4.3 (a) Let r have the parametrization z(t) = x(t)+iy(t) , a<t<b. When we evaluate the complex line integral we must make the substitution dz

[ax + ik]dt at at we must find the restriction of

When we want to integrate the one-form h(z)dz dz to r,


dz

~ at h'>t = [ax at

+ ~at ~]dt

and then exchange the geometrical volume-element dt with the "Riemannian" volume-element dt Both integrals therefore reduce to ([f(X,Y)+ig(X'y (bl Notice that dh(z) =

)][~~ + i~]dt

~~ dz + ~~ dz

but h(z) is holomophic and therefore ~~ 0 (Cauchy-Riemann's equation). Consequently we get dh(z) = ah dz az The formula is now a simple consequence of Stokes' theorem

fr
No. 8.5.1

h'(z)dz

= J dh(z) = J

ar

h(z)

= h(B)-h(A)

(c) According to exercise 7.6.9 h(z)dz

is a closed form. Therefore we get

r d[h(z)dz] = 0 Jr h(z)dz = fa~h(z)dz = J~


J
a~

o
J cpAd*d1/l
~

* [cpd1/l]

CPA*d1/l = J d[cpA*d1jJ] J an
~

= J dcpA*d1/l
~

J~dCPA*d1jJ - f~CPA*A1/I = (-l)n-l[f~*d1jJAdCP


(-1)n-l[<d1/lIdCP> + <A1/IIcp>] Which shows (8.32). If you interchange cp
a~

+ fn*A1/I ACP]

and 1/1

you get + <ACPI1/l>]

J * [1/Idcp]
a~

= (-1)n-l[<dCPld1/l>
*[1/Idcp]

Then you subtract these two identities and get (8.33):

J *[cpd1jJ] - f
If we put
Cp=1jJ

an

= (-1)n-l[<A1/IICP>

<1jJIA~J

'1']

in (8.32) we finally get

<dCPldCP> + <CPIACP>
If

= (-l)n-lJ

* [cpdcp] an

cp is a harmonic function that vanishes on the boundary,we conclude <dCPldcp> = 0

464

II

Therefore d~=O so that must vanish throughout ~

is constant, but since it vanishes on the boundary it

U
~=~

No. 8.5.2 (a) This is simply the result of exercise 8.5.1 formula (8.32) when we put - :5-(3)dx + ~ + ~dz) (b) d~ r r r r
*d~

- :5-(4:YAdz + "ZdzAdx + ~Ady)


r r r r

Using this we get <p*d</> = - .l;.[xdyAdz + ydzAdx + ZdxAdy] r The integrals are now easily computed: - lsin9d9Adlp
r

IJ -

aM

1 Sin9deAdlp = - E

Jrr J2rr Sin9d9d~


~O

_ 4rr
E

9=0

No. 8.6.2 (a) The Lagrangian density is a scalar field. Consequently

~=1h+ aL aIE=O a</>'" alai)


field,while

"'"
0jl'"

is a scalar field too. From exercise 6.9.3 it now follows that

~~ is a scalar

~ a (ajl</> )

are the contravariant components of a vector field.

(b) The left hand side is the Hilbert product of

and

dW :

ArJ.(a ~);.:gd4x = <AldW> = J*AAdW J ~ rJ. ~


Using the generalized theorem of partial integration (section 8.5) this is rearranged as

But here the boundary term vanishes since

~=O

on the boundary

an. Using that

d*A = *&A
(see 7.62) we finally get

f ArJ.(arJ.~)/=gd4X = <AldW> = <5AI~> = -J arJ. (/=gArJ.)~d4x


~ ~

Moral: When one of the factors vanishes on the boundary then adjoint operators". 0

n&

and

are

No. 8.6.3
(al

SI =

qJ:2~ ~~dA
J
~

qJ:2[J~Av(X)04(X-X(A))d4X]~:dA
J A (xl J ll(x)/=gd 4X n ll

A~(Xl[qfA2~~ ~4(X-X(A))dA]/=gd4X Al I-g

465

(bl The dynamical quantities to be varied are the position of the particle, and the gauge potential, All (xl :
x~(Al
+

x~(Al+EY~(Al
B~

A~(xl + A~(X)+EB~(x)

=0

on

all.

We then proceed in

4 steps:

1. Variation of Sp: This has been discussed in section 6.7: T2 d2 xa dSp a dx~ dxv = (~+ mr ~v---;rr ---;rr)Ya(T)dT IE=O TI 2. Variation of SF: This has been discussed in exercise 8.6.1:

a.e- J
J

F- = dS a [;:gF~vlB (x)d 4 x de: 1 E=O Il ~ v

3. Variation of s:} with respect to the particZe trajectory

q(>~[X(A)+EY(A)l(a:: + e:~)dA
i.e.

(where we have performed a partial integration on the last term) A


q Al

J J

[~ _ lfuL]dx~
axv
ax~ dA

v dA
Y

2 dx~ V Fv~iiJ\ Y dA

IT2 ~S---;rr dx S dT Ya
TI

Al

where we have re-introduced the proper time

4. Variation of SI with respect to the field: Using the result obtained in a) we easily get
I -dS d-- = s 1 s=o

J ;:gJv(x)B
Il v

(x)d 4 x

Collecting all the results we finally obtain

o = dS

dSls=o

But

Ya and Bv are arbitrary, so this is only consistent if (8.37) and (7.85) are satisfied.

466

II

No. 8.7.3 (a) To get the equations of motion for B we perform the variation
A~A+e:U

i.e.

We then obtain H(e:) = !<BIs> + e:<dtJls> + !e: 2<dt.J1du> + 1<D1/I1D1jJ> + !fie:<u1/JID1/J> - !fie:<D1/IIlJll!>

+~2e:2<lJll!1lJll!>
Thus we must demand dH . o = dE = <dt.J Is> + ~[<lJll! 101/1> - <D1/I1lJll!> 1 1e:=0 If we remember that this as follows , we can easily rearrange
-

o = <u1&B

~(~D1/J

~
-

This leads to the equations of motion

- &B = ~(~D1jJ
1/1 i.e. We then obtain
~

1/1~)

(b) To get the equations of motion for 1/1 we perform the variation 1/1 + e:$

Thus we must demand dH o = ~e:=~ !<D1/IID$> + 1<D1D1/J> +

JaI1/l12(1/I au - + 1/1) - e:

= l<D1/lID> + 1<D1D1jJ> + <al~121/11> + <1~1/I> _ Here half of the terms involve and the. other half involves As usual we may formally treat them as independent variation. We therefore get

o = <DID1jJ> + <$12~1/I>
Now we are almost through. But we should be a little careful about the first term: Here we use that <A</l 1D1jJ> = fAi;poi1/l so that we can rearrange it as
<D 1D1jJ> = <$ I &D1jJ> + ~$ I * CAA *D1jJ) > Thus we end up with the condition:

.rg d 3 x

o = <$I&D1jJ + ~W*D1jJ) + This leads to the equation of motion


- &D1jJ -

2al~121/1>

~CAA*D1jJ)

2al~121/1

Dualizing it we finally get

467

II
=

au 2alwl2wE
No. 8.9.1

9 d*0IJ! - iflAII*OIJ!

9 Cd - itiA)*OIJ!

D*OIJ!

The verification of the current-conservation is done by a brute force calculation:

-= a (Fg Jl1) =....!L a v-g 11 ;:g 11


1 =_.....9....

J- I i4
l1 dx dT dT

(X-X(T))dT

=_....!L ~

l1 rdx a- tS 4 (X-X(T))dT dT dXl1(T)

but using the chain rule, this can be rearranged as

I=g

J~ tS 4

(X-X(T))dT

= _....!L [tS 4 (X-X(T) )]T=_ = 0

;=g

T=-OO
as

since

is fixed and Ixo(T) I

00

ITI

00

No. 8.9.4 (a) Let $ be a test-form of degree 0 (i.e. a smooth scalar field with compact support). Using (9.63) and (9.66) we get

&J[] = J[d] = q frl

= q[(T=-)

- (T=-OO)] = 0

since has compact support. Consequently &J is represented by the O-fUnctional. (b) Let be a test form of degree o. Using (8.101) and (8.103) we get

(*)

-&E[] = -E[d]
But observe that

= -lim <Eld>
E+o

ME

i E=~E----L..dW 471Eo r3 471EO

where

wei) =1. r

Inserting this into lim o E+o

(*) we get

-& E[] = ~

<dwld>M
E

Now we use Green's identity

(8.32) and rearrange it as

= ~ lim
7IEo E+o But
$(~)

[f *dW - <Awl>]

aM

E ME so the last term vanishes. In the first term we integrate

W is
with

harmonic in
(O):

over a sphere shrinking to a point. In the limit we can therefore safely replace

Here we obtain

*dW = - L(xdylldz
r3
so that we get (ME

+ ydzlldx + zdXNly) = -sin

6d611d$

is outside the sphere, so it gets the opposite orientation):

468

II

But then we have shown ,


-&E[CP] = ~ o

cp(O) ,

which means that

-&E

is represented by 3- times the a-functional. o

469

II

chapter q DIRAC MONOPOLES


9,1 MAGNETIC CHARGES AND CURRENTS
Previously we have only been considering conventional electrodynamics according to which magnetic charges are excluded. This is in accordance with experiments, where electric fields are always generated by electrically charged particles, whereas magnetic fields are generated from electric currents. Never-the-less there is apriori no reason to exclude magnetic charges, and as shown by Dirac*\n 1931 their existence would have interesting theoretical consequences. Especially the magnetic charge will necessarily be quantized due to quantum mechanical effects. Current ideas about the fundamental forces and the origin of the universe also strongly suugest that magnetic monopoles were in fact created in the early history of the universe. Most of these monopoles would have annihilated each other again ( in the same way as matter and anti-matter annihilate each other ), but a small fraction may have survived, allthough they will be extremely difficult to detect mainly due to their large mass. Even if magnetic monopoles do not exist the underlying mathematical model is very interesting and it has had a great impact on our understanding of gauge theories. Let us consider a pOint particle which serves as a source for the electromagnetic field. We expect a singularity at the position of the particle (like the singularity of the Coulomb field). Therefore we cut out the trajectory from Minkowski space, i.e. if the point particle then the basic manifold is L Furthermore F satisfies Maxwell's equations (7.88)

r is the worldline of

= M'

f. We know then

that the electromagnetic field is represented by a smooth 2-form F on L.

iF

Consider now the magnetic flux through a closed surface surrounding the point particle. To fix notatiOn let S be an inertial frame and consider a three dimensional regular domain magnetic flux ~ through the boundary

contained in a space

slice relative to S, cf. fig.170. We have then preciously shown that the

an

is given by

fanF

(cf.

(8.51)

).

*) ~antized singularities in the electromagnetic ~ield, Proc. Roy. Soc. A133,(6o).

470
Space time diagram - One dimension suppressed.

II

XO,~~
I / ___~.
x2

We want to show now that, as a consequence of the Maxwell equations, the magnetic flux has the following three basic propert:ies:

1} It is the same for aZZ ctosed suzofaces SUI'rOunding the point particle. As a consequence it can be interpreted as the magnetic charge of the pointparticle (in the same way as the electric flUX throough a c Zosed suzoface
Fig. 170

~(xl
2}

reproesents the electric charge within the suzoface).

It is independent of time, i. e. the magnetic charge is consel'Ved. 3} It is independent of the obserovero, i.e. the magnetic charge is a Lorentz scalar (oro to be proecise: a pseudo scalar, since the flUX depends upon the orientation).
Space time diagram - One dimension suppressed

Fig. I7la

F:ig. l71b

To show property 1 we consider two closed surfaces Q 1 and Q2 contained in the same space slice, cf. faces will be denoted
fig.17~The

region between the two sur-

W. Notite that its boundary, aw, consists of the

two surfaces QI and Q2, but that W induces opposite orientations on Q 1 and Q2. From Stokes' theorem we now get:
o

= fW dF = faw F =

fQ2F -

fnlF

:i.e.

~I

= ~2

By repeating the same argument we can similarly show property 2 and 3. (This requires the consideration of two closed surfaces contained :in A point particle can of course in principle carry both an electric charge q and a magnetic charge g. If the part:icle carr:ies both electric and magnetic charge :it is called a dyon. diffrent space slices as shown on fig.17Ib).

471

II

We want now to enlarge our considerations and consider a system where we have included a smooth distri.I:AItion of magnetic charges and currents. We kIx:1N that the electric charges and currents may be comprised in a smooth I-form which acts as a source for the electromagnetic field through the Maxwell equation, sequences: a. Taking the exterior derivative at both sides we get, d*J charge. b. If
~

d*F =-*J.

This equation has two important con-

0, which

is the continuity equation, i.e. the conservation of electric is a three-dimensional volume contained in a space slice

then the electric flux through

is equal to the eZectric charge


d*F =

contained in

~:

Q
O

-f n

*J =

In

fd~ *F
K where

With this in mind we now introduce a magnetic four-current k is the magnetic charge density and

(k ,k ,k

the magnetic

current. The magnetic four-current is going to act as a source for the electromagnetic field. Consequently we must modify the ordinary Maxwell equation, dF

= o.

Now, if we insist that the modified equation

should guarantee the conservation of magnetic charge, and the identity of the magnetic charge contained in a volume with the magnetic flux through its surface, then we are forced to give it the following form: (9.1) dF=-*K In this form the symmetric role played by the electric and magnetic charges is displayed very clearly.

Exercise 9.1.1
Problem: Prove the following equivalences
_1_ 0a (Fg~aB)

(9.2
(9.3

r-g

-k B iff jB iff

daFsy+dsFya+dyFaB

FgEaByokO

_1_ d (FgF aB )

r-g

daFBy+dBFya+dyFaB

;.:q

.0 -gEaByoJ

- -

--

According to exercise 9.1.1 we can translate ( 9.1) valent covariant formula: 9.4
1 d (r-g~ll\!) Fgll

into the equi-

To summarize, we have established the following scheme for the extended Maxwell equations involving both electric and magnetic charges:

472 Conventional vector analysis Vs


( 9.5
)

II

Covariant expressions aaFSy+aSFya+ayFaS = _ FgE aSyo


KO

Geometric formula

= Pm

dF =-*K

as + VxE = -k at

...

r
L

_l_a (Fg;aS) = S K ] r-ga

[6F

=-K]

VE = ~ E P
0

* aaFsy+aSFya+ayFaS =
JO aSyo [_l_a (FgF aS ) = Ja] FgS

( 9.6

_ r-gE

d*F =-*J

aE at

C2~xS =-~ J EO

-t-

[6 F =

J]

Using singular differential forms we can also associate electric and magnetic currents with point particles. These currents are then represented by singular I-forms

and

defined throughout the


F

whole of Minkowski space. Similarly the electromagnetic field is represented by a singular differential form Maxwell equations satisfying the modified

- dF = *K - d*F *J on the whole of Minkowski space. The components of the singular currents are comprised in the following scheme:

Conventional vector analysis


p
-t-

Covariant expression
1 o"(x-x(TdT dT --= q JdX ll

( 9.7

= =

qo3(~-i(t) q~o3 (~-i(t


JIl

r-g

PM
( 9.8
)
+

... ... = gM o 3 (x-x(t


+
3
-7

k = gMvo (x-x(t

KIl

1 (f[ --o"(x-x(TdT = gM JdX

Il

r-g

Within the framework of distributions the currents are simply given by


( 9.9

= qr

and

where

is the world line of the particle.

Next we consider the dynamics of charged point particles. They interact with the electromagnetic field and thus experience forces. To

473 nates in what follows. We then already know that an

II

simplify calculations we will restrict ourselves to inertial coordicharged particle experiences the Lorentz force so that its equation of motion is given by
(1. 26)

But if the particle is magnetically charged too it should experience a force analogous to the Lorentz force. It is this force we are going to determine. Although you might guess it correctly using a symmetry argument (see the illustrative example at the end of this section) we will derive it from a more general argument. The interaction between

the magnetically charged particles and the field should be constructed in such a way that the energy and momentum of the total system, i.e. fields and particles, are conserved.
1. 6)

(Compare the discussion in section N par-

Therefore we consider a system consisting of a field and ticles. The qn n'th and the magnetic charge gn

particle is supposed to carry the electric charge (so we admit the possibility of

dyons). We have previously determined the energymomentum tensor of this system (section 1.6) given by
(1.36)
N L

The contribution from the particles is dx S n Jp n a(T)----d o'(x-x (TdT T n

n=l

and the contribution from the field is given by


(1.41)

We must choose the interaction in such a way that


3 TaB

-3 TaB

Okay! Let us get started. As in section 1.6 we get

N
(1.39)

n=l

Jo'(X-X (T~
n
1"

dP a

de

Then we look at the contribution from the field:

(*)
Fa JY + (3 Fa )F YS +
Y

where we exchanged tions.

3 F YB

with

JY

according to the Maxwell equa-

Worked e~ercise 9.1.2 Problem: Show that the generalized Maxwell equations (3 Fa )F YB + ln aB 3 [FyoF 1 B Y __ ~ _ J _ _ yo

474

II

According to exercise 9.1.2 we can rearrange the expression Fa JB_Fa KB B F B B We then insert the expressions 9.7,9.8) netic currents and finally arrive at
_ d TaB

(*)

as

for the electric and mag-

N [ dx B TaB = L q Fa _n_ B F n=l n B dT Comparing this with dBT aB p you see that energy-momentum is conserved, provided the following identity holds dpa dxB n Fa n (9.10) dT qn BdT

-a

Thus magnetic charges gives rise to a force very similar to the Lorentz force except that we have interchanged the role of electric and magnetic fields. We may summarize this in the following scheme:

(9.11 )

(9.12 )

. ..
F

Conventional vector analysis

Forces arising from the interaction between charged particles and the fields Lorentz invariant equations of motion

F = q(E+~xB) = g(B-~xE)

m--crr

B d 2 x a = FaB~ q dT

d 2xa = *a dx B -gF BdT ~

We can also write down the equations of motion

(9.11,9.12)

in co-

variant form. As we have seen in section 6 . 7, th. 4 this requires the introduction of Christoffel fields corresponding to the fictitious forces present in a non-inertial frame. Thus we are led to the following covariant equation of motion dx\.l dx v d 2 x tl a *a a dx B (9.13 ) (qF B- gF B) dT - mr m~ dT dT
\.IV

(Compare this with

(6.46)

ILLUSTRATIVE EXAMPLE: CHARGE ROTATIONS Having completed the discussion of magnetic charges you should observe a curious fact. The equations of motion for the fields and particles exhibit a special kind of symmetry: They are invariant under the combined transformations:
(9.14 )
(9.15 )
Cosa [ Sina Cosa [ Sina

-Sina
Cosa

-Sina
Cbsa

475

II

(Observe that *F' is really the dual of F' Observe also that the transformation rule for the charges is a consequence of the transformation rule for the field, when we identify charge with flux!) To check this symmetry we first rearrange the equations of motions for the particles:

You can now easily check the invariance by broute force. Then we observe that because the world lines of the particles are unaffected by the transformation the currents will transform in the same way as the charges:

] -+[

~:

]
= _

[~~::
* [K]

-Sino. Coso.

][ ~ ]
J

When we now rearrange the Maxwell equations on the form


J we immediately see that this is invariant too because F and [ K] transform [ *F with the same matrix. J This symmetry is usually referred to as charge symmetry. Decomposing t~e tensor F we obtain the following charge transformation of the field strengths E and
(9.16)

d[ *F F ] = f-*K ] L~J

[: ]
E

[
(E + l:l)
+2 ;t2

C~so.
S~no.

-Sino.] [ Coso. E

!]

Observe that the energy and momentum of the system are unaffected too by charge transformations. It is enough to check this for the electromagnetic field, but here it is obvious as the energy and moaentum densities are given by
( 1. 4) 2

E.o "2

and

-+

= EoE )(

-+

B+

The meaning of charge symmetry is presently not well understood, but it has one amusing consequence which is in fact responsible for its name. On the classical level it is completely impossible to distinguish between electric and magnetic charge. What you call electric and what you call magnetic is only a matter of convention. You can redefine all charges by performing a charge transformation. If you e.g. choose 0. =:! you find:
2

qn
versa ..

g'n

gn

i.e., what was before called electric charge is now called magnetic charge and vice

Exer-aise 9.1.3
PrOblem:
(9.17)

Introduce the complex valued differential forms:


F
J

F +

iF

(Complex field strength) (Complex current) (Complex charge)

K + iJ

Q
(9.18)

g +iq

a) ShOW that the complex field strength is anti-self-dual, i.e. *F = -iF

b) Show that the equations of motion can be rearranged as B d 2 X o. (9.19) dF =-*J Re(QiFo. B)dd~ ~
(9.20)

c) Show that the charge symmetry transformations reduce F + F' = eio. F J -+- J' = eio. J

to

Q + Q'

476

II

Show also that the Lagrangian of the free electromagnetic field is not invariant under charge rotations. (This is another example of a symmetry of the equations of motion which is not present at the Lagrangian level. Cf. the discussion in section 2.2)

9.2 THE DIRAC STRING


When we want to include magnetic monopoles in the Lagrangian formalism we immediately run into difficulties. The electromagnetic field strength

is no longer derived from a global gauge potential

A,

i.e. it is no longer exact. If we insist on introducing gauge potentials it can only be done at the expence of singularities in the gauge potentials(i.e. Dirac strings, cf. the discussion in section 7.8) which were not present in the electromagnetic field strength. Somehow we must learn how to handle these singularities. ILLUSTRATIVE EXAMPLE: THE DIRAC STRING AS A PHYSICAL STRING
Consider once more the pure monopole field

It can be derived from the gauge potential (cf.(7.91)):


~

r(r-z)

-K(Cos8 + l)9tp

with

o
However, A is singular at the origin (r=O) and on the positive z-axis (r=z). The singularity at the origin reflects the singularity in the pure monopole field. The singularity at the positive z-axis constitutes the string, which was not present in the pure monopole field. We will now investigate this singularity and propose a PhYsical interpretation. To be able to control the singularities we regularize the gauge potential
~

R(R-z)

with

477
Observe that is smooth throughout the whole of space. Bowever it has a tendency to "peak" at the origin and on the string. The corresponding electromagnetic field is given by

II

They satisfY Maxwell's e~uations, provided we introduce a current J ~XB = 1 7 C2 J o We may consider this to be an (extremely idealized!) model of a semi-infinite solenoid generating the magnetic field! In the limit +0 the semi-infinite solenoid becomes infinitely thin and we Fig. 172 have a string of concentrated magnetic flux, which is spread out at the origin as the monopole field. Thus we have modified the originally pure monoPole+fi:ld by superimposing the string with the magnetic flux. The singularity of the A-fleld at the position of the string now reflects the singularity of the magnetic field in the modified monopole field! Observe, that if we compute the magnetic flux through a closed surface, we now get zero, because the flux of the monopole field is exactly compensated by the singular flux through the string.

Worked exercise 9.2.1


Problem: (a) Show that the regularized magnetic field is given by

B - ~ [; _ 2 (2R-z) kl - 4n RT R3(R-z)2 J
(b) Show that the modified monopole field is given by
~

Fig. 173
A

B = +0 lim B = 4":& n r

~a(x)a(y)8(z)k

where 8(z) is the Heaviside stepfunction and along the z-axis.

is the unit-vector

The associated current j which is responsible for the concentrated magnetic flux along the string is extremely singular:
a (x) a ' (y)

:23
o

~xB =

-gM
[

-a' (x) a (y)

where a' (x) is the derivative of the a-function. The magnetic field associated with the string has some nice properties. It corresponds to a n8ga~ive magnetic charge -9M at the origin, since implies

BSTRING

= -9MO(X)6(y)0(Z)k

~.BSTRING = ~a(x)6(y)d~~Z) = -gMa (x)6(y)6(z)

This cancels exactly the positive magnetic charge

9M

associated with the pure mono-

478

II

pole field! Thus when we consider the modified monopole field (9.23) we see that the magnetic monopole has disappeared.

Consider the electromagnetic field properties

produced by a magnetic mono-

pole and an electrically charged particle. This field has the following

1)
(9.24)

2)

- dF = * K fnF = gM

(where (where

K is the magnetic current) gM is the magnetic charge)

where

is any sphere surrounding the monopole. Both of these proF

perties make it impossible to generate

from a global gauge potential

A.

Dirac cured these diseases in much the same manner as in the pre-

ceding example: First he chosesa string extending from the monopole to infinity. This string sweeps out a two-dimensional sheet
L

Worldline

in spacetime. The string is

chosen completely arbitrarily, except that it is neVer aZ}owed to cross the


trajectory of the charged particZe.

This is known as Dirac's veto and it will play a crucial role later on! To cure the diseases of F mentioned above Dirac now introduces a singular electromagnetic field, which onZy lives on the string. From a physical point

the electrically charged _____-------,particle

____---,of

rg
Fig. 174

of view Dirac's veto is then clear. If the string crosses the trajectory of a charged particle, the particle would be strongly influenced by the singular electromagnetic field associated to the string. But the string is a fictitious object, which must not disturb the electric charged particles. Now, although the sheet
L

is not compact and consequently not a

regular domain in the strict:sense, it can still be decomposed in an interior region and a boundary. The boundary is simply the trajectory of the magnetic monopole. The interior region is a two-dimensional submanifold, submanifold, int L ,
a~

and the boundary is a closed one-dimensional (xO,X i ,X 2 ,X 3 ) on Minkowski space. Then we a on int E,

rM

Fix a coordinate system

may consider the coordinates as scalar functions


p~xa(P)

Furthermore we may consider the

a-function,

-.L a~ (x-x(P

r-g

479
as a scalar function on int~ Using this we can construct the following two-form on
int~

II

Space time diagram

int~

Being a two-form we can then integrate it over the two-dimensional domain


int~

[XO (P) , J
)(

and obtain the quantity *as (x) = -gM f l i t (x-x(pdx ~ ~a r-S S 1:-a "aX

(9.25)

r-g

The integral obviously depends on the point x in the a-function and on the indices a and

h
x
Fig. 175

S.

We have written Sas(x) , because it forms the contravariant

the quantity in the form

components of a tensor. This is shown in exactly the same way as for the singular current (8.59) of section 8.8. The integral -gMf ~1 ..
~ ~a ~B

r-g

<5 (x-x (P) )

dx "dx

is a geometrical quantity, independent of the coordinate system we choose on parameter integral as


int~

Usually we parametrize the sheet by a space-like where

Al ,

-OO<AI<O,

and a timelike parameter

A2 ,

Using such a parametrization we can rearrange the

(9.26)
with

sa (x)

* B

-gMf u-- <I (x-x (,\ , A

r-g

It

[axa ax B- aP" axa ""fiT axB] cIA ""fiT aP"

dA

U {(A I ,A 2 ) I AI<O}. We will use such an explicit representation of the integral whenever it is preferable. Clearly the tensor field with components SaB(X) is a skewsymmetric tensor field. Consequently SaB(x) are the contravariant components of a 2-form. We have written this 2-form as *5, i.e. as the
dual of another 2-form 5 . As -5 = *(*5) terized by the covariant components: this 2-form 5 is charac-

(9.27)

aP"UI\UI\

ax ..,

~o

I .. ,2

Remark: Within the framework of distributions the sheet ~ itself can be considered a singular 2-form. This 2-form is closely associated with 5. If U denotes an arbitrary testform we get through a formal computation:

480

II

<*slu>

~fR,;aS(X)Ua8(X);=gd'X
-HR,gMf uo (x-x(A ,A
, 1 2

[axa ax8 axa ax 8] ~i>:TW-WaIT Ua8 (x)dA

dA d x

2,

Within the framework of distributions we therefore have

(9.28)
We summarize the main properties of the singular form following lemma Lemma 1 (Dirac's lemma) The singular form 1)
(9.29 )
(9.30)

in the

has the following properties:

It vanishes outside the string.


dS

2)
3)

*K

fns

=-gM

for any cZosed surface

surrounding the mono-

pole. Proof:

(1) (2)

If

lies outside the sheet


(9.2~

L,
u~e

then the

o-function

o'(x-X(A 1 ,A 2

automatically vanish. we that it is equivalent to

To check the relation


_1_

(Fg a8)

;=ga

= KS

(chain-rule)

(Stoke's theorem) = (3) Finally we must


CCIlIpUte

th= flux through a

closed surface sur-

rounding the monopole.

But a closed surface

surrounding

the monopole is the boundary of a regular domain

W con-

481 taining the monopole. Consequently we get using (9.29) Ins =

II

laws

Iw ds

-lw*K

=-gM

Worked Problem:

e~ercise

9.2.2 Prove (9.30)by an explicit computation of the integral.

Remark:

We can also reformulate (9.29 ) as

5*S = - K Within the framework of distributions the boundary operation coincides with the codifferential and we therefore get &*S = -g~l:

= -gMal: = ~rM

= - K

Using Dirac I s lenuna we can now "cure the diseases" of therefore had the properties (9.24) dF=-*K InF=gM

F.

It was

generated by a magnetic monopole and an electrically charged particle and

We now choose an arbitrary string extending from the monopole to infinity. Associated with this string we have a weak form properties: (9.29,30) S with the

dS

= *K
F + S is exact, although singular, and we which generate

Consequently we see that The gauge potential A

therefore can find a global gauge potential A,

F + S
F ,

will, of course, be singular too. It will be singu-

lar at the position of the monopole, reflecting the singularity of and it will be singular at the string, reflecting the singularity of S Formally S represents a concentrated magnetic flux flowing towards the monopole. Hence we may formally interpret F + S as a concentrated magnetic flux flowing towards the monopole position along the string and then spreading out to produce the monopole field. However, it should be emphasized that S has no physical meaning. The position of the string can be chosen completely arbitrarily and the introduction of S is a purely formal F! trick, which cures the diseases of

Fig. 176

482

II

9.3 DIRAC'S LAGRANGIAN PRINCIPLE FOR MAGNETIC MONOPOLES


We are now in a position where we can state the Lagrangian principle of

Dirac~)Dirac's
S

action consists of three pieces:

(9.32)

= SpARTICLES

+ SINTERACTION + SpIELD

where SpARTICLES

SINTERACTION

Jr

A dxpa dA

a dA

SpIELD

Notice that the interaction term only contains a coupling between the electrically charged particle and the field! However, if we look a little closer at
(9.33)

SpIELD
P

we see that it contains information about the


-aA-S v ~ ~v

monopole trajectory because


~v

=aA
~

Hence when you vary the trajectory of the monopole, you will have to vary the sheet which terminates on the trajectory. But that will force
S~v to vary, and thus SpIELD contains the coupling between the monopole and the field. That the above action in fact gives the expected

equations is the contents of the following famous theorem:


Theorem 1 (Dirac's theorem) All equations of motion for a system consisting of monopoles, electrically charged particles and the electromagnetic field can be derived from Dirac's action, provided you respect Dirac's veto, i.e. netic strings are never cally charged particle.
Proof: The proof is long and technically complicated and you may skip it in a first reading. First we list the equations of motion which we are going to derive d2 x a dx ~ dx v a dx S e + ra e e ] _ qP 6--e(1) ( 3) me [ --;rrzdF =-* K ~ v liT dT
allo~ed

the mag-

to cross the worldline of an electri-

---err -

(2)

d*F =-* J

(4)

Then we list the dynamical variables to be varied:

*) The theory of magnetic poles, Phys. Rev. 74 (1948) 817

483
(a)
(b)

II

(c)

Xea(A) xma CA) All (x)

Trajectory of electrically charged particle. Trajectory of monopole. Gauge potential.

Observe, that the string coordinates X(A I ,A 2 ) are not considered as dynamical variables. They are fictitious coordinates and shOuld be completely eliminated in the end of the calculations. Okay! Let us go to work: Eq. (1) is not a dynamical equation. It is a purely kinematical equation which is built into the model from the beginning: F + S = dA but by construction we have:

dF + dS = dS = *K

dF

-dS

Eq. (2). Performing the variation

dS I dEIE=o

A A + EB Il Il Il fni=gJa(X)Ba(x)d'x
~

we get:

which leads to the desired equation of motion:

Eq. (3) follows from the variation of the trajectory Il(x) . Performing the variation XIl(A) ~ XIl(A) + E~(A) we get from a previous Ealculation (exercise 8.6.3) :

dS I dEl E=O

-a) dA ([aAa_S]dXCX + A 3L. BY dA adA ax aAa dx B a dxa ~S - 8 (IT' Y - ax8 (IT' ya]dA q J[aA ax dAS)dXCX -B - - - - - Y dA qJC) dXa dA q Al

From which we conclude: v 8 dx ll dx \ d2 B me g a8 ( ~ + r Ilv<fr <fr)

484

II

sO we got an extra term. This is the first time We use Dirac's veto! According to Dirac's veto the sheet will never cross the trajectory of the electrically charged particle. Thus S B vanishes on the trajectory of the electric charged particle and we simply throw i% away!

Eq. (4) follows from the variation of the monopole trajectory and it is by far the most complicated step. ' Performing the variation! ~ll (A) ~ ~ll (A) + ~ (A) we get: d2X B dS p m d!=O =-~gaB~+

So this caused no problems! But then we must investigate To do that we must proceed carefUlly. First we observe that when we vary the monopole trajectory, we must vary the sheet too:
~(Al,A2) + ~(Al,A2)+YV(Al,A2)

Final

The variation ~(Al,Al) should satisfY the boundary conditions:


O=YV(A 1 ,A 2 )

on the initial and final space-slice

i.e. the deformed sheet terminates on the deformed monopole tra177 jectory! but apart from these boundary conditions we can choose ~(Al,A2) arbitrarily. It will also be convenient to rearrange the expression for SFIELD! We have previously found (cf. (7.56)):

YV(O,A 2 ) =~(A2)

So we may rearrange the field action as follows

the calculation is simplified to the computation of F d!=O where dS

'" d [ J FllVd!=o gm a J
61+6 2+6 3

(x-x-y)

~ (d (Xll+yll)
dAr

485

II

63 =

gmJF~vJO"(X-X)~ ~
61

d;l.ldA 2d"x

In the first term

we rearrange the integral

Consequently we end up with

( *)

61 = 9

---.l!..':'. (x) m axa


62 :

*v aF

ax~ axv "5:T ~ y<Xd;l. IdA2


a
A

We then attack
62

= 9mJ[J;~v(X)O" (X-X)d"XJ~ ~ = 9mJ F ~V (x)

dA dA

~~aXV

arr a57 dA 12 dA

Using that

we may rearrange this as follows (observe that we cannot use partial integration due to the term
(fXJ ') all.

In exactly the same way we can rearrange the

( ***)

a * ~ ax~ 63 = gm aTZ"(F~vy )tfrdA 1 d;l.2-

Consequently 61,62 and 63 have been split into two groups of similar terms. Three of the terms obviously match together. Performing the following index substitutions: 2. term ( **) a + ~ v + v ; ~ + a 3. term ( ***) ~ + ~ a + v ; v + a these may be rearranged as

* * * a~~ x 9mf (a a F~v +a ~ Fva +a v Fa~ ) arr

,~v~ ~T dA 1 d;l.2 a;l.~

but according to the Maxwell equation already established in (2) we knOW that: * * * ~ B aaF~v+a~Fva+avFa~ =-v-gea~vBJ But vanish~s on the string, due to Dirac's vetd Thus the three terms cancel automatically, and we are left with two remaining terms:

dele=o

dS F

= gm arr(F~vy )~;I. dA

Ja

* ~~

ax\!

a * ~v ax~ 1 2 + gm aTZ"(F~vy )~A dA

486

II

because the boundary of the sheet is nothing but the monopole trajectory. Believe it or not: We are through! dSp. . f dSp Combining the above expression for a:;:;r-- Wl th the preVlous one or a:;:;r-~IE~ ~IE~ we get

From which we deduce

d 2 iS mmgaS [ ~ +

r ~v-~ (f[" = -gmPaS(f[

di~ diVl

diS

This concludes the famous proof of Dirac's theorem.

9.4

THE ANGULAR MOMENTUM DUE TO AMONOPOLE FIELD


Consider a system consisting of a pure monopole and a pure electric

charge and suppose that they are at rest at some time


...

...

t=O

E~g

--(9.11)

.... ,
g
x

"q -.-~
...
E

-178

...
B

Fig.

Then they will stay at rest forever! The electrically charged particle will experience the Lorentz force: where and

...... E,B

q(E + V

B)

are the field strengths produced by the monopole. But because the particle is at rest. Similarly the monopole

E=O

...

;;xB=O

will experience the force:


(9.12)

487

II

which vanishes for the same reason. Next we observe that the chargemonopole pair create an electromagnetic field with a non-vanishing momentum density
(1. 42)

In fact,

Ex B

is circulating around the axis connecting the mono-

pole and the electric charged particle. For a monopole at rest at the origin the momentum density is responsible for an angular momentum density
-t-

-+-+

rxg

and consequently the electromagnetic field created by the charge monopole pair carries a total angular momentum given by

J =

Jjd3~

Jrxgd3~

(Bypassing we observe that the angular momentum is independent of the choice of the origin as long as it is placed on the axis connecting the monopole-electric charge pair. This is due to the fact that perpendicular to the axis). If we introduce a Cartesian coordinate system with the monopole at the origin and the z-axis pointing towards the electrically charged particle, then a computation, which we leave as an exercise, shows that the components of this angular momentum are given by
(9.33 )
-+

is

J3

= 41T

-gMq

Worked exeraise 9.4.1


Problem: Compute the angular momentum of the electromagnetic field created by a charge-monopole pair.

So the electromagnetic field does carry a finite angular momentum along the z-axis and the size of the angular momentum is independent of the separation between the two particles. This is the first hint that quantum mechanically the magnetic charge charge q gM and the electric are not independent of each other. In fact, we expect some-

thing like the following naive quantum-mechanical argument to be valid: The size of the angular momentum along the z-axis should be quantized as
~

, where

is an integer.

(At this point we cannot ex-

clude half-integer spin, because the angular momentum carried by the field is not of the conventional orbital angular momentum type). Thus we expect
(9.34)

integer

(Dirac's quantization rule)

But this is very beautiful, because this could provide us with an explanation of why electric charges are quantized in multiples of electron

488

II

charges. If there exists just one monopole, then, continuing our very speculative line of argument, every electrically charged particle would interact with this monopole. Due to the quantization of the angular momentum created by the charge-monopole pair the electric charge must therefore be an integer multiple of 2nft
gM

The above considerations can be extended to dyons too. Since the angular momentum is invariant under charge rotations, we can rotate one of the dyon-charges into a pure monopole charge (cf. the illustrative example in section 9.' ): (g, ,'1,) ... (gi ,'1i) = (~,O) (g2'~) ... (g2,.2) It follows that the angular mome~tum.dens~ty comes from the magnetic field created by the first dyon and the electrlc fleld E2 created by the second dyon. As above we therefore get

S,

where rO is the unit-vector pointing from the first to the second dyon. In terms of the original variables the angular momentum is given by j = g,'12 -g2 '1, ~

1\

4n

and we therefore expect the combination (g,'12-g 2 '1,) to be '1uantized.

Now we want to investigate a little closer the angular momentum carried by the electromagnetic field. For this purpose we consider a static spherically symmetric monopole field. You may think of it as being created by an "infinitely" heavy monopole placed at the origin. We then want to investigate the motion of an electrically charged particle in the monopole field. We shall treat the situation in the non-relativistic approximation because then we can later on quantize the motion of the electric charged particle using the Scrodinger equation (Cf. the discussion in ch. 2). Newton's equation of motion is given by
~

d2

= qvx = IGT

...

qgM ~

r vXro

Introducing the parameter, I< = 4 gM , which eventually will become the 4n angular momentum of the electromagnetic field, we rearrange this as (9.35) We proceed to analyse the consequences of eq. (9.35) in 5 steps.

Step 1: In the first step we look for constants of motion. We multiply both sides of eq. (9.35) with ~, thus obtaining:

i.e.

i.e., the kinetic energy is conserved. This is because the monopole

489 field performs no work on the particle, since the force


stants of motion:

II

is always

perpendicular to the velocity. Thus we have found the following aon-

( 9.36)
~:

v (= speed) and

T = ~mv2(= the kinetic energy) (9.35) with

Then we multiply both sides of eq.

r,

where-

by we obtain

But now we can use that d

~-~

2(r2) _ d 2 r2

Integrating this we get


(9.37)

r2 =v 2 t>+at+13 t=o the particle is as close as

Let us fix the time-scale so that at possible at the origin, i.e.

r(O) = d(= distance of closest approach) Then eq. (9.38) This has an important consequence. If the speed
bound states.
(9.37) reduces to

vio , then the par-

ticle comes from infinity and returns to infinity. Thus there are no [The case v=O where the electric charge stay at rest is a very special situation. In fact, it is unstable: The slightest perturbation and the electric charged particle will move to infinity! It is therefore irrelevant when we investigate the scattering of charged particles in monopole fields.]
~:

Then we consider the orbital angular momentum

t =

rxmv

Since the monopole field is spherically symmetric we expect that there should be a "total angular momentum" which is conserved. First we observe that the orbital angular momentum is not conserved: dt(rxmv) = vxmv+rxrnat = KrX(vxr') = K where r =

d ......

. . ...... dv

...... r

rv-(;.v)r r2

r r
d
A

is the unit vector pointing towards the charged particle.

To rearrange the right hand side we observe that

d~ = dt ~) = ----r~~

d ( ... ,

rv-r -dt

... ... dr

so that we finally obtain:

490

II

d ~~) 'dt(rXmv

dr Kdt

But then we have found the following constant of motion:


(9.39)

rxmV-Kr

...

This means that the total angular momentum of the system must be identified with

j . The second term,-K~, is, of course, nothing but the

angular momentum of the electromagnetic field. Step 4: nents


(9.40)

Since we have decomposed and Kr we immediately get


J

into two orthogonal compo-

t
K

But

is a constant, and

is conserved. Consequently

is con-

served too. So, although the orbital angular momentum quence: Consider the system at t(-oo) t=-oo and and

is not con-

served, we find that its size is conserved! This has a curious conset=O , then we get t(O) mv(O)d

= mv(-oo)b

Particle Distance of closest approach Impact parameter (distance of closest approach for a free particle)

Tangent at infinity

Fig. 179

where

is the impaat parameter. But


b

and

are conserved. Thus

we finally obtain:
(9.41)

Observe, that if we point directly towards the monopole, i.e.

b=O,
IlOve

we will hit it! But this is a very exceptional situation: If we point directly towards the monopole, the electriCally charged particle Will freely because
...

vxB

But the slightest perturbation and the particle will react to the force from the monopole field and no longer hit the monopole. Consequently this is an unstable situation and we shall neglect it.

491
~:

II

Finally we observe that multiplying both sides of the equar we get i.e. Cos(J,r)
+
~

tion (9.39) with

(9.42)
But and J is conserved and
K

=J

-K
~

is a constant. Thus the angle between

is conserved and therefore the partiale is aonstrained to mave on a aone with j as axis:

Worked exeraise 9.4.2 da Problem: Compute the differential scattering cross section dn for the scattering of electrically charged particles in a monopole fielu.

9.5

QUANTIZATION OF THE ANGULAR MOMENTUM


As we have verified, the magnetic strength gM of the monopole en-

ters into the expression (9.39) for the angular momentum of a charged particle in a monopole field. When we quantize the motion of the charged particle we know that the angular momentum becomes quantized, and we expect this to liminate the possible values of the possible values of of angular momentum:
K A A A
K

To determine

we must therefore determine the operators

J l ,J2 ,J 3 . For a charged particle moving in a magnetic field we have previous(2.65)):

ly determined the Hamiltonian (Cf. eq.

(9.43
where

H=
A

JL(-i~V-qA)2
2m
~gauge ~tliU

is the gauge potential generating the magnetic field. ConA


~

sequently we must first find

producing the monopole

492

field. But here we run into the wellknown trouble that no globally defined gauge potential can produce l:he monopole field, i.e. A will necessarily contain singularities. However, as we have seen in section 9.2, we may concentrate the singularities on the z-axis. We can still make a choice: The singularities form a string which can be either infinite or semi-infinite. As we shall see, the actual choice of the string has consequences on the quantum mechanical level and we shall therefore work out both cases. The semi-infinite string was originally introduced by Dirac, and we shall speak about the Diraa formalism. The infinite string has been especially advocated by Schwinger and we shall speak about the Sahwinger formalism. Let us collect the appropriate formulas in the following table:
(9.44) Sahwinger formalism: (9.45) Diraa formalism:

'-/

- ~cose,,1P
41T

zy _ --Y--_r(r:z)
= K

qA

r(r-z)

Observe that the Hamiltonian (9.43) becomes a singular operator, but there is nothing to do about this for the moment. We will have to live, for some time with the singularities along the z-axis and we will have accept similar singularities in the wave function w(r,t). Next we construct the operator corresponding to angular momentum. On the classical level we know that ~ ~ ~ 7 j = ~ rxmv - Kr =; r x (p-qA) Kr (9.46) This corresponds to the operator (9.41) ~ r x (-ifiV-qA) - Kr = -ifirxV - ( rxqA + Kr)

*0

Consequently we have found the following candidates for the angular momentum operators:

493

II

(9.48)
1\

Schwinger formalism:

(9.49)

Dirac formalism:
K-L

J
1\

l =-ifl(y,}Z
2

- Z~) ay -

KX 2 +y2 Kxl:y2

xr

31 =-ifl(Y~ ay az - z~) 32
=-in( z-1ax

r-z

J
1\

- x~) =-ifl(Z~ ax dz

- x~)
dZ

- K~

r-z

3 =-ifl(

x"~ - y"~)

x~ ~3 =-ifl( , "y - Ya"x)

+ K

Before we continue the discussion we recall some general aspects of the angular momentum. According to Dirac the components of the angular momentum are always represented by three Hermitian operators ~1'~2'~3 satisfying Dirac's commutation relation:
(9.50)

If the system is spherically symmetric they must furthermore commute with the Hamiltonian, i.e.
(9.51)

[ft;~i]

This guarantees the conservation of the angular momentum, since according to the quantum mechanical analogue of (2.61) we have ifl dt

d~i

[ft;~.]
~

= 0

In many problems with spherical symmetry we can use the operators of orbital angular momentum
(9.52)

which trivially satisfy Dirac's commutation rules (9.50). But we cannot use them as angular momentum operators in this particular problem because they do not commute with the Hamiltonian (9.43). On the contrary the above operators (9.48-49) not only satisfies Dirac's commutation rules (9.50), but they commute with the Hamiltonian (9.43) as well. The verification of this is left as an exercise:
Exeraise 9.5.1
Problem: (a) Let

f(X)~ be a differential operator. Show that:


[f(X)a"x;g(x)]

= f(X)X

i.e. the commutator is a multiplication operator. (Hint: Let both sides operate on a testfunction ~(x)). (b) Show that the operators of angular momentum 9.48-49) satisfy the commutation rules with (c) Show that the operators of angular momentum 9.48-49) satisfy the commutation rules (9.50) and (9.51) as required by Dirac.

494

II

We pr=eed to investigate the sp:ctrum of the angular m:mentum. Using the commutation relations (9.50) one can determine the possible eigenvalues for the angular momentum operators.
A A A

(For details see e.g. Schiff [196S]).


A

First we introduce the square of the total angular momentum: (9.54) It follows that states (9.55b) (9.55b) Furthermore it m (9.56)
j

J2 = J~ + J~ + J~ J2 commutes with J2 and J


3

,J2

and

and consequently

we can diagonalize 1jJjm

simultaneously. If we label the eigenrnfI1jJ.

we get:

=
=

j (j+l)f'1.21jJ.

Jm

the possible eigenvalues are given by -j,-j+l, . ,j-l,j

Jm

O,t,1,t,2,f,

Finally it is preferable to introduce the raising and lowering operators (9.57) They satisfy the commutation rules (9.58) and as a consequence the operators J+ and J genfunctions with a fixed j. (9.59)
=

connect the different ei-

fl/j (j+l)

m(m+l) 1}!. - m(m-l) 1jJ.

= fi/j(j+l)

Jm Jm

In the particular example of a charged particle in a monopole field we can now use the explicit form (9.50-51) for the angular momentum operator to determine which of the possible eigenvalues (9.56) that are actually realized. Recall that classically (9.40) j2 = 12 + K2 This equation has a quantprn mechanical analogue We have decomposed the quantum operators in the following way:

=~

(-iflV-qA) -

K; = t - Kr

But then we can square this operator relation:

Worked ea:erais!il,. 9. 5. 2

Problem: Let Show

L b.lO. the operator


L'r and r'r,

'r, = ~x(-iI'lV-q.Jt)

that<!>t~e operat~r4Products

both vanishes.

495

II

According to exercise 9.5.2 been established:


(9.60)

both operator products

automatically vanishes. The following operator relation has therefore

~2

~2
j2
and

Here

is a C-nurnber and

r?

/I

therefore have the same

eigenfunction~:

is a positive operator, becau/lse But 1 2 3 are Hermitian operators. Thus the eigenvalues of L2 and we conclude
(9.61)

)m/l = L2+L2+L2
/I /I

r?1jI.

= [j(j+l)h 2-K 2 l1jl.

)~
1

/I

/I

L ,L

and

L
3

are positive

i.e.

j(j+l) .::

K) 2 (..:11

This is our main result because this shows that in a monopole field j=O is not allowed. Consequently the minimal value of j is non-trij and vial and the system therefore possesses an intrinsia spin! To determine the exact spectrum, i.e. the allowed values of m together with Xhe j2
eigen~unctions

Yjm , we must now use the explicit

expressions for

and

It is convenient to use spherical coor-

dinates. After a long but trivial computation you obtain the following explicit expressions:

Schwinger formalism:

~2 J
(9.62)
J
3

_~2r __l__ JL(s;ne~)


11

LSine ae

3e

Dirac 3:2
(9.63)

formalism: l'l2[ I a ( . eJ,. +_1 32-J_2iflK___ I ___ JL+2lC2 ___ 1 ___ Sine ae s~n-ae) SinTGW . I-Cose 3IP . l-Cose
K

-ifl JL +

alP

The Schwinger formalism is the most complicated one. But let us start with it to get a feeling for the general machinery. We look for eigenfunctions of the form gate the eigenvalue equation (9.55a) i.e.
J31j1. flml/J. 3)m :lm -ifl a\llY jm (0,1P) = mflY

yjm(e,IP)

First we investi-

jm

(0,\Il)

496

II

But this shows that we can factorize y, in the following way )m IP (9.64) Y (0,1P) = Pjm(COs0)e i m jm Since Yjm (0,1P) ,must be a smooth function we demand that m is an integer, so that e 1mIP is periodic with the period 2n You should observe that e imlP is still singular at the z-axis where it is disconthen eimlP=l eimlP=_l) tinuous. (If we approach the z-axis along the line ~O but if we approach the z-axis along the line IP-~ then Usually we get rid of this singularity by demanding the z-axis, because then

O=Pjm(Cos0) on Yjm itself will be smooth. In our models, however, we have singularities on the z-axis from the beginning since
the gauge potential A

...

itself is singular on the z- axis. We shall therem can only take integer values we (9.55b) . Inserting

fore neglect the problem. Because

see that onl.y bostmia states are possibl.e. Then we must use the second eigenvalue equation the expression (9.64) and introducing (9.65) 2 [ - :x [ (1-x );x ] + x=Cos0
-

we obtain: ] j(j+l) Pjm(x)

m2+2nix+ (K) 2 l-x 2

o
has

We should then look for regular normalizable solutions of this equation on the interval [-1,+11. It can be shown that eq. (9.65) m and
j
K

regular normalizable solutions if and only if through the relation (9.66)

both integers or both half-integers. Furthermore

Ii are either is constrained

' )

hl 1'1

in agreement with the previously obtained result (9.61).

Wopked exepaise 9.5.3


PrOblem: Consider the differential e~uation (9.65) on [-1,1]. Determine under what circumstances it has regular normalizable solutions and show that such solutions are given by
-Hm+!'5.)

PJ'm(x)

= N.(l-x)
Jm

~ (l+x)

-Hm-!'5.)

~ ~(l-x)
dxJ-m

j-m

(j +!'5.)

~ (l+x)

(j _f.)

~]

In the Schwinger formalism integer valued. Consequently

we know from the beginning that

is

must be integer valued tOQ' To summa-

rize we have obtained the following results:

Theopem 2 The Sahwingep fopmal.ism l.eads to a bosonia speatpum:


K ,

m')'K

ape al.l. integeps

Fupthepmope IKI is the intpinsia spin of the state. The eigenfunations ape given by the fopmul.a

497 (9.64)
where

II

Pjm(COS0)eimq)

(9.67 )

Njm(l-x)

-l:;(m+~) 11

(l+x)

-l:;(m-~) l1d j-m[


dx

j-m (I-x)

(j+~) fl

(l+x)

(J"-~l fl

Then we briefly discuss the Dirac formalism

. Here the spectrum is (9.55a) we get

more easily obtained. From the eigenvalue equation

-i ~Y" alP Jm (0,n) ,0/ = (m-~)Y" fl Jm (0,n) ,0/ But this shows that we can factorize (9.68) Since ei(m-fi)1P Yom as follows:
)

i(m--)IP
fl

Yjm (0,1P) = Pjm(COs0)e

has to be periodic we conclude that

ger valued. But from the general theory, we know that Either (j,m'K) are half-integer valued, or (j,m'K)

is inte(m-K) m is either are all integer

half-integer valued or integer valued! We have now two possibilities: valued. Thus we have the possibility of a fermionic spectrum! To check that all the listed combinations are admissible we must investigate the second eigenvalue equation (9.55b). Inserting the expression (9.68) tion (9.69) [ _ and introducing x=Cos0 we obtain the differential equa-

~[ l-x 2 ~] ax ( ) ax
exactl~

+ m +2ll1fix+.fi" I-Xi

2 K (K)2

] -j (j+l) Pjm(x) = 0

But that is

the same equation as the one we analyzed in the

were both half-integers or both integers! Thus they are all admissible and we conclude
Theorem
:3

Schwinger formalism .! (See eq. (9.65) ). But there we found that K there existed regular normalizable solutions, provided m and fi

The Diraa

formalism

has both a fermionia and a bosonia speatrum:

are all integers or all half-integers Furthermore

IKI

is the intrinsia spin of the state. The eigenfunations

are given by the formula:


(9.68)

where
(9.67 )

PJ"m(x) = N" (I-x) Jm

-l;(m+K)

(l+x)

-l:;(m-K) d j - m [ (j+K) (j-K)] --"- (I-x) (l+x) dxJ-m

498

II

9.6

THE GAUGE TRANSFORMATION AS A UNITARY TRANSFORMATION


As we have seen, the choice of the gauge potential

has consequences

on the quantum mechanical level. This may be somewhat surprising: Usually the choice of a gauge potentialis unique up to a gauge transformation and gauge transformations leave physics unchanged. When discussing monopole fields you should, however, be careful! The crucial point is that here different choices of the gauge potential need not be related through a global gauge transformation. E.g. the spherically symmetric monopole field has been represented by the two gauge potentials:

(9.44-45)
Formally we have

and

A2 =-*(COS0+l) dIP

A2 = Al - *dlll
but III is not smooth throughout space time. It makes a jump somewhere between

and

2rr

and therefore

Al

and

A2

are not related

through a global gauge transformation. On the quantum mechanical level ,things behave a little different. Here gauge transformations are represented by unitary transformations, as we will now explain: Suppose the state of a quantum mechanical system is represented by the Schrodinger wave functions
~(r,t)

and the various physical quan-

tities are represented by Hermitian operators transformed wave function

PI,P2,P3,

etc. If

is a unitary operator then the same state can be represented by the

~ '(r,t) = U~(r"t) and the various physical quantities by the transformed operators

PI' = Ul? IU- I


The transformation:
(9.70)

p;

UP 2 ij- I

is called a unitary transformation and it leaves all matrix elements invariant


<~2Ipl~l> ~ <~2Ip'l~i>
~ ~

<~2Iu UPU ul~l>

~t~t~

<~2Ipl~l>

Consequently a unitary transformation leaves physics unchanged, which justifies that


~.

= U~

represents the same state.

Consider the Hamiltonian

Jl(-inV-qA)2+q$
2m

representing a charged particle moving in an electromagnetic field. We now introduce the following unitary operator

II

(9.71)

It generates a unitary transformation which transforms the wave function in the following way (9 . 72) l.nX r,t

.g

(+

1}J(~,t) ~ l/J' (~,t) = e

l/J(r,t)

Furthermore the operator, P = -inV, representing the conjugate momentum transforms according to the rule itx(r,t) (9.73) Finally the operator ding to the rule (9.74) l. I1at' e
(-i~v)e

-i* X(~,t)

-qVX- inV = P-qVX

.'" a l.I1 at

representing the energy transforms accor. a -itx(~,t) (l.nat)e

... a

i*x(f,t) e

qat + l.Ilat'

lx

... a

Using this we see that the transformed wave function the transformed Schrodinger wave equation: ifl.;tl/J' =

l/J' (~,t)

solves

{2~[-ifl.V-q(A+VX) ]2+q(~
A~A+VX

tt) }l/J'

But this is obtained from the old Schrodinger equation provided you perform the substitutions: (9.75) and

This justifies our claim that quantum mechanically gauge transformations correspond to unitary transformations! There is however one important difference between the classical situation and the quantlw mechanical situation: Classically we have always demanded that

X(~,t)

should be

a smooth function. Quantum mechanically this is no longer relevant. All - + . gox (t: t) we should demand is that U(r,t) = el.rr ' is a smooth function. Therefore the jump of X need not be single valued, but can make jumps provided *X is proportional to

2~ ! Thus the unitary transforma-

tion is slightly more general than its classical counterpart. With this in our mind we return to our discussion of the spherically symmetric monopole field. The choice of the two gauges (9.44) (9.45) Schwinger formalism: Dirac formalism: Al = ~osedlP A2 =~(Cose+l)dlP

suggests that we look at the following "unitary operator" -l.fi:IP e This "unitary operator" is not well-defined unless it is a periodic (9.76)
U
.K

function of

~,i.e.

500 is an integer! When

II

is an integer we have

previously seen that both choices of the gauge are admissible and therefore we have shown that the Schwinger formalism and the Dirac formalism are unitary equivalent in this case. When
K h

is a half-integer we have previously seen that only the Di(9.76) which formal-

rac formalism is admissible. The unitary operator ever, that for instance U = e 2ifi~

ly transforms it into the Schwinger formalism breaks down. Observe, howis well-defined in that case. It -lr(COse+l) d~ ~ -lr(COse-l) d~ corresponds to the "gauge" transformation

i.e.

where we have moved the semi-infinite string from the positive z-axis to the negative z-axis. Thus, although we cannot convert the semi-infinite string into an infinite string, we can still move the string around, using unitary transformations. Finally you should observe that although the unitary operator (9.76) is periodic for suitable values of gular at the z-axis.
K ,

it is still singular at the z-

axis. This is inescapable, because the gauge potentials themselves are sin-

9.7

QUANTIZATION OF THE MAGNETIC CHARGE

We conclude this chapter by presenting two other arguments for the quantization of the magnetic charge:

1. argument:

On the classical level the position of the string is choYI and Y2 .

sen completely arbitrary. Consider two di'fferent strings They give rise to two different gauge potentials:

AI: (-'h, All and A2 : (-<P 2 , A2 ) , which are singular along their respective strings. Each of these gauge potentials generate a Hamiltonian,. which in the non-relativistic lirni t is represented by HI

= 2~(-inv-qAI)
~ ~l1at
,

+ q<P I

respectively

Thus we get two quantum mechanical descriptions of the

..

system:

respectively

in~ at = H2 '/' 't'

But since the position of the string is of no physical importance, the two situations must be indistinguishable, i.e., they must be unitary equivalent. The two gauge potentials are connected through a singular gauge transformation, A2-A I=dX , where X is singular along the strings. Thus the two descriptions are only unitary equivalent provided eKp[~<i~,t)]

501

II

is a single valued function outside the strings. The jump of X must therefore be quantized, i.e.
(9.77)

This jump can now be computed using the following trick: Consider a sphere

s2 which we split into two semi-spheres 2 2 s+ and S_ by the equator y, cf. fig.18l
Then we get

fYdX = fYA2 -fYAl = fYA2 f-y Al


+

Using Stokes' theorem we therefore get

S2 (Notice that A2 is smooth on the semisphere

fydX

JS! dA2

J dAI
s:

F =

s:

and similarly for AI). Thus is simply given by the

the jump of

magnetic charge! The quantization rule (9.77) can therefore be rearranged as

1r = n21T
which is precisely equivalent to (9.34)! 2. argument: While the first argument was very formal the second argument will be more physical. It is based on path-integral techniques and is closely related to the Bohm-Aharonov effect. Consider two paths with the final point the figure. Then r r B and
l

and

which connect the initial point

in our space diagram (see fig. 182). Suppose r form the boundary of a surface

furthermore that the paths pass on each side of the string as shown on
l 2

which

is hit by the string. sions). The amplitude

(Strictly speaking we should make a space-time

diagram, but it is not easy to "visualize" this situation in 4 dimenK(BIA) for a charged A to B is
Space diagram (time suppressed)
~r(BIA)

particle to propagate from

the sum of all the amplitudes specific path K(BIA) = r ,cf. (2.17).
=

for the particle to propagate along some

J~r(BIA)D[r]

Ie~

2:.S (B;A)
r D[r]

When the charged particle moves in an external electromagnetic field it picks up a change in phase given by: .!:gJ A dx a (2.33) efl. r a
A

Fig. 182

502
The contrib.ltion fran

II

r 1 am r 2

will interfere

am

this interference is given by:

(9.78)

Using the notation of section 9.2 rearranged as follows

the integral in the exponent can be

~rA

= ~dnA = fndA = fnF+S = fnF+fns


.!.sf F .!.sf S e fl n e fl n

Consequently the change in phase due to the external field is given by


(9.79)

The first term is Okay. The charged particle ought to be influenced by the monopole field, but the second term must not contribute, since otherwise we could experimentally determine the position of the string using the Bohrn-Aharonov effect! We are therefore forced to put
(9.80)

e fl.

.!.s f

n =

But fns through (9.34)

n,

is the magnetic flux concentrated in the string which passes i.e. it is equal to g. Consequently eq. (13.42) implies

It is instructive to compare the quantization of the monopole charge with the fluxquantization of a magnetic vortex in a superconductor (section 2.12). In both cases it is the magnetic flux which is limited to a discrete set of values. Furthermore the basic mechanism behind the quantization is also very similar. In both cases the magnetic field interacts with electrically charged particles represented by a Schrodinger wave function ~(t,t) . In the case of a monopole the phase of this wave function can make a quantized jump when we go once around the Dirac string, while in the case of the vortex the phase can make a quantized jump when we go once around the vortex. There is however one essential difference between the two cases: In the case of a monopole we have b'een giving very general arguments to show that the jump of the phase is identical to the magnetic flux. Since we believe that a consistent description of this world must necessarily include quantum mechanics, we therefore conclude that if we should ever succeed to find a magnetic monopole its charge will be quantized. In the case of a magnetic flux string this is no longer so! We can easily construct flux strings where the flux is not quantized. (Observe too that otherwise the Bohm-Aharanov effect would disappear) . It is of course still true that the phase of a wavefunction can only make a quantized jump when we go once around the fluxstring,but this jump need not in general be related to the magnetic flux in the string. When a mag-

503

II

netic fluxstring happens to lie in a superconductor this will have two very special consequences: a) The length of

will be constant and nonzero inside the super~

conductor. [The length represents the density of the Cooper-pairsl b) The gauge covariant derivative of essentially the Cooper-currentl vanishes. [It represents This implies the following rela-

tion between the phase ~ and the gauge potential


(9.81)

A:

Conditions a) and b) are characteristic for the so-called oFdered media and it is only through the interaction with an ordered medium that the magnetic flux in a flux string becomes quantized!

SOLUTIONS OF WORKED EXERCISES:


No. 9.1.2

(d Fa )F YS +ln ai3 d (FyoF ) 4 i3 yo S Y l(d Fa )FYS+l(d Fa )FYS+lnaSFYo(d F ) 2 S Y 2 S Y 2 S yo

+
2 0

S~o

+
2

S~Y;

Y~o

l(d Fa )Fyo+l(d Fa )FoY+lnaSFYo(3 F


Y Y 0 2

S yo

2n

1 as

[(doFSy)F

yo

+(3 y F So)F

oY yo

+(dSFyo)F

yo

2n

1 as

[doFi3y+dyFoS+3SFYOlF

But according to (9.2 ) we can exchange

with

o
No. 9.2.1

a)

A trivial calCulation gives


3R _ y. dY - R 3R
3z

=~
R

Using this We obtain

d [ 1] xFR-Z). 3 [ dX R(R-z) = - R (R-z) 2 I dy R(R-z)

1 ] __
-

y (2R-z) R3 (R-Z) 2

2[_1_]
dZ R(R-z)

J;,.
R'

b)

We can now calculate the curl! The x- and y-components offer no difficulties, but the z-component is a rather long story. If you split off the term z/R 3 and USe the relation x 2 +y2 = R2 _z 2 _2 you should, however, obtain the result listed. In the limit ~ the first term reproduces the monopole field, so it is the remaining term Which produces the string. It is preferable to decompose the second term in the following way:

504

II

E2 (2R-z)
~STRING

= =

BE

2E 2 (R-z) + E2 Z gM 2E2 A gM E2 Z A - 4n R 3 (R-z)k - 4n R 3 (R-z)2k

Now, choose a point on the ~ositive z-axis: (?,?,zo)' I~ the limit where E~O we can replace R-z by ~. Thus on the posltlve Z-axlS the two terms behave 2zo like gM 1 1 (El) (El) 2)
1)
1T

Thus the first term is approximately constant on the string, while the second term diverges. Consider the two terms as a function of the cylindrical radius p. They have a behaviour somewhat like the behaviour indicated on the figure 183.

El

Fig. 183a

Fig. 183b

Clearly the first term vanishes in the limit where E+O, while the second may very well produce a a-function. To investigate this consider it as a function on the plane perpendicular to the string. We want to calculate the integral:

!:~

..3...

4nJR2R3(R_Z)2dxdy

In the limit E+O the main contribution comes from a region close to the string. But then we can safely perform the SUbstitutions

Using polar coordinates we therefore get

. gM E2 Z 11m - 4n J R2 R 3 (R-z)2 dxdy


E~O

The latter integral is actually independent of E as you can easily see by performing the sUbstitution P=Et The integral then becomes a standard integral with the value ~ and We finally get . gM E 3 Z 11m - 4nJR2R3(R-Z)2 dxdy = -gM
E~O

But then the limit is a two-dimensional a-function! For a point on the negative zaxis we know that R-z~2tzl and therefore the string terms vanishes in the limit
E-+O.

505

II

No. 9.8.2 To calculate the integral explicitly we first find a coordinate system in Minkowski space where the coordinate expression for SaS(x) is especially simple. This is done in the following manner: Consider a space-slice. Then we USe the string itself as the positive x 3 -axis. In these special coordinates the position of the monopole is always

Time

intE

i
Fig. 184

(X I ,X 2 ,X 3 ) Xl

(0,0,0) x3 ~ 0

and the string always occupy the positions:

= x2 = 0

We say that these coordinates are adapted to the string. We can now evaluate the components of S ,

SaS(x) = gm I UU
As

~ ~ (x-x ~ (A , I ,2) axY dx e ,A ) EaSyoaIT aIT

dA dA

a parametrization of the sheet we simply choose

XO

= A2 , SaS

i l

=0

x2

= 0,

i 3

-AI

Then the expression for

reduces to

SaS(X)

= -Eas30gmIuo'(x-i(AI,A2))dAldA2
and the

If X3 is negative the sheet can never cross the point X = (xo,xl,x2,X3) integral automatically vanishes. If X3 is positive we get:
Iuo~ (x-x(A1

,A2})dA l dA2 =

Iuo(xo-io (AI ,A2))O(XI)O(X2)O(X3_i3 (AI ,A2))dA l dA2


o(X I )O(X2)

Consequently We have shown

SaS(X) = -gmEaS30o(XI)o(x2)0(X3)
Now it is trivial to compute the magnetic flux associated with face. If the surface is hit by the string, we can USe (X I ,X2 ) surface in a neighbourhood of the string:

S through any surto parametrize the

InS =
No. 9.4.'

I-

SI2 d Xl dx 2 =-gMIo(x l )o(x 2 )dx l dx 2 =-gM

The angular momentum in the direction of then given by:


A

fa
A

(where
]

fa

denotes a unit-vector) is

roO

~. ~ oE)(ror) ~ ~ -M 0 OE (r =47fEoJ-r-30
r

r[f

d3

To evaluate this integral we observe that

506

II

Inserting this relation we get

f .j = - ~4 _ 0 _ _ (ij'E)d 3x + ~4 ij. _ 0 _ o nor nor

I('? .;)

I [(f.;) E] d 3x
where ;, is the position of

In the first integral we use that

ij'E =

JLQ3(;_;,)
Eo

the charged particle. In the second integral we use Gauss' theorem to convert it to a "distorted" flux integral: Jij.[('?o:;t) E]d 3X

o which thus vanishes automatically. Consequently

~fcoseSinede~

I' .j = -K(f 'r') o 0 This shows precisely that the component of j perpendicular to that the component of j along t, is given by -K.

;,

vanishes and

No. 9.4.2 Consider a beam of electrically charged particles carrying a specific anergy E . A particle with impact parameter b will be scattered at a certain angle e=e(E;b) . Let us briefly recall the definition of the differential crOss section. For a given scattering angle e we consider the values of the impact parameter b corresponding to e (Remark: There may apriory be several different values of b producing the same

en.

bi -

".\----------r

Fig.

185
e and
s+~e

Then we consider a small ring between the area

. On the unitsphere it represents

~n = 21TSine~e

The values of b corresponding to this ring will itself form a number of rings from b i to bi+~bi. The total area of these rings is given by
~O

E21Tb.~b. 1. 1.

The differential cross section

~ is then defined as
21TSine~e

(9.82)

dOte) =
dQ

beam area = detector area

l:21TMb

= Ebl~1 d(Cose)

where we sum over all values of b corresponding to the given value of e. We know that a given particle moves On a cone. It is customary to characterize this cone by the angle X, where 1T-X is the angle spanned by the cone. This angle can be expressed in terms of t and K Considering the triangle shown on fig. 186. we immediately get

507 X Cot 2 2
mvb

II

= lKf = lKf

As we shall see, the scattering angle e depends only on X We can therefore rearrange eq. (9.82) as (9.84) It remains to determine e as a function of X . Here it is profitable to use an adapted coordinate system with the symmetriaxis as the third axis, i.e. j is pointing along the third axis. The position of the particle ;(t) can now be expressed in terms of the radial distance r and the azimutal angle <p Using this we get
....

v = dt =

dr

....

drA ~

+ dt Xr

~ ....

On the other hand we obtain from eq. (9.39) that

jx;
and consequently (9.86)
v

= (rxm;)x; = r2m;
(~.;)~ + ~x:t
mr

- ;(~)

By comparing the expressions

(9.85)

and (9.86)

we then deduce

Using eq. (9.38) we can now integrate the above equation for choose <p( 0) = 0 Remember also that b=d): <p(r ) = ~ dt dr o b mr2 dr
0

<p (Obviously we can

= ~[~ mv 2

arcsin

~] ro

where + refers to positive time and - refers to negative time. This leads to the following asymptotic values (9.88) (Since <p
lirrujl(t)
t~co

has the total increase

1\<p=~=4 mVb 2
we see that for S/2=2n the particle will move n Now we can finally express the scattering angle (9.86) we get for large
_~( --00) :

times around the symmetry-axis.) e in terms of X . From eq.

It I
~(+w)
and

and therefore the scattering angle

is identical to the angle between

508

II

1\

v (-"')

Fig. 187

1\

1\

r(-"')

-r(-"')

Thus we obtain cose

-;(+00);(-00)
-Cos2 ~[Cos~(+oo)cos~(-OO)+Sin~(+OO)Sin~(-oo)1-Sin2 } 2Cos2

-~- .Sin2r~ .!L] 2 lmvb 2

Using that

~
mvb

= 122+K2 = 11+tan2 X

=_1_

Cos X
2

we finally obtain the desired formula cose = 2Cos2 If we put

f Sin2 [ cos ~xJ 2

!.
I; =_2_

Cos X
2

this implies that

d(~~se)

=2Si~os}[-Sin21;+I;Sin~cosI;1
We have derived the following marvelous for-

Okay, substituting this into eq. (9.84) mula for the differential crOss section

.!L I; =_2_ Cos} Having gone that far let us make some comments. First you should compare it with the Rutherford formula for Coulomb scattering

(9.92)

do
<ill

lIe 2 )2 =1I\2E

1
Sin"S!.
2

Observe that the energy dependence ].s different in the two cases:

(~) monopole ~ i
and

1 dO) ( <ill Coulomb ~ #

Then we consider forward scatterin~, where e is very small. From eq. (9.89) that to lowest order 0=X Furthermore we observe that Cos1;;"'O

we see

509

II

In the case of forwards scattering we can therefore use the following approximation

_1_

Sin4&
2

which has a structure very similar to the Rutherford formula. For the case of backwards scattering. things are more complicated. On the one hand w~ noW get.cont:ib~t~ons from several values of X. On the other hand ~ b~comes slngular, l.e. lnflnlte at a series of angles en converging towards n . Th~s is due to the term
Sin2s-~Sin~Cos~
I

which can no longer be neglected. The denominator becomes zero when tans

=~
~

For positive ~ this equation has an infinite number of solutions these solutions corresponds a specific value of e. The angles en are called rainbOW angles.

'

. To each ~f nfor which ~


d(l

=m

'

No. 9.5.2
a) b)

t-r

~t
=

~.(tX(-iflV-q!)]

[;x(-inV-qA)].;

-in;- [;xV] -ili.CtxV]-x:


to component expressions:
=

IT
No. 9.5.3 The analysis of the eigenvalue equation (9.65) is most easily performed using the ralslng and lowering operators J = J l iJ2 Neglecting the normalization for a moment we know that
JP jm
a:

P jml
K

In the Schwinger formalism the raising and lowering operators are given by

J
P

=-~+ a~ mc:~: h]
a:

=-ll/l-?

[a: ~:5]

From this we obtain the following recurrence relations jml

Il_X2[~ mx+~ ax l-xTJ

P.

Jm

These recurrence relations can be further rearranged using the identity P'+fP = eff(x)dx In this case

~ [eff(X)dx p]

510
and therefore ff(x)dx

II

= -~(m+~)ln(l-x)
'( K '(

- ~(m- K)ln(l+x)

The recurrence relations are then rearranged as P.


~l

a:

(l-,(J' (l-x)' mfi) (1+,J m+r;

-K)t[

d (l-x)+;; m+ f;" (1+x)+;; m - f;" P. dx ~

-'(

K)

-'(

K)

Ey induction we then deduce

( * )

P.

Jm+n

(x)

a:

(1_x)2(n+m+i;" (l+x)2(n+m-i;"

K)

K)

( ** )

( jm- n x)

a:

(l _x)~(n-m -~) (l+x)Hn-m +i)

This is our main result! From the general theory we know that vanish for sufficiently large n. But this implies that

Pjm+n

and

P.

Jm-n

and
Q. :

(l_x)Hm+i)(I+X)Hm-i)p. (x) Jm Pjm(x)

= Q.

Jm

(x) and

are polynomials. Eliminating Jm

we get the following relation between


K K

Qjm(x) = (I-X)m+il(I+X)m-KQjm(x) But as Q. and Q! are polynomials this is only consistent if are both I~tegers! Jm Using the relations and we see that m and and
K m-J;

f;"

are both either half-integers

or integers!

From the recurrence relations (*,**) we can also obtain explicit expressions fr the functions P. (x) . We know that P .. (x) is killed by the raising operator J+ Therefore we get Jm 1 JJ C'(l-X)' J +i;" (l+x)' J-i;" JJ . IKI This is a regular normalizable solution provided J~ we then obtain the rest of the functions:
P .. (x) =

, (.

K)

'(.

K)

. Using the lowering operator

511

II

chapter 10
SMOOTH MAPS - WINDING NUMBERS
10.1 LOCAL PROPERTIES OF SMOOTH MAPS
In this chapter we are going to examine maps from one manifold to another and unfortunately it is going to consist of some rather technical investigations, but there are at least two main reasons for doing this anyhow: First of all maps actually play an important role in physics. When we e.g. discuss space-time symmetries we consider certain maps from space-time into itself, such as rotations or translations, and we then investigate what happens to field configurations during such symmetry transformations. (This will be explored carefully in the sections 11.4-11.6). It is also worth noticing that field configurations themselves are maps from space-time into a field space. E.g. a complex scalar field, say a charged Klein-Gordon field, is simply a smooth map from space-time into the complex numbers, and properties of such maps may contain non-trivial informations about the system we are considering. (Examples of such non-trivial properties will be given in section The second reason is that it turns out that we can use maps to transfer differential forms from one manifold to another in such a way that the various operations of the exterior calculus, i.e. the exterior derivative, the wedge product etc., are preserved. This gives a much greater flexibility in the computation of various quantities. Okay, as we can not talk us out of it, let struct a map from one manifold to another, explain what we mean by Let f f be a continuous map.
M

10.8-10.9 ).

and

N'n

be dif-

ferentiable manifolds. We want to consider what happens when we conf : M m ~ N n. We want to being smooth at a point

P.

Consider a point Po in M and the corresponding point


Q = f(P ) O O

i'n

N.

I f we

introduce coordinate systems covering

Po

and

QO' then f is represented by an ordinary

Fig. 188

512 Euclidean function: (y1, ... ,yn) = f(x1, ... ,xm) Let U be an open neighbourhood around
~

II

neighbourhOOd around

because

f- 1 (U) is an open O is continuous. But this shows

Q ' Then

that the Euclidean representative f is well defined in a neighbourm hood of (X 1 , .. ,x ), where (x01, ... ,xom) are the coordinates of 0 o

PO'
smooth at Thus i t has meaning to ask if the Euclidean representative f (x 1 , ,x m). [If f is not supposed to be continuous is

then the Euclidean representative f need not be defined at pOints close to (x 1 , ,x m) and you can no longer wOrk out partial deri-

vatives!] This motivates the following definition:

Definition 1 m A aontinuous map, f: M ~ N n in M,


is said to be smooth at a point
P

is an ordinary smooth m EuaZidean map at the aorresponding aoordinate point (X 1 , .. ,x ).

i f the aoordinate representative

Observe that if just one coordinate representative is smooth, then they are all smooth, because different coordinate systems depend smoothly upon each other. Consider now a smooth map matrix in [ayi/axj] f. Then a coordinate representative

will have partial derivatives of arbitrarily high order. The Jacobi local properties of is of particular interest to us as it controlls the f. In a neighbourhood of a point P(Xb , ,x m)

we have the Taylor expansion

~ AX] + higher order terms

ax]1
~n

Xo
P with a linear

Thus when we approximate

a neighbourhood of

map, this linear map will precisely be generated by the Jacobi matrix. we can make these vague ideas more precise in the following way: Consider the tangent spaces show that f T P (M) and T f (P) (N ). We will now generates, in a canonical fashion, a linear map
f

* :

TP

(M )

~ T f (P) (N )

which controls the lOcal properties of the Jacobi matrix!

and which is represented by

TO see how this comes about, we proceed in the following way: Let ~P be a tangent vector at curve A(t).
P.

Then

~P

is generated by a smooth

513

II

----------------------i.~ t-axis
But we can use the smooth map a smooth curve
th~ transport of

Fig. 189 A(t) into A'

f on

to transport the curve

A' (t) = f(A(t))

N.

The transferred curve

generates a tangent vector that

uQ
P

Vp'

at Q = f(P), which we define to be Q However this only makes sense if we can show A(t) we chose to xi = xi (t) ,

is independent of the particular curve vp'

generate with

TO see that this is actually the case, we work out the coLet A be given by the parametrization xi(O).
i

ordinates of coordinates

U : Q

corresponding to

Then

~P

is characterized by the

dx

dt!t=O is characterized by the parametrization The curve A' = f(A) yi = f"i(xi(t)). Using the chain rule, we therefore find the following coordinates of
b~

... u

=~
dt\t=O

d i

But this clearly shows that the coordinates of coordinates of

~P'

U Q

only depend on the A. Furthermore

not on the particular curve

we immediately get that the induced map,


f* : Tp(M) ~ TQ(N)

is represented by the Jacobi matrix:


(10.1 )

[~]\

Especially it is a linear map! The induced map satisfies a natural composition rule:

514

II

Theorem 1 Let there be given three differentiab~e manifo~ds and two smooth maps: L...!.... M.JL. ,"I. Then
(10.2)

M,

and L

(gof)*

g*of*

Proof: This is an immediate consequence of the chain rule. Let us introduce three coordinate systems: (x 1, ... ,Xl) covering (y 1, ... ,ym) covering and (z1, ... ,zn) covering P Q
R

in L f (P) g(Q)
L) ...

in M in
N
N)

(gof) * is the linear map T p ( Then Jacobi matrix, i az axj or equivalently, i az ~ ayk axj

TR (

generated by the

We conclude this section with a discussion of the behaviour of vector fields. Let there be given a smooth vector field We know that f generates a pointwise map f* : Tp(M) ~ Tf(P)(N) Does this mean that vector field f*(V)? f* actually maps the vector field No, not in general! V onto another Let us explain where the
M

V(x)

on

troubles may come from: (a) If f is not surjective, there are pOints on
(b)

at which

we do not attach any tangent vector. If f is not injective, there exist two points which are mapped into the same point two tangent vectors' V(P ) Q.

P and P 2 1 Consequently the

and V(P 2 ) may very well be map1 ped into two different tangent vectors at Q = f(P ) = f(P 2 ) 1 N, the resulSo when we transport a vector field from to M ting "vector field" will in general be a "multivalued" vector field which is not "globally defined on N". (Consider, for instance, a constant map f which maps the whole manifold M into a single pOint). But even if f is bijective, we may be in trouble. Consider the coordinate expression for the transported vector field: b i (y1, ... ,yn) =

~.
axj

a j (x 1 , .. ,x n )

515

II

We must demand that it depends smoothly on the coordinates (y1, ... ,yn). If the inverse function f- 1 is smooth too, then everything works out fine and we get
(10.3)

2L j
ax

If-1 (y)

. a j (f- 1

(y) )

where the right hand side is smooth. But if in trouble! We must therefore demand that but that i t is a diffeomorphism too! f

f-

is not smooth we are

is not only bijective

Conclusion:

Only a diffeomorphism generates a well-defined trans-

port of vector fields from one manifold to another.

We proceed to study the local properties of induced map stems around f. P Let us define the rank of f*[T p ( M)]. and f(P) f the dimension of the vectorspace

in terms of the P f to be is

at a point

(Using coordinate sy-

is is well known that the rank of

equal to the rank of the Jacobi matrix case we can then define:
Definition 2 A smooth map map
f f

Df(P).) As in the Euclidean

is regular at a point

provided the induced

has maximal rank.

We can introduce still a useful concept. Let

f: M

be a

smooth map and let Q be a point in N. Then we consider the preimage f- 1 (Q) (whic~can be empty if f (M) does not cover Q):
Definition 3 *) We say that in the pre image of critical value. preimage is empty.)

is a regular value i f

is regular at all points then it is called a

Q.

If

is not regular,

(Remark: Q

is also counted as a regular value i f the

*)The above definition differs slightly from the definition used by the mathematicians in the case where Dim M < Dim N. See e.g. Guillemin/Pollack [1974].

516

II

Remember that the tangent spaces have the same dimension as the underlying manifold. We can then distinguish between three cases: Dim M = Dim N , Dim M < Dim N , Dim M > Dim N We will now discuss the three cases in some more detail: (a) Dim M = Dim N As the first case we consider the especially important one, where M and N has the same dimension n. are vectorspaces of the same dimension Then Tp( M) and T Q { N) n, and f is regular at a point P exactly when f* maps T P ( M) isomorphically onto T Q ( N) This means that the Jacobimatrix Df(P) is a regular square matrix, i.e. it has non-zero determinant. But then the inverse function theorem from analysis tells us that f actually maps an open neighbourhood U of P onto an open neighbourhood V of stricts to a diffeomorphism ,f: U ly nice locally! Q
~

v.

and furthermore that f reThus f behaves extreme-

Fig. 190 The inverse map f IV (f*) where


(10.4)

-1

then generates the reciprocal map:


M)
~ Tp( N)

-1

: TQ(

This is a trivial consequence of the composition theoEem (10.2) since the composite map of flU and flU reduces to the identity map. We can also consider a regular value Q in N. If f- 1 (QO) O is non-empty we can find Po such that f(PO) = Q and by definition O f is regular at PO' It therefore maps an open neighbourhood U of Po onto an open neighbourhood V of Q This argument can be O strengthened: Suppose f- 1 (QO) is finite and put {P ,,P k } 1 Then there exist disjoint open neighbourhoods
f- (QO)

-1

U i

of

Pi

and open

517 neighbourhoods
f:Ui~Vi

II restricts to diffeomorphism

Vi

of

so that

But here we can safely replace V1 ""'V k by their intersection V ~ V. which will be an open neighbourhood of Q We O i=1 l. 1 can then shrink U to U1 U n f- w) and f will now restrict i i to diffeomorphisms,
f: U!
l.
~

Thus we have shown the following useful Lemma:

Lemma 1 Let f:Mn~Nn that f-l(Q)

be a smooth map and let Q be a regular value such say

is finite,

f-l(Q) = {P1, ... ,P k }. Then there exists a single open neighbourhood V of Q such that decomposes into disjoint neighbourhoods
f:Ui~V

U1"",U k of

P1"",P k

and

such that the map f restricts to diffeomorphisms:


,i=l, ... ,k

Fig. 191 Remark: In the above lemma it is essential that It is in g~ false when f-1(Q) is infinite. f- 1 (Q) is finite.

n m (bl DimM < DimN : A smooth map f: M ~ N (m < n) which is everywhere regular is called an immersion . Let f be such an immersion. In analogy with the discussion of Euclidean manifolds we expect f(M) to be a submanifold of N. Locally everything works out all right and we can transfer coordinate systems from M to f( M). But globally we may be in trouble because f need not be injective so that f( M) can have self intersections. (See figure 192 ). Consequently we must demand that f is injective. But that is not enough. We still have to worry about the topology since the inverse map f- 1 : f( M) ~ M need not be continuous. We eliminate this by demanding that f is a homeomorphism, i.e. f is injective and both f and f- 1 are continuous. Then it should come as no great surprise to you that the following theorem holds:

518

II

Fig. 192

Fig. 193

Theorem 2 Let
(1) (2)

f: ~ ~ ~ (m < n) be f is an immersion f is a homeomorphism,

a smooth map such that


f:

then

f(M)

is a submanifold of N and

M ~ f(M)

a diffeomorphism.

A smooth map which satisfies both the above properties is called an embedding and we say that M is embedded in N. (c) DimM > DimN: A smooth map f: M N n (m > n) which is everywhere regular is called a submersion. Now let f be a smooth map and let Q be a point in N We consider the pre image f- 1 (QO)' cf. O fig. 193, wh~ch consists of all solutions to the equation
f,1 (x 1, . ,x

m)

i. e.

fn(x 1, . ,x m )

We have previously discussed how to construct Euclidean manifolds using equations of constraints (See section 6.3 ). Using those techniques we can generaliz;e theorem 2 , section 6.3 to the following theorem:

Theorem 3 Let f:!.f'",~ (m > n) be a Smooth map. If Q is a regular vaZue in N, then either (1) f-~ (Q) is empty or (2) f- (Q) is an (n-m)-dimensional submanifold of M.
We recapitulate the preceding discussion in the following scheme (See also figure 192 and 193 ). f: M m
~

n N

If

is a regular map then f it is called an Immersion

Such a map can be used to construct submanifolds in


N M

m < n m > n

of dimension m of dimension m-n

Submersion

519 In the preceding


~iscussion

II

we introduced the notion of regular

points and regular values. They were important in the characterization of the behavior of smooth maps. But to use the theorems just obtained it will be necessary to control the existence of regular points and values. For a specific map we can of course compute the Jacobi matrix and check its rank and thereby determine the actual positions of the regular pOints and the regular values. But interestingly enough it turns out that it is possible to make general statements which holds for any smooth map. Consider for simplicity ordinary functions f: R
~

R.

Then a regular fl (x)

pOint is a point where

O. It

is easy to construct smooth functions which are everywhere regular, e.g. fix)

= x.
f(x)

It is equally easy to con-

struct smooth fUnctions which are nowhere regular, e.g. the constant function
=
~~----------

o.

Thus we cannot hope

____~~x
Fig. 194

for general statements about regular

points. But consider now regular values. Even in the worst case of a constant function there is only a single critical value, so the critical values seems to be very exceptional. This turns out to be a general feature of a smooth map. Consider a smooth map,f: ~~Nn. The points in M can be divided into two types: (a) Critical pOints Similarly the pOints in (b) Regular pOints

can be divided into 3 types: (b) Images of only regular points (c) Not an image of any point

(a) Image of a critical point

A famous theorem of Sard now states that the set of criticaZ


points is mapped into a zero-set of vaZues is a zero-set in N. N, i.e. the set of criticaZ

Equivalently we can say that almost (For a complete discussion of

every pOint in

is a regular value.

zero sets on manifolds you should consult e.g. Spivac [1970] or Guillemin & Pollack [1974]). Remark: By abuse of notation we will call a smooth map regular even if it does possess critical points as long as the critical points are all isolated. E.g. the function indicated on fig. 194 will thus, slightly incorrect, be referred to as a regular function.

520

II

10.2 PULL BACKS OF CO-TENSORS


Let f:
M

m~

Nn

be a smooth map. We have seen that we run into

difficulties when we try to transport vector fields from one manifold to another. We will now investigate what happens when we try to transfer other objects. It is instructive to consider maps first: Suppose <p: R
~

is a smooth map into fo<p: R


~N

Then we can clearinto


N. This 1s

ly push it forward to a smooth map


M

connected with the transport of vectors. A smooth map <p: R


M

represents a

,1
R <pof: M
~

,-

,.

curve on

M.

But smooth curves gene-

rate tangent vectors and that is the basic mechanism behind the transport of tangent vectors as explained in section

,. ,.

,. ,/ fo <p

10.1 .
Suppose then that <p: N
~

is a

smooth fUnction on R on

N
M.

This can be pulled back to a smooth function This can now be used to transport covectors. COnsider a covector which generates
W,
ill

, ,,
M

<pof ,-

,
f

,. ,.

,~ f,

on N <p:
W

Then we
N

can find a smooth function Le. <pof:


d(<po f)

R
I

d<p

and
If

this smooth fUnction is pulled back to the smooth function we can show that
w
Ni~R

is indepen-

dent of the particular function sen we can therefore pull back denoted
f w.

to

d(<pof)

<p choThis also allows us to which will be

determine the coordinates of the pulled-back covector

The above considerations suggest that it is natural to push forward tangent vectors and to pull back covectors , but it even turns out that pull-backs have nicer global properties than vector transports. This we will now investigate in some detail.

521

II

We start by reinvestigating the pull-back of covectors: be a co-vector field on vector field on on


M.

Let

w
p

N.

We want to construct an associated cof *w. Consider a point T p (M) is transported to a unique

which will be denoted T f (P) ( N).

Then a tangent vector in

tangent vector in

This motivates the following definition:

Definition 4
(10.5)

Fig. 195

------------------+.
* T P (M ) f,,-,:
~

R
~

Clearly

is a linear map since

f*: T p (M )

Tf(p) (N) is line:r but then we have constructed a globally defined covector field f w on M. It remains to be shown that it is smooth. is characterized by To do that we introduce coordinates. Then f *w the component functions
(10.6)

bi(x)

-+ cf *wle,>

-+ cwlf*e i>

a.

(y(x~. ~

ax

av j . k =

a.(y(x
J

lY.! .
ax~

But the y-coordinates depend smoothly on the x-coordinates since f is a smooth map, and the coefficients a.(y) are smooth functions, since This shows that the w is a smooth covector field on N. right hand side depends smoothly on x. So the miracle has happened. We never get in trouble when we try to pull back covector fields! Clearly the above analysis may be extended to arbitrary co-tensor fields:
J

522
Theorem 4 Let

II

f:

Mm ~

Nn

be a smooth map. Then


~T(O,k)(M)

generates a map

f * : T (O,k) ( N)
i. e.

f*

pulls back a smooth M. If


T

c"tensor field on is a smooth


-+

to a smooth co-

tensor field on

cotensor field on
-+

N,

then

the pull-back is given by,

f T (V 1 , .. ,vk) ; T (f*v1 , ,f,..vk) which can be written out in components as follows,


(10.8)

(10.7)

*-+

-+

(f *T) .

(x)

1.1l.k

= T.
and

(y (x) )

~ i
ax
1

J 1 Jk

Observe that if tes on a manifold!

M =N

is the identical map, then we

recover the usual transformation formula for the exchange of coordinaThe pull-back has several simple properties:
Theorem 5
(a)

(10.9)
(b)

is linear, i. e. * * * f (T+s) = f (Tl + f (S)

* (AT)

* <T)

commutes with the tensor product, i. e.

(10.10)

f * (TS) ;
L

* (T)
then

* (S)
L, M, N and two smooth maps:

(c)
(10.11)

If we have three manifolds

.5...

(gof)*

= f*og*

Proof: Only the composition rule is worth consideration. It can be derived directly from the composition rule for vector transports (10.2): * ~ ~ (gof) T(v1'. ,vk)
-to

def
-+

T[,(gof) *v1'' (gof) *vk1

... :t f * (g * T)[V , .. ,v 1 k 1

T[(g*(f*v 1 ),,g*(f*v k )]

(g T)[f*v 1 ,,f*v k l

-+

-+

If you prefer, it can of course also be derived from the component expression for the pull-back using the chain-rule. We shall esepcially use the pull-back in two cases: a) Metrics: If to a cotensor
g

is a metric on f *g

N,

then

is a symmetric, non-

degenerate cotensor field

field of rank 2. Consequently we can pull it back of rank 2 on

M.

It will obviously be symmetric, but it need not be non-degenerate!

523

II
f g

(If e.g.

is constant,then
M.

*g

vanishes everywhere). Thus

needs not be a metric on


Exercise 10.2.1

Problem: Let N be a Riemannian manifold with the positive-definite metric g. Show that f*g is a metric on M if and only if f* is everywhere injective. Exercise 10.2.2 Problem: (a) Let M and N be manifolds of the same dimension. Suppose N is equipped with a Minkowski metric and ~et f: M N be a smooth regular map. Show that f g is a Minkowski metric on M 4 Let N be the Minkowski space R with the usual metric and let M be the real line R As the smooth map we consider f(t)

(b)

(t;t;O;O) f *g

Show that f* is everywhere injective, but that vanishes identically.


b) Differential forms:

If

is a differential form on

N,

then

T is by definition a skew-symmetric cotensor we can pull it back to a cotensor metric, i.e.


M

field. Consequently The pull-back of a dif-

field, f *T, which will be skew-sym-

f*T

is a differential form too!

ferential form is particularly simple. Let us introduce coordinates on and


N.

(7.8)

Then we may decompose T as 1 i1 ik T = k! T. . (y) dy l\ l\ dy


~1" '~k

From the transformation rUle (theorem 4) we get that rized by the components: T . . . . . (y(x) ) ~ i J1 Jk ax 1 Le. it can be decomposed as
(10.12)
f *T =
j1

f *T is characte-

jk ~ i k ax

1. " k.

. J 1 . Jk

( Y (x

~
i

i dx 1

i
l\ l\

dx

ax 1 T by performing the

But then you see that substitution

f*T

is obtained from

dx j axj so as usual the formalism produces the simplest possible answer.


(10.13)

dyi

Before we proceed with the investigation of differential forms, we will briefly discuss how one can extend the transport of tangent vectors to a transport of arbitrary tensors. One must then restrict to diffearorphisns.

524
Definition 5
A diffeomorphism, f: Mn ~ N n,

II

generates a pull-back of tangent

vectors from

to

M,

f*:TQ(N)

... T

-1
(Q)

(M)
Q

If
( 10.14)

is a tangent vector at
f

in

then

is given by

*;;Q =

(f -1 ) * ;;Q '

and in coordinates it is characterized by the components,


( 10.15)

... i [f* Ql

But once we can pull-back both tangent vectors and covectors we can clearly push forward tensors of arbitrary type.
Definition 6

A diffeomorphism ,f:
of type

generates a transport of tensors

(p,q)

from

to

N:

T f(P)
If given by
T

(p,q) ( N)
at the point

is a tensor of type

(p,q)

then

f*T

is

.... * * *.... *--jo (f*T )(w ' ,w; Y1""'v) = T(f w , . ,f W; f v , .. ,f v ) q 1 ... 1 ~ q ...1 ~ and in coordinates it is characteri~ed bY,the components i i ~ ~ k 1 q ~m 1 kp (f-1(y~ .~ k k L R.1 iq J Jq (10.17)
--jo

.aL2.
ax

1 ,

ax

ay

ay

Observe that in this way smooth tensor fields are transported into smooth tensor fields. Of course we could equally well pull them back using the inverse map. Theorem 5 can now be generalized as follows
Theorem 6 Suppose f: M n N n is a diffeomorphism inducing the tensor transport f.: T (p,q) ( M) ~ T (p,q) ( N ). Then

(a) (10.18 ) (b) (10.19 ) (c)

f*

is linear, i. e.

f* (T+S) = f * (T) + f*(S) f* (AT) Af* (Tl f* commutes with the tensor product, i. e. f*(T~) = f * (T) GI f* (S) f* commutes with contractions, i. e.

525
(10.20 )
(d)

II

C C f*( T ) = [f*T1 If we have 3 manifolds


and 2 diffeomorphisms ,

L , L ~

of the same dimension then

M ~ N ,

(10.21 )

(gof)* = g*of*

Proof: There is no need to go through all the details. It suffices to observe that when f: M

is a diffeomorphism we can use


M

f
~

to
M

transport a coordinate system from is a coordinate system on on f*T

to

N,

i.e.

if

cj>: U

M then

focj>:U

N
yi

is a coordinate system

N.

Using such associated coordinate systems the coordinate exf reduces to the identity map

pression for

xi

and

and

get identical components.

fo~.Because

In the rest of this section we restrict our considerations to differential of the exterior calculus the pull-back of differential f: M

forms is a very powerful tool. Consider a smooth map generates pull-backs:

N.

It

f*:J\k(M)~J\k(N)

k=0,1,2,

which we know are linear (theorem 5). We will investigate that part of the exterior calculus which only depends on the manifold structure and not ,e.g., on a metric. Thus, for the moment, we are concerned only with the wedge product, the exterior derivative and the integral of differential forms.
Theorem 7 The pull-back commutes with the wedge product and the exterior derivative, i.e.
(10.22) (10.23)
f

* (T"S)

(f * T)

"

(f * S)

f*(dT )

d(f*T)

Proof: To avoid drowning in indices we work i t only out in the case of one-forms. * * * * * * (a) f (A"Bl = f (AB-BA) = f (A) f (B)-f (B)f (Al so the result is a trivial consequence of theorem 6b. f

* (Al"f * (B)

526
(b)

II

[f * (d(A)lkR.
(NB!)

[dAl ij axk ax'!'


j

~~= (~_ aAi)~~=~~_aAi~


ay~ a~ axk axR.
axk axR. axR. axk a ~_ a * a * -R. (Ai k) - ---l<f Ali - axR.[f Alk ax ax ax"

_a_(A } . Y':) axk Jaxi

= [d(f*A )lkR.

Observe that there is a subtle point involved in these rearrangements: If you actually carry out the differentiations involved, you will pick up two extra terms involving the double derivatives ~ k 1 and ~ 1 k ax ax ax ax but since partial derivatives commute they cancel automatically. U
Mathematicians have a very efficient way of representing computation rules like those of theorem 7. Consider especially the rule concerning the exterior derivative

( 10.23)

*od

= do f

When you pull back a differential form and take the exterior derivative it does not matter in which order you perform the operations. This is clearly an example of a commutative law as you know it from elementary number theory. Now consider the following diagram where we have represented the various maps involved in (10.23) as arrows. If you take an f* element in the upper right corner you can map it along the arrows and evidently there are two possible ways of mapping it down to the lower left corner~where it can arrive either as d (r'T) or as f*(dT), but according to (10.23) they are identical so the end result is independent of which f* route you actually follow in the diagram. This'is expressed by saying that the diagram is commutative.

..

..

As another example consider the composition rule expressed in theorem 6d. Here we have 3 maps involved corresponding to the following diagram and the computation f~ M , g * rule in theorem 6d simply states that this L - -..... ....---N diagram is commutative. (gof)* When we corne to integration of differential forms we must be some-

what more careful. We want to find out what happens to an expression

like
rank

J rl
k

when we pull it back. Here and

T is a differential form of
N

is a k-dirnensional orientable regular domain in

527

II

We must therefore also investigate what happens to n when we pull that back: In general f- 1 (n) will not be a k-dimensional orientable regular domain of nal!

for the following reaSOns: It need not be

compact, i t need not be a submanifold, and it need not be k-dimensioTo see this consider the following example
Examp Ze

1: Let f (t,x)
N

= R 2, t 2 _ x2

and put (0,0).

Then

f: M

is a smooth map although i t is not regular at

The set n [-1,1] is a 1-dimensional regular domain in N, and f- 1 (n) is the "strip" between the hyperbolas as shown on the figure.

N +l

-1
Fig.

196

It is not compact, and it is not one-dimensional.

0
n'
diffeomorphi-

Consequently we must proceed a little differently. We must try to find a regular domain

n'

in

such that

maps

cally onto the regular domain


ExampZe 2

n.
R

Consider the following example: and consider the smooth map

Let f(x)

M =

_x_ 1+x2 which has the following graph. Then

n=

[-~,~]

is a 1-dimensional re-

gular domain in N The pre image f- 1 (n) = R is not a regular domain, but

n'

[-1,1]

c::

is a

1-dimensional regular domain in

which is mapped diffeomorphi-

Fig. 197

528 cally onto f n by f.

II

(In one dimension this essentially means that f to n').

is monotonic when we restrict

Exercise 10.2.3
Problem: Reexamine example 1 and try to find a suitable l-dimensional regular domain n' which is mapped diffeomorphically onto n = [-1 ;1].

Okay, let integrals

be a given regular domain such that we can find n. Then we can compare the

n'

which is mapped diffeomorphically onto

f*T J n'
Theorem 8 Let

and

T be a differential form on

of rank
M,

k.

Let
f

be a maps

k-dimensional orientable regular domain in

such that

n
in

diffeomorphically onto a k-dimensional orientable regular domain


N.

Then if
f

preserves the orientation reserves the orientation

(10.24) if
f

Proof (outline): The proof will proceed in two steps. Step 1: Here we assume that intn is a simple subrnanifold. Then we
(~,U)

may cover it with a single coordinate system the positive orientation:

which generates

~: U

bij., intn c: M;i.e.x i

= ~i(A1, ... ,),k)

u
,. '"
N

Fig. 198

AS

maps

intn

diffeomorphically onto
int f(n):
c: N

int f(n)

we can use

to

generate a coordinate system on


f ,: U

~, int f(n)

529 In these coordinates the restriction of by the Euclidean function f to intn

II is represented

lli=Ai;

i = 1 , .. ,k

i.e. the identity map. If f preserves the orientation, then (fo~,U), 1 i.e. the (11 , .. ,llk)_ coordinates generates a positive orientation on reserves the orientation, then (fo~,u) generates a nef (n). If f
gative orientation and this costs a sign when we evaluate the integral

using these coordinates.


Okay, working out the two integrals in these especially adapted coordinate systems we get for the case where the orientation: f actually preserves

(f T). . (x('\)) - 1 U ~1"'~k . a,\

ax

ax

i. J d'\ 1 d,\k

J1
(y (x))E.Y...-

J J

1 . J1 T. . (y(x('\))) EY1 U J1 J k . aA J1 T. . (y(ll)) ~ 1 u J l' Jk all

ax

Jk ~1 ~~ 1 . ik ax . a,\ Jk ~ d,\1 . d,\k


j

. a,\k.

aA dll 1 dll k =

~ k
all

JT f(n)

StfrP 2: Hfrrfr intn is an arbitrary manifold. Wfr can thfrn COvfrr it with a vositivfr orifrntfrd atlas intn c

iEI

~.(U.)
J. J.

Using a partition of unity Wfr now cut up T in small pifrcfrs so that frach littlfr pifrcfr is frfffrctivfrly conCfrntratfrd in onfr of thfr simplfr manifolds ~(U.). From stfrP 1 thfr linfrarity of f* and thfr linfrarity of thfr intfrgfal Wfr now dfrdUCfr thfr dfrsirfrd formula for thfr gfrnfrral casfr.1

If we restrict ourselves to diffeomorphisms we can especially push differential forms forward. The statements of theorem 7 and 8 now carryover trivially for the transport of differential forms using diffeomorphisms, i.e. the transport commutes with the wedge product and the exterior derivative and i t preserves integrals.

530

II

10.3

ISOMETRIES AND CONFORMAL MAPS


In the preceding section we considered pull-backs of differential

forms. This time we will concentrate on metrical aspects of smooth maps. Suppose

(M n'(/1)

and

(N n'(/2)

are differentiable manifolds (/1 and (/2' If f: M .... N (/2 to a

of the same dimension and with metrics new metric manifold f f *(/2 on (/1

is a smooth regular map we have seen that we can pull back

M
and on

We have now two metrics on the same initial f * (/2

M and we can then investigate various metrical aspects of

by comparing

Consider e.g. a smooth curve smooth curve f (rJ


N

on

M.

It is transferred to a
N

Now suppose we want to compare the lenghts of corresponding arcs. Then we get

((.)~
L11 ~~~~(
(~r,

A A2 . B dP)dA { Arclength of P P 2 on M}= ~A J 2/(/1 dA dA 1 A1 df(p) _ dP and using that we furthermore get f*dA A } - f A21r, df(P) df(p) (/2 (~, ~) dA .-{ ArC~ength of f (P 1 ) f (P 2) on N A2 1 [Arclength of. P P dp f*dA)dA = fA /f*(/2(~r, ~r)dA = on M rela~ive1t~ = 1 lthe metric! f*(/2 .

------ar- -

J</(/2(f*~r,

----

But then we see that we only have to consider the initial manifold

where we can compute the arclengths of (/1 and f *(/2

P7P2

relative td the two

metrics

\\.

The simplest possible behavior of a smooth map and f *(/2 coincide.

is

whe~.
\

(/1
\

\\
Definition? A smooth regular map, f: it preserves the metric, i.e.
M

n ~ N n , is called a local isometry i f

* f(/2= (/1 It is called a globaL isometry i f it is a diffeomorphism too.


Observe that isometries preserve arclengths. Isometries are however very special maps so it is usefull to discuss a broader class of

531 maps where we still have some controll of what is going on.
Definition 8

II

~ N , is called a conformal!!!EE. i f it rescales the metric, i.e. there is a strictly positive scalar field Q2(x) on M such that
A smooth regular map, f:

* f!12

n 2 (x)!11 of notation we call a smooth map regular

Remember that by abuse ted.

even in the presence of critical points as long as they are all isolaLet us first give some general remarks about isometries and conformal maps to acquaintain you with these new concepts. Suppose a pcint in map f* P is

As we have seen a smooth map

generates a linear Tf(p) (N).

from the tangent space f


... -+

Tp(M)

to the tangent space

Consider first Riemannian manifolds where the tangent spaces are Euclidean. When is an isometry then

so that vectors. When

f*

!11 (u,v) = f !12 (u,v) = !12 (f*u,f*v) preserves the inner product between two tangent vectors,

-+-+

-+-+

i.e. an isometry preserves the length of a vector and the angle between f is a conformal map then

2 -+-+ * -+... -+-+ n (p)!1 1 (u,v) = f !12(u,v) = !12(f*u,f*v) so this time the inner product between two tangent venctors is rescaled.

Evidently the lengths are no longer preserved, but angles are since Cos and this is

(ii,v)

=
rescaling. Thus a conformal map

preserves the angles between arbitrary tangent vectors. But the converse also holds, i.e. any angle preserving map is a conformal map. In fact we can even show
Lemma 2 A regular map, f: M - N, Ves the right angles.

is conformal i f and only i f it preser-

Proof: Clearly it suffices to show that if two Euclidean metrics determine the same right angles, then they are propcrtional. Let

be a

532 fixed vector. Any vector posant parallel to sant orthogonal to


~ ~.

II

now has a unique decomposition in a com-

and a compoObserve that

this decomposition is independent of the choice of metric because they detennine the same right angles. It is well-

known that the parallel compos ant is given by

... a
g' (b,~)
g' (b,~)
g (b,~)
(['

Fig. 200

We thus get

g' (~,~)
Thus the ratio

Le.

(~,~)

g (~,~)

g'
g

(b,i::) (b,i::)
....
a, i.e. the metrics g and g' are

is independent of proportional.

b.

Since the metrics g and g' are symmetric coten-

sors it must also be independent of

Then we consider manifolds with Minkowski metrics. Now angles are no longer well-defined objects, but observe that the null-cone structure is preserved by a conformal map, i.e. time-like vectors are mapped into time-like vectors, null-vectors into null-vectors and spacelike vectors into space-like vectors. This time we can then show Lemma 3 A regular map, f: M ves the light-cone structure. Proof: Clearly it suffices to show that if two Minkowski metrics generate the same light-cone structure, they are proportional. Let time-like vector and
~ ~

N,

is conformal i f and only i f it preser-

a space-like vector. Then the line A1 A2

a+

be a
A~

intersects the light-cone in two different pOints corresponding to the two values A2 On the other hand solves the equation and A1 and

533

II

o = g (~+Ab, -:;'+Ab) = g
positive and
g (b,b)

(-:;.,-:;.) + 2Ag (-:;',b) + A 2g (t,b)

(Observe that this actually has two roots, since


II (~,~) -+ -+ g (b,b)

g (a,a)

......

is strictly

is strictly negative). We then get

,.,
"1

"2

and But here 1.1 conclude that


II

1.2

are independent of the choice of metric so we

g'

1(-:;',-:;') (b,b)

Le.

g I (c,c)/g

(c,c)
k ......

is independent of

... c.
I

Let us put it equal to a


(~,~)}

constant k. g I (u,~)

Then we finally get O;{g I (u+~, 'll+~) - g I ('ll,'ll) - g

= ~g
Exercise 10.3.1

(u+v, u+v) - g (u,u) - g (~,~)}

-+-+

-+-+

= kl

('ll,~)

Introduction: Let RP+q be the pseudo-cartesian space equipped with the metric i i i i

A dilatation is a map represented by An inversion is a map represenr,ed by

y
Y

AX

A translation is a map represented by T Problem:

<xl x';> xi + a i

(a) Show that a dilatation is a conformal map with the conformal factor

n 2 (x) =
formal factor

1.

(b) Show that an inversion is a conformal map with the con-

1
---

(x) = <xlx>2 (Strictly speaking we must restrict ourselves to the manifold M = R1 {xl<xlx> ,. O} since the inversion breaks down on the "cone" <xix> = 0).

.>+,

(c) Show that the. transformation

C = ITr is given by i x1+a 1 <xlx> y 1+2<alx>+<ala><xlx> (which is strictly speaking only well-defined on the manifold RP+~{xI1+2<alx>+ <ala><xlx> # O}). Show furthermore that it is a conformal map with the conformal factor (1 + The transformation C transformation. (Hint: conformal maps f and g formal factor given by

n2 (x)

2<alx> + <ala><xlx2 is called a special conformal Show that the composite of two is again conformal with the con-

534

II

Wor'ked exercise 10. 3 . 2


Problem: Show that the stereographic projection from the sphere onto the plane rr:S2'{N}~ R2, is an orientation reversing conformal map with the conformal factor
n2(8,(j))

=~ 2
Sin

where

is the polar angle.

Okay, by now you should feel comfortable apout isometries and conformal maps. We proceed to investigate various concepts which can be derived from the metric. n be a manifold with metric Let M
g

Then we have previously in(See

troduced an equivalence relation between tensors of various types.

section 6.9) In coordinates this corresponds to the raising and lowering of indices using the components of the metric tensor. Thus the cotensor

T~j = g~ Tki etc. Now when we use a diffeomorphism to transport tensors from one manifold to another we should be careful. Suppose f:(Mn,gl) ~ (N n ,g2)
Tij is equivalent to the mixed tensor with components is a diffeomorphism and that then there is no reason why sors on T f.T and and T' are equivalent tensors on f* commutes

'k

f*T'

should be equivalent ten-

(with respect to

g2)

But we know that

with tensor products and contractions. If for instance

we therefore get (f*T') j Consequently we conclude


i

Lemma 4 Suppose f:(Mn,glJ + (N n ,g2J is a diffeomorphism. If T and T' are equivalent tensors on M with respect to gl, then f*T and
f*T'

are equivalent on

with respect to
f

f*gl

Thus we see that unless

is an isometry it will not respect the gl and

equivalence relations induced by the initial metrics M and N. scalar field


~

g2

on

This observation is of vital importance in physics. Consider a and let

(M,g)

be the Euclidian space

R3

(with the

standard metric) representing the physical space. The energy density is then given by

~ai~

ai

~
metric~

But here we have used the equivalence relation induced by the From the field itself we can only construct the covector
d~

with the

535 components

II components

di~'

So when we use the contravariant

di~

it is implicitly understood in this expression that we have used the

metric components to raise the index. Thus it would be more correct to write out the energy-density as
H

= ~gijd.~d.~ l. J

This is a very common situation in physics: Indices are contracted using the metric. Now observe that the metric is a fixed geometrical quantity. It is physically measurable, e.g. the arclength of a curve in a physical space can be measured with great accuracy in the laboratory. We are not free to exchange this metric if we want to compare the predictions of the theory with the experimental results. Suppose now that we have been given a diffeomorphism of space into itself f : R3 ~ R3 Then we can investigate the transformed field configuration formed field configurations at corresponding points and But they are only identical if f is an isometry since the metric is
~'

f*~

E.g. we can compare the energy densities of the original and the trans-

fixed. This distinguishes the isometries from a physical point of view: They leave various physical quantities invariant, i.e. they act as
symmetry transformations.

We have previously discussed time-like geodesics on a manifold with Minkowski metric. (See sec. 6.7). We will now extend the concept of a geodesic. Motivated by the discussion in section 6.7 we define
Definition 9 A geodesic on a manifold paramatrized by
(6.49)
x~
M

with a Minkowski metric

is a curve

= X~(A)
~ +

which satisfies the geodesic equation

d2x~

r~

~i3 dA

dx~ dx 8 = 0
dA
~

The parameter fine parameter.

A involved in the geodesic equation is called an

Observe that the affine parameter is only determined up to an affine parameter shift

536

II

A = as+b
new parameter at
P Then

(aiO)

i.e. a geodesic satisfies the same equation (6.49) with respect to the s .
P

Consider a point

in

M and a non-trivial tangent vector

-+

vp
P

there is exactZy one geodesic which passes through

and has tangent vector

~p,

To see this we introduce coordinates

around

P.

x~

and

~P

by

In these coordinates P is represented by coordinates ai A geodesic through P .with tangent vector d2xi
j

then has to satisfy the geodesic equation

-a-vwith the boundary conditions

i
jk

"d."I""d."I"

dx

dxk

and But this second order differential equation has a unique solution by the well-known uniqueness and existence theorem for ordinary differential equations. If we perform an affine parameter shift

as+b

(afO) vp
-+

the tangent vector placed by


P

is re-

a~p

so we have acP

tually shown that to each point and each direction at correspond exactly one geodesic.
Worked exercise 10.3.3
Problem: (a) Let r be a curve on M parametrized by xCJ. = x(l(s). Show that is a geodesic ir and only if it satisfies an equation of the rorm

(10.25)

A(s)

a::

(b) Show that the corresponding arrine parameter


(10.26) A

J:

A is given by

expU:2A(SI)dSl ]dS 2

(up to a linear change or the parameter)

The affine parameter is not only characterized by the simplest possible form of the geodesic equation. It is also distinguished by the following property

537

II

Theorem 9 Let vector


A

be an affine parameter on a geodesic. Then the tangent has constant length, i. e.


(1(v(>.) ,V(>.))

~ = ~~

is constant

along the curve.

Proof: d dx el dx i3 , dALgai3CDl CDlJ

= 2g ai3
a dX [ B ]dX" gaSCDl r II" CDl

dx ll dX" , CDl d>' J due to the identity

Here we can exchange a dx [ By ldx" gaBCDl ~g dllg"y CDl Thus we get a a 2 i3 d [ dx dXS] dX [d x d'\ gai3CDl CDl = 2g aB CDl CfP" +

ll dx dX"] II" CDl CDl

and this last expression vanishes automatically for a geodesic.

So the affine parameter is a natural parameter on a geodesic! We can now divide the geodesics on (a)
Time-~ike

M into three classes:

geodesics, where all the tangent vectors are

time-like so that (b) (c)

(The affine parameter can then be normalized dx el dxS gaSCDl CD\ = -1 ). where all the tangent vectors are null-.

Nul~-geodesics,

vectors.
Spaae-~ike

geodesics, where all the tangent vectors are

space-like. so that We can then show


Theorem 10 (1) (2)

(The affine parameter can then be normalized dxel dx S gaSCD\ CD\ = +1 )

Isometries map geodesics into geodesics and preserve the affine parameter. (This is valid fo-::' Riemannian manifolds too). ,\ is an affine Conformal maps preserve null-geodesics. If parameter on the null geodesic affine parameter on

r,

then

fn2('\)dA

is an

f(r),

where

n2

is the conformal factor.

Proof: The first proposition is almost trivial since isometries preserve arc lengths and since geodesics are in general characterized as extrernizing the arclength. Special care should however be paid to nullgeodesics, but here the result will follow from the second proposition.

538 To prove (2) it suffices to consider a single manifold consider a rescaling of the metric
g
+

II

M and to
r is a

n2 g

= g.

Now if

null-geodesic then it especially satisfies the geodesic equation (6.49) Under a rescaling

d~~

r~

~e dI" dI"

dx~ dx~

=a

ga~
the Christoffel field

g~e

n2(x)ga~

r~~e

given by (6.47) is changed into

r~~e

= r~~e

+ o~~aelnn + o~~a~lnn - g~ea~lnn

Thus in the rescaled metric we get

d2x~ -~ dx~ dx il _[d~~ ~ dx~ dX B ] dx~ d ~. dx~ dx il ~+r ~~dI"dI"- "(IT2+r ~ildI"dI" +2dI"d"lnn-[a lnnlg~~dI"dI"
Here the first term vanishes because get r is a geodesic, and the third term vanishes because the tangent vectors are null-vectors. Thus we

But according to exercise 10.3.3 this shows that parameter given by

is a null-geodesic

with respect to the rescaled metric. However it gets the new affine

~(,,)

J"exp o

[Js 2dds (lnn 2 )ds


0

]dS 2

o
M with metric
g This follows

-------

In many applications it is preferable to control the set of all possible global isometries: Consider a manifold and suppose <p is a global isometry. Then so is <p -1 immediately from the composition rule (theorem 5):

(<P-I)*g

= (<P-I)*<p*g =

[~I<p]*g

id*g

= g

By the same argument it follows that if <PI and <P2 are global isometries, then so is the composite map <P20<Pl Thus we conclude that the set of global isometries form a group, which we call the isometry group of the manifold M, and which we denote Isom[M,g] Let us now turn our attention to space time, i.e. Minkowski space. Here we can determine the isometry group explicitly. It will be preferable to restrict to an inertial frame. Then the following holds:
Theorem 11 A diffeomorphism, it is on the form,
(10.27)
<p

M ~ M, is a global isometry i f and only i f

y~

A~~x~+ba

where

(A~~)

is a Lorentz matrix.

539

II

Proof: The proposition is intuitively clear since an isometry preserves geodesics, i.e. it maps straight lines into straight lines, and thus it must be an affine map of the form

yrJ. = ACl. xi3+brJ. S


with respect to an inertial set of coordinates. From this the rest follows easily: The pulled back metric
~*g

is characterized by the components

Le.
Consequently the metric is preserved if and only if

n
i.e. if and only if
in section

= 'A+~

'A
(Compare the discussion

is a Lorentz matrix.

6 . 6)

.0
theorem 3) which has exactly the same form altheo-

The transformation (10.27) should be compared with the Poincare transformation (ch. 6 though a different meaning! The Poincare transformation (ch. 6

rem 3) was a coordinate transformation: The points are fixed while their inertial coordinates were exchanged. The new transformation (10.27), on the contrary, moves the points while keeping their inertial coordinates fixed. It is called an active Poincare transformation, We have thus shown that the isometry group of Minkowski space time is the Poincare group. If we start out with the Euclidean space n R equipped with the usual Cartesian metric, then the isometry group is called the EucZidean group of motions. In Cartesian coordinates an isometry will then be represented by a linear map, (10.28) where

y, = AX, +b, is an orthogonal matrix, i,e.

A+'A

Exercise 10.3.4 Problem: Consider the unit sphere 52 in R3 . (a) Show that an orthogonal transformation in R3 = AX, where AEO(3) 1 generates a global isometry on 52 . (b) Show that the geodesics on 52 are the great circles. (c) Let r be a geodesic and ~(r) its image under a global isometry ~. Show that the common points of r and ~ (r) must be fixpoints of ~

(d)

540 II Show that a global isometry on 52 belongs to either of the following 3 types: (1) The identity map. (2) A reflection in a plane through Origo. (3) A rotation around a line through Origo.

10.4

THE CONFORMAL GROUP


We will now study conformal transformations in the Minkowski The basic "defect" of conformal transformations Nevertheless one can generate them The linear repre-

space in more detail.

in Minkowski space, as compared with isometries, is that they do not act as linear transformations. frc:rn linear transfonnations in a higher dirnens ional space. topic of this section. pseudo-Cartesian space duct: It turns out to be instructive to consider the general case of a q RP x R equipped with the standard inner pro-

sentation of the group of conformal transformations will be the main

Naively we should only consider conformal transformations, which are smooth everywhere. This turns out to be too restrictive. In that case the only non-trivial conformal transformations are the dilations. We will therefore admit conformal maps which break down at a suitable subset. The standard example of such a conformal transformation is the inversion: (10.29) (cf. exercise 10.3.1). origin: <xix> It breaks down at the null-cone through the 2 y2 = xit furthermore In some sense the inversion therefore We would

= O.

From the relation

follows that the closer a pOint is at the null-cone through the origin, the farther away the image is. maps the entire null-cone through the origin to infinity. cone at infinity".

therefore like to enlarge the pseudo-Cartesian space by adding a "nullIn the enlarged space the inversion will then be a diffeomorphism, which exchanges the null-cone through the origin with the null-cone at infinity.

541

II

To get an idea of what that actually means let us take a quick look at the Euclidean space Rn. In this case the "null-cone" is just the origin itself. We must therefore add a single point to Rn, so that the inversion can exchange thi~ point with the origin. For this purpose we consider the unit sphere Sn in Rn 1 Using a stereographic projection from the north pole we can then map Rn onto Sn - {N}.
N

ints from infinit

p-axis
Fig. 203

P2= Cote

n Thus we can 'replace R by Sn - {N}. The north pole now represents the point at infinity. Furthermore the stereographic projection is a conformal map (cf. exercise 10.3.2), i.e. the metric on the sphere and the transferred metric from Rn are conformally related. From the point of view of investigating conformal transformations we can therefore equally as well work on Sn. Now suppose we identify the Euclidean space Rn with Sn - {N}. What does the inversion look like on the sphere? As indicated on figure 203 it is easy to see that the inversion corresponds to a reflection in the equator. On the sphere the inversion is thus a nice diffeomorphism, which exchanges the south pole (i.e. the origin) with the north pole (i.e. the point at infinity). Remark: The sphere Sn is a compact manifold. By adding a point at infinity we have therefore compactified the plane Rn. In the mathematical oriented literature, the sphere is therefore called the one-point compactification of the Euclidean space.

space

Motivated by this example we now return to the pseudo-Cartesian RP x Rq . First we enlarge the pseudo-Cartesian space by adding One time like, labelled u, and one spacelike,

two extra dimensions:

labelled v. In this way we obtain the pseudo-Cartesian space R1 + p x Rq + 1 in which a typical point will be denoted w = (Uix1, ... ,xp;y1, ... ,yq;v) =
(UiZCliV)

The goal is to embed the pseudo-Cartesian space RP x Rq as a suit1 able subset of R + p x Rq + 1 (similar to the embedding of the Euclidn ean space R as a sphere Sn inside Rn + 1 ). This will be done in a tricky manner! In the enlarged space R1 + p x Rq + 1 we introduce the null-cone through the origin:

K.

In the first step we then embed

542

II

P R

q R

isometrically as a subset of

K.

Define

section of the null-cone the bijective map

~
(

RP

and the hyperplane q x R ~ M constructed in the following way


Z ; <zlz> 2
1)

M to be the interu-v = 1, and consider

~ (z)

<ZIZ> + 1 2

By construction

is an isometry.

To see this we observe that system on


M

actually generates a global cordinate vectors are given by

The basic frame

Consequently the metric coefficients of the induced metric on duces to __ ab ++ ab ++ gab = <eale b > = (+)z Z + <ealeb>(!)z Z = <ealeb> Since the embedding is an isometry we can simply identify

M re-

RP

Rq

with this particular section M, which henceforth will be denoted q M(R P x R ). q ) will ultimately be replaced by another The section M(R P x R section of the null-cone cone
K
A

K, but before we proceed with the construc-

tion we must take a closer look at the conformal structure of the nullgenerator of
~
=

K is a null vector

(u,i",v)
~~

It generates a line
R,+

on

K,
(\(+00

= \(u,t,v)

called a characteristic line. A given characteristic line !~ will intersect the section M(R P x Rq ) at most once, and there are characteristic lines which do not intersect M(R P x Rq ) at all. They correspond to the lines which are parallel to the hyperplane they are generated by null-vectors are precisely the null-vectors where u-v = 1, i.e. u=v. But these

~ = (u,t,v), where

Thus there is a one-toone correspondence between characteristic lines missing M(R P x Rq ) Cartesian space and points on the null-cone through the origin in the original pseudoRP x Rq (cf. fig. 204 a). Consequently the exception-

t2

= O.

al lines represents points on the null-cone at infinity! Consider now two local sections N2 on the null-cone K. N1 and Suppose furthermore that the characteristic lines intersect N1 and

N2

at most once.

Then we have a natural map


11 :

N1

N2

543

II

Fig. 204a

Fig. 204b

obtained by projection along the eharacteristic lines (fig. 204b). The basic observation is the following one:
Lemma 5 The projection along characteristic lines, conformal map. Proof: . d uce new coord lnates In R'+p x Rq+' ,. W e nee d two radial It is preferable to lntro variables r"r Z and p+q homogeneous variables a' , ... ,eP,.p , . .pq: r,
Z
IT:

Nl

N , 2

is a

z Z u + (x') + z
+ (y' )

...

+ (xP ) + (yq)

8'

x'/u, .

,a P

xp/u yq/v

rZ Z

z
+

...

.p' = y'/v, . ,q

(As usual we have troubles with the homogeneous coordinates, which break down when u=O or v=O. Since we are only interested in a local result, we will simply assume that u=O and v=O does not intersect N, or N ). The null-cone K is then Z characterized by the equation If we put the common value equal to r we can introduce the following intrinsic coordinates, (r,e', .. ,eP,.p', ,.pq) on the null-cone K. A characteristic line is then given by a fixed set of the homogeneous coordinates. Furthermore the local sections can be parametrized as follows:
r = f

(e ' , .. , eP , 1 , ,.pq)

r = g(e' ,

,a P ,' , . ,q)

We can therefore use (e', ,eP ,' , . ,.pq) as intrinsic coordinates on N, and HZ. With this choice of coordinates the projection map IT is simply represented by the identity map!

544

II

We must now determine the various mrtrics inyolved in the game. Consider first q the complete pseudo-Cartesian space R +P x R + with the canonical metric de 2
= -

du

1 2 p 2 1 2 q 2 + dv2 - (dx) - - (dx) + (dy) + + (dy )

In the new coordinates this is reexpressed as

The null-cone
ds

K, where
2

= r 2 = r,
. .

now gets the induced metric:


kl

IK

r [ - e ij (6)d8~d8J + d
N1
C

(<jd<j> d<j> ]

Finally the two sections


N1

and ij

N2

are equipped with the induced metrics

dS~
dS~
=

f2(8,<j[ -

j (e)de i d6 + dkl(<j>ld<j>kd<j>l] j + dkl(<jd<j>kd<j>l]

HZ

g2(6,8)[ - C (e)d8 i d8 ij

Consequently

i.e.

Exercise 10.4.1 q RP x R as the interIntroduction: In the above discussion we have embedded 1 using the bijective section M1 of the null-cone K and the hyperplane u-v map <z1 Iz 1>+1 <j>1 (z1) = ( 2
q We could equally well have embedded RP x R as the intersection COlle Jl and the hyperplane u+v = 1 using the bijective map 1 + <z2Iz2> 1 - <z2Iz2> $2 (z2) = ( 2 z2 2) Problem: Show that the projection along the characteristic lines,
1T:

of the null-

M1 '" M2

'

corresponds to an inversion in ' RP x Rq

Having obtained this len'ma we can now apply it to project q M(R P x R ) into a suitable section of K (analogous to the sphere in the Euclidean case). This new section should then include the "nullcone at infinity", i.e. each characteristic line should intersect it exactly once. Unfortunately we run into a slight technical problem: In general it is not possible to find a single section, which is intersected exactly once by each characteristic line. We shall therefore adopt the following strategy:

Let r and 1 Cartesian space


r 1

545 II denote the radial variables in the complete pseudo2 1 p q +1 : R+ x R

2 =u2 12 + + (x p2 + (x) )

Denote by N the intersection of the null-cone the hypersphere

Clearly straint

N is a submanifold defined by the following equations of conr


1
~

Topologically N is therefore a product of the two unit spheres sP q + 1 Consequently N is a hyper-torus: R in R1+ p and sq in

sP

sq

The hyper-torus sP x sq is a nice section on K , but each characteristic line will actually intersect it twice in antipodal points (see fig. 205). Consequently each point in the originaZ pseudo-Cartesian space

P R

q R

is represented by a pair of antipodaZ points on

sP

sq.

Fig. 205

q If follows from lemma 5 that the projection map, ~: M(R P x R ) ~ sP x sq, is a conformal map. From the pOint of view of investigating conformal transformations we can therefore equally as well work on sP x sq, except that we must restrict ourselves to transformations which maps pairs of antipodal points into pairs of antipodal points. q P x R In coordinates the projection from sP x sq to R is given by
~(u,z]l,v)

u-v

546 consequently the pOints where u and v

II

coincide are "sent to infin-

ity". But according to our previous analysis these points are in a one-to-one correspondence with the null-cone in RP x Rq . Thus q by adding a "cone at infinity". P x R sP x sq is obtained from R Because sP x sq is a compact subset of R1 + p x Rq + 1 it is often q P x R referred to as the confopmal compactification of R
q Consider once more the Euclidean space R . In this case the conformal compactq ification reduces to soxS , but the zepa-dimensional sphepe SO only consists of two points, u = l. In this case (and only in this case!) the conformal compactification therefore breaks up into two disconnected components {+l} x sq and {-1} x sq Thus we need not double count the points in the conformal compactification because we can cimply throwaway the component {-l} x sq ! But then it is superfluous to enq large the space with f time-like coordinate, i.e. we simply enlarge R to the q Euclidean space R +, and use the unit sphere sq in this enlarged space as the conformal compactification. It is now easy to show that the point at infinity corresponds to the north pole and that the projection along the characteristic lines corresponds to the stereographic projection (cf. fig. 206). Thus we have a nice simplified picture in the Euclidean case.

Conformal compactification of the line.

Fig. 206 Using the conformal compactification of the pseudo-Cartesian space q it is easy to construct conformal transformations on RP x Rq x R O(1+p; q+1) consisting of pseudo-orthogo1 R + p x Rq + 1 . Each matrix in O(1+Piq+1) sP
x

RP

Consider the matrix group nal matrices operating on

generates a conformal transformation on

sq

in the following way:

Notice first that a pseudo-orthogonal matrix S preserves the inner product in R1 +p x Rq + 1 Especially it maps the null-cone into itself. Consequently it maps the hyper-torus S[Sp
x

sP

sq

isometpi-

caZZy onto a new subset,

sql, of

K.

To get back to the hyperIn this way we

torus we then project along the characteristic lines!

547 II have constructed a mapping of the hypertorus into itself which we denote by
1[(S']:

Alternatively we can describe 1[[8] in the following way: Each pair sP x sq generates a unique charac{p , -p}, on of antipodal pOints, teristic line formation line and ~p

S
sq.

.9- p on the null-cone K The pseudo-orthogonal transmaps this characteristic line into another characteristic

l'.S(P)
x

The image of

is then the intersection between

.9- S (P)

According to lemma 5 the combined transformation

1[[S]

is now a conformal transformation of hyper-torus into itself. From the construction follows immediately some basic properties of
1[[S].

First it maps a pair of antipodal pOints into a pair of anti-

on

podal pOints, i.e. it can also be considered a conformal transformation RP x Rq . Next the assignment of a conformal transformation to each pseudo-orthogonal transformation constitutes a representation of the pseudo-orthogonal group, i.e.
= = 1[[ S2S1]
-

= 1[[ S2]

Notice, however, that the correspondence between pseudo-orthogonal transformations in O(1+p;q+1) and conformal transformations in q q RP x R is not one-to-one. This is because a point in RP x R responds to a pair of antipodal points on rate the identity. matrix: sP
x

cor-

sq.

A pseudo-ortho-

gonal matrix which interchanges antipodal points will therefore geneThere is precisely one such pseudo-orthogonal

T.

It follows that each conformal transformation is in fact

generated by a pair of pseudo-orthogonal matrices mations by matrices! The conformal transformations generated from

{S, -S}.

This is

the price we have to pay when we want to represent conformal transforO(1+p,q+1) evidently C(p,q).

constitute a group, known as the conformaZ group and denoted by


In anaZogy with the conformaL compactifioation of now represent each conformaL transformation in "antipodaL" matrices in O(l+p,q+l).

RP

Rq

we can

C(p,qJ

by a pair of

Let us investigate the structure of the conformal group a little closer.

548

II

Suppose that the pseudo-orthogonal transformation S actually preserves the additional coordinates (u,v), i.e. that it is on the form

o
(10.30a)

5 o
~

s E O(p,q)

Then the corresponding transformation on ZCl


RPxR q

RP

Rq

reduces to

(u; (u-v) zCl ;v)


sP x sq

~ '"

(u; (u-v) s (3z (3 IV)


CI

sP x sq

Consequently the conformal group

C.(p,g)

contains the group of

origin preserving isometries 0 (p,g). In the remaining investigation it suffices to consider pseudo-ortho-

gonal matrices of the form

*
S

[:

:]
+ --2-

They will be divided into four general types <c Ic> c\l I c\} c\} I
1 -

-2-

<clc>

(10.30b)

(I)

C].!
-2-'-

<cle>

-<clc> c" 1
2 -2-

(10.30c)

Ii (A)

-sinhA

1 + --2-

<clc> ].!

<clc>

coshA

- -r-

c <clc>

-ev

-2-

c" <clc> 1

Each of these types constitute a representation of a particular subgroup of e(p,g), i.e.

The first type,


(10.29)

S(I), represents the inversion ].! z


~

z].! <zlz>

The second type,


(10.31)

T(c].!), represents the group Of translations z]'!


~

z].!+ c].!

The third type,


(10.32)

D(A), represents the group of diZatations z].! ~ eAz ll

549

II

Finally the fourth type, formal translations

C(c~), represents the group of special con-

(10.33)

Let uS check the last of these statements just to illustrate the principle: A point z~ in RP x Rq is represented by the point [u,iu-v)z~;v] on the hypertorus sP x sq. By the pseudo-orthogonal transformation C(c~) this is mapped into the point [u ' ; z'~;v'] where u' <clc> (u+v) u + <clz>(u-v ) + ---2--

z,1l = (u-v)z~ + c ll (u+v) v' In P R


x

<clc> v - <clz>(u-v ) - ---2-- (u+v)


(u-v)z~ + c~(u+v) (u-v) + 2<clz>(u-v) + <clc>(u+v)
z~ + cU~

q R this corresponds to the point


z,~
~,

u-v

+ 2<clz> + <clc> u+v

u-v

It remains to determine the factor (u+v)/(u-v). Since on the hyper-torus, it follows that u2 + (u_v)2 x 2 = v 2 + (u_v)2y2 From this relation we especially get u 2 _ v2 = (u-vt{y2-x 2 )
i.e.

[u; (u-v)z~; v]

is a point

<zlz>

----2

u2 _v 2

(u-v) Inserting this the induced mapping from


z~ ~

=-;:;::v
q P x R R into itself finally reduces to

u+v

z~ + c~<z Iz> 1 + 2<clz> + <clc><zlz>

We can also summarize the above findings in the following way:


Pseudo-orthogonal transformations in D1+P x Dq+l: Transformations preserving both and v Transformations preserving Transformations preserving u-v u+v u conformal transformations in P R x

Pseu~o-or~ho&onal ~ransformations: ,"S" E O(p,q) z ~ S SZ

Translations:

zCl

zCl + c Ci

Special conformal transformations: CI zCl + cCl<zlz> z ~ 1 + 2<cl z> + <cl c><zl z> Dilatations: Inversion
z CI _ eA z (l
z~

Transformations in the u-v-plane Reflection in the u-z-hyper-plane

zCI ~ <zlz> .

550 the conformal group zation: C(p,q).

II

By now you should feel comfortable about the general structure of It has the following simple characteri-

Lemma 6 On a pseudo-Cartesian space

P R

Rq

the conformal group,

C(p,qJ,

is the smaZZest group containing the isometry group and the inversion.
Proof: First we make the trivial observation that we can generate special conformal transformations using only translations and the inversion. This is due to the identity

(cf. exercise 10.3.1). dilatations.

It remains to show that we can also generate

But that follows from the identity

a 1 = _c = a = a = _c a D(1+<clc= T(1+<cicC(c )T(c )C(1+<CIC

(The verification is rather tedious but straight forward. quickest way to verify it is to show that
_c a a a 1 _c a

Perhaps the

[T(1+<clcC(c )T(C )] (x) = [D(i+<clcC(1+<Clc>lJ (x) where we work directly with the conformal transformations on P R
x

Rq).O

We conclude this section with a few remarks about whether or not C(p.q) in fact contains all possible conformal transformations. The answer is affirmative if the dimension of RP x Rq is greater than two, though we will not prove it here. The two-dimensional case is an exceptional case because any holomorphic map of the complex plane into itself is in fact conformal. This will be discussed in section 10.6. Never-the-Iess C(2) still depotes the conformal group constructed aC(2) is essentially the Lorentz group bove. Notice that

0(1.3)!

Exercise 10.4.2 Problem a) Show that the conformal group C(2) of the complex plane into itself:
w = yz + 0

consists of Mobious transformations


(ao

az + S

and

w =

yz +

~ 0

- Sy

1)

551
b) In the following we represent the Mobius transformation

II

w(z)

(az + S)/(yz + 8)

a r Ly ~]

by the 2x2-complex matrix:

Show that the matrix group SL(2,C) (consisting of 2 x 2 complex matrices with determinant 1) constitute a prepresentation of C (2) (consisting o of orientation preserving Mobius transformations).

10.5 THE DUAL MAP


It still remains to investigate that part of the exterior calculus which also depends on a metric (and orientation), i.e., the dual map and its related concepts, like the inner product between differential forms, the co-differential 5 and the Laplace-Beltrami operator
~.

The dual map plays a special role, because it is connected to the dimension of the manifold. Let a dual map * Then * a differential form of rank manifolds with dual maps *1

M be a manifold of dimension
Okay, suppose *2 and let and f *1(f*T)
M M

m k
N

with into

converts a differential form of rank (m-k) and m


+

and n
N

are

be a smooth but they


N

map. Then we would like to compare can only have the same rank if the same dimension with a metric
g

f*(*2T)
M

M and

are of the same dimension! and have

For this reason we will assume in what follows that n. To begin with we let

be an n-dimensional orientable manifold

We will of course assume that f: M


g
~

M is an n-dimensio-

nal orientable manifold, but a priori we do not assign any metric to

M.

We will let

be a smooth regular map. Then we can pull f*g on

back the metric map *f on


12

to a metric

(cf. exercise 10.2.2).


E

Equipped with this metric we have a Levi-Civita form

M.

f We can then show the following theorem:

and a dual

Th~or~m

Let th~ regular map sing). Then


(10.34)

Mn ~ Nn

be orientation preserving (rever-

and

commutes

f*E = Ef

(anti-commut~s)

(f*E = -E ) f
with

th~

dual map, i.e.

(10.35)

Proof: It is sufficient to check the case where tion. Let P be a point in f preserved the orienta-

As of P

is regular, we know that it

maps an open neighbourhood neighbourhood

diffeomorphically onto an open

of

Q in

552

II

Fig. 207 We may safely assume that ted coordinate system

is in the range of some positively orienBut then we can use


[fo~,Ol

(~,O)

as a f . yi

positively oriented coordinate system covering in these coordinates, then ponents are unchanged f

V.

If we express

reduces to the identity map:

= xi

But that is not all! When we pull back the metric, then the metric com-

as are the components of any differential form identify them using the map f.

we might like to

pull back. Thus locally the two manifolds are indistinguishable if we

We now address ourselves to the more realistic situation where we have been given two orientable manifolds with metrics and If f: M *1
~

N
and

is a smooth regular map, then we want to compare the *2 on f*g2

dual maps

M and
on
~nd

N.
M
g2

But observe that

g2

is

pulled back to a metric dual maps generated by two metrics


gl

f*g2

and the relationship between the are completely controlled by

theorem 12. We are thus really left with the problem of comparing the and f*g2 This reduces the problem to the following: We consider just a single manifold

and see what happens with

the dual map (the co-differential, the inner product etc.) when we exchange one metric with another. Of course everything turns out to be simplest when we do not exchange the metric at all! When we work with isometries we can therefore immediately take over the results obtained in theorem 12. Combining this with theorem 7 and 8 we then get

553
Theorem 13 Let (a)
(10.35)
(b)

II

f : M f f

be an orientation preserving local isometry. Then

commutes with the dual map, i.e.

*l [f*T] = 1*[*2T].
commutes with the co-differential and the Laplace-Beltrami 5f*T If f operator, i.e.

(10.37 )

f*&T

hf*T = f*hT.

(c)

is a global isometry it preserves the inner products <f*Tlf*s>

= <Tis>
f is a conformal map. It

Almost equally simple is the case when

will be preferable to consider at first a single manifold what happens if we perform a rescaling of the metric.
Theorem 14 Let (a)
(10.38)

and see

g
E

and n M and

g' = Q2(X)g Then


E'

be conformally related metrics on the

manifold

are conformal ly related:


E' = Qn(x)t

(b)
(10.39 )

Let

T be a differential form of rank


are conforma l ly related: *'T = Qn-2k(x)*T

Then

* 'T

and

*T

Proof:
(a)

is an g' (x)

nxn

matrix. Thus we have:

DetC'

Det[Q2(x)C]

Q,n(x)DetG

from which we immediately get

RlXT
(b)

Ea

"'an

When we compute the components of

*T we must raise the index.


gij.

This involves the contravariant components of the metric, l But they are the components of the reciprocal matrix

G-

Therefore they transform with the reciprocal factor: g,ij


1

Q-2(x)gij
blCl bkc k

Using this we immediately get

k ;gr('XT
Qn-2

E al a 1 k.

~k

bl b

g'

'g'
g
blCI

T c c

k
T

(x)-, Ig (x)

al"

'an - k b l "b k

"'g

bkc k

Cl"

'ck U

554 Combining the results of theorem 12 and 14 we then finally obtain:

II

Theorem 15 Suppose f*g2


(a)

(Mn,gl)

and Then

(N n ,g2)

are orientable manifolds. Let

be an orientation preserving (reversing) conformal map, so that

,,2 {xl gl

(10.40) (10.41)

f*[E21 = "n{X)El f*[*2T1


n

(f*E2

-"n(X)El) -"
n

(b)

"n-2k(X)*1[f*T1
is even and

(f*[*2TJ
k

n-2k

(x)*1[f*T1) ac-

Observe that if

:r

has rank

"2

then

tually commutes (anti-commutes) with the dual map

(10.42)

f*[*2TJ = *df*TJ

{f*[*2T1 = -*df*Tll

10.6 THE SELF-DUALITY EQUATION


We have previously introduced the concept of self-dual and antiself-dual forms (see ex. 7.5.3 and ex. 7.6.9). They can only be constructed in spaces of even dimension and they satisfy the equation (10. 4 3) rank T = ~ dim M *T = AT but due to the identity (7.43) A is constrained to very special values
i.e.
A = {

**T -T Observe that in the latter case the self-duality equation only works
if

l i

if

**T

for compZex valued differential forms. For future reference we collect the various possibilities in the following scheme:

(10.44)
Dimension
2 4

,
T I-form 2-form

The self-duality equations Minkowski metric *T = T *T = i T

Euclidean metric *T = tiT *T = T

Problem:

E:nercise 10.6.1 (a) Let n = dim M be even and let rank T = ~. Show that T can be decomposed uniquely into a self-dual and an anti-self-dual part
(b)

T=T +T. Show that the above decomposition is an orthogonal decomposition, i.e. (T+ I T-)= 0, in the case of a Riemannian metric.

555

II

Next we observe that the self-duality equation has the following important invariance property:

Theorem 16 The self-duality equation is conformally invariant.


Proof: or anti-self-dual then

T is a self-dual n rank T = '2 so that a conformal map commutes (anti-commutes) with the dual map.
This is an immediate consequence of (10.42). If dim M

=n

is even and

'0

Observe that an orientation reversing map actually maps a solution to the self-duality equation

*T

AT

into a solution of

*T

= -AT

The self-duality equation is by far the most important example of a conformally invariant equation and it comprises many well known partial differential equations in disguise. To get familiar with it we will study in some detail the historically most famous self-duality equation: ILLUSTRATIVE EXAMPLE: THE CAUCHY-RIEMANN EQUATIONS.
Here we consider maps from

R2

into itself:
, 1.e. {WI
W2

= w2 (x ,y )} = WI(X,y)

Such a map generates a complex valued l-form dw = dWI + idwt , and we will demand this to be anti-self-dual *dw = -idw If you write out the components you will immediately find that this is equivalent to the following partial differential equations
(10. 4 5)

2..:'i.=
dY

_~ dX

which are nothing but the famous Cauchy-Riemann equations! Here it is preferable to introduce complex numbers, i.e. to identify in the canonical fashion and consider complex functions
i.e.
dW dWz a?lz 3zd
w = w( z)

R2

with

Such a complex function generates the complex-valued l-form dw = + (For the notation you should consult exercise 7 .6.9and 7.4.3~ But here dz is antiself-dual and d~ is self-dual, so this is precisely the decomposition of dw into a self-dual and an anti-self-dual ~art. (Compare with exercise 10.6.1) But then dw is anti-self-dual if and only if ~ = 0 which is the complex form of the CauchYZ Riemann equations. To summarize we have thus shown (compare with exercise 7.6.9) :

Lermza 7 A complex function w = 'W( zJ is holomorphic (anti-holomorphic) if and only if it generates an anti-sel/duaZ (self-dual) l-form dw.

556

II

Let us now characterize the holomorphic fUnctions geometrically. Here we have the following well-known characterization:

Theor-em 17 A non-t1'iviaZ compZex funation w = w( z} is hoZomorphia (anti-hoZomo:rphia) if and onZy if it is an o1'ientation pr-eserving (or-ientation r-eversing) aonfoT'mal map.
Proof: We know that the Cauchy-Riemanns equation are con formally invariant, i.e. if w is any holomorphic fUnction and f: C ~ C is an orientation preserving conformal map then w( f( z) ) is a holomorphic fUnction too. Applying this to the identity function w( z) = z which is trivially holomorphic, we see that f is holomorphic too. To prove the remaining part of theorem 17 we consider the standard metric on C given by

(10.46)

iT

= cfx!sdx

+ dyedy

= ~(d2PIiz

+ d7M.z)

If f: C ~ C is any smooth map represented by the complex function therefore obtain the following pulled back metric
f*g =

= f(z)

we

1f (c/.w0!i + IhQdw)
+
1 dW 2

dW 2 = (ldZI

dZ I ) g +

d d 3W (az -az)dzxdz + (a;: W W

dW - d:z)dv<dz

When

f
f*g

is holomorphic (anti-holomorphic) this reduces to

= 1a;: 1

dW 2

(f*g =

1 ~zl

'w

g)

which shows that in both cases f is a conformal map! (Strictly speaking we should cut out the points where w is stationary, i. e. 3w = o. They correspond to the points where the map is not regular, and for a non~ivial holomorphic function they are isolated). n

One of the reasons for the importance of the Cauchy-Riemann equation to physics, lies in the fact that any solution to the Cauchy-Riemann equations is automatically a harmonic function, i.e. a solution to the Laplace equation in 2 dimensions:
(10.47) d*dq,

Le.

To see this well-known result in our new formalism we let w =w(z) be a holc:rrorphic function. Then rlw is anti-self-dual and we get d*dw = d[-idw) = -id", = 0 Thus the real and imaginary parts Wl and W2 are harmonic functions. (The same argument holds for anti-holomorphic functions.) This is a common feature in many applications of the self-duality equations. They are brought into the game to simplify the search for a solution to a second order equation. In this case it is the Laplace equation
(10.47) d*dq,

557

II

which is trivially solved by solutions to the self-duality equations.

(10.48)
thus easier to handle!

*dcp

idcp

But the self-duality equations are only first order equations, and are Interestingly enough it turns out in many applications that the "relevant" solutions to the second order equations automatically solves the first order equations! ("Relevant" means solutions that satisfy appropriate boundary conditions, integrability conditions or other conditions imposed by the problem at hand.)
Remark: In the example concerning the Laplace equation this is precisely known to be the case, i.e. not only is the real part of a holomorphic function automatically a harmonic function, but all harmonic functions are generated this way. Let us also take a look at this well known result in our new formalism: Suppose cp is a real harmonic function and take a look at the complex valued l-form
2

l1 az d z
cp

It is necessarily anti-self-dual and because

is harmonic it is a closed form

d(2 l1 ~P dZAdz = 0 az dz) = 2 azaz


But then it is locally generated by a complex valued function f,

(* )

df

= 2l1 az dz '
cp

and by lemma 7 this function is holomorphic. It remains to be shown that real part of f . Now observe that since cp is real we have

is the

(**)
Combining (*) and

df = 2#dz

an. -

(**) we now get

~d(f+f) = ~dz + ~Z = dcp


i. e.

cp

~(f+f)

(up to an irrelevant constant).

Exercise 10.6.2
Problem:
(10.49)

(a)

(The wave equation). Show, using theorem 15 directly, that the Klein-Gordon equation for a massless scalar field cp in (l+l)-dimensional space time,

d*dcp =

[(w

a2

axz-)CP

a2

0]

(b)

is conformally invariant, i.e. the wave equation is conformally invariant. Suppose cp is a massless scalar field generating a self-dual (antiselfdual) field strength:

*dcp
(c)

dcp

[*dCP = -dcp]

Show then that it automaticalJy solves the wave equation. Show that the self-duality equations reduce to,

= -

[~ = ~]

and that the complete solution is on the form,

558 <P(x,t)
(d)

II
[ <p (x, t)

= f(x-t)

f (x+t)

1,

where f is an arbitrary smooth function. Let <p be an arbitrary solution to the wave equation. Show that it is on the form
<p = <PI + <P2 where d<PI is self-dual and d<P2 is anti-self-dual. (This shows that the general solution to the wave equation is on the form

(10.50)

<P(x,t) = f(x-t) + g(x+t)

After this long digression on the Cauchy-Riemann equations we conclude with the most important example of a conformally invariant field equation in classical physics:
Theorem 18 Maxwell's equations for the electromagnetio field are oonformal invariant in (3+1)-dimen?ional spaoe-time.

Proof: This can be proved directly using theorem 15 (compare with exercise 10.6.2 and 10.6.3 below). It is however instructive to reduce the Maxwell's equations directly to the self-duality equation, although it is a little bit tricky: First we complexify the field strength, cf. exercise 9.1.3 (10.51) *F On the other hand: If
F

F = F+i*F

In this way we produce an anti-self-dual 2-form

*F-iF = -iF is an anti-self-dual 2-form then it is neces-

sarily of the form (10.51). Maxwell's equations can now be cast into the following complex form: (10.52 ) *F -iF
rtF

=0
F is

Here the second equation is invariant under arbitrary smooth transformations, while the first one is invariant under orientation preserving conformal transformations. (Under an orientation reversing map F'
i.e.

mapped into a self-dual form *F' But then

= iF

FT

is anti-self-dual and has the same real part, so the F' will still solve Maxwell's equations.)

real part of

'II:Bl'oi,se 10.6.3 Probl.em: Show, using theorem 15 directly. that the Maxwell equation for the massless gauge potential. A in (3+1 )-dimensional space time.

rl*dA = 0

[ (a )J a)J) -~ A

v (a)J A )J ) = 0 1

is conformal.l.y invariant.

559

II

ILLUSTRATIVE EXAMPLE: CONFORMAL MAPS ON THE SPHERE


In the previous example we have discussed conformal maps of the plane into itself and we have seen that they were given by the non-trivial holomorphic functions, f:C ~ C. This time we will consider conformal maps from the sphere into it'tself, f:S2~ S2. It is worth observing that this example is closely related to the preceeding example, since the stereographic projection from the sphere into the plane is itself an (orientation reversing) conformal map. (See exercise 10.3.2). You might think then that the results obtained in the previous example can be carried over trivially, i.e. the conformal maps from the plane into itself are transferred to the conformal maps from the sphere into itself ueing the conformal stereographic projection. But as we shall see this is not quite true, so let US take a closer look. Suppose we have been given a conformal map, f:C ~ C. We would like to lift it to a map 1:S2 ~S2. But here we can get into trouble, since the lifted map is apriori not well-defined at the north pole (which corresponds to infinity in C).

C
Fig. 208

g.

Consequently f can only be lifted ir it has a well-defined limit at infinity, since only in this case we can extend 1 by continuity at the north-pole. Consider then a map from the sphere into itself, g:S2~ S2. We want to project it down to a map g:C ~ C. But here we should be careful about the points which are mapped into the north pole. They give rise to singularities in the projected map In the case where g is conformal, i.e. g holomorphic (anti-holomorphic) this means that g will get poles. Combining these observations you might now guess the answer:

Theorem 19 An orientation preserving (reversing) map g:S2~ S2 only if it is projected down to an algebraic function
(10. 53 )
~here

is conformal if and

w(z) = p(z)/Q(z)

w(z)

p(z)/Q(z)

P,Q aPe arbitrary polynomials.

We now show this in detail: 2 The basic idea is to treat the sphere S as a complex manifold. We already know what a real manifold is (cf. the discussion in sec. 6.1-6.3), and the corresponding definition of a complex manifold is a straight forward ~neralization: 2n with Cn. Let U be an open subset of C and f:U ~ C a smooth We identify R function of n complex variabels: f(zl, ... ,zn). We say that f is holomorphic (i.e. complex-analytic) it either of the following two equivalent conditions are fulfilled: 1) i=1., ... ,no
2)

The function f can be expanded in a convergent power series around each point (z~, ,z~) in U:

560
n

II

f ( z 1 , ... , z )

Pl",Pn Pl",Pn

~ E

(1 1 Pl n n Pn z -z ) (z -z )
0 0

Consider furthermore a mapping ~ from an open subset U of en into e :

~(zl, ... ,zk) = (~~zl, .. ,zn); ... j(zl, .. ,zn)


We say that the mapping ~ is holomorphic provided each of its components ~l, . ,~k are holomorphic fUnctions. Suppose now that M is a real differentiable manifold of dimension 2n. A complex coordinate system on M is a smooth regalar homeomorphism ~ of an open subset U in en onto an open subset ~(U) in M. Furthermore two complex coordinate systems are compatible provided the transition functions are holomorphic maps. With these preparations we finally arrive at the main definition:

Definition 10 A complex manifold is a real differentiable manifold M equipped UJith an atlas of compatible complex coordinate systems.
Notice that in the complex case the concept of a smo6th function has been replaced by the concept of a holomorphic funtion. We can now show that the sphere 8 2 is a complex manifold. As complex coordinate systems we use stereographic projections from e into S2. It suffices to consider two such coordinate systems: one corresponding to a stereo graphic projection from the north pole and one corresponding to a stereographic projection from the south pole (cf. fig. 20 9).

-' ' : ---' /4


rm
w
/'

w-plane

"

Rew

w "'-r'"

w ~ Tg[~Jexp[-i~J

"

Standard coordinates on the Riemann sphere. z-plane Re z

z ~ Cot[~Jexp[i~J

Fig. 209
The transition function is then given by

~21(Z) = 1. z
which clearly is a holomorphic ,function: Equipped with these coovdinate systems the sphere 8 2 is thus a complex manifold known as the Riemann sphere. The complex coordinates generated by stereographic projections from the north pole and the south pole are referred to as standard coordinates. As in the real case we say that a continuous mapping, f: M ~ N, between two complex manifolds is holomorphic, provided it is represented by a holomorphic map when we int~oduce complex coordinates. We can now show, using the same arguments as in the complex plane, that a smooth mapping from thB Ri"mann sphilra int~ i tae Lf i$ confOI"maL if and only if it is holomorphic. All we need is therefore a closer examination of the holomorphic maps from the Riemann sphere into itself. Rema~kably enough we can now classify completely the holomorphic maps from the Riemann sphe~e into itself. This requires some preliminary knowledge about holomorphic functions, which can be found in any elementary textbook on complex analysis:

561

II

Suppose f is a holomorphic function with an isolated singularity at the point z (i.e. f is discontinuous at zo). Then f can be expanded in a Laurent series ar8und z : o 'i: a (z-z )n f(z)
n=-co n
0

Since f is discontinuous at z at least one of the coefficients a corresponding to a negative integer n is non~vanishing. We distinguish between twoncases: Either there exists an integer k such that a ~ 0 while a = 0 when n<-k, or there exist an infinite number of negative in~egers n such ~hat a is non-vanishing. In the first case we say that f has a pole of degree k at z ~ In the second case we say that f has an il30lated singularity at z . 0 Notice that if f has a pole of degree k at ~ then we can rewrite it on the form
o

f(Z)=~ (z-z _ r'"


o

where g(z) is holomorphic even at the point Zo ( i.e. the singularity can be removed by multiplying f with a polynomial). If, on the other hand, f has an isolated singularity at z , then the famous theorem of Weierstrass states that f maps each punctuated n~ighbourhood of Zo

into a dense subset of c. We will also need Liouville's theorem which states that a bounded holomorphic function defined on the ~hole oomplex plane is necessarily constant.
The basic properties of holomorphic maps from the Riemann sphere into itself can now be summarized in the following way:

Lemma
A

non-trivial holomorphic map properties: (a) The preimage of a point is (b) Using standard coordinates w = w(z), which has only a are poles.
Proof:
(a)

f:S2~S2 is characterized by the foll~ing


necessarily finite. it is represented by a holomorphic function, finite number of singularities, all of which

(b)

In terms of local complex coordinates the preimage of the point Wo corresponds to the zero-set of the holomorphic function w(z) - w Consequently the pre image c8nsists of isolated points. But in a compact set, like S2, a subset consisting of isolated points can only be finite. Suppose f is represented by the holomorphic function w = w(z), where we we have introduced standard coordinates by means of a stereographic projection from the north pole. A singularity of w(z) corresponds to a point z where w(z ) = ~, i.e. to a point in the preimage of the north pole. Byo(a) there egn only be a finite number of points in this preimage. Since f is continuous it maps a small neighbourhood of z into a small neighbourhood of the north pole. Due to Weierstrass theor~m it consequently cannot correspond to an essential singularity of w(z).

Using this lemma we can then easily prove the following powerfull theorem:

Theorem 20 Using st~~d coordinates a given holomorphic map tram the Riemann sphere into itself, f:S ~ , is represented by an algebraic jUnotion, i.e. it is on the form
- P(z} f{ z } - Q( z}

where P,Q are polynomials.

562

II

Proof: We already know that f is represented by a bolomorphic function w(z) with a finite number of poles. If we denote the poles by z. and the corresponding degrees by k we can remove the singularities by multiplyi~g w(z) with tbe polynomial i Q(z) = n(z-z. )ki i 1 Consider now tbe function p(z) = Q(z)w(z) By construction it is a holomorphic function without poles. Thus it can be expanded in a Taylor series I a zn p(z)
n=o n

which is everywhere convergent. If P is non-trivial tbe corresponding function p(~) = Q(~)w(~) = E a z-n z z z n will have a singularity at z = O. According to lemma 8 the function w(i/z) cannot have an essential singularity at z = O. It follows that P(l/z) has a pole of some finite degree k, i.e. an = 0 when n>k. Consequently p(z) itself must be a polynomial

10.7 WINDING NUMBERS


AS an application of the preceding machinery we now specialize to

the case where same dimension can now show

and Let

are compact orientable manifolds of the f : ~ ~ Nn be a smooth map. Then we know

that almost all points in

are regular values (Sard's theorem). We

Lemma 11 Then either Let Q be a regular value in N. or it consists of a finite number of points.

f-I(Q)

is empty

Proof: The proof is somewhat technical and may be skipped witbout loss of continuity. If ~1(Q) is non-empty we consider a point P in it. Then f* maps Tp(M) isomcrpbically onto T (N) and there exists open enighbourhoods U,V around P and Q, so that f ~ps U diffeomorphically onto V (compare the discussion in section 10.1). Especially P is an isolated point in the pre image ~1(Q) . On tbe other band, ~1(Q) is a closed subset of M. As M is compact, r-1(Q) must itself be compact. If f-I(Q) was infinite, it would then contain an accumulation point, which by definition is not isolated. (Eacb neighbourhood of the accumulation point contains other points from the preimage. ) Therefore t-1(Q) can at most be finite. 0

563

II

Let us now introduce coordinates which generate positive orientations on M and N. Let Q: (yl, .. ,yn) and P : (Xl, .. ,xn ) be chosen so that Q
p

lies in the pre image of the regular value f. maps U of


Tp(M)

Q,
,

i.e. we f

f(P)

Since

isomorphically onto

TQ(M)

know that the Jacobiant, maps a neighbourhood V of Q.

Det[dyi/dxj]IP' is not zero~


p

Furthermore
~

diffeomorphically onto a neighbourhood f


~

If the Jacobiant is positive, then

U V

preserves

the orientation and if it is negative then orientation. Since the preimage f-l(Q) number of points, we can now define:
Definition 11
f
M

f: U

reverses the

can at most contain a finite

The degree of a smooth map is the integer


(10,54)

~ Nn

at the regu Zar vaZue

Deg(fiQ)

p.Ef-l(Q)
~

j Sgn I dyi /dX I I P ,

Here

sgn[ dy i/ 3xj]

l Q

according to whether Deg (f j Q)

f*

preserves or rever-

ses the orientation. Cons'equently tive number of times If we let Jacobiant and p q p empty, it is understood that

simply counts the effecis is covered by the map f . I f C l (Q) Deg(fjQ)

O. f-l(Q) f-l(Q) Q with positive with negative ranges through

denote the number of points in denote the number of points in and q

Jacobiant, then

need not be constants as

the compact manifold sional example. presented


N

This is easily seen already from a I-dimen,

E(1;amp tel: Consider the following map, f : S I ~ S I

where we have re-

by a periodic function:

211
1

p-q

1 1 1 1 1

2
3

2
1

+
-;------------------~2-11----~

M Fig. 210

564

II

But the following deep theorem holds, which we shall not attempt to prove: Lemma ]0 values in (Brouwer's lemma)
f

The degree of a smooth map. N.

N.

is the same for all regular

This common value is then simply denoted the effective number of times The integer Deg(f) since i t counts how many times

Deg(f)

and it measures f.

N
M

is covered by

M by the map
N

is therefore referred to as the winding number-, is wound around In the lite-

rature it is also referred to as the Brouwer degree. Lemma If Proof:


If f

11
f: M -

fails to be surjective. then the winding number is

o.

fails to be surjective there exist a point is empty. Such a point Deg(fiQ) = 0 Q

so that

f-l(Q)

is therefore trivially a regular

point with

At this point we take a look at some illuminating examples: Example 2: by: f(cp) Cos ncp [ Sin ncp i.e. n If we introcp' = ncp . As an example in dimension n=l we look at M = N = Sl f given .

If we introduce polar coordinates, we may consider the map

This is obviously a smooth map with winding number Example J: In n=2 dimensions we look at

M = N = S2. f

duce polar coordinates, we may consider the map f(8,cp) : [:::::::] f3(8,CP) sphere circle
N

given by: (8' ,cp' )=(8,ncp).

Sin8 Cos nCP] Sin8 Sin ncp i.e. Cos8

This is clearly a smooth map outside the poles and it maps the first n 8 times around the second sphere. In fact, it maps each little 8
0

in

M n

times on the corresponding little circle in

565

II

If

was a smooth map, n. UnIn

it would therefore have winding number fortunately necessarily f is not

smooth~

general it is singular at the poles. This cannot be seen from the above coordinate expression since the polar coordinates themselves break down at the northpole and southpole. This is a subtle point, so let us investigate it in some detail. We may introduce smooth coordinates at the northpole, say the standard coordinates x and y They are related to the polar coordiS

nates in the following way:

x
y

SinS Coscp SinS SinCP

Arcsin~

Arctg

If

is smooth, then the partial derivatives af ax and


-+

af ay

will depend smoothly upon (x,y) .


sic coordinates of
-+

For simplicity we compute the extrinthat are tangent vectors to the second Coscp CosS Sincp SinS Sincp CosS Coscp SinS and

ex

and

sphere. Using that: dS ax

dCP ax

~l
ay

as
as

ax

-1 acp ax acp

dCP ay

2Y 2Y

l-

we easily find the following extrinsic components of

~y

~x

Cos<PCos ncp + n SincpSin ncp] [ Sin<PCos =l CoscpSin ncp - n SincpCos ncp ;~y = SincpSin -CoscptgS -SincptgS

n CoscpSin ncp] n CoscpCos ncp

To be continuous at the northpole pendent of depend cp, when we put cp S = 0 explicitly on

(S = 0)

these vectors must be inde0 n or Cos 2 cp+Sin 2 cp=1 l, then the

But the two first components Hence if

unless they reduce to n = l


n n

This miracle only happens for above map is not


smooth~

For

1
-1

it reduces to the identical map it reduces to the antipodal map

of a sphere onto itself. For

of the sphere onto itself. They are obviously smooth.

566 II Exeroise 10.7. 1 Introduction: Let X: [o,~l + [o,~l be a surjective increasing smooth function such that all the derivatives vanishes when S = 0 or S = ~ . (You can e.g. put X(S)
~

XeS) with

= p,Joe
A

S ___ 1_

t(1T-t) dt
e- t(~-t) dt
1

11

= fll o
(a)

Problem:

Show that the following map is smooth:


g(S,~)

=
[

SinX(S) Cos n~l Sinx(S) Sin ~ CosX( S)

Fig. 212

(b)

Show that the only critical values are the northpole and the southpole.

The examples above can clearly be extended to produce smooth maps Sn ~ sn with arbitrary winding numbers.

Consider once more two arbitrary compact orientable manifolds and


If
N

M
f*T

of the same dimension is a n-forrn on


N

and let

f :

M~

be a smooth map.

T
M

then we can pull it back to an n-forrn and

on

We want to compare the integrals:


JMf*T f NT

Since

f ,

need not be a diffeomorphism, they are not necessarily idenf is characterized by an integer, the winding number which tells you how many times

tical. However, Deg(f) by

N are effectively covered M. We can therefore generalize theorem 8, section 10.2 .. The proof is omitted since it requires a greater machinery than we have dev~loped yet. (See e.g. Guillemin/Pollack [1974.

Theorem 21
(10.55)

(BrouUJer's theorem)
JMf*T = Deg(f)fNT

Remark: Let Q be a regular value in image is finite:

with a non-empty preimage. Then the pre-

f -1 (Q) = {Pl"",Pk } According to lemma 1, section 10.1, we can find a single open neighbourhood V of Q, and disjoint open neighbourhoods U of Pi such that f restricts to diffeoi morphisms:
i

= l, . ,k

With these preparations we consider a differential form T of rank n which vanishes outside V. Then f*T vanishes outside Ul U.. U Un Therefore an application

o~~:;o:em~8'fse;:~o: l~.~il~:1
i=l Ui i=l ax]

Pi

fv T

567
But

II

JVT

is independent of

Consequently we get

J~*T

= Deg(f,Q)JNT = Deg(f)JNT

where we have used Brouwer's lemma. This shows that the formula works for a differential form with a suitable small support.

Using Brouwer's theorem we can now construct an integral formula for the winding number. Let there be given a Riemannian metric on Then it induces a volume form, the Levi-Civita form: E = IgdxlA .. Adx n . Because lume: (10.56) If or (10.57) Deg[f) f: M
~

is compact, i t can be shown to have a finite positive vo-

is a smooth map we therefore get: JMf*E

Deg[f)J

Deg[f)Vol[N)

This is the formula we are after! To be able to apply it, we must construct the volume form first. Consider especially the sphere
Lemma '12 Let

Sn-l.

(X1, ... ,X )

be the extrinsic cartesian coordinates on the

unit sphere

Then the volume form with respect to the induced metric is given by: a a an _ 1 (10.5B) n - <n=1TT a .a x l dx 2 AAdx l n

Proof: The induced metric on rounding Euclidean space vectors at a point


+ +

Sn-l n R

is the metric inherited from the surLet

(~l' .. '~n-l)

be a set of tangent

on the sphere. What we must show then is that

n(vl, ... ,v _ l ) is the Euclidean volume or the parallelepiped spanned n+ + n by vI' . ' v n - l in R (Compare exercise 10.7.2 below)'. lOr simplicity i.e. for the tWo-sphere. (x l ,x 2 ,X 3 ) Then the ra1 dial vector ~ = also has coordinates (X ,X 2 ,X 3 ) Let (~,;) be two tangent vectors at P with the coordinates (U l ,U 2 ,u 3 ) and 1 2 3 (v ,V ,v ) . We then get: Let P be a point with the coordinates we prove this only in case n

= 3,

OP

568 This is the three-dimensional volume of the parallelepiped spanned by


+

II

r,u

and

But

is a unit
u
+

vector, which is orthogonal to and

~.

Consequently the three-

dimensional volume of the parallelepiped is equal to the two-dimensional area of the parappellogram spanned by

~'
a

and

~. We have
is the

thus shown that by


+

Q(~,~)

area of parallellogram spanned u,v


+

and we are through.

Fig.

213

Consider a compact orientable manifold M of dimension n and cp : M ~ Sn be a smooth map which we parametrize as: n cpa = cpa(X) , l: [cpa(x) )2 = 1 Here we have used the extrinsic coo~anates of the sphere. The winding let number of (10.59)

cp

Deg[cp]

is now given by the formula: 1 1

Vol(Sn) M

cp*n

n:Vol(Sn) M aoa n

a a a cp o dcp l A. Adcp n

This formula, which is based upon a particular choice of volume form on the sphere, has turned out to be very useful in various Remark: It can be shown that, (10.60) where (10.61) r is the r-function characterized by the recurrence relation:
r(~) '"
applicat~ons.

liT,

r(l) '" 1 ,

r(x+l)

= xr(x)
3~

We list a few particularly useful cases: (10.62) Vol[Sl) '" 2~ ; Vol[S2)

4~; Vol[S3)

8 2

Exercise 10.7.2
Problem: Consider the sphere Sn and let us introduce intrinsic coordinates on the sphere (e l , ... ,en). They induce canonical frames in the tangent+sp!ces: ~l""'~n' The induced metric is now characterized by the components: gij = eiej and the volume form is given by
E

1 n Igde A Ade

1)
2)

Show that

Ii = volume
Let
-+

of the parallelepiped in

Tp(S

E(Vl, ... ,vn )

(vl' ... ,v ) be any set of tangent vectors and show that n -+ n -+-+ = volume of the parallelepiped in Tp(S) spanned out by v" ,v ' n

n) spannedby ~ + el, ,e n .

569 (Hint: Use the Legendre identity: Det(A . E.) = Det(A ij )Det(B ij )
1

II

where Aij are the cartesian components of components of Bi ).

A.
l

and

Bij

are the cartesian

Exercise 10.7.3 Problem: (a) Consider the unit sphere Sl and introduce the usual polar coordinate ~. Show by explicit calculation that the volume form (10.58) reduces to n = dep. (b) Consider the unit sphere S2 and introduce the usual polar coordinates (S,ep) Show by explicit computation that the volume form (10.58) reduces to n = SinSdSAd<p.

Remark: It is important to observe that the concept of a winding number is only well-defined for smooth maps between aompaot manifolds. If so that the local Brouwer degree is not well-defined. But even if it is well-defined it need not be constant on tempted to consider
N.

is not compact then the preimage may consist of infinitely many points, You might still be

an integral expression like 1 a o al an (*) fM n.'V 0 IN Ea '" an A Ad. o But here the integral need not converge and even if it converges, it

d.

need not reproduce an integer! Consider for instance the smooth map : R2 ~ S2 where denotes stereographic projection from the center (cf. figure 214).

Points at infinity

Fig. 214 It covers the lower hemisphere. Points on the upper hemisphere have Brouwer degree 0 while points on the lower hemisphere have Brouwer degree 1. Similarly the integral intuitively clear since (*) takes the value
~
~

~.

All this is

covers half of the sphere. So you may be to this map and that is okay is not an integer.

tempted to attribute the winding number with me, as long as you recognize that

570

II

10.8

THE HEISENBERJ FERR(J>'INlNET


As an exemplification of the machinery built up in this chapter

we will now look at a famous model from solid state physics. We have previously studied superconductivity ( see section 2.12 and 8.7). This time we will concentrate on a ferromagnet. A single atom in the ferromagnet may be considered a small magnet with a magnetic moment proportional to the spin. At high temperature the interaction)energy between the local magnets is very small compared with the thenreU energy. As a consequence the direction of the local magnets will be randomly distributed due to
~l

vibrations.

FOr sufficiently low temperature, how-

ever, this interaction between the local magnets becomes dominant and the local magnets tend to lign up. We thus get an ordered state characterized by an order parameter, which we may choose to be the direction of the local spin vector. Thus the order parameter in a ferromagnet is a unit vector. If we introduce Cartesian coordinates it can be represented by a triple of scalar fields [$1 (X);$2(X) ;$3(X)] subject to the constraint $a$a

1.

CONTINUUM LIMIT
$Cx)

[$IC~);$2(~);$3C~)]

=p

I V$a. v$a
R2

d2x

We are going to study equilibriun configurations in a ferromagnet. Consider a static configuration $(~), where $(~) is a slowly varying spatial function. In the Heisenberg model for ferromagnetism one assumes that the static energy functional is given by
(10.63)

corresponding to a coupling "between nearest neighbours". To obtain the field equations for the equilibrium states we must vary the fields $a. But here we must pay due attention to the constraint $a$a

= 1.

The order parameter

is not a linear vector field, i.e. we

are not allowed to form superpositions!

571

II

The field equations for the equilibrium configurations are given by


(10.64a)

1 1

or equivalently
(10.64b)

Proof: We incorporate the constraint by the method of Lagrange multipliers. Consider therefore the modified energy functional

~[~a(X);A(X)]

~<d~ald~a>

<A(x)I~2(X) - 1>

Now all the fields,~a(x) and A(X), must be varied independently of each other. Performing the variations A(X)
~

A(X) +

~(x)

and

we obtain the "displaced" energy functional

~[]

~[O]

J<d$ald~a>

2~<d$ald$a>

2<A(X)I~a(x)$a(x
+ <~(X)I~2(X)

+ 2<A(X)I$2(X

-1>

As usual we then demand

o = d~[l - J<d$ald~a> dI=O-

2<A(X)I~a(x)$a(x

<~(X)I~2(X)
- 1>

- 1>

<$aIJ&d~a + 2A(X)~a(x

+ <~(X)I~2(X)

This leads to the Euler-Lagrange equations

As usual the equation of motion for the Lagrange multiplier degenerates to the equation of constraint. The constraint can now be used to eliminate A(X) from the equations of motion. Differentiating the constraint twice we obtain

(*)
Consequently

where we have used the equation of motion and the identity (*). Inserting this back in the Euler-Lagrange equations we finally obtain:

(The equivalent version of this equation is obtained by dualizing it!) From now on we restrict ourselves to the case D

2, which has the

most interesting topological properties. Thus we consider a two-dimensional ferromagnet, where a spin configuration consequently corresponds

572

to a map

$:

R2 ~ S2

This suggests that we introduce a topological quantity, the winding number Q, which tells us how many times the sphere is covered (10.65)

Q[~] =
'"

.l..Je: il:ad$b"d~C 811 abc'" '"


R2 Furthermore it need

But here we encounter the usual problem, as the integral need not be well-defined because R2 is not a compact manifold. the preceeding paragraph ). not be an integer even if it is well-defined ( cf. the discussion in We must therefore inwoke a boundary condition and fortunately we have in this model the natural choice: The only spin aonfigurations that are physiaally relevant are those with a finite energy, i.e.

This leads to the boundary condition (10.66) i.e. $a lim


p +
00

d$a

must be asyptotically constant lim $a(x)


P'+OO

(10.67)

$a
0

Exactly which components the constant vector at infinity has we cannot say. All unit vectors are equally likely to be the one Nature chooses. But observe that once we have fixed the unit vector at infinity we can not change it without breaking the condition of finite energy. In what follows we choose the north pole as the asymptotic value of the order parameter: (10.68) lim $a(X)
p +
00

[0;0;1]

The boundary condition (10.68) has the important consequence that we can now compactify the base space. As usual we perform a stereographic projection of the plan,e into t..ile sphere, 11:5 2 R2. Notice that the North pole corresponds to the points at infinity. We can now lift a spin configuration in the plane, $a(x), to a spin configuration on the sphere,

~a

= 1I*$a.

We then extend the lifted field ~a

to the north pole by con-

tinuity

cf. the discussion in section 10.6. As a consequence of our boundary conditions we therefore see that a spin aonfiguration with finite energy aan be lifted to a map 52 ~ 52, whiah maps the north poZe into the north pole. This shows that the winding number is well-defined for all

573

II

the permitted configurations. The space E consisting of all the smooth finite energy configurations therefore breaks up into disconnected sectors

En' where each sector is characterized by an integer n:


n

(10.69)

= 8~J
S'

abc$ad$bAd$C

Remark: As a consequence of (10.42) the energy functional is conformal invariant:

H$2[$l

!J<dWaldW~s2 = ~JJs2*~aA~ = ~JIs2*rr*(d$a)Arr*(d$a) -~JIS2rr*{*d$aAd$a} = !JIR2*d$aAd$a = !J<d$ald$~R2 = ~2[$l

Consequently we can solve the field equations (10.64) on the sphere, whenever it is advantegous, since any solution to the equations of motion on the sphere automaticallyprojects down to a solution of the equation of motion in the plane.

We proceed to investigate various equilibrium configurations. Let us first determine the vacuum configuration for the Heisenberg Ferromagnet. The classical vacuum is characterized by having vanishing energy density. This implies that d$a vanishes, i.e. $a is constant. Thus the classical vacuum corresponds to a spin configuration where all the local spin vectors point upwards. Clearly it has winding number zero. We then proceed to examine the non-trivial sectors. The fundamental problem is whether we can find ground states for the non-trivial sectors, i.e. whether we can determine a configuration in the sector En which has the lowest possible energy among all the configurations with winding number n. We call such a ground state a spin
~ave.

(Notice that

the energy functional (10.63) only contains a "kinetic" term. Although the model is based upon scalar fields it thus corresponds precisely to the eXceptional two-dimensional case, where Derrick's scalings argument does not apply, cf. the discussion in section 4.7 ). Observe that by defihition a spin wave is a local minimum for the energy functional (10.63) and consequently it represents a solution to the second order differential equation (10.64). It is a remarkable property of the Heisenberg Ferromagnet that we can actually explicitly determine all the spin waves. The first step in the analysis of spin waves consists in a reduction of the second order differential equation for the spin wave to a first order differential equation. This reduction is due to a
decomposition of the energy functional
Bogomo~ny

(10.63). Consider the quantity

d$a, i.e. ai~a. It carries two indices: A space-index, i, referring to

574

II

the physical space, and a field index, a, referring to the field space. Let us concentrate on the field index for a moment. Consider the vectors They are tangent vectors on the unit sphere in field space. We may now introduce a duality operation in this tangent space. As usual for a twodimensional vector space it corresponds to a rotation of tangent vector the duality operation is given by (10.72) #d$a
~TI.

We denote

it by # and observe that in terms of the Cartesian components for the

= ~abc$bd$C

Observe especially that #2= -1, and that # preserves the inner product in the tangent space. Worked exercise 10.B.l Problem: (a) Show that

(10.73) (10.74)

<~ald~a> = <#d~al#d~a> = <*d~al*d~a> (b) Show that the topological charge, i.e. the winding number, is given by

n[$l = ~<*d~al#d~a>

(c) Show that any field configuration automatically satisfies the differential equation (10.75)

d#d~a

~a(#d~bAd~b)

(Compare this with (10.64b) ! )

From exercise 10.8.1 we now easily obtain the desired Bogomolny decomposition (10.76) But then we conclude a) The energy in eaah seator En is bounded
(10.77)
belo~

(with II

If

<

I > )

by

b) A aonfiguration

~ith ~inding

number n is a spin wave (i.e. a

ground state for the seator En) i f and only i f it satisfies the first order differential equation:
(10.78)

&a {#d$a *dcp _#d$a

(when n is positive) (when n is negative)

(This is kno~n as a double self-duality equation).

575

II

Exercise 10.8.2 Problem: Show, by explicit c0mputation, that the first order equation (10.78) implies the second order equation (10.64). The next step will be to show that every solution to the double self-duality equation (10.78) corresponds to a well-known geometrical object. In our case it turns out that we get the following nice characterization of spin waves: A spin configuration,

$:R2

S2,

is a spin ~ave exactly when it

generates a conformal map from the plane into the sphere. Proof: Let ~1'~2 be the canonical frame vectors generated from the Cartesian coordinates in vectors

R2.

Notice that they are lifted to the tangent

01$,02$.

If

is conformal this forces

01$,02$

to be orthogo-

nal vectors of the same length. We leave it to the reader to verify that the converse holds, i.e. that the above property actually characterizes the conformal maps from the plane into the sphere. Now suppose as follows But from these equations it follows immediately that al$,02~ thogonal vectors of the same length. Consequently are or-

is a configuration with winding number

which sol-

ves (10.78). This first order differential equation can be rearranged

is a conformal map.

On the other hand it is not too difficult to show that any conformal map actually solves (10.78). We leave the details as an exercise: Worked exercise 10.8.3 Introduction: Let $:R2~ 82 be a smooth map and introduce the following vectors in field space:
i.e. i.e,

..p,

, ,

'-

'-J J

,0,$ ,0,$
+

$xo,$
'-

Problem: (al Show that if

$ is

'-J J

$'0,$
'-

a conformal map then the following holds

Pl,i\ = Ql'Q2 = PI'QI = (b) Show furth!;.rmor!;. that if PI = P2 = 0


or

$ is

P2'Q2

Pl'Q2

P2'QI

a conformal map then either

This concludes the proof:

As an example we notice that the stereographic projection itself is a conformal map and thus it represents a spin wave with winding number -1 (it reverses the orientation!) Observe also that the stereogracan be used to lift any spin configuration ~:R2.... R2 to phic projection

576

II

a spin configuration ~:S2~S2. Consequently there is also a one-to-one correspondence between spin waves and conformal maps from the unit sphere into itself which maps the north pole into itself. In the final step we introduce aomplex analysis. In the present ,case it is completely trivial. It is well-known that the sphere is isomorphic to the extended complex plane
S2
...

* =C

u { co}

and that the ortentation preserving conf<mnal maps are holCllOrphic, cf. the discussion of the Riemann-sphere in section 10.6. But holomorphic maps are necessarily algebraic, i.e. they are on the form (10.79a) FUrthermore w(co) w(z)

= !:ill Q(z)

with

P,Q polynomials.

implies that

DegP>DegQ. Thus we have finally

explicitly construced all spin waves with a negative winding number.*) Similarly a spin configuration with a positive winding number corresponds to an anti-holomorphic map on the sphere, i.e. it is on the form: (10.79b) w(z) =

~m

with

P,Q

polynomials, DegP>DegQ

It would also be nice to find a simple formula relating the winding number to the polynomials P and Q. That is easy enough: The map w=P/Q is a smooth map and according to Sard's theorem there are plenty of regular values. Let Wo be a regular value. Then the preimage consists of all solutions to the equation P(z) - woQ(z) = map preserves the orientation. Thus this means that

o.
FUrthermore a holomorphic DegP and w has winding number

As DegP>DegQ it has Degp distinct solutions.

has winding number -DegP.

Remark:

Rather than relating spin waves to conformal maps one can relate them directly to holomorphic (or anti-holomorphic) functions. Using a stereographic projection IT from the unit sphere in field space to the complex plane we can project the order parameter down to a single complex field given by
W

= -~l_iji3

1.\1

~2

+ i-'l'-I.\l_iji3

In the same way "he tangen" vector a,$a on the unit sphere is projected down into the tangent vector diw in the complex pl~e, i.e. d$a ~* dw Since IT is conformal it preserves the right angles. Furthermore it reverses the orientation so that ~* -id.w #d$a ]!.* -idw i.e.
1

Similarly it follows from the linearity of

IT*

that

e: i } }

Aa

1f*

e:, .d.W
1J J

i.e.

*d$a JJ..* *dw

*) A.A. Belavin and A.l1. POlyakov,"lIetastable states of two-dinensional isotropic ferromagnets", JETP Lett. 22 (1975) 245.

577

II

Putting all this together we therefore see that the double self-duality equation (10.78) is projected down to the ksual self-duality equation for a holomorphic function *dw = -idw

Okay, this concludes our discussion of spin waves, i.e. the ground states for the various sectors. We might still ask if there are other finite energy solutions to the full field equations (10.64) Apriori there could be local minima lying somewhat above the ground state, or there could be "saddle points". But Woo*) has investigated this problem and using complex analysis he has shown that the answer is negative:
Any finite energy solution to the seaond order equation (10.64) automatiaally solves the first order equation (10.78) Worked exeraise 10.8.4 Problem: a) Show that a spin configuration represented by the complex valued function w(z) has the energy density
(10.80)
b)
2 {ow H = (l+ww)Z

oZ

oZ-

aw

oZ

oW

ozJ
and

Show that the corresponding equations of motion are given by (l+ww)

( 10.81) c) ( 10.82)

OZdZ

~2W

zw
=

illi: ~
oZ dZ

(1+WW)o;2~
=

2w

~ ~

Consider the following two complex-valued functions

awaw

awllw

fez)

oZ oZ (l+wwP

and

g(z)

az oZ (l+ww)Z

d)

Show that fez) is holomorphic and that g(z) is anti-holomorphic provided w(z) solves the equation of motion Show that if f has no poles, then w itself must either be holomorphic or anti-holomorphic.

In the above exercise we have almost proven that a solution to the full field equation (10.64),which has finite energy, is automatically represented by a holomorphic or an anti-holomorphic function. There is however one possible loop-hole: The quantity

f (z)

(l+wwF
a"pole~

az

oW oW oz

might have a pole, or equivalently the energy density might have Consider, e.g., the non-admissible solution
w(z)

zls

*) G.

\\00,

"Pseudoparticle configurations in two-dimensional ferromagnets",

J. Math. Phys. 18 (1977) 1264

578

II

(It is non-admissible because it is multi-valued!) Neglecting the branch cut for a moment we notice that it produces the energy density

1
H = 2Izl(l+lzl)2

Consequently the energy density has a "pole" at Thus we cannot exclude the possibility that f(z)

z=O . But it is still might have a pole.

integrable (as you can easily see if you introduce polar coordinates). One must then investigate the solution to the first order equation (10.82) very carefully to exclude that possibility too, i.e. near a pole of like f(z) we must show that w(z) is necessarily multi-valued w(z) =

z~. FOr further details you should consult the original

paper of Woo.
Exeraise 10.8.5 Introduction: Let F be an ordinary function: F:R1R and let w(z) be a complex valued field characterized by the energy density
H =

F(wW){~ ~ az az

~ az ~} aZ

In analogy with the Heisenberg ferromagnet we will only be interested in static configurations with a finite energy. Problem: a) Show that the equations of moti~n are giv~ El 2 F d_W =-F'w ~ ~ and F ~ =-F'w ~ ~ azaz dZ dZ azaz dZ az b) Show that

fez) = F(ww) ~ az ~ az

is a holomorphic function provided w(z) solves the equations of motion. c) Specialize to a static massless complex Klein-Gordon field in (2+1) space-time dimensions and try to characterize the static solutions with finite energy. (There are none except for the trivial solutions w(z) = constant!) Hint: Show that it corresponds to the case F=l.

10.9

THE EXCEPTIONAL f4-MODEL

As another interesting example we will consider a model in two spacedimensions, which on the one hand is related to the Ginzburg-Landau model for superconductivity, on the other to the Abelian Higg's model (cf. the discussion in the sections 8.7-8.8). It will be based on the static energy functional:

(10.83)

579 ,cf. (B.60) and (B.BB).

II

For the moment we leave the potential unspe-

cified except that we assume that it is gauge invariant, i.e. it is a function of 1~12

and furthermore that it only vanishes at the non-

zero value I~I= ~o' The field equations for a static equilibrium configuration are given by (10.B4a) (10.B4b)

au
D*D~

2~2~[

- &B

i~[~D~ - ~D~J 2

cf. the worked exercise B.7.3. As usual we consider only finite energy configurations. This leads to the boundary conditions: (10.B5) lim
p -+
co

lim
p +
..,

D~

lim
p +
co

As in the case of ordinary superconductivity we have furthermore flux quantization, i.e. (10.B6)
<P

JR2 B

21f
e

where n is related to the jump in the phase of the Higgs' field when we go once around the flux tube. So far nothing is new. We will now try to see if we can find the groundstate configurations for the non-trivial sectors ny decomposition of the static energy functional. We start by guessing a reasonable set of first order differential equations. Guided by the (l+l)-dimensional models and the Heisenberg Ferromagnet we try the ansatz: (10.B7a) (lO.B7b)
B
*D~

(where nfO). As

usual the investigation of the ground states will be based upon a Bogomol-

v'2u[~]
iD~

= *hU[~l

(i.e. the self-duality equation) .

They lead to the following pair of second order differential equations (lO.BBa)

-D*D~

-iD2~

-eB~

-eV2U[~1

(where we have used exercise B.7.2).

(*)

-5B = -5*{2U[~1

=-

*dV2U[~1

But from the self-duality equation we get


*d~

iD~

ie~*A

*d~

=-iD~

ie~*A

Consequently we get

5BO
~*d$ + $*d~

II

Inserting this into equation (*) we can rearrange this as


1

(lO.BBb)

- 5B

=-

v'2U[$)~2 i[cjlDcjl - cjlDcjl)

au

Compairing (lO.BBa-b) with (lO.B4a-b) we see that the first order differential equations are compatible with the second order differential equations provided the potential satisfies the identity

This restricts U to be a fourth order polynomial of 'the form

(lO.BB)

U[cjl) = 2ClcjlI2_ cjlo)2

with

Icjll < cjlo

Notice that the above potential represents a spacial case or the potential energy density in the abelian Higg's model where we have put e2 (lO.B9) A= 2 The abelian Higg's model based upon the potential energy density (lo.BB) is known as the exceptiona l $"-mode l. In superconductivity it corresponds to the case where (10.90)
(K is the Ginzburg-Landau parameter (B.67). The identity (10.90) follows immediately from (lO.B9) when you perform the substitutions A ~ 8

and e ~ q/h ). Okay, in the exceptional cjl"-model we thua have the possibility of reducing the field equations (10.B4) to the first order equations (lO.B7). This suggests the following Bogomolny decomposition: H

~<B+V2U[cjl)EIB+V2U[cjl?E> + ~<*Dcjl+iDcjll*Dcjl+iDcjl> + a rest term

All we must check is that the rest term is a topological term, i.e. that it only depends on the "winding number" n. We leave this as an exercise to the reader:
Worked exercise 10.9.1

Problem: (a) Show that the winding number is given by (10.91) n = ~ <*dcj>lidcj (This should be compared with (10.74)!)
11'1'0

(b) Show that the exceptional cj>"-model possesses the Bogomolny decomposition (10.92) H

= !\lB+/2U[cjl)EW

~1*Dcjl+iD$1I2 : n1lcjl~

(with

II

1\2= < I> )

We therefore conclude as usual:

581 (a)
(10.93)

II is bounded below by

The energy in the sector

H ~ Inl1f<p~

E n

(b)

A configuration with winding number n is a ground state i f and only i f it satisfies the first order differential equation:

(10.87 )

In the above discussion you might feel that the first order equation (10.87a), which preceded the Bogomolny decomposition, was sort of "pulled out of the air". We did try to justify it by referring to our earlier experience with (l+l)-dimensional models, but you may not find that very convincing. We shall therefore present another argument, which allows one to make reasonable guesses in such a situation. The argument is closely related to the reasoning behind Derrick's scalings argument. We know that a pure scalar field theory in two dimensions cannot possess stable non-trivial static solutions. In the above model it is thus the precense of a gauge field which makes it possible to stabilize a static configuration. Let us look at this in some detail: Consider the scaling transformation

D" (i{)

'=

A~

It is a conformal transformation woth the conformal factor

QZ(x) =

A2

Let us furthermore introduce the scaled configuration

From (10.40-41) we now get

(or by working out the coordinate expression directly)

2J*dAAdA = AZ<dAldA> <dAAldAA > ; J*DtdAADtdA = AZJDt[*dAAdAl = A Z Z R R D" CRz)

<DA<PAID,,<P A > = J*DtD<PADtD<P = JDt{*D<PAD<P} = J*D<PAD<P


RZ RZ DA CR2)

<D<pID<P>

JU[<PA1 ; A-ZJU[Dt<P1Dt ; A-ZJDt{U[<Pl(} ; ,,-2JU[<Pl ; ,,-2JU[<Pl t RZ R2 RZ RZ DACR Z)


This leads to the following displaced static energy:

H[<PA;A,,] ; l"z<slg>

!<D<PID<p>

A-ZfU[<P]E Z R

582

II

A stable configuration and especially a ground state must now satisfy

Thus the condition of stability leads to the "viriral theorem": A stable configuration satisfies the ldentity: (10.88) Thls should be compared with (4.75). In analogy wlth the (l+l)-dimenslonal case we now suggest that a ground state not only satlsfles this ldentlty globally, but that it ln fact satisfles it point-wise, i.e. that B = 2:v'2U[cjl)
E

Thus we recover preclsely the ansatz

(lO.87a).

So all we have got to do is to solve the first order equations (10.87). Cnfortunately that ls a very hard task, slnce one cannot wri[sing advanced analysls it can be shown te down the explicit solution.

that the most general solution with n flux-quanta depends upon 2n arbltrary parameters correspondlng to the center posltlons of n flux tubes each carrying a unit flux. Taubes and Jaffe) If one speciallzes to spherical symmetric configuratlons, the analysls becomes very slmple. If we make the ansatz (10.89) A(x) = eA(p)dcp
n

(Ior details about the machlnery requi-

red to analyse the general solution you should consult the book of

it is easy to see that thls represents a flnite energy conflguration with n flux quanta provided (p) and A(p) tlons (10.90) lim
p -+

satlsfies the boundary condl-

4i'(p)

0;

lim' 4i'(p)
p -+
co

1;

p -+

lim A(p) 0

0;
p

lim A(p)
-+
co

1,

cf. the dlscussion in section 8.7.

Worked exercise 10.9.2


Problem: Show that the field equations (10.87) reduce to n dA 1 2 2( ~2) &b = ncjl(l-A) ~ (10.91) p dp = 2e cjlo l-cjl ; p~ when we spacialize to the ansatz (10.89)

It follows fran standard theory of first order differential equations that there is preclsely one solution (cjl(p) ,A(p)), which interpolates between (0,0) and the "equilibrium point" (1,1) .

583

II

SOLUTIONS OF WORKED EXERCISES:


No. 10.3.2 In polar coordinates the stereographic projection.is.giv~n by P = 2Cot~ ; = ~ The Jacobi matr1x 1S g1ven by

(;5) [- ";'~
= G1

:1

(Especially we observe that the Jacobiant is negative, i.e. the stereographic projection is orientation reversing.) The metrics gl and g2 are characterized by the components
=

[1 0 e]
0 Sin 2

= G 2
=

[1

Fig. 215

The pull-backed metric is then characterized by the components


s+ -

f*g2

D21ii2D21

["~"
a

o
4Cot

28

1 "2

_1_ Sin 4il. 2

-------

No. 10.3.3 If we perform a general parameter shift and so the geodesic equation is transformed into d Xll + rll d s2
2

A = A(S)

then

as

dx

ds

B ll dx = dx ~ ds ds A' (s)

Thus the most general, covariant equation for a geodesic is on the form with On the other hand any smooth function A(s) = with A(s)

-~ A( s) A' (s)

can be written on the form

~ = A' (s)

[lnA'(s))'

584

II

No. 10.8.1 (a) The first equality is a consequence of the fact that # preserves the inner product in field space, since it corresponds to a rotation. The second equality follows from theorem T chapter 8. (b) (c) trr<*d$al#d$a> =

8~J**[d$a]"#d$a

_~Jd$a"abc$bd$C

~JabC$ad$bAd$c =n[$]

d#d$a '" d[ b $bd$c] '" b d$bAd$c


a cae a c

Notice that Fa = b

d$b"d$C

is the vector in field space obtained by taking

the cross product of the two tangent vectors d$b an~ d$c. Consequently it must be proportional to the radial vector $a and since ~a is a unit vector we

Fa = $a($bFb) = $~b

abc

d$CAd$d= $a(#d$dAd$d)

n U

No. 10.8.3 + + + (a) Let us write out Pl,P2,Ql and !1 = 02~ - ~'01~ Ql '" 02~ + ~'01~ Then we get

o
+ + + + ~ ~ PlQ2 = P2Ql = -2g1~ d2I But for a conformal map ~:R2~ we know that 01$,02$ are orthogonal tangentvectors of the same length, so this forces the right hand side to vanish. (b) We start with thx following important remark: If 01$ (or d2~) vanishes, then aLL fOur tangent-vectors Pl,P2,Ql and Ql vaand

nish.
To see this let 01$ vanish. Then we get P2 = -$'02$ and Q2 = $'d2$ ~ut+th$Y are o~thogonal vectors so.this forces 02$ to vanish too. Clearly Pl,P2,Ql and Q2 must now all van~lh.+ + + To prove (b) we now observe that Pl,P2,Ql and Q2 are all tangent vectors in the same tangent plane to the unit sphere. As they are mutually orthogonal at least two of them has to vanish. But apparently there are more possibilities than t~ose lilted i~ b). + + Suppose PI and Ql van~shes. Then 32$ = !(Pl+Ql) hal to vanish too and by the above remark they iherefo~e all vanish. Similarly PZ=Q2=O make them all vanish. Suppose then that PI and Q2 vanish. From Q2=0 we get 31$ = $.32$ and when we insert this in the expression for PI we obtain = = d2~ - $K($XdZ$) = 232$ So again 02~ has to vanish, and therefore they all vanish. Similarly P2=Ql=0 make them all vanish. This leaves with the two possibilities listed in (b).

o PI

No. 10.8,-4 '; (a) From section 10.3 we know that the stereographic projection is a conformal , map with the conformal factor \12 S 4 (a) 1 = i l l 2 = Cl+wW)2 Consequently we get 1dlWl2+ 13 2wI 2 !(I Ol$1 2 + 102$1 2) = 2 (1+wW) 2 Thus the energy density may be re-expressed in terms of complex variables as follows:

585

II

aiWa? 2(1 +WW)2

Ow aw} 2{aW OW +~ az ~ az (l+wW) 2

(b) The corresponding Euler-Lagrange equations can be written in complex form as aB aw


=

.1.f3B } + .1. LaB}


azla~ azla~

~~ = fzf}~}
la az

~{a~}
'a aZ:

az

az

E.g. the first equation reduces to 2(l+wW),w aw ~ + ~ ~ (1+WW)4 az oZ: oz az _1_ QH + J.. _1_ az (1+wWl'az a~ (1+wW1 2 az 2 w - 2(1+ww ~ aW. oW ow ya -)[~+w~l2R (1+wW)2 azaz - 2 (1+wW) [azw+waz1ai + (1+wW azaz alw az dZ
(1 +

= 1-

aw

wW)"

which after a long, but trivial, calculation reduces to a2aw OW (1+wW)a!~ = 2waz az: (c) We must show that f satisfies the Cauchy-Riemann 2 2 1 -)2 a w aw + ow a w _ aw~ 2(1+ww-) M = ( +ww az azaz az (1 + wW)"

mz

az mz

2 2 __ 1_ _ {(1 -) a w '>.-: ow~ (1+ww-) a w _ 2w awa~ } - (1+wW)3 +ww azaz - ~w azaz + az:az azaz Inserting the equations of motion we see that the right hand side vanishes automatically. In the same way we can show that g(z) is anti-holomorphic. (d) If fez) has no poles it is an entire function, i.e. a global holomorphic function. As the energy density vanishes at infinity we conclude that
az 0 11m = (1+w.W) '" p"''''' Consequently !f(z)! ... 0 at infinity and therefore f(z) must be bounded. But a bounded entire function is necessarily constant (Liouville's theorem) Therefore fez) must in fact vanish identically. But then the equations of motion (10.81) reduces to aw/a'l. = 0 or aw/dz = o.
p ...
co

!~I -~-O 11m (1+w.W) -

IQRI

No. 10.9.1 (a)

<*d~lid~>

-iJ~Ad~
R2

-iJd[~d~l
R2

lim Po"'''''

-iJ~d~
p=po

2n1f~~

(b)

Expanding the normS we immediately get

111 B-v'2U[~h:,,2
=

+ l"I<D~-:i1J~1I2

!<sls> + <vUf<PfEI~E> - <B1VZ0T0E> + !<D~ID~> -1<*D~liD~>

Consequently the rest term is given by A + B = <B1v'2U[~1> + 1<*D~1:i1J~> We will now determine when this is a topological term. Using that we can expand the B-term as D=d-ieA

586
Here B1 is the topological term we are after B1 Furthermore B,

II

= ~<*d~lid~>

= nIT~%

vanishes automatically since B,

= ~ie2<~*AI~A> = -~ie2fl~12AAA

but

AAA

Thus we are left with B2 and B3 which we can rearrange as B2+B3 =

-~ef(~d~+~d~)AA = -~eJdl~12AA

Here we can safely replace 1~12 with 1~12_~~ since the constant drops out anyway when we take the exterior derivative. Thus we get B2+B3

= -~eJd(I~12_~%)AA = -~efd[A(I~12-~%)1

~efB(I~12_~%)

The first term is converted to a line integral at infinity and it now vanishes sincel~12_~% vanishes at infinity. To summarize the rest term has thus been broken down to

fBV2U[~1
R
J2U[~1

J~Be(I~12-~%)
R

nTI~~

But here the first two terms obviously drops out provided

~e(~~_1~12)

i.e.

U[~l

~e2(1~12_~ij)2

(In this way we have actually recovered the


No. 10.9.2.

~'-model!)

Let us first look at the equation (10.84a) B = ~e(~ij_I~12) E On the other hand we get

B =

dA

ndA

e dO

dPA~

epdp

ndA
E

By comparison we thus obtain the first field equation. The other one is a bit tricky. First we observe that (cf. exercise 7.5.5) *dp

= -pd~
i~

and

*~

~dp p

Then we get by a trivial calculation


D~ = ~oe

dpdp + ~o~e

~ in~. (

In l-A)~

The two sides of the self-duality equation therefore reduces to


*D~

..
iD~

in~ = l~oeapdp

- ~o~e

~i~

n(l-A)d~
~

By equating the coeeficients of dp or field equation.

we now finally obtain the second

587

II

chapter

II

SYMMETRIES AND CONSERVATION LAWS


11.1 CONSERVATION LAWS
We start by investigating the general structure of conservation laws in physics. They are concerned with a vectorfield equation of continuity (see (7.88: (11.1) The vectorfield integral (11.2) is interpreted as a "charge". with (8.53). Consider now a tube as shown on fig. 216, and let us assume that the current

satisfying the

&J = 0
J

(FgJll) [ -L~ Fg II
J

0] .

is interpreted as a "current" and if through

main in a spaceslice, then the flux of

n is a don , given by the

Q=-f*J,
(The minus sign is conventional. Compare

vanishes outside the tube.' (The width of the tube can be very large and the fact that the current vanishes outside the tube in practice means that "the current vanishes sufficiently fast at spatial infinity".) We can then show the following lemma: xl

Lemma 1

(The tube lemma)

A smooth "current" face

which satisfies the continuity equation

(11.1), wiLL produce the same flux

-fn*J

through any closed hyper sur-

intersected by the tube, or equivalently, the "charge"

contained in a hypersurface wiLL be the same for aLL hypersurfaces interseoted by the tube.

588 Proof: Let n 1 and n2 be two hypersurfaces intersected by the tube. We will assume that they are far apart. Together with the wall of the tube they then constitute the boundary of a 4-dimensional regular domain W. Be careful about the orientations relative to theorem we now obtain

II

W! From Stokes'

o
(If

= Jw*&J= Jwd*J = Jaw*J = J n2 *J-J n1 *J'


n1 and n2 are not far apart we n3 which n1 and 02 and use Fig. 217

introduce a third hypersurface are far apart from transitivity).

Okay, having constructed the machinery we can now apply it to some important cases: Example 1: CONSERVATION OF ELECTRIC CHARGE.

The electric current is represented by a co-closed I-form. Let us prove that this leads to the conservation of electric charge. To avoid convergence problems we assume that the current is confined to a tube in space-time. Suppose S is an inertial frame of reference, and let

Fig. 2l8a

Fig.2l8b

0 1 and O2 be three-dimensional regular domains contained in space slices relative to S such that n 1 surrounds the charges and currents at time (8.53 Q{ttl = t1 and similarly for t1 02 at time t2 {see fig.
2l8~.

Then

the total charge at time

, respectively

t2 , is given by (Cf. Q{t2)

-J n1 *J

respectively

-J 02 *J

589

II

But according to the tube-lemma they are identical, i.e. the charge measured by an observer in S is independent of time. S1 and n1 and S1 S2 S2 are two be a ren2 a reS2 , Actually we can prove something more. Suppose

different inertial frames of reference. Then we can let gular domain contained in a space-slice relative to gular domain contained in a space-slice relative to As we have seen the electric charge measured in is given by the integral Q1 = S1 Q2 =

(see fig. 218b).

,respectively

-J n1 *J

respectively

-J n2 *J

But by the tube-lemma they are identical. Thus the electric charge is independent of the observer too. We have now given the promised proof of the first half of Abraham's theorem (section 1.6): Theorem 1 If

is a co-cLosed l-form, which vanishes sufficientLy fast at


Q =

spatiaZ infinity, then the "charge"

-J n *J

JOdx 1dx 2dx 3 xO=t O

is independent of time and the observer. Example 2: CONSERVATION OF ENERGY AND MOMENTUM. symmetric tensor of rank 2

We start by observing that the energy and momentum densities of a field configuration are represented by a (Cf. the discussion in section 1.6). At first thought you may therefore think that the analysis of energy and momentum falls beyond the scope of exterior calculus, which deals only with skew symmetric cotensors. On the other hand, we want to define the total energy and the total momentum of a field configuration. This involves integrals of the form

J
xO=tO rization of these integrals.

TCLo d 3 x
geometricaZ, characte-

and it would be nice to have an invariant, i.e.

To do that we "mimic" the discussion of electric charge. We select an inertial frame of reference rating the basic tangent vectors eo ,e1 ,ez ,e3 These constitute four vector fields, which loosely speaking specifies the direction of the CarteSian axes and the direction of time. By con-+ -+

S. This inertial frame of reference

is characterized by a set of inertial coordinates


-+
-+

590

II

tracting the cotensor four new l-forms


(11.3)

with these four vector fields, we obtain

which in our inertial frame are characterized by the components Po : TasO~ = Toe It follows immediately that Pi Po Pi : TaSO~ = TiS represents the energy current, while

represents the momentum current along the xi-axis. Observe, that

although the currents PO,Pl,P2,PS are purely geometrical quantities, they depend strongly on the choice of the inertial frame of reference

in question. In terms of the inertial coordinates associated with that the energy momentum tensor
(1. 38)

we know

satisfies the continuity equation

and we therefore conclude

~
where (Pa)S ,the current

S (p a ) S

= 0

are the contravariant

components of

' Consequently

Pa is conserved, which can be expressed in the geometrical form &P a = 0 Thus PO,Pl,P2,PS all give rise to conserved quantities. If 0 is a space-slice characterized by the equation xO=tO , then the integrals
(11.4)

are all time-independent provided the energy and momentum currents vanish sufficiently fast at spatial infinity. Let us interpret these integrals. Using that dxo=O along n we get p = f (p )odx 1 dx 2dx s = a xO=t O a Observing that in an inertial frame _Too=T ly obtain that
(11.5) (11.6)


the total energy

final-

-p

!Toodxldx2dx3 xO=tO x =t

P. l.

of

Tiodxldx2dxs

the total momentum along 1 the x -axis (Do not be

Consequently the total energy and momentum is conserved.

confused by the signs. If we "raise" the index a we get [pO,pi] = [total energy, total momentum along xi-axis] with the correct signs as usual!) We will then investigate what happens when we exchange the inertial frame of reference. Suppose 52 is another inertial frame of reference

591

II

connected to

81

through a Poincare transformation


yr:J.

Ar:J. )3+b r:J. S It follows that the new basic tangent vectors
+

are connected to the old

e {If

through the formula {cf.

{~

are connected to the (6.12

where

AS r:J.

is the reciprocal Lorentz

matrix. The new basic tangent

vectors give rise to new energy and momentum currents


(ll. 7)

(I) S

AS
r:J.

Consequently the new currents depend linearly upon the old currents. Let
~1

and

~2

be two three-dimensional volumes, such that 81 and 02 (see fig. 218b).


p (2)
r:J.

~1

is

contained in a space-slice relative to relative to 82 ,

in a space-slice

From the tube-lemma we then get


r:J.

-J ~2 * (2) P

-J *
01

(2)r:J.

Using (11.7) this is rearranged as


(1l.8)

P (1)S (P O ,P l ,P 2 ,P 3 ) interpret

AS
r:J.

Formally this shows that the components energy momentum vector transform

of the total (Po ,PI ,P2 ,P 3 )

like a Lorentz co-vector under a

Poincare transformation. However, we can not

as the components of some co-vector for the following reason: It has no foot point, as it is not attached to any specific event in spacetime~

To summarize we have now given the promised proof of the second half of Abrahams theorem (8ection 1.6):
Theorem 2 Suppose

is a symmetric co-tensor Of rank 2 in Minkowski space,

that is conserved (i.e.

~STr:J.S=O

in an inertiaL frame).

Let

be an

inertiaL frame of reference generating the basic vector fieZds

~o, .. '~3
(1) (2)

Then the foLlowing holds:


Pr:J.
2

The l-forms The "charges"


Pr:J.

= -JO*Pr:J. =

J T dx l dx dx 3 xO=tO r:J. 81
and

are conserved, i.e. independent of time. (3) If two inertiaL frames of references ted through a Poincare transformation

82

are connec-

592
yc!.

II

AC!. 'xi3+bC!. i3

then the "charge" transforms as a Lorentz-coVariant quantity.

i. e.
(11.8)

i3 P (ll i3 c!.

Remark: In the theory of general relativity spacetime is curved and therefore we can not choose a global inertial frame. The energy-momentum currents then no longer exisi and this lead to difficulties in the interpretation of various quanti ties like "the total energy of a closed system".

11.2 SYMMETRIES AND CONSERVATION LAWS IN QUANTUM MECHANICS


Now that we understand the formal aspects of conservation laws we turn our attention to symmetries. To get a feeling for the machinery will investigate symmetry transformations in ordinary quantum
WE

mechanic~

We will also indicate how they lead to conservation laws in quantum mechanics, but these are of a very trivial type involving no geometry, so this aspect is only included for physical reasons. We start by conSidering a single particle moving in a potential V{i) The system is completely characterized by its Schrodinger wave
~(t,i)

function

Let us neglect dynamics for a moment and just con-

sider the wave function as an ordinary complex-valued scalar field defined on the Euclidean space R3 Suppose now that (fA)AR is a family of diffeomorphisms. Then we can push forward the wave function (ll.9)

~J,.{i)

= (fJ,.)*~{i)
{UJ,.)J,.R' where

But as this operation is linear we may equivalently say that we have constructed a family of linear operators, (ll.lO) If fUrthermore
UJ,.~ (f.J,.)*~

fJ,.

is a family of isometries
~ion

, then

UJ,.

preserves

the inner product (see theorem 13,


A

10.5):

<&J,.~21&J,.~1> = <{fA)*~2 I (fJ,.)*~l> = <~2 I~!>


Consequently {UJ,.JJ,.R is a family of unita1y operators. Observe that

~J,.~
UJ,.~

is normalized too, i.e. as a wavefunction, i.e.

Jl6J,.~12 dV
UJ,.~

1 , so we can interpret

represents another possible state

of the system. Such families of unitary operators have been studied intensively both from a physical and a mathematical point of view. Before we go into a further discussion, let us recapitulate a few basic facts from elementary quantum mechanics. Suppose T is a Hermi-

593
A

II

tian operator representing a physical quantity T (e.g. the energy refl2 ... '" presented by H = - 2m~+V(x) or the momentum represented by p=-iflv) When we measure this quantity the outcome of the experiment is a number, and the possible numbers we can measure are precisely the eigenvalues of the operator T. Let $n be a complete set of normalized An' eigenfunctions with the associated eigenvalues
T$ n = An $ n

If the state of the system is characterized by the wave function can decompose it on the above set with Here Jan J 2 is the probability of measuring the number mean vaLue of the physical quantity T is given by
(11.11)
<.T> = Jij?r$dV = LJa

we

An

and the

J2 A n T generates a family of uni-

Now observe that a Hermitian operator tary operators given by (l1.12)

This family is actually a one-parameter group, i.e. it satisfies UA2 + A1 = U A2 UA1 and the Hermitian operator T is called the infinitisemal generator of this one-parameter group. On the other hand, if we have been given a one-parameter group of unitary transformations U ' then a famous A theorem of Stone guarantees that it is actually on the form (l1.12)

U A

=exp[-kAT1 T (Of course fl has only been intro-

for some Hermitian operator

duced for convenience. It will not occur in mathematical references!) Thus there is a bijective correspondence between one-parameter groups of unitary operators and physical quantities represented by Hermitian operators. In the preceeding discussion we have seen that a family of isometries generates a family of unitary operators. If we assume that the family of isometries is a one-parameter group, i.e. it satisfies f A2 + A1 = f A2 0fAl then i t will actually generate a one-parameter group of unitary transformations. In the end we will therefore find that isometries in space are linked up with various physical quantities! When {fA)AR is a one-parameter group of isometries we can associate a vector field

to this group. This vector field is the geo-

metrical equivalent of the infinitisemal generator in physics. Geome-

594

II A
~

trically it is generated by the curves that fA{P) pass through P at fA


i

fA(P)

at

A=O .

(Observe P).

A=O

thus generating a vector at


+

If we introduce coordinates ,

will be represented by a

yi = f~{xj)

and the components of the vector field (ll.13)


i _

are thus given by

a - dA I A=O We call this vectorfield the character~8tic vector field assooiated with the one-parameter group with respect to the suffix IA=O
8

5!Y..

~f

isometries.

NB! In what follows we will have to do a lot of differentiations A at A=O. For typographical reasons we will drop A=O. so in the rest Of this ohapter it will aZways be under-

tood that differen tiations a:r-e to be \~arried ou t at

Using the characteristic vectorfield we can now construct the infinitisemal generator

T:
i
A

d -flAT dACe 1/J)

!~

-<d1/J I;;:>
Consequently we have shown The infinitisemal generator rator given by (1l.14)

is the first order differential ope-

We can also consider veotor partioles. Quantum mechanically they are characterized by a complex-valued vector field complex valued scalar field). But the vector particle at the point (ll.lS)
-+

tt{i)

(rather than a

tt{i)
+

still has the usual interpre-

tation. Its absolute square measures the probability density of finding

P{x),=1/Ji(x)1/J (x)
A ...

--+i+

Again we can use a family of isometries, {fA)AER' to generate a


~

group of unitary transformations, UA1/J = (f )*1/J , w~ere the transformed A wavefunction is characterized by the components 1/J~{i) We proceed to determine the infinitisemal generator T: iA i d A ~ i d :t i -KI'1/J ;rr{U A1/J) =dA[{fA)*'I'] d A axj 1/J A (f_A(x i.e. in components we get '"
d

[~j

~]

(11.16)

T1/J

=-ifi[(d.1/Ji)a

595

II

This is a bit

surprising~

By construction we know that the quantity covariant quantity, but the separate
derivative~)

on the right hand side is a

terms are not covariant quantities. However we have a skew-symmetrization involved (in analogy with the exterior Thus we have incidently discovered how to construct a new vector field out of two given vector fields, the commutator or the bracket.
Exercise 11.2.1 Problem: Let u,'t be two given vector fields on a manifold M characterized by the components ai(i) and bi(i) . Show that

(11.17)

(a.bi)a j - (a.ai)b j are the components of ~other. vector field, which We denote by [~;;]
J J

USing the notation of exercise 11.2.1


The infinitisemal generator rator given by

~e

have thus shown:

is the first order differential ope-

(11.18)
Exercise 11.2.2 PrOblem: Show that the bracket satisfies the following rules:
(1) Skew-symmetry: (2) Bi-linearity:

[~I~]

(3) Jacobi-identity:
EXerci8e 11.? 3 Problem: Let Tl,T2
A

= -[~;~] [~;~1+~2] [~!~l] + [~;;2] [~;A~] [~;[~;;ll + [~;[;;~ll + [;;[ii;~]] = 0

be first-order differential operators, i.e. of the form (for a scalar particlei (for a vector particle) is given by the first order differential (for a scalar
partic~)

(a) (b)

T1/J =-ift<d1jJI~>

t =-ift[~;+]

(11.19) (11.20)

Show that the commutator [Tl,T2] operators (a) [Tl,T2]1/J _fi 2<d1jJ1 [~1;~2]>
(b)

[T1 ,T2]$ = -1i.2[it;[~1;;:2ll

(for a vector particle)

We are now in a position where we can investigate some symmetry transformations. In quantum mechanics the dynamical evolution is conA

trolled by the Hamiltonian

H , and a unitary transformation

is

called a symmetry transformation if it commutes with the Hamiltonian,

i.e.

[U,H]

This has the well-known consequence:


If
1/J

is a solution to the equations of motion, then


U1/J.

80

i8 the

transformed stat!!

596

II UH$ HU$

Proof:

Putting

$'=U$

we get

o
T
T

But symmetry transformations also lead J*T$dV


T

to conservation laws. If

is a Hermitian operator representing the physical quantity is conserved if J*T$dV

then

is the mean value of this physical quantity, and we say that is constant in time. We can then easily

show
If
U

is a one-parameter group of symmetry transformations, then

the infinitisemaL generator

is conserved.

Proof: From that TH=HT

UAH=HU

we get by differentiation with respect to A A i.e. the infinitisemal generator commutes with the Hamilto-

nian too. But then we obtain

i~d~I*T$dV

I{-i~:t$)T$dV
-fH$T$dV +

I*T{ind~$)dV
= I*[T,Hl$dV

I~H$dV

In our case U is generated by applying the space transformations A fA Consider the Hamiltonian (l1.2l) We know that fA
H
2 + =- n 2mll+V(x)

commutes with the Laplacian (theorem 13, section 10.5)

which leaves us with

Le.
A

UA commutes with

i f and onLy i f the potentiaZ

is invariant

under the transformation group

fA' i.e.

{fA)*V = V

11,3 CONSERVATION OF ENERGY, MOMENTUM AND ANGULAR MOMENTUM IN QUANTUM MECHANICS


Okay, it is time to look at some simple applications. In the present purpose), translations and rotations, {Cf.

R3

the

isometry group consists of reflections (which are uninteresting for (10.28.

Translations: In Cartesian coordinates a one-parameter group of trans-

lations is given by (11.22)

597 where a i

II

obviously are the components of the characteristic vector field. Since the components a i are constant, we get that the infini-

tisemal generator (l1.23)

reduces to the first order differential operator

T =-ifla j

~
ax]

(This is valid for both scalar particles and vector particles!) For translations along the Cartesian axis we get especially
Tl
A

=-lTI ax

,a

T2

=-iTI ay

T3 = -lTIa-z

,a

i.e. the infinitisemal generator of translations represents the operator of momentum! Furthermore we see that the momentum is conserved if the potential is translation invariant, i.e. if V is constant. This is certainly reasonably enough: In the presence of a non-trivial potential the 'particle will experience forces, and therefore its momentum will change.
Rotations: Consider first as an explicit example rotations around the

z-axis. This one-parameter group is generated by the matrix

"-I::;:

-::~:
-1

~l

i ..

f,(i,)-~~,
-y
x

By differentiation we get the characteristic vector field

o o
which is circulating around the z-axis.
Exercise 11. J.1 Problem: Show that the rotation matrix

:]

[~] [
-1

RA
with

can be rearranged as,

o
o o

Sz

o o

so that the characteristic vector field is given by,

~I

Sz~1 .

Consider the case of a scalar particle. Then the corresponding infinitisemal generator (ll.14) is the operator of orbital angular momentum around the z-axis (11.24) We get similar results for rotations around the z-axis and the y-axis, so we see that the infinitisemal generators for rotations are the operators of orbital angular momentum.

598

II

In the case of a vector particle things are slightly more complicated. Here the infinitisemal generator is given by the commutator, and we get an extra term

i.e.
(lL25)
J

- Yax ~] z =-in[x.1... ~y

+ in

5z

(Compare with exercise 11.3.1). The first term is again the orbLtal angular momentum, but the second term is something new. It can be interpreted as the spinoperator for the vector particle. Observe that the operator of ih 5 can only have the eigenvalues O,h i.e. the spin proz jection along the z-axis can only take the values O,l (in mUltiples h). So we see that the vector
p~rticle

posses an intrinsic spin

of 1 unit. (11.26)

(Compare with the discussion in section 3.6). The operator

t+

i.e.

Ji

= Li +

8i

is then interpreted as the operator for the total angular momentum. Observe that the angular momentum is conserved in both cases when the potential is rotationally invariant ( 1. e. when it only depends upon the radial distance r).

Exercise 11.3.2 Introduction: Let us introduce the skew-symmetric matrices components (11.27) Problem: (a) Show that (Si) k = Eijk
j

s.,
~

characterized by the

AS.

(b) (11.28) (c) (11.29) (d) (11.30)

e generates rotations arougd the xi-axis. Show that t~e matrices Si' i=1,2,3 satisfies the commutation rules
~

, [So ,S.] = E"


~

~J k

Let a. be the corresponding characteristic vector fields. Show that they s~tisfY the commutation rules
[;;:. ;;;:.]
~

Show that the operators of angular mom~ntum get the following commutation rules
~ ~ ~

E"

~J k

tK

[Ji,J j ] = i~Eijk J k (It should be emphasized that they hold both for scalar particles and for vector particles!)

Notice that it follows from exericse 11.3.2 that the operator of angular momentum for a ticle,
(9.50)
scal~r

particle, (ll.24), and for a vector par-

(ll.26), both satisfies Dirac's commutation rule

599

II

Exercise 11.3.3 Problem: Consider a vector particle. Show that the operator of angular momentum commutes with a spin-orbit coupling of the form , 4-4A Vl(r)L'S , 4where L is the operator of orbital angular momentum and S the spin operator.

...

It follows from exercise 11.3.3 that a Hamiltonian of the form A.fI2 ... S (11.31) H = - 2m~ + V{r) + Vl{r)L' is rotationally invariant. Coulomb potential and (This is the Hamiltonian for an electron V{r)
3

moving around a proton if we put

e2 V l (r) = 2m 2 c 2 r

-e 2 /r

corresponding to the

corresponding to the spin-or-

bit coupling). You can easily show that the orbital angular momentum and the spin are not separately conserved, because they do not commute with the spin-orbit coupling. But according to the above exercise 11.3.3 the total angular momentum will be conserved! The above discussion of the angular momentum can be generalized to more complicated systems. In the above cases the particle interacted with a scalar potential. We can also investigate what happens, when the particle interacts with a gauge potential, i.e. it interacts with the electromagnetic field. Ear simplicity we consider only the interaction of the electromagnetic field with a scalar particle. An electro-static field, like the Coulomb field, is generated from the electro-static potential ~(~), which is completely analogous to
f ...

the potential V(r) in the above discussion! In that case the operator of angular momentum is therefore given by (ll.24). A monopole field, on the other hand, is generated from a operator of angular momentum is given by (9.47) The correction term to the orbital angular momentum is determined by the similar correction term in the corresponding classical expression, \ cf. (9.39). (singular) vector potential A{~). In this case we have previously seen that the

~owing way too: The vector potential A entering in the Hamiltonian (9.43)

The correction term to the orbital angular momentum can be understood in the folis not spherically symmetric. When (fA) is a one-parameter family of rotations we therefore have (fA)~'; A But it generates the spherically symmetric monopole field, B = dA,and consequently we have i.e.

and

(fA)~

are gauge equivalent:

600
Differentiating with respect to

II

A at A = 0 we therefore conclude (cf. (11.18)):

( 11.32) -q[ii;it] = dx with X = q. ~~A [A=O We interpret X as the infinitisemal generator of the gauge transformation necessary to compensa~the rotation. It remains to identify X The characteristic vector field a associated to rotati'ons around the unit vector ft is given by
~

a = n x r

The corresponding operator for the angular momentum around the n-axis is according to eq. (9.47) given by

ft.~
We can now show
Lerruna 2

ft.!

+ fi,[-;xqA - Kr]

The infimtisemaL generator of the gauge transformation necessary to compensate for the rotation is identicaZ to the correction term to the orbitaZ anguLar momentwn, i. e. ( 11. 33 ) X = fi [-;xqA - Kr]
Proof: We rewrite the From this we get Using that a.a while
~

expressio~

= -(ftxr).qA -Knr = -q[ii'A


+ a.a
J

for

X as

=0
k. j + n J a.(1S..)]
~

a.A j i gM xk ~ a/ = 41f E ijk ? we can rearrange the above expression as gM i i aX = -q[aj(a.A ) - Aj(a.a ) + 41f(
~

E a ~Jk

S r

Here the last two terms cancel automatically When you work them out, and we therefore end up with

11.4 SYMMETRIES AND CONSERVATION LAWS IN CLASSICAL FIELD THEORY


Symmetries of a classical field theory has far reaching consequences as we have learned from the discussion in sections 3.9 and 3.11. Let us now examine it once more in a slightly more general framework. In what follows we will oniy consider diffeomorphisma of the manifold. Usually we pull back differential forms, but when we restrict ourselves to diffeomorphisms, we can equally well push them forward. Consider a field theory consisting of a field configuration and the action functional:
~a

Sn[~a]
Here
~a

JnL[~a]E

can stand for a collection of scalar fields or vector fields f we can push forward the spacetime region
~a'

or something more complicated, say a cotensor field. using a diffeomorphism

and the field

configuration

601 Definition 1 A diffeomorphism


(11.

II

is a symmetry transformation if

34)

Sn[<Pex 1 "" Sf<n) [f*<Pexl

for arbitrary regular domains <Pex .

and arbitrary field oonfigurations

This especially implies that f*<p ex is an extremum for the action, i.e. it is a solution to the field equation, whenever <Pex is. Following the arguments of the preceding section we now extend our considerations to families of diffeomorphisms: Defini tion 2 A one-parameter family of diffeomorphisms , (f)""ER' is a oolleotion of diffeomorphisms with the following properties: (1) fo = the identity map (2) For eaoh point P the ourve )"+f)..(P) is smooth, i.e. f).. depends smoothly upon )..

Observe that the one-parameter family (fA)AER generates a tangent vector field on M inacanonical fashion. To each point P we have associated a smooth curve f)..(P) passing through P at A=O , thereby generating the tangent vector df).. (P) +
(11.35 )
~

---crx-

ere it is understood that differentiation is carried out with respect to )..""0 !). The corresponding vector field ~ is called the oharaoteristio veotor fieZd associated with the one-parameter family of diffeo( H

morphisms (f)")AER' Suppose <Pex is a field configuration on Minkowski space and let us introduce coordinates (x~) on the region of space time in question

fA CU)
f-

A'
I U I

11
f-

'I
~l'

1- -

Fig. 219

-"- .1

602

II

Then

fA

is represented by the coordinate map A=O; i.e.

reduces to the identity map at


Ithat dy~/dA

y~ = fA~(XV) which fo~(XV) = x~ . Observe

are the components of the characteristic vector field

The space time region n corresponds to the coordinate domain U , and the displaced region to fA(U) . The field configuration ~a is represented by ordinary Euclidean functions ~a(X~) and so is the ordinary Euclidean transport of ~ a , i.e. (fA)*~ a are represented by for A=O functions too which we denote ~ A (x~) , reproducing ~a(X~) A a (The explicit form of ~ a will be calculated later on, when we need
it) .

Let us now specialize to a one-parameter family of symmetry transformations. Then we can finally state and prove the celebrated Noethers theorem:
Theopem 3 (Noetheps theorem) If a fieLd theopy with the canonicaL enepgy momentum tensor

~~v

(section 3.2) admits a one-papameter famiLy of symmetpy transformations

(fA)AR
(11.36)

then the cuprent

is conserved.

Here

aV

ape the components of the chapactepistic vecA=O!)

tor field,and ~aAthe components of the displaced field configuration.

(As usual we calculate the derivative at

Proof: The proof is quite technical, though interesting. You may skip it in a first reading. First we notice that the invariance implies

The displaced action can then be eValuated as

Here we run into the technical difficulty that not only the integrand, but also the integration domain depends on A. However,when we calculate the integral using n for all the coordinates we are free to use the same ccoPdinate domain U in R integrals involved. So let us transform the integral back to the A-independent coordinate domain U , performing the substitution y = fA(x) in the integral. This is legal because fA is a diffeomorphism. We then get Sf
A

(l)[(fA)*~aJ

J L[q,aA(y(x;k~~a\y(x;gai3(y)l,l-g(y(x))I~1
U
ay~ax

dnx

(It is important to notice that y We can then finally perform the


(11.

still depends on

371

0 - ~I

dS

A=O

J{[~ dL
a

~fferentiation and V dcj>a ClL d dX a ~ + a(d cj dA(~ ~ a Cly dX

A!). obtain:

7aA(y(x))

603
+

II

L[~ (I-g(y(x))

+ I-g(x)

~ Ifxl ]}dnx

Here we need some identities which we leave to you as an exercise:

Worked exercise 11.4.1


Problem: Remember that a) Let a

denotes the components of the characteristic vector field.

I~I

denote the Jacobiant. Show that

(11 . 38)

~Iffl =~ (~) = aIJ alJ dA ox dA axlJ

b)

i.e. it suffices to differentiate the trace of the Jacobi matrix, which in turn gives the divergence of the characteristic vectorfield. Show that

(11.39)

Here it is instructive to divide the terms into two groups. Those containing V and those containing a

= fU{A + B}/-g dnx A


~d

We want to rewrite

B separately ~ divergences. Consider first A: aL d~a A=a;p~+~alJ~ a IJ a Using the covariant equations of motion (8.35) A is immediately reduced to
oL d~a

( 11.46)

A =

----.T-g

a- (.T-g _ _ oL_.~) IJ a(dlJ~a) dA

Then we turn our attention to the second group: B = - aca o~ ) IJ a

av~a

0 a
IJ

+ La alJ + L----- a (.T-g)alJ + a'dL (avgall)a v IJ .T-g IJ gall v a).T-g) +

[o(~~~a) v~a + olJvL ]alJa +{L I~g

a!~1l

avgall}aV

But here you recognize an old fellRw in the first paranthesis. It is nothing but the canonical energy momentum tensor ~ IJ (Cf. (3.17)). The second term is due to the fact that we work in a covariant fonallsm where .T-g need not degenerate to a constant. We need still a useful identity:

Worked exercise 11.4.2


Problem: Show that the fOllowing identity holds, provided of motion:
( 11.42)
~a

solves the equations

604
Using (11.42) We can finally rearrange the B-term as: (11.43) B=

II

--La (;::gllaV )
I-g Il

Okay, it was a long tour de force, but now A and B has been rearranged as divergences and we obtain the identity which we have looked forward to find: (11.44)
As

aL ~ = dS dA = Ju[ 1 a {/-g(~ d" I-g Il lJ"'a

--=

+ aV Tv ll ) ~ I-g dnx O

U is arbitrary, this is consistent only if the integrand vanishes.

This is the main theorem! To apply it we must gain some familiarity with the current, especially the last part. Let us try to calculate d<PaA/d" explicitly. We consider first the case where
-1 .

<Pa

is a collec-

tion of scalar fields. Then by definition (f,,)*<P(x)

<P(y

(x,

i.e. <p a (y-1(y(x) = <Pa(x) " and the last term vanishes

<P~(f"
so

(x)

(f,,)*<Pa(y(x

<P~(fA (xl) does not at all depend on

automatically! We thus conclude:

Theorem 1/

(Noethers theorem for scalar fields)

If a field theory, involving only scalar fields parameter family of symmetry transformations lowing current is conserved:
(11.4S)

<Pa' admits a one-

(f"}"tR' then the fol-

where

o ... aV~ lJ i.e. J = T(a,') v is the canonical energy momentum tensor, and
Jil

... a

is the cha-

racteristic vector field associated with

(f"},,ER' <Pa=Aa' the last term

However, in the case of a vector field, i.e. no longer vanishes. By definition we get f*A (xl Il ax = --aylJ
v

A (y v

-1

(x

i.e. ax S -1 =-AS(y (y(x) aya axS/aya

<pa"(f,(X)= (f,l*Aa(y(x
A A

=~A

B aya

(x)

So in this case we get an extra factor

which does depend on

" ! If we perform the differentiation and use Exercise 11.4.1 (b) once more we get

i.e. this time we have a non-vanishing contribution to the current

605
which now looks as follows: (11.46)
J

II

aT,}

VO

3(3 A)
~

3L

A"

SOa

as

C<

that it can actually be rewritten on the form

We will return to this peculiar current in section 11.6 and show JI1 = T~ a v where T~
v

is the true energy momentum tensor (Cf. the discussion in section 3.2) . In the above discussion we have only considered symmetry transformations generated by space time transformations. We can however easily extend our discussion to internal symmetries as well, where an internal transformation is characterized by only affecting the field components. Suppose we have a one-parameter family of internal transformations, T,: <I> (x) ~ <I> A (x) "a a is a symmetry transformation provided it leaves the ac-

We say that T (11.47)

tion invariant, i.e.

for any space-time region Q. Then we can immediately take over the conclusions obtained in theorem 3:
Theorem 5 (Noether's theorem for internal symmetries) If a field theory admits a one-parameter family of internal symmetry transformations,
(l1.48)
J

(TA) .. ER' then the current d<l>A (x) 3L a d(d~<I>a) ----;rr-

is conserved. (As usual we calculate the derivative at A=O).

The standard example of an internal transformation is the phase transformation <I> (x) duces to
~

exp[iA]<I>(x)

~(x)

~ exp[-iA]~(x)

where <I> is a complex scalar field. The Noether current (11.48) then reJI1 _ dL d<l>A ~.M.A - 3(d~<Id)" + 3(d~~)dA
=l.'t'~

. ['" dL

11

which is nothing but our old friend (3.67) : We can also generalize the concept of a symmetry transformation. In the preceeding discussion we have assumed that the displaced action is identical with the original action. From a dynamical pOint of view, however, two actions will lead to the same dynamics provided they just differ by a boundary term (i.e. their Lagrangian densities differ by a divergence). We may therefore say that a space time transformation is a symmetry transformation in the generalized sence provided it satisfies: (11.49)

606 II If we have a one-parameter family of symmetry transformations in the generalized sence, (f) tion of the Noether current. This time the displaced action is not constant, so we must first subtract the surface term, which can be rearranged on the form

A AER

, we must now be a little careful in the deriva-

This produces the following additional term to (11.44): dS;n dA

-J
n

a].1{dAA[~al}v=g

dX].1

dnx

Consequently the Noether current must now include an addtitional term of the form dX\

-cD:"" [~al
To conlude we have therefore shown:
The most genepal Noethep cuppent consists of thpee tepms: (aJ A piece,
(11.50)
1

o
JIl = aVT Il
V '

which comes fpom the displacements of the space-time points. (b) A piece,
(11.51)
Il _ _

"_L_ d~~
a
II

"(

) dI-

which comes fpom the displacement of the field. (Notice that it is genepated paptly fpom the space-time symmetpies and paptly fpom the intepnal symmetpies involved.}
(c)

A piece,

(11.52)

J~

which comes from the boundapy tepm in the dispZaced Lagpangian.(As usual we calculate the depivative at A=O}.

11.5 ISOMETRIES AS SYM~lETRY TRANSFORMATIONS


We can now investigate the consequences of the transformations in the isometry group. Consider Minkowski space equipped with a field configuration
~a.

The dynamics of

~a

is controlled by an action

principle, i.e. we have a Lagrangian density constructed out of the fields and their first order derivatives

L = L[g'~a,d<l>a]
The crucial assumption is that
on

L[!1,<I>a,d<l>al
~L~

itself is a scalar field

M, so that the action integral

is covariant. This guaran-

607

II

tees that the field theory becomes covariant, i.e. the equations of motion will be covariant etc. Now, using a global isometry pull back the field
f*~a

we can

~a

and_obtain the transformed configuration

. The important thing to recognize then is that


f*L[g'~a,daJ

(11.53)

L[g,f*a,df*aJ

This is not true for an arbitrary diffeomorphism! It only works because f is an isometry (in general we get Consequently we get (11.54)
SQ[f*~aJ

L[f*g, ... J on the right hand side) .

JQ*L[g,f*a,df*~aJ

JQ*f*L[g'~Ct,d~aJ

Jnf*(*L[g'~a,d~aJ)

Jf(n)*L[g'a,d~aJ

Sf(Q) [~aJ because f is an isometry and therefore commutes with the dual map. We have thus shown

A Poincare transformation is a symmetry transformation for any fieZd theory, where the Lagrangian density is a scaZar fieZd constructed entireZy out of the fieZds and their first order derivatives. soZution by appZying a Poincare transformation.
The converse is not true. A specific field theory may very well admit a larger symmetry group than the isometry group. See exercise 11.5.1 below.

If

~a

is

a soZution to the equations of motion, we can transform it into another

Worked exercise 11.5.1 (Compare with exercise 10.6.2 and 10.6.3) Problem: (a) Show that the theory of a massless scalar field in (l+l)-dimensional space time is conformally invariant, i. e. the conformal transformations are symmetry transformations. (b) Show also that the theory of electromagnetism in (3+1)-dimensional space time is conformally invariant.
I

Now we are ready

\0

apply the machinery from the preceding section.

All we have to do is to extract some suitable one-parameter families out of the full Poincare group. In what follows we restrict to inertial coordinates.

ExampZe 1: The group of transZa~ions: A one-parameter group of translations is given by


(11.55) where

a~

are the components of a fixed four-vector

~. Clearly this

constant vector field is the characteristic vector field associated

II 608 with the one-parameter group. Consequently we get from Noether's theorem that the current

is conserved, i.e.

o = a J~ = aV[a
~

v ~]
a~

(The last term in (ll.4G) vanishes because

is constant). As

aV

is arbitrary, we conclude that the canonical energy-momentum tensor itself is conserved. As a result we see that invariance under the transZation group impLies conservation of the canonicaZ energy and momentum. Example 2: The group of Lorentz transformations.

By definition a Lorentz matrix is a 4x4-matrix (6.32)

ATnA = AnA T = n

satisfying
=

i.e.

ATrl

nA- l

This is not a useful characterization when we want to construct oneparameter groups of Lorentz transformations. In analogy with the oneparameter groups of unitary transformations (section 11.3) we will try to construct one-parameter groups of Lorentz matrices on the form exp[Aw] ded with some matrix exp[w]
w

Here it is useful to observe that

is a Lorentz matrix, provi-

w satisfies

w n -nw i.e. nw is skew-symmetric. If we write the matrix elements of w as w~ and if we raise and lower indices using the metric components, \) (Obthen nw is characterized by the "co-variant" components serve that w~ are not the components of a tensor!) ~~ Whenever nw is a skew-symmetric matrix we can therefore construct
a one-parameter group of Lorentz transformations as
(11. 56) y, = exp[Aw]x, By differentiation of this we get (11.57)
=

=t=

-~ a, d)'

= wX 1

i.e. the characteristic vector field nents


(11.58) a
V

..,.
a

is characterized by the compo-

Exercise 11.5.2 Introduction: Let uS introduce 6 matrices labelled where is skew-symmetric in (po). These matrices are goin~Oto be_chara@~erized by the "co-variant" components n n -n n i.e. n itself is characterized by the components: ~p va ~o vp po

n ,

(npo)~v = 8~pnvo 8~on~p

609

II

Problem:

a)
b)

Let (ijk) be an even permutation of (123). Show that exp[AQ .. J generates rotatiQns around the k'th axis through an angle A .1J Show that exp[An kJ k = 1,2,3 generates a special Lorentz transformation a18ng the k'th axis with the velocity v given by the relation Cosh A = _1_ Il-v 2 The parameter A is called the rapidity and unlike the velocity v it is additive under composition of special Lorentz transformations in the sawe direction.

J~

Let us co.nsider scalar fields first. We then get fro.m theo.rem 4 that aV~ ~ is co.nserved, i.e.
v

o
As
W

~ a J~ ~
~

vp

(0 xpv~)
~

Vp

is an arbitrary skew-symmetric matrix we co.nclude that

(11.59)

~ a [xp~V~_Xv~p~]
~

ConsequentZy invariance under the group of Lorentz transformation implies the conservation of the angutar momentum!

It is co.nvenient to. intro.duce


(11.60) Then ~pv~

~pv~

xpv~_xv~p~

co.mprises the currents o.f the angular mo.mentum density,

but o.bserve that it is net a co.variant expressio.n, net even if we restrict o.urselves to. inertial frames. Fro.m the co.nservatio.n o.f the angular mo.mentum we new deduce

o =
Here the with
(11.61)

a MPV~
~

0 P~v~_o V~P~+xp(a TV~)_Xv(o ~p~)


~ ~

last

two. terms vanish since

v~ T

is co.nserved and we end up

i.e. the cano.nical energy mo.mentum tenser is

b~rn

symmetric. There is

co.nsequently no. need to. repair it. In the case o.f a scalar field the
canonical energy mom9ntum tensor is the true energy momentum tensor!

In a geo.metrical language we start o.ut with the inertial frame o.f

epo

reference co.mpo.nents
(11. 62)

S . We then asso.ciate to. (pO)

6 vecto.r fields

is skew-symmetric in

. Here"" e po

e po ' where is characterized by the

..,.

i.e.
tens~r

When we co.ntract them with the energy momentum served currents:


(11.63)
i.e.

we o.btain 6 con-

(J

po

)~ = x p ~ 0 ~-x 0 ~ p ~

which are interpreted as the currents o.f angular momentum relative to.

610 the xP-x o -plane. If

II

denotes a spaceslice relative to

then

the associated charges (11. 64)


J

po

-J n *J po
x p -x 0 -

are interpreted as the total angular momentum relative to the plane.

Worked exercise 11.5.3


Introduction: inertial frames of references SI and S2 and the assOangular momenta JI , J2 given by the above formula inertial coordinat~~ (xa)Poand (ya) associated with are assumed related through the Poincare transformation ya = Aa )l + ba B Show that the basic vector fields and ~2 are related through po the formula -+1 .... lJ v\,l +2 - b ~2] e ejl\.,A pA 0 + [b P~2 po o p 0 Show that the total angular momentum measured in Sl and S2 are related through .. v J2 JI v'W + [b p2 - b p2] po P0 0 p IN A pA 0 p2 where is the total energy momentum measured in S2 jl Consider two ciated total (11.64). The SI and S2

Problem:
(11.65)

a)

b)
(11.66 )

Then we turn our attention to vector fields. Here the conserved current is given by (see
(~1.46)

and (11.58))

Jjl

=w

(p&Vjl _ aL AV) vp x a(a A ) jl P

which leads to the conservation of the slightly more complicated angular momentum (11. 67)

~PVjl _ [ pOVjl_ VOPjl] xT xT +

ra(aA)A aL p
L

jl V

aL a(aA)A jl P -

vl

Here the first term represents the conventional orbital angular momentum of the field, but what is the origin of the last term? From a physical viewpoint the main difference between a scalar field and a vector field is that a scalar field carries no spin, while a vector field carries spin 1, (compare the discussion in Section 3.6). This is reflected in the fact that the scalar field carries no space-time index, while the vector field carries one. This again leads to the different behaviours under symmetry transformations. formation property of Aa It is the nontrivial transwhich on the one hand allows us to construct

eigenfunctions for the rotation operator with non-trivial eigenvalues, while it on the other hand generates the last term in (11.67). Now, for a field carrying spin we expect the total angular momentum to be composed of a spin part and a contribution from the orbital angular momentum:
J

L+S

611

II

and it is only the complete angular momentum which is conserved, not the separate terms. Okay, this suggests then that the last term is interpreted as the spin density.

11.6 THE TRUE ENERGY-MOMENTUM TENSOR FOR VECTOR FIELDS


We have just seen that the conserved current (11.46) for a vector field has a complicated structure, involving not only the canonical energy momentum tensor, but also a spin-contribution. We also know that the canonical energy momentum tensor for a vector field fails to be symmetric, and has to be repaired (cf. the discussion in sections 3.5 and 3.7). We will now extract the true energy momentum tensor from the expression (11.46) Jil
=

v v il _

a (alJA a )

dL

A a as S a

The main clue is the observation that the expression (3.17) for the canonical energy-momentum tensor is not covariant, i.e. are not the components of a tensor. factor
o jJ Tv

(Remember we are working in arbi-

trary coordinates on an arbitrary manifold!). The trouble lies in the avAa' where we have differentiated through a vector field which immediately destroys the covariance (compare the discussion in section 7.1. For a vector field it is only the field strength FaS = aaAs-asAa which is covariant.) However by construction the complete expression for the current JjJ is covariant, and it would be nice to have this covariance explicitly built into the formula (11.46). This is where the last term comes into the game. It can be rearranged as follows:

Wopked exepcise 11.6.1


Problem: Prove that

---~--- are the components of a skew-symmetric tensor.


a (ajJAa)

Using the covariant equations of motion (8.35) we now rearrange the

612

II

Inserting this we can finally decompose the current into two parts
(11.69)

Jjl

jl
J (2)

av -A v aA jl

[-0

-l-aa[~o(~LA
r-g
a (a A ) (Ava ) jl Ct S

jl a

V )AVa ]

which are manifestly covariant. From exercise 11.6.1 we learn that


(11. 70)

are the contravariant components of a 2-form therefore given by


(11. 71)
(2 )

and the last part is

from which it immediately follows that it is conserved. Furthermore it does not contribute to the total charge since

Q
(2)

= -fn*

J
(2)

=-f n*5S = fan*s

where we have used Gauss' theorem. Here the last integral vanishes provided the vector field dies off sufficiently fast at infinity. But then we can simply split off the second part. If we furthermore introduce the abbreviation
(11. 72)

jl
Tv
=

-0 -aAAv jl
=

3L jl 0(0 A )FVCt + 8 v L

jl a

we can now
(11.73)

express the first component as

J
(1)

T(~i)

We want to show that

TjlV

defined by (11.72) is the true energy mojlV

mentum tensor. Consequently we must show that it is symmetric, conserved and produces the same energy and momentum as does cussion in section 3.2). Since (Cf. the disis conserved is con-

and

J J

are conserved it follows that

too. From translat66al invariance we therefore get thAe) TjlV served. Furthermore TjlV and ~jlV produce ved angular momentum

does not contribute to the total charge, so same total energy and momentum. Finally i t folTjlV

ffi~

lows from Lorentz invariance that

produces the following conser-

i.e.

o=
Consequently
TjlV

0 jl M

PVjl

T VP - T Pv

is a symmetric tensor.

613

II

Exercise 11.6.2
Problem: a) Consider the theory of electromagnetism based on the gauge potential A . Show that (11.72) rep~oduces the true energy momentum tensor (1.41). U b) Consider the theory of a massive vector field W . Show that (11.72) reproduces the energy momentum tensor given by e~ercise 3.7.1).

Using the true energy momentum tensor for vector fields we can now reformulate Noether's theorem in the case of vector fields:

Theorem 6 (Noether's theorem for vector fields) If a field theory based upon a vector field Au admits a one-parameter family of symmetry transformations (fA}AtR > then the following current is conserved:
(l1.73)

JU
(1)

aVT U
v

i.e.

J
(1)

where T given by (11.72) is the true energy momentum tensor and where ~ is the characteristic vector field associated with the one-parameter family. Exercise 11.6.3
Problem: Consider a mass-less scalar field in (l+l)-dimensional space time respectively the electromagnetic field in (3+1)-dimensional spacetime. (a) Show that the conserved current and charge associated with the invariance under dilatations are given by (11.74)

(D

=1

or 3)

(b) Show that the conserved currents and charges associated with the invariance under special conformal transformations are given by (11.75) (c) Show that the dilatational charge transforms as a scalar under Lorentz transformations and that the special conformal charges transform as the components of a Lorentz vector under Lorentz transformations. How do they transform under general Poincare transformations?

Exercise 11.6.4
Problem: (a) Show that the energy momentum tensor of a conformal invariant field theory is trace-less, TU = 0 . (Hint: Consider the conservation law u corresponding to dilatational invariance.) (b) Consider a field theory possessing an energy-momentum tensor which is trace-less, TU = 0 . Show that the charges (11.74) and (11.75) are u conserved.

614

II

11.7 ENERGY-MOMENTUM CONSERVATION AS ACONSEQUENCE OF COVARIANCE.


In this section we will finally take up the problem of how to construct the true energy momentum tensor from a more general pOint of view. We saw in section 3.11 how gauge invariance could be used to construct a general expression for the electromagnetic current. In that case we have a Lagrangian depending on a charged field Ajl
~
~

coupled to an exter-

nal gauge potential Ajl . Performing a variation of the external field, Ajl + EOAjl , the corresponding variation in the action will be liOAjl' oS dS(E) dE [10=0
8A

near in

and we then showed that the coefficient in front of identical to the current: (3.76)
dL

jl

actually was

aA jl

Furthermore the gauge invariance of the action allowed us to prove that this current would be conserved as a consequence of the dynamics of the charged field tion
~a.
~

In the present case we are considering an arbitrary field configuraTo construct the Lagrangian, which has to be a scalar field, we furthermore need the metric tional of the general form

g. We then construct an action fUnc-

S = JL[g,$a,d$al~dxlA . Adxn
(Observe that the metric is involved in the construction of the volumeform too!) The basic idea is to treat the metric as an extermal field and see what happens when we perform a variation in the metric: gjlV
~

gjlv+EOgjlV . The corresponding variation of the action will be liogjlV and can therefore be written on the form oS dS(E) - dE: 110=0 Tjlv(x)

near in (11.76)

~JTjlV(X)Og

jlV

(x) Fgdnx

The coefficient

is a symmetric tensor and as we shall see it

reduces to the true energy momentum tensor 'in the now well known cases of scalar fields and vector fields. To obtain an explicit expn,ssion for TjlV in terms of the Lagrangian we need some useful identLties involving the derivatives of the metric.

Worked exeroise 11.7.1


Problem: Deduce the following identities (11.77) a) djl[lnDetM(x)] where M(x)

= tr[M-l(X)djlM(X)]
is an arbitrary matrix function.

II

(11.78) (11.79) (11.80)

b)
c)

d)

Using (11.79) we now get


OS = _ _ d_ ILFgdnx

dEjE=O
dL f[ agjlV

~gjlVL]og

I1v

Fgdnx

from which we read off:


(11.81)

TjlV

2~ + gjlV L ag jlV

As an example we take a look at the electromagnetic field. Here the Lagrangian density is given by (cf. (3.36))
L

= -!F F PO = 4 po

-!F F g pa oS 4 po as g TjlV defined by (11.81) will auto-

Note that find

is gauge invariant, so

matically become a symmetric gauge invariant tensor. Using (11.80) we

aL ag jlV

= -!F

4 po as

[_gI1PgvagOS_gPagI10gVS]

=-!Fjl F av 2 a

Inserting this into (11.81) we then end up with TjlV = _ Fjl F av _! g jlV F F Po a 4 po and
tha~

is precisely the true energy-momentum tensor for the electromag-

netic field (cf. 1.41). Motivated by this example we will call the tensor defined by eq. (11.81) the metl-ic energy-momentum tensor.
Exercise 11.7.2 Introduction: We consider a theory consisting of a single scalarfield ~ and assume that the Lagrangian density is constructed as a function of the two scalar fields

and

gaSaa~aS~

(d~jd~)

i.e. L = L(~;gaSaa~aS~) Problem: Show that the metric energy-momentum tensor (11.81) coinsides with the canonical energy momentum tensor (3.17) [Hint: Show that

II 616 Exercise 11. 7. J Introduction: We consider a theory consisting of a single vectorfield A and assume that the Lagrangian density is constructed as a function o the two scalar fields,

e
where

= ga8A A = (AlA) and a 8 FaS is the usual field strength, aAS-OSAa'


L = L(ga SA A 'gaPgSoF

FaS
i.e.

F ) a 8' a8 po . Problem: Show that the metric energy-momentum tensor (11.81) coincides with the true energy-momentum tensor (11.72). Worked exercise 11.7.4 Introduction: We consider a system consisting of a single relativistic particle and assume that the Lagrangian density is constructed as follows (11.82) Problem:
I dxU dxS L =-m ,<gaS(x)TA aT

1 --=--0 (x-x(A))dL

I-g

where Xa(A) is a parametrization of the particles world line. a) Show the La ran ian densit (11.82) leads to the conventional action: S =-m as(x(A))dI"dI"dA. b) Show that the metric energy-momentum tensor (11.81) coinsides with the conventional energy-momentum tensor (cf. the discussion in section 1.6): a 8 1 IlV ' T = m dT dT --=--o"(x-x(T))dT .

dxU dx

Jdx dx

I-g

As shown by the exercises 11.7.2-4 the metric energy-momentum tensor coincides with the true energy-momentum tensor for a system consisting of particles, scalar fields and vector fields. Interestingly enough we can now give a general proof for the conservation of the metric energymomentum tensor based upon the covariance of the action. The proof is somewhat similar to the proof of Noether's theorem in section 11.4. Especially it is somewhat technical and you may skip it in a first reading:
Theorem The metric energy-momentum
(]1.81)
~nsor

T IlV

= 2~
og\.lv

+ gllVL

will be conserved for any field theory based upon a covariant Lagrangian( i.e. the Lagrangian is a scaZarfield constructed from the metric, the fields and their first order derivatives). Furthermore the conservation of the energy-momentum tensor can be expressed in a covariant way
, as

(11.83)

617
Proof: From covariance we get

II

* * * * Sn[~~l = JfA(n)L[fAg;fA~a;dfA~alfAE
where fA is an arbitrary one-parameter family of diffeomorphisms. For the present purpose we choose the diffeomorphisms flo. so that they map the interior of n into itself, and such that flo. reduces to the identity on the boundary. Observe that this especially implies that the characteristic vectorfield ! vanishes On the boundary. Introducing coordinates we now get

Sn[~~] = Jlf[g~)y);~~(y);~~(y)]l-i(Y)dny
(Observe that the coordinate domain this time is independent of A from the beginning). We assume now that ~ (x) is a solution to the equations of motion. This has the important consequence, ~hat ~A does not contribute to the variation of the action integral. (Observe that the figld this time really is varied. In the case of a scalarfield we get for instance: ~A(y) = ~ (f~l(y))). When we differentiate we therefore need only take into considera~ion the~new variations coming from gA
IN

We must then

with

x = f~l(y) . Differentiating exercise 11. 4.1b) A

at

10.=0

we get (using

dg)lV ~ = -( a a )g dA jl av

Inserting this we find TjlVas/-gldny jlV Performing a partial integration on the first term we now end up with

o =

r [Tjlv;.:ga Va jl Ju

+ rB

o = JU[/~gav(l-gTSV)
As the variation as

+ rSjlVTjlV]aS/-gdny

is arbitrary this is consistent only if

(11.83)

-1-0 (l_gT SV ) + r S TjlV r-g \) jlV

=0

Finally we specialize to an inertial frame. Then the identity (11.83) reduces to

av TSv = 0
which is the standard expression (1.38) for the conservation of the energy-momentum tensor.

From theorem 8 we learn too that the covariant expression for the divergence of a symmetric rank 2 tensor is given by

618
(*)

II

-1-a (!=gTBV)
I-g v

rB T~v
~v

This should be compared to the covariant expression for the divergence of a skewsymmetric rank 2 tensor:
(**)

-1-a
I-g
V

(l_gF Bv )

In fact (*) is valid for a skewsymmetric tensor too, since in t~at case t~e co~tribu tion from the last term vanishes due to the symmetry of the Chrlsto~fe~ fleld ln the indices (~v) . Thus (*) is valid for arbitrary rank 2 ~ensors: Th~S lllustrates a general principle: The covariant expression for the partlal derlvatlves.of a tensor will contain the Christoffel field, but once we restrict to skewsymmetrlc tenso:s the contributions from the Christoffel field vanishes automatically and we end up wlth the usual expressions from the exterior algebra.

u.g

SCALE INVARIANCE IN CLASSICAL FIELD THEORIES


we will show how conformal transformations can be
The com-

In the final sections

combined in a natural way with an internal transformation. and interesting class of classical field theories.

bined transformation then acts as a symmetry transformation for a large As a preparation we consider scale transformations (dilatations) n-dimensional Minkowski space. scale transformation is given by Using inertial coordinates such as a in

(11.84)
Under a scale transformation a scalar field transforms as

The displaced action is therefore given by

Changing variables,
5 D (Q)

ya
D*q, =

AXU, this is rearranged as


L[q,(X);1" -

[ ] Jn

ax~

q,(x}] A d

nn

Consequently a scale transformation is a symmetry transformation provided:

(l1.8S)
The standard Lagrangian of a scalar field is on the form
L = -

l(a )J q,} (a~q,) -

U(q,}

619 Usually the potential energy density


U(~)

II

only consists of mass term, If

lm2~2, but in general it can be an arbitrary positive function.


n must be 2.

the theory is scale invariant, it follows from the kinetic term that Furthermore the potential energy denSity must vanish. Thus we are left with the trivial example of a massless free scalar field in (l+l)-dimensions! Let US now combine the scale transformation with the internal transformation,

Here

is a real scalar, called the scaling dimension of the field. The displaced action is then changed

This transformation is very similar to the phase transformation in the theory of complex scalar fields. to SD(I1) fA -d
D*~l

= JI1Lf).. -d ~(x)

-d-1 n n ; A a\l~(x} lA d x

Consequently the combined transformation is a symmetry transformation provided (11 .86)


=
Lf~ (x);

a\l <I> (xl

If the theory is scale invariant it follows from the kinetic term that
(11. 87)

--r

n-2

(Scaling dimension of a scalar field)

This number is then referred to as the scaling dimension of a scalar

field.

The potential energy term must furthermore satisfy the homo-

geneity property: AnU[ ;..--2-- ~l= U[~l


1 -

2-n

Substituting on the form: (11 .88)

and

x = A

'2

we therefore see that

must be

2n n U(x) = U(1)x - 2

n 2. 3

E.g. in 4 dimensions this means that a term like

is admissible.

Notice, however, that a mass term is never admissible! A


IJ

The same analysis can be applied to a vector field placed field is given by

The dis-

620 This leads to the displaced action:

II

A scale transformation is thus a symmetry transformation provided


(11.89) If the theory is scale invariant it follows from the kinetic term, 1 F F~v that 4" ~v ' d = n-4 (11.90) (Scaling dimension of a vector field) . -2-

This number is then referred to as the sea Zing dimension of the vector

fieZd.

Furthermore a potential energy term must satisfy 2-n )..nU [)..-2-- A 1 = U[A 1
~ ~

This is the same condition as before, i.e. in 4 dimensions this means that a term like, U[A 1 = . ~. (A A~)
~

1fT

is admissible, while e.g. a mass term is never admissible. Notice that with the above choices of scaling dimensions, the conditions (11.86) and (11.89) actually coincide!
~a

If we in general let

denote a collection of scalar fields and/or vector fields, the

preceding arguments thus show that a theory is scale invariant provided the Lagrangian satisfies the homogeneity property (11. 91 )
)..n

a' This leads to the following rule:

L [ )..-2- ~.

2-n

)..

2" d~ C< 1

A theory in Minkowski space is scaZe invariant provided each term in the Lagrangian scaZes Zike )..-n. The scaZing of a term is found by scaZing each fieZd Zike f~-W/2, and each derivative Zike A- 1 .
As an important application of this rule we consider the case of a complex scalar field interacting with the Maxwell field through the rule of minimal coupling. modified to In this case the standard kinetic term is

The first term presents no problem, but the two last terms scale like

)..(4-3~/2
iour in

and n=4

)..4-2n.

Thus they only have the correct scaling behavWe therefore conclude:

dimensions.

621 Minimal coupling is only scale invariant in 4 dimensions.

II

In standard physics literature the above criterium for scale invariance is often stated in a slightly different way using dimensional analysis of the individual terms in the Lagrangian. In units where ~ = c = 1 the action S itself is dimensionless. [In general S and ~ has the same dimension so that S/~ is dimensionless. Cf. Feynman's rule (2.20)]. Any physical quantity T will now have a dimension, denoted [T] , corresponding to a power of a length. E.g., the volume element dllx has dimension n. Each term in the Lagrangian density consequently has dimension -no Consider first the kinetic term - !3,,$alJ $ or - ~ F FIJV ~ 4 IJV Since a derivative corresponds to a division by a length, it reduces the dimension by 1. Consequently
- n

= 2 ([<I>]

- 1)

i.e. the dimension of


(11.92)

<I>

(and similarly of
2

AIJ)

is given by
= -2-

[$] =~

2-n

Notice that the dimension of the field coincides with the scaling behavior of the field (cf. (11.91. Except for the kinetic term all other terms will now contain coupling constants, which mayor may not carry a dimension. Consider e.g. a mass term !m2 $2 In that case we get 2-n -n = 2([m] + [<1>]) = 2([m] + -2-) i.e.
[m] = -1.

Consequently the mass parameter m has dimension -1 in any number of space time dimensions. Any coupling constant, which carries a dimension therefore generates a "mass-scale", i.e. a suitable power of the coupling constant has the dimension of
mass.

As a famous example we consider Newton's constant of gravity. As we have not yet developed a field theory of gravity, we will use Newton's second law, m--=Gdt 2 r2 as a starting point. Notice that in units where e=l time will have the same dimension as length. Furthermore the mass has dimension -1, so that we get
[G] =

ix

mM

[~] 2
dt

+ 2[r] - [M]

-1 + 2 + 1 = 2

The mass scale in gravity is therefore given by


(11. 93) G-!

(/

= 5.5 10- 8

kg)

which is known as the Planck-mass. On the contrary a dimensionless coupling constant can not in any way be related to a mass: It simply generates a pure number. Consider, e.g. the coupling constant in electromagnetism. In four dimensions, A has dimension -1. It follows from consideration of the covariant derivative, IJ d - ieA ,that e is dimensionless. Reintroducing ~ and c, the pure number itlJrepres~nts, is given by
(11.94) 2 1 e ftc ~m

which is known as the fine structure constant. Returning to the criterium of scale invariance, we can now reformulate it in the following way:

622

II

A theory is scale invariant precisely when it has the following two properties: 1) It contains no mass terms; 2) AU coupling constants are dimensionless. (By abuse of language a theory with these two properties is often referred to as a massless theory).
Given a scale invariant theory there exists a corresponding conservation law. This can be found from Noether's theorem, where the dis$a(x) transforms placed configuration now includes the action of the internal transformation. Consequently a general field configuration as follows

where

is the scaling dimension.

We thus get an additional contrie

bution to the conserved current coming from the factor group of dilatations is given by

-Ad

in front.

Since the characteristic vector field associated with the one-parameter

aV
the conserved current reduces to (11.95 ) This is a nice except for one little detail.
J~

Unlike the case of

ordinary space-time symmetries, the current is no longer on the form

= aVTv~'

This is very analogous to what happened in the case of

Tector fields, where we also obtained an additional spin contribution and consequently had to repair the canonical energy momentum tensor (cf. (11.46. This suggests that even the true energy-momentum tensor needs to be repaired in prder to absorb the additional term in (11.95). Actually it can only be repaired in the case of scalar fields. This is due to a different structure in the two cases. case the additional term is simply a gradient: (11.96 )
3L

In the scalar

- d$

~
~

= d$3 ~ ~
3L

In the vector case we similarly get (11.97 ) (cf. exercise 11.6.1). _ dA

a 3(3~Aa)

But this is not a gradient!

623 To see this let us temporarily assume it was a gradient, A F].lCl. ~ a].lX[A 1 CI. CI. In that case the functional Tn[ACl. 1= la].lACl.F].lCl.dnX where a ary term].l

II

is an arbitrary constant vector, can actually be converted to a pure boundn 1 faX[A ]n].ld - x an].l CI. is independent of variations in ACI. perform such a variation, ACI. ~ ACI. + EBCI.

Tn[A 1 = CI. Consequently the functional Tn[ACI.] they vanish on the boundary. If we

as long as

[A (a].lACI._a].lACI.) + E{B (a].lACI._aCl.A].l) + A (a].lBCI._aCl.B].l)} + E2B {a].lBCI._aCl.B].l}ldnx, CI. CI. CI. CI. We therefore obtain dEdiT = a].l6 E=O Performing a partial integration this is rearranged as o= a aCl.A].l + g].lCl. a AV]B dnx ].l v CI. This is only consistent provided ACI. satisfies the identity aCl.A].l = g].lCl. a AV v Taking the trace you immediately see that aCl. AV must vanish! This contradicts our assumption that (*) was supposed to hold for an arbitrary configuration!

o=

f[-

As we shall see later, this difference between the scalar case and the vector case also has the consequence that scale invariant scalar theories are actually conformal invariant whereas scale invariant vector theories are not (except in the trivial four-dimensional case). In the following we shall therefore confine ourselves to the scalar case. As usual we modify the energy-momentum tensor by adding a term satisfying the requirements (3.18). tion is a bit tricky. This time, however, the construcIt is based upon the properties of a particular
d~f.

tensor constructed directly from the metric: (11.98)


gCl.~y6 gCl.yg~6

gCl.6g~y

Notice the following important symmetry properties (11.99a) (11.99b)


g~Cl.y6 gCl.~y6

gCl.~OY
gy6C1.~

=-

gCl.~Yo

[.'.w-'''''''' i. ,h. pair of indices

first and last

symmetry between the first and the last pair of indices

Using this we now construct the following correction term (11.100)

624 From (11.99b) it follows that

II

e~V

is symmetric in

and

Furthermore (11.99a) implies that

a~()~v
eOo = _ which shows that
e~v

= 4(~-!/)
(n-2)
i

apr

aAa~gA~Pv</>21
e io

Finally we get by explicit computation that


4 (n-1 )

a a i ",2
'I'

does not contribute to the total energy and

momentum. Thus (11.100) is an admissible correction term! Using the tensor gA~PV we can also construct a conserved current: (11.101) (Notice that the bracket is skew-symmetric in expression can be rearranged as
J~
6

A [g

and

~).

This

4lIl=1T
X 6~v _

(n-2)

Xv

[a a gA~PV",2]
A P
'I'

4lIl=1T

(n-2)

Av P

a (gA~PV",2)]
'I'

(n-2) 4

a~</>2

On the other hand scale invariance led to the following conserved current (cf. (11.95-96)).

J~
unwanted term:

- Xv

T~v

(n-2) ~~",2 + ~ " 'I'

By adding these two conserved currents we thus precisely cancel the

Notice that since 0

J~

is conserved, we get

a (x T~V) = T ~ + xva~ T~v ~ v ~ But here the last term vanishes trivially because

j~

T~v

is conserved.

Consequently the modified energy-momentum tensor is traceless! Let us summarize the main outcome of the investigation:
Any scale invariant scalar theory in ordinary Minkowski space pos-

sesses a conserved, symmetric, traceless energy-momentum tensor


(11.102)

T~v

= T~v

(n-2) [ 4(n-1)

~v

[]
The associated conserved

known as the conformal energy-momentum tensor. current


(11.103)

J~

x T~v

is known as the scale current.

625

II

11.9. CONFORMAL TRANSFORMATIONS AS SYMMETRY TRANSFORMATIONS

Now that the group of scale transformations is under control we return to the problem of conformal invariance in a classical field theory The main difficulty is that we have to combine the conformal transformation with an internal transformation. ternal transformation in some detail: suppose that f:M
~

Let us first look at this in-

M is an arbitrary manifold equipped with a metric g and M is a suitable diffeomorphism. We would like to assign
f. If
g

a scale to the transformation

and

f*g

happened to be pro-

portional, the constant of proportionality would define the square of the scale, but in general there need not be any simple connection between
g

and

f~g

Therefore we turn our attention to the volume

form instead. (11.104)

Since

and f*c

f.

both are n-forms, they are necessar-

ily proportional:

= Jf(x)E
Jf(x) to the f preserves the orientation, then

Consequently we can always attach a volume expansion transformation f. Notice that if according to (10.34)

f*t reduces to Ef (where ' f is the volume form generated from the pull-backed metric f*g). This justifies the name of J f (x). The scale " associated with a scale transfonnat1.on can now in -the general case be replaced by the scale "f(x)

= [J f

(X)]1/n If f is an isometry,

Let us look at a few illustrative examples. then f*g

=g

and consequently
J
f
(x)

= 1

according to whether larly, if f consequently (11.105) according to whether (10.40) )

preserves or reverses the orientation. f*g

Simiand

is a conformal transformation, then

= Q2(X)g

preserves or reverses the orientation (cf. Jf(x). From

We can also work out a coordinate expression for ( 1 0 8 ) we get

626 [f* tl (x)


1 . n

II

Ig((x

E.

J 1 .. I n

~
ax 1

fg((x) ) where

I~I ax
f. Consequently

is the Jacobiant of the transformation

the volume expansion is given by: (11.106) f,g(f(x)), I~I

rIglX)I

ax

Notice that in ordinary Minkowski space expansion has been denoted by capital Suppose f 1 , f2

Jf(x) j I).

simply reduces to the

Jacobiant when we use inertial coordinates.

(That is why the volume

From the coordinate expression follows an important group property. are successive transformations z Then Jig (f 2 (f 1 (x))) I (11.107) J f of (x) 2 1

= f 2 (y)
~
v'lg(f (x)) I 1 J f2

= f1

(x)

I~!I

v'lg(f 2 (f 1 (X)))1 v'lg(fl(X1

~ I~~I~I

(f (x)) J (x) 1 f1

We are now in a position where we can generalize the internal transformation associated with a scale transformation. or a vector field. Under a diffeomorphism T f Let
~a

denote a

general field configuration consisting either of a set of scalar fields it will then not only be The compushed forward but also multiplied with an additional scale. bined operation will be denoted by (11.108) As usual d [Tfa 1 (x)

f ; din J -1 (x) [f*~a] (x)


f

and it is given by

denotes the scaling dimension of the field Jf(x)

~a.

In the An. f is an

case of a scale transformation, we know that case of ordinary scale transformations. time transformation. The operation (11.109) Tf

= nn(x) =

Consequently (11.108) reproduces the conventional transformation in the Notice too that when orientation preserving isometry, it just reduces to an ordinary spacesatisfies the important group property

627 To check this we notice first that

II

(cf.

(10.17)).

Then we apply the right hand side of (11.109) to a


~a

field configuration [T f
2

[T f <P 11 (x) 1 a

Jd~~
f2

(x) [ (f 2 ) * (T f <Pa) 1 (x) 1

Jd/n(x)[(f) -1 2 f2

* (Jd/n(x) -1
f1

[(f )*"'N 1 )1(x) 'I'~ 1

Using the group properties (11.107) and (10.21) this reduces to

(If you think this proof is too abstract, you should try to work it out in coordinates!). The essential content of the group property f belongs to a group of transformations on constitutes a representation of this group. Tf (11.109) is that whenever

then the operation

Notice also that transformation

T f acts linearly upon the field ~a' For a given field theory we can now investigate if a space-time f is a symmetry transformation. Using the now famil-

iar techniques we calculate the displaced action:

substituting

= fIx)

this is rearranged as

Using (11.106) this is finally rearranged as (10.110) Sf (Il) [Tf<Pa 1

Notice that apart from a scaling of the Lagrangian the following substitutions have been performed

628

II

(11.111)

a[jl~~
(11.112)

[f*<P ) (f (x) ). _ _ fa. axv


'\tv (x) ...
gj.lV

(f

(x) )

(11.113)

(f (x) )

The first substitution, associated with the potential term U[<Pa.) , is quite harmless. E.g., in four dimensions the potential term U[<p] = ~ <p4 will be invariant under arbitrary space-time transformations! The second one-,-- associated withthekinetic-term -,<d<Pa.1 d<Pa.>' will however cause us trouble. When we square the derivative, we get additional terms which are linear in aj.l<pa.' These terms will prevent the displaced action from being identical to the original action. All we can hope is therefore that they differ by a surface term, i.e. that the displaced Lagrangian is essentially equivalent to the original Lagrangian. The condition for f being a symmetry transformation is consequently axV a[Tf<pJ (11.114) Jf(x)L[(Tf<Pa.) (f(x ; - - - v - (f(X);~v(f(x) '" L[<pa.;aia.;gj.l) ayj.l ax (Here '" denotes that they may differ by a pure divergence) We can now apply this to investigate conformal invariance. For simplicity we restrict ourselves to consideration of scalar fields. Furthermore we need only worry about the kinetic term. If f is an orientation preserving conformal transformation we have previously seen that n Jf(x) = Q (x) From the group property (11.107) it furthermore follows that J -1 (f(x
f

= Q-n(x)

Consequently the substitutions (11.111) and (11.112) reduce to


<p(x) ... Q-d(x)<p(x)

and

Consequently the kinetic term is replaced by Qn{_,g


j.lV

(f(x

dyj.l

ax P ax(J d d d d --- [Q- ap<p + <papQ- ][ Q- a <p + <pa Q- ]} ayV (J (J

629 But since


f

II

is a conformal map we know that P axcr g].lV (f (x)) ax = Q-2(x) g Pcr(x) ay].l ayV

Thus the displaced kinetic term reduces to


_ 1

~ "

nn-2gPcr

[n-d~ ~

"Opo/

n-d][ -d -d <pap" Q acr<p + <pacr Q ]

Inserting the scaling dimension (11.87), this finally boils down to

The first term is precisely what we are after, but the last term is not a manifest pure divergence. We therefore rearrange the displaced kinetic energy density as follows: (11.115) It follows that a conformal transformation with the conformal factor Q2 (x) is a symmetry transformation precisely when (11.116a)
1

I-g

a].I (Fg

a].llnQ)

= 2-n a].I InQ


-2~

a].llnQ

Le.
(11 . 116b)

[] InQ = 2;n (dlnnldlnQ)

The latter condition is far from automatically fulfilled, but it tHrns out to be fulfilled for all conformal transformations in the ordinary flat Minkowski space. Since the conformal group is generated In that case we know from exercise from isometries and the inversion, it suffices to check that the inversion is a symmetry transformation. 10.3.1 that the inversion has the conformal factor

,,(x)
Using that that a].llnQ

= x- 2
= 1

= -2x- 2 x].l

and

Fg

it is now trivial to verify In a curved space

(11.116) is satisfied.

Notice, however, that the above argument

only applies in ordinary "flat" Minkowski space. at least not when you use the standard Lagrangian!

time the theory of a massless scalar field is not conformal invariant,

Worked Exercise 11.9.1 Problem: a) Consider a conformal transformation with the conformal factor g2(x). Show that the kinetic term of a vector field is displaced as follows:
(11.117)

b)

630 II Let nx(x) be the conformal factor associated with the one-parameter group of special conformal transformations xJl + AbJl<xlx> 2 + 2A<blx> + A <blb><xlx>
Show that

(11.118)

and

(aVlnn ) A

IA=O

c)

(Hint: Use exercise 10.3.1). Show that the kinetic term associated with a vector field is not conformal invariant when n~4. (Hint: Show that the correction term cannot be a pure divergence when we apply an infinitesimal special conformal transformation).

Exeraise 11.9.2
IntrOduction: Problem: In this exercise we restrict to the four-dimensional ordinary Minkowski space. Show that minimal coupling is conformal invariant, i.e. show that the gauge invariant kinetic term - a a - !(aa - ieAa)~(a + ieA )~ is conformal invariant.

We can summarize the above findings

(including the results obtained

in exercise 11.9.1 and 11.9.2) in the following way:

In the four-dimensional Minkowski spaae any saale invariant theory based upon saalar fields and veator fields is aonformal invariant. In an arbitrary dimension different from four it is only the saale invariant theories based upon saalar fields, whiah are aonformal invariant.
This explains among other things why we could not construct a conformal energy-momentum tensor in the case of vector fields course in the trivial four-dimensional case). ing one: (except of The reason is the follow-

tion of a conserved symmetric energy-momentum tensor the associated scale current is given by llv T

Suppose a theory is scale invariant and allows the construcllv such that

SV = x
Il

Notice especially that the conservation of the scale current implies llv is traceless. that From the vanishing trace it now follows tri-

vially that the four currents

Sp

given by
2x x
P Il

[SJ'J =

x I x >gllP

)T llv
But these are preCisely

are conserved as well (cf. exercise 11.6.4).

the currents aSSOCiated with the special conformal transformations!

631
not only scale invariant but actually conformal invariant.

II

In a massless scalar theory this is as expected, since such a theory is However, in the case of a massless vector theory we do not expect any conserved currents associated with the special conformal transformations since such a theory is not conformal invariant!
It is instructive to derive the conserved current (11.119) directly from Noether's theorem. Then we have to be careful since the displaced action differ from the initial action by a divergence term (cf. (11.115)): n-2f 1 ( r- 2 ~ ) v_g r- dn x ~ ~a~ v_g~ a Inn

This must consequently be subtracted from the displaced action. The Noether current associated with a one-parameter group of conformal transformations is therefore given by: 0 aL d~A J~ = a"T ~ + ~ dA (f),(x)) v ~ IA=O The displaced field configuration is given by

~A(fA (x)) = Jd~~( fA (x) )[ (fA )*<pHf A(x))


f
rrsiD~

this the Noether current reduces to

(11. 120) which generalizes (11.95). In the case of special conformal transformations, where the one-parameter group is given by xV+AbV<xlx> yV = 1+ 2A<blx>+A 2 <blb><xlx> we get the characteristic vector field: a V = dy = b~<xlx> - 2x~<blx> dA1A=o Furthermore we know from exercise 10.3.1 that Q A Consequently dQ A = _ 2<blx> dAIA=o Inserting this the Noether current (11.120) reduces to - 2x <blxT~v + 22-n <blx> a~[ b
o 2
v

0A

is given by

= (1

+ 2A<blx> + A2 <blb><xlx-1

~2 1
<
x>

As in the case of scale transformations we thus get an additional term, which we must somehow absorb in the canonical energy-momentum tensor. In analogy with (11.101) we therefore consider the current

J~e=dn-21la,a {[b <xix> n- A P \l


Since

- 2x

<blx':lgA~P\l~2}

gA~pv[b <x\j> _ 2x <blx>] = <xlx>[b~gAP _ bAg~P] _ 2<blx>[X~gAP _ XAg~P] v \l which is skew symmetric in ~ and A, it follows that J~ is trivially conserved. Performing the differentiations involved we can now rearrange the expression for J~ as follows:

632

II

J~ = e

[b <xix> - 2x <blx>le~v+ v v

~<blx>2a~[~12 1
2 <b x>

2(n-1}

3(n-2)ap[(b~xP

x~bP)~21

By adding these two conserved currents we almost cancel the unwanted term:

j~ = J~ + J~
e

= [b \! <xix>

- 2x

<blx>lT~v+ 3(n-2)a [(b~xP - x~bP)~21

2Tr1=TT

Furthermore the additional term constitutes itself a current which is trivially conserved. We can therefore throw it away, whereby we finally recover (11.119)!

SOLUTIONS OF WORKED EXERCISES


No. 11.4.1 (a) The determinant is given by a.

I~I

=E:

a.l . a.n axl

.&2

Using Leibniz' rule we therefore get

d~ 1;51
If we put
d

= a.l ... a.n

M:~:l):~;

A=O

this reduces to

dA!A=O

I a a.1

(b)

We start with the identity

h-1L
aya. axil

\!

a.

o\!

Il A we get v

Differentiating this with respect to

~axV) a a.
dA\aya. For A=O

ax ~Ai:) = 0
+

aya. dA\aill

this reduces to

i.e.

~.axV)
dA!A=o\ayll Remark: When trix:

~2L)
+ dAIA=o'axJl

= 0 0 the Jacobi-matrix is close to the identity ma-

A is close to

633

II

(~]
If we work out the determinant and drop all higher order terms the off-diagonal contributions automatically cancel, and we get

I~I_ 1 n . lax S - 1+ 1++ n + hlgher order terms


and this is where the trace in (a) comes from! No. 11.4.2

.....La (I-::gfi
I_g)l

)l)

(Fg)L [.....La ,-- v


v-g

av L]

In the first term we can use the covariant equations of motion (8.35). In the second term we can interchange a and a . The first two terms can then be reduced to -a L + -~- a g Q' from which the result follows. V "ga.S v a." ______ _ No. 11.5.1 (a) Let f be a conformal map, and put
~=f*~

aL

)l

o
According to th.15,sec.10.5 we have,

f*[*~l = Qn-2(x)*f*[d~1
where we have used that n=2. We then get,

= *~,

<~I~>Q
(b) Let f

fQ*~A~ = fQf*[*d~A~l
ff(Q)*d~Ad~

= <d~ld~>f(Q)'
= *dA' ,

be a conformal map, and put f*[*dAl

A'=f*A. From th.15,sec.10.7 we now get,

= r,P-4(x)*f*[dAl
fQ*d~'~dA'
ff(Q)*dAAdA

where we have used

n=4 We then get, <dA' IdA'>Q

= fQf*[*dAAdAl = <dAldA>f(Q)'

No. 1.1.5.3 (a) We know that

(NE! Observe that we are using indices). Consequently we get


+2

covariant indices, not the usual contravariant


+2 +l
v

epa

=ypea

+2

- Yaep = ypesAa - Yaea.Ap

~l~a

)~IAS_(x AS+b )~IAa. ( xa.Aa.+b p p S a S a a a. p

634
(b)
By definition we have

II

and po = T(e po ") and (b) follows immediately from (a).


J

-+

No. 11.6.1

The Lagrangian density for a vectorfield depends only on strength F~v = a~~-avA~ We therefore get

a~~

through the field-

_ _ a_L_ _ _ dL aF"" _ _ aL p 0_ p 0 aL aL dL ala A ) - aF ~ - aF [o~0a. 0a.cV = ~-~=2~ ~ a. po ~ a. pa ~a. a.~ ~a.


No. 11. T. 1

(a)

To prove a) we observe that

a~lnDetM(x)/x=xo

d DetM(x) = MetM(xo)/X=X o

a [DetM(x) ] ~ Det!;J(xo ) /x=x o

=-1 = a~Det[M (xo)M(x)]

But M-l(x )M(x) is a matrix function which reduces to I at ing the ar~ument from exercise 11.4.1 we now immediately get

x=xo' Repeat-

(b)

From (6.47) we get ra. ag _, a.p ~ ~a. - ~g ax~ and


=-1 G

Introducing the matrices as ra. = 'TrG-la G


~a. ~ ~

we can reformulate this

which by a) can be converted to

(c)

Let the matrix we get from a) dA

ga.~

depend in an arbitrary fashion on the parameter " .A. ln / g/ = _1_.A. dA dA

A . Then

~ga.~dga.~ = ~

;-rgr

;-rgr = _l_.&[ dga.~ irgT dga.~ dA


~ga.~/fiT and
oa. y we immediately get:

From this we conclude that the symmetric matrices must be identical. (d) Differentiating the identity
ga.~~y

o =~
agpa

a a.~

~y

+ga.jlo~oY p a

a~ =~ g +ga.poY ag ~y a

pa

and multiplying with

g~Y this is converted to


~

o =~
gpa

a a.~

+ ga.Pg

D
______ --'

635

II

(b)

aL

a~V

dx].l dxv = ;m -dT dT


')

1-g

6 4 (x-X(T))dT -

~g

].IV

from which we conclude aL- + g].Iv L 2ag].lV

= mfdX].I dT

1 dxV --=-o4(x-X(T))dT dT ;_g

n U

No. 11.9.1

for a conformal transformation


(*)

(11.112)

reduces to

a <I> (x) ].I a.

axP{Q-tX)a[f*<Pa.](f(X)) + [f*<I> ](f(x)) aQ-dl. ay].l axP a. axP f

In the case of a vector field we have aa [f,.Aa.lCf(x)) = 2:..A (x) aya. a Therefore (*) leads to the following displacement rule: a a a A ~ axP{n'll [ax A 1 + ax A a n- d} ].I a. ay].l P aya. a aya. a p Using that and this can be rearranged as P a A ... ax axa{n-da A + A a Q-d _ Q-dA ax< ~} ].I a. ay].l aya. Pa a P KayA axPaxa

A skew-symmetrization in (].Ia.), which on the right hand side corresponds to a skew-symmetrization in (pa), then leads to the following displacement of the field strength: P a F ax { n -~ ].Ia. a ... x --- ---p + (Aaapn- d- Apaan- d ) } ay].l aya. pa The displaced kinetic energy term therefore looks as follows

636

II
T

_ !rf(x)gll"(f(x))gUS(f(x))dX ~ ax ax x 4 ayllayUay"ayS

a A

x{Q-~
since f

pa

+ A a Q-d _ A a Q-d}{Q-~, + A a,Q-d _ A,a Q-d} a p pa AT TA AT


p

is a conformal map we know that

gil" (f(x) )gUS(f(x)) ax ax ax ax dyll ayUay"a?


Furthermore d = (n-4)/2 (b) According to exercise
10.3.1

gPA (x)gaT (x)


(11.117).

and the displaced kinetic term now reduces to

QA(X)
Thus

(1+2A<blX>+A z<blb><xlx 2Abv + 2A 2 <blb>x" (1+2A<blX>+A 2 <blb><xlx

we have 1

from which (e)

follows immediately.

If the kinetic term is conformal invariant the displaced kinetic term must differ at most by a divergence term. In terms of inertial coordinates we must therefore have

(**)
I f we let fA

represent the group of special conformal transformations we get from

(b)

that

Thus we see that FIl"A b must be a divergence (for an arbitrary choise of b"), but that is impossiblella~cording to the general argument g~ven below (11.97).0

637

INDEX OF SUBJECTS,
Abelian gauge theory 58 Abelian Higgs' model 445-50 Abraham's theorem 27,589,591-92 Abrikosov vortex 78 Action of a point particle 33 86 Action of a relativistic field Action of a relativistic particle 294 Action of a relativistic string 443 Adapted coordinate system 396 Affine parameter on a geodesic 535 d'Alembertian 20,364 Ampere's law 5 Angular momentum 28 Angular momentum of a charged particle in a monopole field 489-90 Angular momentum of an electromagnetic field created by a charge-monopole pair 487 Angular momentum operators for a charged particle in a 492-93,495 monopole field Angular momentum of a relativistic string 444 Approximative ground state 142 Arc-length 291 Atlas 256 Backlund transformation 150 451 Bag Basic form 333-336 BCS-theory of superconductor 73 Bianchi's permutability theorem 151 Bion 148,210,211 Bloch wave 235 140 Bogomolny decomposition Bogomolny decomposition in a (1+1)-dimensional scalar field theory 141 Bogomolny decomposition in a ferromagnet 573 Bogomolny decompos~tion in the exceptional ~4-model 580 Bohm-Aharonov effect49-53,78-79,502 Bohr-Sommerfeld quantization rule 200,203 Bosonic spectrum of a charged particle in a monopole field Boundary Bound states (in the sine-Gordon model) 148-49 Bracket 595 Breather 147,210-11 Brouwer degree 564 Brouwer's lemma 564 566 Brouwer's theorem Canonical energy-momentum tensor 89 Canonical frame vectors in the tangent space 270 Canonical frame vectors in the dual space 303 Canonical identification of tangent vectors and covectors 305-06 Canonical identification of mixed tensors 316-17 Canonical quantization of point particles 68-72 Canonical quantization of relativistic fields 93 Canonical system 65 Cartan's lemma 399 Cauchy problem 109-13 Cauchy-Riemann's equations 365,555 Caustic 173 Characteristic line on a null cone 542 Characteristic vector field associated wtih a one-parameter family of diffeomorphisms 594,601 Charge conservation 7,119-21,588-89 Charged field 116-19 Charge rotation 474-75 Christoffel field 295,373 Circularly polarized light 106 Classical approximation of a propagator 44 Classical vacuum 130 Closed domain 400 Closed form 341 Co-closed ~~f8r.. 364 Co-differential 360-62,421-22 Co-differential of a weak form 459 Co-exact form 364 Coherence length 433-34 Coleman's formula for a quadratic path integral 195 Colour field 451 Commutation rules for wedge products 331-32 Commutative diagram 256 Compact subset of an Euclidean 263,397 " space Complex' coordinate system 560 Complex Klein-Gordon field 113-16 Complex manifold 560 Complex valued differential form 365 Composite particles in non-linear field theories 160-64 Confinement 450-51 Conformal compactification of an Euclidean space 541 Conformal compactification of a pseudo-cartesian space 546 Conformal energy-momentum tensor 624

638 Conformal group 547-51 Conformal invariance in field theories 625-32 Conformal invariance of Maxwell's equations 558 Conformal invariance of the self-duality equation 555 Conformal map 531-33 Conformal map on the sphere 559-61 Conjugate momentum of a 65 point particle Conjugate momentum of a relati vistic field 90 Conservation laws 587-92 Conservation laws associated with conformal symmetry 613,631-32 Conservation law, topological 134 Conservation of angular momentum in field theories 609-11 Conservation of angular momentum in quantum mechanics 597-600 Conservation of electric charge 588-89 Conservation of energy and momentum 589-92,614-18 Conservation of momentum 608 in field theories Conservation of momentum in quantum mechanics 596-97 Constraint equation (in gauge theory) 111 Continuity equation 7,21,381 Contraction of differential forms 356,359 Contraction of tensors 310,313,314 Contravariant components of a tangent vector 272,306 Contravariant quantity 274 Cooper current 75 Cooper pair 73 Coordinate line 270 Coordinate system 254 Correspondence principle in 203 quantum mECc):1anics Cotensor " 308 Coulomb field 14,456-57 Covariant components of a " tangent vector 306 Covariant quantity 274 Covector 300-305 Critical value 515 Cross-product 3,351-52,366 Cross section for scattering in a monopole field 506-09 Curl 5,366 Cyclic coordinates 307,383 Cylindrical coordinates 290 Cylinder (as a product manifold)265 Decomposition of differential forms 332 Decomposition of cotensors 311 Degree of a smooth map 563 o-(delta-) function in Euclidean space 18 o-(delta-) function on a manifold 455 Derrick's scaling argument for scalar field theories 158-59 Derrick'S scaling argument for the exceptional $4-theory 581-82 Determinantal relation 196,197-98 Determinant of a differential operator 194 Determinant of a metric 279,348 Diffeomorphism 515 Differentiable manifold 258 Differentiable structure 257,258 Differential 304 Differential form 336 Dilatation 533 Dilute gas approximation 233 Dirac formalism for monopoles 493 Dirac string 18,383-84,476-81 Dirac's action for a system including monopoles 482 Dirac's commutation rule for the angular momentum operators 493 480 Dirac's lemma Dirac's quantization rule for the magnetic charge 487 482 Dirac's theorem Dirac's veto 478 Dispersion relation 84,90 Distribution 458 Divergence 5,366 Double dualisation 354-55 220-222 Double square well Dual map 352-55, 551-54 Dual map in Minkowski space 374 Dual map in R3 367 Dual map of a weak form 460 Dual vector space 301 Dyon 470 Eckhardt potential 162 Edge condition for a relativistic string 444 Einstein-deBroglie rule 48 Electric current in a gauge theory 120 Electric flux 429,430,432 Embedding 518 Energy current 24,88 Energy density 24,87 Energy-momentum tensor 22-28,320-21 Energy-momentum tensor, canonical 88-89 Energy-momentum tensor, conformal 62). Energy-momentum tensor, metric 615 Energy-momentum tensor of a point particle 23 Energy-momentum tensor of electromagnetic field 26,99-100,320

639 Energy-momentum tensor of massive vector field 108 Energy-momentum tensor, true 89,611-13 Energy split in a double well 232,245 Equation of constraint 111 Equation of continuity 7,21,381 Euclidean action 213 Euclidean group of motions 539 280 Euclidean metric Euclidean propagator 212,214-16 Euclidean vacuum 216 Euler-Lagrange equation for a point particle 33,35 Euler-Lagrange equation for a relativistic field 87,91 Euler-Lagrange equation on covariant form 425 Even form 332 Exact form 341 Exceptional ~4-model 578-82 Exterior derivative 336-42 Exterior derivative in Minkowski space 375 Exterior derivative in R3 371 Exterior derivative of a differential form 337 Exterior derivative of a function 302 Exterior derivative of a weak form 459 Extrinsic coordinates of a tangent vector 271 Faraday's law 5 Fermionic spectrum of a charged particle in a monopole field 497 Ferro-magnet 570 Feynman's propagator 42,167-7 2 Feynman's propagator in canonical formalism 71 Feynman's propagator in momentum space 171-72 Feynman's propagator of a free particle 46 Feynman's propagator of a harmonic oscillator 174,181 Feynman's propagator of a timedependent oscillator 188-93 Feynman's formula for the path-integral 182 Feynman's principle of the democratic equality of all histories 43,93 Feynman-Soriau's propagator for the harmonic oscillator181,187 Fictitious forces 296,299 Field quantum in a free field theory 89-90 Field quantum in a non-linear field theory 94,130-31 Fine-structure constant 621 Flux integral 416 Flux-quantization in superconductors 77 -78,502-03 Four-velocity 293 Free field theories 89-90 Free particle wave function 49 Functional derivative 91 Galilei's principle 298 Gauge covariant derivative 60,69 Gauge group (in electromagnetism) 57 Gauge phase factor 50 Gauge potential 7-11,381 Gauge scalar 59 Gauge theory, basic ingredients 61 Gauge theory of electrically charged fields 116-18 Gauge transformation 10,498-99 Gauge transformation of the action 38 Gauge transformation of the first and second kind 118 Gauge transformation of the propagator 56 Gauge transformation of the Schrodinger wave function 55-57 Gauge transformation on covariant form 307,381 58 Gauge vector Gaussian approximation of the path integral 238 Gauss' law in electromagnetism 5 Gauss' theorem in 3-space 6 Gauss' theorem on a manifold 417 Geodesic 296,535-38 Geodesic equation 296,535 Ginzburg-Landau parameter 436 Ginzburg-Landau theory for superconductivity 432 Gradient 5,366 Gravitational forces 299-300 Green's identities 423 Ground state in quantum mechanics 219-23 Ground state for a double well 232 Ground state for a harmonic oscillator 178 &riound state for a non-pertubati ve sector 138-43 Ground state for a periodic potential 235 Group property for the propagator 46 Hadron 445,450 Hamiltonian 65 Hamiltonian density 87 Hamiltonian formalism for a point particle 65-68 Hamiltonian formalism for a relati vistic field 90-9 4

0<

640

Hamilton's equations for a 66 point particle Hamilton's equations for a relativistic field 90,92 Hamilton's principal function 200 364,422 Harmonic form Heat operator 214 Heaviside function 18 Heisenberg's commutation rules 'for a point particle 71 Heisenberg's commutation rules for a relativistic field 93 Heisenberg's model for a ferro570 magnet Heisenberg's uncertainty principle 176 Helicity of a photon 104 Hermite polynomials 179 Hermitian operator 70 Higg's field 445 Hilbert product of differential 420 forms Hodge duality 354 Holomorphic function 365,555-56,560 560 Holomorphic map Holomorphic map on a sphere 561 252 Homeomorphism Honey-comb structure 334-35 Immersion 517 Impact parameter 490 266 Implicit function theorem 282 Induced metric 285 Inertial coordinates Inertial frame of reference 285 Infinitisemal generator for a one-parameter group of unitary transformations 593-95 Infinitisemal propagator 64 Instanton 216 Instanton in quantum mechanics217-19 Instanton, relationship with tunnel effect 223-25 Instanton gas 232 Integral of a differential form 403- 1'1 Integral of a scalar field 413 Integral of a simple form 418-20 Interior of a regular domain 397 Internal symmetry 116,605 Intrinsic coordinates of a tangent vector 272 Intrinsic spin of a charged particle in a monopole field 495 Inversion 533,540 Isometry 530 Isometry group on a manifold 538 Jacobiant Jacobi matrix 344-45 252

k-form Kink Kink sector Klein-Gordon equation Klein-Gordon field Kronecker delta

327 137 138 95 95-98,427-28 315

Lagrangian density 86 Lagrangian density for the electromagnetic field 100 Lagrangian density for the KleinGordon field 95 Lagrangian formalism and the exterior calculus 424-29 Lagrangian formalism for a monopole 482-86 Lagrangian formalism for a point particle 32-38 Lagrangian formalism for a relativistic field 86-90 Lagrangian formalism for a relativistic particle 295 Lagrangian formalism for a relativistic string 442-44 Lagrange multiplier 571 Laplacian in 3-space 5 Laplacian on a manifold 363,422 Laplace's equation in 2-space 556-57 Legendre transformation 202 Leibniz' rule for differential forms 340 Leibniz' rule in R3 372 Levi-Civita form 350 Levi-Civita symbol 348,350 Light cone 281 Light cone coordinates 150 Linearized equations of fluctuations 153 Linearly polarized light 105 line element 293 Line-integral6,415 Liouville's theorem 561 Locally finite family of functions 405 Lorentz force 4 Lorentz force, generalized 474 Lorentz force in quantum mechanics 50,53-55 Lorentz group 287 Lorentz matrix 287 Lorentz transformation 287 Lorenz gauge 10,112

Ma);he~~c charge 469-70,472 Magnetlc current 471-72 Ma,gnEltic flux 11-14,429,432 Magnetic monopole 13-14,318-20,381-84,469-503 Magnetic string in the abelian Higgs' model 448-49 Magnetic string in a superconductor 78,437 -41 Manifold 258 Manifold defined by an equation of constraint 268

641 Massive vector fieldl07-8,112-13,428 Massless field theories 622 Maximal atlas 257 Maxwell's equations 5,376,381,428 Maxwell's equations including monopoles 472 22,428 Maxwell field Mehler's formula 179 Meissner effect 72,435 Metric 276-84 Metric components 277 Metric energy-momentum tensor 615 Metric on a sphere 281-83 Metric on Minkowski space 288-89 Metric, reciprocal components 278 Minimal coupling 69,117,621 Minkowski metric 280 Minkowski space 284-90 Mixed tensor 312 Mobius strip 346 Momentum density 24 Monopole 13-14,318-20,381-84,469-503 Monopole confinement 450 Monopole in the abelian Higgs' model 449-50 Multilinear map 308 Multi-soliton solution in the sine-Gordon model 153 Nambu string 442-45,452-54 Nielsen-Olesen vortex 448,453 Node-theorem for eigenfunctions in the discrete spectrum 157 Noether's theorem 602-06 Noether's theorem for internal symmetries 605 Noether's theorem for scalar fields 604 Noether's theorem for vector 613 fields Non-linear field theories 9 4 ,127-31 Non-pertubative sector 132 Null-geodesic 537 281 Null-vector Odd form 332 One-parameter family of diffeomorphisms 601 One-parameter group of isometries 593-94 One-parameter group of unitary transformations 593 One-point compactification of the Euclidean space 541 Order parameter in a ferromagnet 570 Order parameter in a superconductor 432 Orient ability of a boundary 401 Orient ability of a manifold 345 Orientability of a regular domain 401 Ortochronous Lorentz transformation 288

Paracompact 406 Partial integration 413 Partition of unity 406 Path-cum-trace integral 199 Path integral 42,181-88 Path integrals and determinants 194 Pauli's exclusion principle 73 Penetration length 435 Phase ambiguity in the path integral 180-81 ~'-(phi-to-the-fourth) model 127 Plane wave 135 Planck mass 621 Poincare group 539 Poincare's lemma for the co-differential 360 Poincare's lemma for the exterior derivative 338 Poincare's lemma for weak forms 460 Poincare'S lemma in R3 372 Poincare transformation 285,286,538-39,607 Poisson bracket for point particles 67 Poisson bracket for relativistic fields 92 Polarization vector for the electromagnetic field 100 Primitively harmonic form 365,422 Probability amplitude 39-40 Product manifold 264-65 Proper Lorentz transformation 288 Proper time 293 Pseudo-orthogonal group 546 Pseudo-orthogonal matrix 546 Pseudo-Riemannian manifold 280 Pseudo-tensor 350,355 Pseudo-vector 3 Pull-back of cotensors 522 Pull-back of covectors 521 Pull-back of differential forms 523 Pull-back of equivalent tensors 534 Pull-back of integralS 528 Pull-back of metrics 522-23 Quadratic Lagrangian 45,194-198 Quantization of the magnetic charge 500-02 Quark 450 Quark confinement 451 Quark potential 452 Quasi-static solution in a non143 linear field theory Rainbow angle Rapidity Rank of a cotensor Rank of a mixed tensor Regge trajectory Regular domain 509 609 308 312 445 397

642 Regular domain Regularity of a smooth map Regular value Relativistic string Restriction of a differential form Riemannian manifold Riemann sphere 460 515 515 442 408 280 560 Space slice 431 Special conformal transformation 533 Special relativity, basic assumptions286 Spectrum of the angular momentum in the Dirac formalism 497 Spectrum of the angular momentum in the Schwinger formalism 496-97 Spectrum of the Eckhardt potential 163 Spectrum of the Schrodinger operator 157 Sphere (as a manifold) 259-64,281-83 Spherical symmetry in Minkowski space 289-90 Spinning string 444-45 Spin of a photon 103 Spin of a vector field 610-11 Spin of a vector particle 598 Spin wave in a ferromagnet 573 Standard coordinates on a manifold defined by an e~uation of constraint 267 Standard coordinates on a sphere 261 Standard Lorentz transformation 288 Static solution in a non-linear field theory 138-40 Stationary phase approximation 205 Stationary state 71 Stokes' theorem in 3-space 6 Stokes' theorem on a manifold 411 Stratification 334 Submanifold 395 Submersion 518 Superconductivity 72-79,432-441 Symmetry transformation in field theory 601 Symmetry transformation in quantum mechanics 595-96 -Tangent space Tangent vector Tensor Tensor field Tensor product Tidal force Time-dependent oscillator Time-evolution operator Time-like curve Time-like geodesic Time-like vector Topological charge Topological conservation law Topological current Torus (as a product manifold) Transition function Transport of tangent vectors Transport of tensors Travelling wave True energy-momentum tensor Tube lemma Unitary group Unitary matrix 272 269-76 312 314 310,314 299 188 169 291 537 281 133-35 134 134 265 255 512-15
~24-25

Sard's theorem 519 Scalar field 274-75 Scalar product between differential forms 357-59 Scale associated with a diffeomorphism 625 Scale current 624
Scale invariance in a relati-

vistic field theory 618-24 Scale invariance of minimal coupling 620-21 Scale transformation in a relativistic field theory 158-59,581-82,618 Scaling dimension of a scalar field 619 Scaling dimension of a vector field 620 Schrodinger e~uation 32 Schrodinger e~uation in the canonical formalism 68 Schrodinger e~uation in the path-integral formalism 62-64 Schrodinge~ wave function 41,55-61 Schwinger formalism for monopoles 492 Self-dual form 355 Self-duality e~uation 554-58 Self-duality e~uation in a ferromagnet 574 Self-duality e~uation in the excepcional ~'-model 579 Simple form 342 Simple manifold 258 Sine-Gordon e~uation 128 Sine-Gordon model 128-29 Singular differential form 454 Singular gauge transformation 18,384 Skew-symmetrization 328-29 Slit experiment 39-40,49-55 Smooth map between manifolds 512 Smooth map in an Euclidean space 252 Solenoid 15-19,52,78-79,306-07 Solitary wave 135 Soliton 144 Soliton-soliton scattering 145-47 Soliton versus instanton 217 Soriau's formula for the propagator of a harmonic oscillator 181 Space like geodesic 537 281 Space like vector

136 89,611-13 587 58 58

643

Unitary transformation Unit tensor field

498 315

130 Vacuum, classical 132 Vacuum sector 440 Vacuum texture 275 Vector field Vector particle 59 4 -95 270 Velocity vector Virial theorem in (1+1)-dimensional space-time 159 Virial theorem in the exceptional <j>4-model 582 625 Volume expansion Volume form 343-44,351 Volume form on a sphere 567 414 Volume of a regular domain Vortex-string in a superconductor 78,441

Wave equation 557 Weak coupling approximation 199 Weak k-form 458 Wedge product in 3-space 6 Wedge product of differential forms 330 Weierstrass' theorem 561 Wick rotation 212 Winding number 564 WKB-approximation 204-209 World-line 284 World-sheet 442 Wronskian 241-42 Yukawa-potential Zero-mode in a non-linear field theory Zero mode in quantum mechanics 97 158 238

You might also like